You are on page 1of 1852

EAU Guidelines on

Management of
Non-Neurogenic
Male Lower Urinary
Tract Symptoms
(LUTS), incl.
Benign Prostatic
Obstruction (BPO)
S. Gravas (Chair), J.N. Cornu, M. Gacci, C. Gratzke,
T.R.W. Herrmann, C. Mamoulakis, M. Rieken,
M.J. Speakman, K.A.O. Tikkinen
Guidelines Associates: M. Karavitakis, I. Kyriazis, S. Malde,
V. Sakalis, R. Umbach

© European Association of Urology 2021


TABLE OF CONTENTS PAGE
1. INTRODUCTION 4
1.1 Aim and objectives 4
1.2 Panel composition 4
1.3 Available publications 4
1.4 Publication history 4

2. METHODS 4
2.1 Introduction 4
2.2 Review 5
2.3 Patients to whom the guidelines apply 5

3. EPIDEMIOLOGY, AETIOLOGY AND PATHOPHYSIOLOGY 5

4. DIAGNOSTIC EVALUATION 6
4.1 Medical history 6
4.2 Symptom score questionnaires 7
4.2.1 The International Prostate Symptom Score (IPSS) 7
4.2.2 The International Consultation on Incontinence Questionnaire (ICIQ-MLUTS) 7
4.2.3 Danish Prostate Symptom Score (DAN-PSS) 7
4.3 Frequency volume charts and bladder diaries 7
4.4 Physical examination and digital-rectal examination 8
4.4.1 Digital-rectal examination and prostate size evaluation 8
4.5 Urinalysis 8
4.6 Prostate-specific antigen (PSA) 9
4.6.1 PSA and the prediction of prostatic volume 9
4.6.2 PSA and the probability of PCa 9
4.6.3 PSA and the prediction of BPO-related outcomes 9
4.7 Renal function measurement 9
4.8 Post-void residual urine 10
4.9 Uroflowmetry 10
4.10 Imaging 11
4.10.1 Upper urinary tract 11
4.10.2 Prostate 11
4.10.2.1 Prostate size and shape 11
4.10.3 Voiding cysto-urethrogram 11
4.11 Urethrocystoscopy 11
4.12 Urodynamics 12
4.12.1 Diagnosing bladder outlet obstruction 12
4.12.2 Videourodynamics 12
4.13 Non-invasive tests in diagnosing bladder outlet obstruction in men with LUTS 13
4.13.1 Prostatic configuration/intravesical prostatic protrusion (IPP) 13
4.13.2 Bladder/detrusor wall thickness and ultrasound-estimated bladder weight 13
4.13.3 Non-invasive pressure-flow testing 14
4.13.4 The diagnostic performance of non-invasive tests in diagnosing bladder
outlet obstruction in men with LUTS compared with pressure-flow studies 14

5. DISEASE MANAGEMENT 15
5.1 Conservative treatment 15
5.1.1 Watchful waiting (WW) 16
5.1.2 Behavioural and dietary modifications 16
5.1.3 Practical considerations 16
5.2 Pharmacological treatment 17
5.2.1 α1-Adrenoceptor antagonists (α1-blockers) 17
5.2.2 5α-reductase inhibitors 18
5.2.3 Muscarinic receptor antagonists 19
5.2.4 Beta-3 agonist 20
5.2.5 Phosphodiesterase 5 inhibitors 22
5.2.6 Plant extracts - phytotherapy 23

2 MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021
5.2.7 Combination therapies 25
5.2.7.1 α1-blockers + 5α-reductase inhibitors 25
5.2.7.2 α1-blockers + muscarinic receptor antagonists 26
5.3 Surgical treatment 28
5.3.1 Resection of the prostate 28
5.3.1.1 Monopolar and bipolar transurethral resection of the prostate 28
5.3.1.2 Holmium laser resection of the prostate 30
5.3.1.3 Thulium:yttrium-aluminium-garnet laser (Tm:YAG) vaporesection
of the prostate Mechanism of action: 30
5.3.1.4 Transurethral incision of the prostate 31
5.3.2 Enucleation of the prostate 31
5.3.2.1 Open prostatectomy 31
5.3.2.2 Bipolar transurethral enucleation of the prostate (B-TUEP) 32
5.3.2.3 Holmium laser enucleation of the prostate 33
5.3.2.4 Thulium:yttrium-aluminium-garnet laser (Tm:YAG) enucleation
of the prostate Mechanism of action: 34
5.3.2.5 Diode laser enucleation of the prostate 35
5.3.2.6 Enucleation techniques under investigation 36
5.3.2.6.1 Minimal invasive simple prostatectomy 36
5.3.2.6.2 532 nm (‘Greenlight’) laser enucleation of the prostate 36
5.3.3 Vaporisation of the prostate 37
5.3.3.1 Bipolar transurethral vaporisation of the prostate 37
5.3.3.2 532 nm (‘Greenlight’) laser vaporisation of the prostate 38
5.3.3.3 Vaporisation techniques under investigation 39
5.3.3.3.1 Diode laser vaporisation of the prostate 39
5.3.4 Alternative ablative techniques 40
5.3.4.1 Aquablation – image guided robotic waterjet ablation: AquaBeam 40
5.3.4.2 Prostatic artery embolisation 41
5.3.4.3 Alternative ablative techniques under investigation 42
5.3.4.3.1 Convective water vapour energy (WAVE) ablation:
The Rezum system 42
5.3.5 Alternative ablative techniques under investigation 42
5.3.5.1 Prostatic urethral lift 42
5.3.5.2 Intra-prostatic injections 43
5.3.5.3 Non-ablative techniques under investigation 44
5.3.5.3.1 (i)TIND 44
5.4 Patient selection 44
5.5 Management of Nocturia in men with lower urinary tract symptoms 47
5.5.1 Diagnostic assessment 47
5.5.2 Medical conditions and sleep disorders Shared Care Pathway 47
5.5.3 Treatment for Nocturia 49
5.5.3.1 Antidiuretic therapy 49
5.5.3.2 Medications to treat LUTD 50
5.5.3.3 Other medications 50

6. FOLLOW-UP 51
6.1 Watchful waiting (behavioural) 51
6.2 Medical treatment 51
6.3 Surgical treatment 52

7. REFERENCES 52

8. CONFLICT OF INTEREST 81

9. CITATION INFORMATION 81

MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021 3
1. INTRODUCTION
1.1 Aim and objectives
Lower urinary tract symptoms (LUTS) are a common complaint in adult men with a major impact on quality
of life (QoL), and have a substantial economic burden. The present Guidelines offer practical evidence-based
guidance on the assessment and treatment of men aged 40 years or older with various non-neurogenic
benign forms of LUTS. The understanding of the LUT as a functional unit, and the multifactorial aetiology of
associated symptoms, means that LUTS now constitute the main focus, rather than the former emphasis
on Benign Prostatic Hyperplasia (BPH). The term BPH is now regarded as inappropriate as it is Benign
Prostatic Obstruction (BPO) that is treated if the obstruction is a significant cause of a man’s LUTS. It must
be emphasised that clinical guidelines present the best evidence available to the experts. However, following
guideline recommendations will not necessarily result in the best outcome. Guidelines can never replace
clinical expertise when making treatment decisions for individual patients, but rather help to focus decisions
- also taking personal values and preferences/individual circumstances of patients into account. Guidelines are
not mandates and do not purport to be a legal standard of care.

1.2 Panel composition


The EAU Non-neurogenic Male LUTS Guidelines Panel consists of an international group of experts with
urological and clinical epidemiological backgrounds. All experts involved in the production of this document
have submitted potential conflict of interest statements which can be viewed on the EAU website Uroweb:
http://uroweb.org/guideline/treatment-of-non-neurogenic-male-luts/.

1.3 Available publications


A quick reference document, the Pocket Guidelines, is available in print and as an app for iOS and Android
devices. These are abridged versions which may require consultation together with the full text version. All
documents are accessible through the EAU website Uroweb: http://www.uroweb.org/guideline/treatment-of-
non-neurogenic-male-luts/.

1.4 Publication history


The Non-neurogenic Male LUTS Guidelines was first published in 2000. The 2021 document presents a limited
update of the 2020 publication; the next full update of the Non-neurogenic Male LUTS Guidelines will be
presented in 2022.

2. METHODS
2.1 Introduction
For the 2021 Management of Non-Neurogenic Male LUTS Guidelines, a detailed review and restructuring of
section 5.3 Surgical Treatment has been undertaken. Section 5.3 has been restructured to reflect surgical
approach rather than specific technologies and is now divided into five sections: resection; enucleation;
vaporisation; alternative ablative techniques; and non-ablative techniques. The literature cut-off date for section
5.3 is April 2019. In addition, a broad and comprehensive literature search related to Serenoa repens in section
5.2.5 Plant extracts – phytotherapy was performed covering the timeframe between the search cut-off date of
the EU monograph on S. repens [1] and April 2020. A detailed search strategy is available online: http://www.
uroweb.org/guideline/treatment-of-non-neurogenic-male-luts/supplementary-material.

For each recommendation within the guidelines there is an accompanying online strength rating form, the
bases of which is a modified GRADE methodology [2, 3]. Each strength rating form addresses a number of key
elements namely:
1. the overall quality of the evidence which exists for the recommendation, references used in
this text are graded according to a classification system modified from the Oxford Centre for
Evidence-Based Medicine Levels of Evidence [4];
2. the magnitude of the effect (individual or combined effects);
3. the certainty of the results (precision, consistency, heterogeneity and other statistical or
study related factors);
4. the balance between desirable and undesirable outcomes;
5. the impact of patient values and preferences on the intervention;
6. the certainty of those patient values and preferences.

4 MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021
These key elements are the basis which panels use to define the strength rating of each recommendation.
The strength of each recommendation is represented by the words ‘strong’ or ‘weak’ [5]. The strength of each
recommendation is determined by the balance between desirable and undesirable consequences of alternative
management strategies, the quality of the evidence (including certainty of estimates), and nature and variability
of patient values and preferences.
Additional information can be found in the general Methodology section of this print, and online at
the EAU website; http://www.uroweb.org/guideline/. A list of associations endorsing the EAU Guidelines can
also be viewed online at the above address.

2.2 Review
The Non-Neurogenic Male LUTS Guidelines were peer reviewed prior to publication in 2016.

2.3 Patients to whom the guidelines apply


Recommendations apply to men aged 40 years or older who seek professional help for LUTS in various non-
neurogenic and non-malignant conditions such as BPO, detrusor overactivity/overactive bladder (OAB), or
nocturnal polyuria. Men with other associated factors relevant to LUT disease (e.g. concomitant neurological
diseases, young age, prior LUT disease or surgery) usually require a more extensive work-up, which is not
covered in these Guidelines, but may include several tests mentioned in the following sections. EAU Guidelines
on Neuro-Urology, Urinary Incontinence, Urological Infections, Urolithiasis, or malignant diseases of the LUT
have been developed by other EAU Guidelines Panels and are available online: www.uroweb.org/guidelines/.

3. EPIDEMIOLOGY, AETIOLOGY AND


PATHOPHYSIOLOGY
Lower urinary tract symptoms can be divided into storage, voiding and post-micturition symptoms [6], they
are prevalent, cause bother and impair QoL [7-10]. An increasing awareness of LUTS and storage symptoms
in particular, is warranted to discuss management options that could increase QoL [11]. Lower urinary tract
symptoms are strongly associated with ageing [7, 8], associated costs and burden are therefore likely to
increase with future demographic changes [8, 12]. Lower urinary tract symptoms are also associated with a
number of modifiable risk factors, suggesting potential targets for prevention (e.g. metabolic syndrome) [13]. In
addition, men with moderate-to-severe LUTS may have an increased risk of major adverse cardiac events [14].
Most elderly men have at least one LUTS [8]; however, symptoms are often mild or not very
bothersome [10, 11, 15]. Lower urinary tract symptoms can progress dynamically: for some individuals LUTS
persist and progress over long time periods, and for others they remit [8]. Lower urinary tract symptoms
have traditionally been related to bladder outlet obstruction (BOO), most frequently when histological BPH
progresses through benign prostatic enlargement (BPE) to BPO [6, 9]. However, increasing numbers of studies
have shown that LUTS are often unrelated to the prostate [8, 16]. Bladder dysfunction may also cause LUTS,
including detrusor overactivity/OAB, detrusor underactivity/underactive bladder, as well as other structural
or functional abnormalities of the urinary tract and its surrounding tissues [16]. Prostatic inflammation also
appears to play a role in BPH pathogenesis and progression [17, 18]. In addition, many non-urological
conditions also contribute to urinary symptoms, especially nocturia [8].

The definitions of the most common conditions related to male LUTS are presented below:
• Acute retention of urine is defined as a painful, palpable or percussible bladder, when the patient is
unable to pass any urine [6].
• Chronic retention of urine is defined as a non-painful bladder, which remains palpable or percussible after
the patient has passed urine. Such patients may be incontinent [6].
• Bladder outlet obstruction is the generic term for obstruction during voiding and is characterised
by increasing detrusor pressure and reduced urine flow rate. It is usually diagnosed by studying the
synchronous values of flow-rate and detrusor pressure [6].
• Benign prostatic obstruction is a form of BOO and may be diagnosed when the cause of outlet
obstruction is known to be BPE [6]. In the Guidelines the term BPO or BOO is used as reported by the
original studies.
• Benign prostatic hyperplasia is a term used (and reserved) for the typical histological pattern, which
defines the disease.

MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021 5
• Detrusor overactivity (DO) is a urodynamic observation characterised by involuntary detrusor contractions
during the filling phase which may be spontaneous or provoked [6]. Detrusor overactivity is usually
associated with overactive bladder syndrome characterised by urinary urgency, with or without urgency
urinary incontinence, usually with increased daytime frequency and nocturia, if there is no proven
infection or other obvious pathology [19].
• Detrusor underactivity (DU) during voiding is characterised by decreased detrusor voiding pressure
leading to a reduced urine flow rate. Detrusor underactivity causes underactive bladder syndrome which
is characterised by voiding symptoms similar to those caused by BPO [20].

Figure 1 illustrates the potential causes of LUTS. In any man complaining of LUTS, it is common for more than
one of these factors to be present.

Figure 1: Causes of male LUTS

Benign
prostac
Overacve obstrucon
bladder/
Detrusor Others
overacvity

Nocturnal Distal
polyuria ureteric
stone

Underacve
bladder/ Bladder
Detrusor LUTS tumour
underacvity

Chronic
Urethral
Pelvic Pain
stricture
syndrome

Neurogenic Foreign
bladder body
dysfuncon Urinary tract
infecon /
Inflammaon

4. DIAGNOSTIC EVALUATION
Tests are useful for diagnosis, monitoring, assessing the risk of disease progression, treatment planning, and
the prediction of treatment outcomes. The clinical assessment of patients with LUTS has two main objectives:
• to identify the differential diagnoses, since the origin of male LUTS is multifactorial, the relevant EAU
Guidelines on the management of applicable conditions should be followed;
• to define the clinical profile (including the risk of disease progression) of men with LUTS in order to
provide appropriate care.

4.1 Medical history


The importance of assessing the patient’s history is well recognised [21-23]. A medical history aims to identify
the potential causes and relevant comorbidities, including medical and neurological diseases. In addition,

6 MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021
current medication, lifestyle habits, emotional and psychological factors must be reviewed. The Panel
recognises the need to discuss LUTS and the therapeutic pathway from the patient’s perspective. This includes
reassuring the patient that there is no definite link between LUTS and prostate cancer (PCa) [24, 25].
As part of the urological/surgical history, a self-completed validated symptom questionnaire (see
section 4.2) should be obtained to objectify and quantify LUTS. Bladder diaries or frequency volume charts are
particularly beneficial when assessing patients with nocturia and/or storage symptoms (see section 4.3). Sexual
function should also be assessed, preferably with validated symptom questionnaires such as the International
Index for Erectile Function (IIEF) [26].

Summary of evidence LE
A medical history is an integral part of a patient’s medical evaluation. 4
A medical history aims to identify the potential causes of LUTS as well as any relevant comorbidities 4
and to review the patient’s current medication and lifestyle habits.

Recommendation Strength rating


Take a complete medical history from men with LUTS. Strong

4.2 Symptom score questionnaires


All published guidelines for male LUTS recommend using validated symptom score questionnaires [21, 23].
Several questionnaires have been developed which are sensitive to symptom changes and can be used to
monitor treatment [27-33]. Symptom scores are helpful in quantifying LUTS and in identifying which type of
symptoms are predominant; however, they are not disease-, gender-, or age-specific. A systematic review (SR)
evaluating the diagnostic accuracy of individual symptoms and questionnaires, compared with urodynamic
studies (the reference standard), for the diagnosis of BOO in males with LUTS found that individual symptoms
and questionnaires for diagnosing BOO were not significantly associated with one another [34].

4.2.1 The International Prostate Symptom Score (IPSS)


The IPSS is an eight-item questionnaire, consisting of seven symptom questions and one QoL question [28].
The IPSS score is categorised as ‘asymptomatic’ (0 points), ‘mildly symptomatic’ (1-7 points), ‘moderately
symptomatic’ (8-19 points), and ‘severely symptomatic’ (20-35 points). Limitations include lack of assessment
of incontinence, post-micturition symptoms, and bother caused by each separate symptom.

4.2.2 The International Consultation on Incontinence Questionnaire (ICIQ-MLUTS)


The ICIQ-MLUTS was created from the International Continence Society (ICS) Male questionnaire. It is a widely
used and validated patient completed questionnaire including incontinence questions and bother for each
symptom [29]. It contains thirteen items, with subscales for nocturia and OAB, and is available in seventeen
languages.

4.2.3 Danish Prostate Symptom Score (DAN-PSS)


The DAN-PSS [32] is a symptom score used mainly in Denmark and Finland. The DAN-PSS also has questions
on incontinence and measures the bother of each individual LUTS.

Summary of evidence LE
Symptom questionnaires are sensitive to symptom changes. 3
Symptom scores can quantify LUTS and identify which types of symptoms are predominant; however, 3
they are not disease-, gender-, or age-specific.

Recommendation Strength rating


Use a validated symptom score questionnaire including bother and quality of life Strong
assessment during the assessment of male LUTS and for re-evaluation during and/or after
treatment.

4.3 Frequency volume charts and bladder diaries


The recording of volume and time of each void by the patient is referred to as a frequency volume chart (FVC).
Inclusion of additional information such as fluid intake, use of pads, activities during recording, or which grades
of symptom severity and bladder sensation is termed a bladder diary [6]. Parameters that can be derived from
the FVC and bladder diary include: day-time and night-time voiding frequency, total voided volume, the fraction
of urine production during the night (nocturnal polyuria index), and volume of individual voids.

MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021 7
The mean 24-hour urine production is subject to considerable variation. Likewise, circumstantial
influence and intra-individual variation cause FVC parameters to fluctuate, though there is comparatively little
data [35, 36]. The FVC/bladder diary is particularly relevant in nocturia, where it underpins the categorisation
of underlying mechanism(s) [37-39]. The use of FVCs may cause a ‘bladder training effect’ and influence the
frequency of nocturnal voids [40].
The duration of the FVC/bladder diary needs to be long enough to avoid sampling errors, but short
enough to avoid non-compliance [41]. A SR of the available literature recommended FVCs should continue for
three or more days [42]. The ICIQ-Bladder diary (ICIQ-BD) is the only diary that has undergone full validation [43].

Summary of evidence LE
Frequency volume charts and bladder diaries provide real-time documentation of urinary function and 3
reduce recall bias.
Three and seven day FVCs provide reliable measurement of urinary symptoms in patients with LUTS. 2b

Recommendations Strength rating


Use a bladder diary to assess male LUTS with a prominent storage component or nocturia. Strong
Tell the patient to complete a bladder diary for at least three days. Strong

4.4 Physical examination and digital-rectal examination


Physical examination particularly focusing on the suprapubic area, the external genitalia, the perineum and
lower limbs should be performed. Urethral discharge, meatal stenosis, phimosis and penile cancer must be
excluded.

4.4.1 Digital-rectal examination and prostate size evaluation


Digital-rectal examination (DRE) is the simplest way to assess prostate volume, but the correlation to prostate
volume is poor. Quality-control procedures for DRE have been described [44]. Transrectal ultrasound (TRUS)
is more accurate in determining prostate volume than DRE. Underestimation of prostate volume by DRE
increases with increasing TRUS volume, particularly where the volume is > 30 mL [45]. A model of visual aids
has been developed to help urologists estimate prostate volume more accurately [46]. One study concluded
that DRE was sufficient to discriminate between prostate volumes > or < 50 mL [47].

Summary of evidence LE
Physical examination is an integral part of a patient’s medical evaluation. 4
Digital-rectal examination can be used to assess prostate volume; however, the correlation to actual 3
prostate volume is poor.

Recommendation Strength rating


Perform a physical examination including digital rectal examination in the assessment of Strong
male LUTS.

4.5 Urinalysis
Urinalysis (dipstick or sediment) must be included in the primary evaluation of any patient presenting with LUTS
to identify conditions, such as urinary tract infections (UTI), microhaematuria and diabetes mellitus. If abnormal
findings are detected further tests are recommended according to other EAU Guidelines, e.g. Guidelines on
urinary tract cancers and urological infections [48-51].
Urinalysis is recommended in most Guidelines in the primary management of patients with LUTS
[52, 53]. There is limited evidence, but general expert consensus suggests that the benefits outweigh the costs
[54]. The value of urinary dipstick/microscopy for diagnosing UTI in men with LUTS without acute frequency
and dysuria has been questioned [55].

Summary of evidence LE
Urinalysis (dipstick or sediment) may indicate a UTI, proteinuria, haematuria or glycosuria requiring 3
further assessment.
The benefits of urinalysis outweigh the costs. 4

8 MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021
Recommendation Strength rating
Use urinalysis (by dipstick or urinary sediment) in the assessment of male LUTS. Strong

4.6 Prostate-specific antigen (PSA)


4.6.1 PSA and the prediction of prostatic volume
Pooled analysis of placebo-controlled trials of men with LUTS and presumed BPO showed that PSA has a
good predictive value for assessing prostate volume, with areas under the curve (AUC) of 0.76-0.78 for various
prostate volume thresholds (30 mL, 40 mL, and 50 mL). To achieve a specificity of 70%, whilst maintaining
a sensitivity between 65-70%, approximate age-specific criteria for detecting men with prostate glands
exceeding 40 mL are PSA > 1.6 ng/mL, > 2.0 ng/mL, and > 2.3 ng/mL, for men with BPH in their 50s, 60s, and
70s, respectively [56].
A strong association between PSA and prostate volume was found in a large community-based
study in the Netherlands [57]. A PSA threshold value of 1.5 ng/mL could best predict a prostate volume of
> 30 mL, with a positive predictive value (PPV) of 78%. The prediction of prostate volume can also be based on
total and free PSA. Both PSA forms predict the TRUS prostate volume (± 20%) in > 90% of the cases [58, 59].

4.6.2 PSA and the probability of PCa


The role of PSA in the diagnosis of PCa is presented by the EAU Guidelines on Prostate Cancer [60]. The
potential benefits and harms of using serum PSA testing to diagnose PCa in men with LUTS should be
discussed with the patient.

4.6.3 PSA and the prediction of BPO-related outcomes


Serum PSA is a stronger predictor of prostate growth than prostate volume [61]. In addition, the PLESS study
showed that PSA also predicted the changes in symptoms, QoL/bother, and maximum flow-rate (Qmax) [62].
In a longitudinal study of men managed conservatively, PSA was a highly significant predictor of clinical
progression [63, 64]. In the placebo arms of large double-blind studies, baseline serum PSA predicted the
risk of acute urinary retention (AUR) and BPO-related surgery [65, 66]. An equivalent link was also confirmed
by the Olmsted County Study. The risk for treatment was higher in men with a baseline PSA of > 1.4 ng/mL
[67]. Patients with BPO seem to have a higher PSA level and larger prostate volumes. The PPV of PSA for the
detection of BPO was recently shown to be 68% [68]. Furthermore, in an epidemiological study, elevated free
PSA levels could predict clinical BPH, independent of total PSA levels [69].

Summary of evidence LE
Prostate-specific antigen has a good predictive value for assessing prostate volume and is a strong 1b
predictor of prostate growth.
Baseline PSA can predict the risk of AUR and BPO-related surgery. 1b

Recommendations Strength rating


Measure prostate-specific antigen (PSA) if a diagnosis of prostate cancer will change Strong
management.
Measure PSA if it assists in the treatment and/or decision making process. Strong

4.7 Renal function measurement


Renal function may be assessed by serum creatinine or estimated glomerular filtration rate (eGFR).
Hydronephrosis, renal insufficiency or urinary retention are more prevalent in patients with signs or symptoms
of BPO [70]. Even though BPO may be responsible for these complications, there is no conclusive evidence on
the mechanism [71].
One study reported that 11% of men with LUTS had renal insufficiency [70]. Neither symptom score
nor QoL was associated with the serum creatinine level. Diabetes mellitus or hypertension were the most
likely causes of the elevated creatinine concentration. Comiter et al. [72] reported that non-neurogenic voiding
dysfunction is not a risk factor for elevated creatinine levels. Koch et al. [73] concluded that only those with an
elevated creatinine level require investigational ultrasound (US) of the kidney.
In the Olmsted County Study community-dwelling men there was a cross-sectional association
between signs and symptoms of BPO (though not prostate volume) and chronic kidney disease (CKD) [74].
In 2,741 consecutive patients who presented with LUTS, decreased Qmax, a history of hypertension and/or
diabetes were associated with CKD [75]. Another study demonstrated a correlation between Qmax and eGFR in
middle-aged men with moderate-to-severe LUTS [76]. Patients with renal insufficiency are at an increased risk
of developing post-operative complications [77].

MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021 9
Summary of evidence LE
Decreased Qmax and a history of hypertension and/or diabetes are associated with CKD in patients 3
who present with LUTS.
Patients with renal insufficiency are at an increased risk of developing post-operative complications. 3

Recommendation Strength rating


Assess renal function if renal impairment is suspected based on history and clinical Strong
examination, or in the presence of hydronephrosis, or when considering surgical treatment
for male LUTS.

4.8 Post-void residual urine


Post-void residual (PVR) urine can be assessed by transabdominal US, bladder scan or catheterisation.
Post-void residual is not necessarily associated with BOO, since high PVR volumes can be
a consequence of obstruction and/or poor detrusor function (DU) [78, 79]. Using a PVR threshold of
50 mL, the diagnostic accuracy of PVR measurement has a PPV of 63% and a negative predictive value
(NPV) of 52% for the prediction of BOO [80]. A large PVR is not a contraindication to watchful waiting (WW)
or medical therapy, although it may indicate a poor response to treatment and especially to WW. In both
the MTOPS and ALTESS studies, a high baseline PVR was associated with an increased risk of symptom
progression [65, 66].
Monitoring of changes in PVR over time may allow for identification of patients at risk of AUR [81].
This is of particular importance for the treatment of patients using antimuscarinic medication. In contrast,
baseline PVR has little prognostic value for the risk of BPO-related invasive therapy in patients on α1-blockers
or WW [82]. However, due to large test-retest variability and lack of outcome studies, no PVR threshold for
treatment decision has yet been established; this is a research priority.

Summary of evidence LE
The diagnostic accuracy of PVR measurement, using a PVR threshold of 50 mL, has a PPV of 63% 3
and a NPV of 52% for the prediction of BOO.
Monitoring of changes in PVR over time may allow for identification of patients at risk of AUR. 3

Recommendation Strength rating


Measure post-void residual in the assessment of male LUTS. Weak

4.9 Uroflowmetry
Urinary flow rate assessment is a widely used non-invasive urodynamic test. Key parameters are Qmax and
flow pattern. Uroflowmetry parameters should preferably be evaluated with voided volume > 150 mL. As Qmax
is prone to within-subject variation [83, 84], it is useful to repeat uroflowmetry measurements, especially if the
voided volume is < 150 mL, or Qmax or flow pattern is abnormal.
The diagnostic accuracy of uroflowmetry for detecting BOO varies considerably and is substantially
influenced by threshold values. A threshold Qmax of 10 mL/s has a specificity of 70%, a PPV of 70% and a
sensitivity of 47% for BOO. The specificity using a threshold Qmax of 15 mL/s was 38%, the PPV 67% and the
sensitivity 82% [85]. If Qmax is > 15 mL/s, physiological compensatory processes mean that BOO cannot be
excluded. Low Qmax can arise as a consequence of BOO [86], DUA or an under-filled bladder [87]. Therefore,
it is limited as a diagnostic test as it is unable to discriminate between the underlying mechanisms. Specificity
can be improved by repeated flow rate testing. Uroflowmetry can be used for monitoring treatment outcomes
[88] and correlating symptoms with objective findings.

Summary of evidence LE
The diagnostic accuracy of uroflowmetry for detecting BOO varies considerably and is substantially 2b
influenced by threshold values. Specificity can be improved by repeated flow rate testing.

Recommendations Strength rating


Perform uroflowmetry in the initial assessment of male LUTS. Weak
Perform uroflowmetry prior to medical or invasive treatment. Strong

10 MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021
4.10 Imaging
4.10.1 Upper urinary tract
Men with LUTS are not at increased risk for upper tract malignancy or other abnormalities when compared to
the overall population [73, 89-91]. Several arguments support the use of renal US in preference to intravenous
urography. Ultrasound allows for better characterisation of renal masses, the possibility of investigating the liver
and retroperitoneum, and simultaneous evaluation of the bladder, PVR and prostate, together with a lower cost,
radiation dose and less side effects [89]. Ultrasound can be used for the evaluation of men with large PVR,
haematuria, or a history of urolithiasis.

Summary of evidence LE
Men with LUTS are not at increased risk for upper tract malignancy or other abnormalities when 3
compared to the overall population.
Ultrasound can be used for the evaluation of men with large PVR, haematuria, or a history of 4
urolithiasis.

Recommendation Strength rating


Perform ultrasound of the upper urinary tract in men with LUTS. Weak

4.10.2 Prostate
Imaging of the prostate can be performed by transabdominal US, TRUS, computed tomography (CT),
and magnetic resonance imaging (MRI). However, in daily practice, prostate imaging is performed by
transabdominal (suprapubic) US or TRUS [89].

4.10.2.1 Prostate size and shape


Assessment of prostate size is important for the selection of interventional treatment, i.e. open prostatectomy
(OP), enucleation techniques, transurethral resection, transurethral incision of the prostate (TUIP), or minimally
invasive therapies. It is also important prior to treatment with 5α-reductase inhibitors (5-ARIs). Prostate volume
predicts symptom progression and the risk of complications [91].
Transrectal US is superior to transabdominal volume measurement [92, 93]. The presence of a
median lobe may guide treatment choice in patients scheduled for a minimally invasive approach since medial
lobe presence can be a contraindication for some minimally invasive treatments (see section 5.3).

Summary of evidence LE
Assessment of prostate size by TRUS or transabdominal US is important for the selection of 3
interventional treatment and prior to treatment with 5-ARIs.

Recommendations Strength rating


Perform imaging of the prostate when considering medical treatment for male LUTS, if it Weak
assists in the choice of the appropriate drug.
Perform imaging of the prostate when considering surgical treatment. Strong

4.10.3 Voiding cysto-urethrogram


Voiding cysto-urethrogram (VCUG) is not recommended in the routine diagnostic work-up of men with LUTS,
but it may be useful for the detection of vesico-ureteral reflux, bladder diverticula, or urethral pathologies.
Retrograde urethrography may additionally be useful for the evaluation of suspected urethral strictures.

4.11 Urethrocystoscopy
Patients with a history of microscopic or gross haematuria, urethral stricture, or bladder cancer, who present
with LUTS, should undergo urethrocystoscopy during diagnostic evaluation. The evaluation of a prostatic
middle lobe with urethrocystocopy should be performed when considering interventional treatments for which
the presence of middle lobe is a contraindication.
A prospective study evaluated 122 patients with LUTS using uroflowmetry and urethrocystoscopy
[94]. The pre-operative Qmax was normal in 25% of 60 patients who had no bladder trabeculation, 21% of
73 patients with mild trabeculation and 12% of 40 patients with marked trabeculation on cystoscopy. All 21
patients who presented with diverticula had a reduced Qmax.
Another study showed that there was no significant correlation between the degree of bladder
trabeculation (graded from I to IV), and the pre-operative Qmax value in 39 symptomatic men aged 53-83 years
[95]. The largest study published on this issue examined the relation of urethroscopic findings to urodynamic

MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021 11
studies in 492 elderly men with LUTS [96]. The authors noted a correlation between cystoscopic appearance
(grade of bladder trabeculation and urethral occlusion) and urodynamic indices, DO and low compliance. It
should be noted, however, that BOO was present in 15% of patients with normal cystoscopic findings, while
8% of patients had no obstruction, even in the presence of severe trabeculation [96].

Summary of evidence LE
Patients with a history of microscopic or gross haematuria, urethral stricture, or bladder cancer, who 3
present with LUTS, should undergo urethrocystoscopy during diagnostic evaluation.
None of the studies identified a strong association between the urethrocystoscopic and urodynamic 3
findings.

Recommendation Strength rating


Perform urethrocystoscopy in men with LUTS prior to minimally invasive/surgical therapies Weak
if the findings may change treatment.

4.12 Urodynamics
In male LUTS, the most widespread invasive urodynamic techniques employed are filling cystometry and
pressure flow studies (PFS). The major goal of urodynamics is to explore the functional mechanisms of LUTS,
to identify risk factors for adverse outcomes and to provide information for shared decision-making. Most
terms and conditions (e.g. DO, low compliance, BOO/BPO, DUA) are defined by urodynamic investigation.

4.12.1 Diagnosing bladder outlet obstruction


Pressure flow studies (PFS) are used to diagnose and define the severity of BOO, which is characterised by
increased detrusor pressure and decreased urinary flow rate during voiding. Bladder outlet obstruction/BPO
has to be differentiated from DUA, which exhibits decreased detrusor pressure during voiding in combination
with decreased urinary flow rate [6].
Urodynamic testing may also identify DO. Studies have described an association between BOO and
DO [97, 98]. In men with LUTS attributed to BPO, DO was present in 61% and independently associated with
BOO grade and ageing [97].
The prevalence of DUA in men with LUTS is 11-40% [99, 100]. Detrusor contractility does not
appear to decline in long-term BOO and surgical relief of BOO does not improve contractility [101, 102]. There
are no published RCTs in men with LUTS and possible BPO that compare the standard practice investigation
(uroflowmetry and PVR measurement) with PFS with respect to the outcome of treatment; however, a study has
been completed in the UK, but the final results have not yet been published [103, 104]. Once available they will
be included in the next edition of the Guidelines.
A Cochrane meta-analysis was done to determine whether performing invasive urodynamic
investigation reduces the number of men with continuing symptoms of voiding dysfunction. Two trials with 350
patients were included. Invasive urodynamic testing changed clinical decision making. Patients who underwent
urodynamics were less likely to undergo surgery; however, no evidence was found to demonstrate whether
this led to reduced symptoms of voiding dysfunction after treatment [105]. A more recent meta-analysis of
retrospective studies showed that pre-operative urodynamic DOA has no diagnostic role in the prediction of
surgical outcomes in patients with male BOO [106].
Due to the invasive nature of the test, a urodynamic investigation is generally only offered if
conservative treatment has failed. The Guidelines Panel attempted to identify specific indications for PFS
based on age, findings from other diagnostic tests and previous treatments. The Panel allocated a different
degree of obligation for PFS in men > 80 years and men < 50 years, which reflects the lack of evidence. In
addition, there was no consensus whether PFS should or may be performed when considering surgery in men
with bothersome predominantly voiding LUTS and Qmax > 10 mL/s, although the Panel recognised that with a
Qmax < 10 mL/s, BOO is likely and PFS is not necessarily needed.
Patients with neurological disease, including those with previous radical pelvic surgery, should be
assessed according to the EAU Guidelines on Neuro-Urology [107].

4.12.2 Videourodynamics
Videourodynamics provides additional anatomical and functional information and may be recommended if the
clinician considers this is needed to understand the pathophysiological mechanism of an individual patient’s
LUTS.

12 MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021
Summary of evidence LE
Within the literature cut-off dates of the 2021 edition of the Guidelines there were no RCTs in men 3
with LUTS and possible BPO that compare the standard practice investigation (uroflowmetry and
PVR measurement) with PFS with respect to the outcome of treatment. All subsequent RCTs will be
evaluated for inclusion in the next edition of the Guidelines.

Recommendations Strength rating


Perform pressure-flow studies (PFS) only in individual patients for specific indications prior Weak
to invasive treatment or when further evaluation of the underlying pathophysiology of LUTS
is warranted.
Perform PFS in men who have had previous unsuccessful (invasive) treatment for LUTS. Weak
Perform PFS in men considering invasive treatment who cannot void > 150 mL. Weak
Perform PFS when considering surgery in men with bothersome predominantly voiding Weak
LUTS and Qmax > 10 mL/s.
Perform PFS when considering invasive therapy in men with bothersome, predominantly Weak
voiding LUTS with a post void residual > 300 mL.
Perform PFS when considering invasive treatment in men with bothersome, predominantly Weak
voiding LUTS aged > 80 years.
Perform PFS when considering invasive treatment in men with bothersome, predominantly Weak
voiding LUTS aged < 50 years.

4.13 Non-invasive tests in diagnosing bladder outlet obstruction in men with LUTS
4.13.1 Prostatic configuration/intravesical prostatic protrusion (IPP)
Prostatic configuration can be evaluated with TRUS, using the concept of the presumed circle area ratio
(PCAR) [108]. The PCAR evaluates how closely the transverse US image of the prostate approaches a circular
shape. The ratio tends toward one as the prostate becomes more circular. The sensitivity of PCAR was 77% for
diagnosing BPO when PCAR was > 0.8, with 75% specificity [108].
Ultrasound measurement of IPP assesses the distance between the tip of the prostate median
lobe and bladder neck in the midsagittal plane, using a suprapubically positioned US scanner, with a bladder
volume of 150-250 mL; grade I protrusion is 0-4.9 mm, grade II is 5-10 mm and grade III is > 10 mm.
Intravesical prostatic protrusion correlates well with BPO (presence and severity) on urodynamic
testing, with a PPV of 94% and a NPV of 79% [109]. Intravesical prostatic protrusion may also correlate with
prostate volume, DO, bladder compliance, detrusor pressure at maximum urinary flow, BOO index and PVR,
and negatively correlates with Qmax [110]. Furthermore, IPP also appears to successfully predict the outcome
of a trial without catheter after AUR [111, 112]. However, no information with regard to intra- or inter-observer
variability and learning curve is yet available. Therefore, whilst IPP may be a feasible option to infer BPO in men
with LUTS, the role of IPP as a non-invasive alternative to PFS in the assessment of male LUTS remains under
evaluation.

4.13.2 Bladder/detrusor wall thickness and ultrasound-estimated bladder weight


For bladder wall thickness (BWT) assessment, the distance between the mucosa and the adventitia is
measured. For detrusor wall thickness (DWT) assessment, the only measurement needed is the detrusor
sandwiched between the mucosa and adventitia [113].
A correlation between BWT and PFS parameters has been reported. A threshold value of 5 mm
at the anterior bladder wall with a bladder filling of 150 mL was best at differentiating between patients with
or without BOO [114]. Detrusor wall thickness at the anterior bladder wall with a bladder filling > 250 mL
(threshold value for BOO > 2 mm) has a PPV of 94% and a specificity of 95%, achieving 89% agreement with
PFS [73]. Threshold values of 2.0, 2.5, or 2.9 mm for DWT in patients with LUTS are able to identify 81%, 89%,
and 100% of patients with BOO, respectively [115].
All studies found that BWT or DWT measurements have a higher diagnostic accuracy for detecting
BOO than Qmax or Qave of free uroflowmetry, measurements of PVR, prostate volume, or symptom severity. One
study could not demonstrate any difference in BWT between patients with normal urodynamics, BOO or DO.
However, the study did not use a specific bladder filling volume for measuring BWT [116]. Disadvantages of the
method include the lack of standardisation, and lack of evidence to indicate which measurement (BWT/DWT) is
preferable [117]. Measurement of BWT/DWT is therefore not recommended for the diagnostic work-up of men
with LUTS.
Ultrasound-estimated bladder weight (UEBW) may identify BOO with a diagnostic accuracy of 86%
at a cut-off value of 35 g [118, 119]. Severe LUTS and a high UEBW (> 35 g) are risk factors for prostate/BPH
surgery in men on α-blockers [120].

MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021 13
4.13.3 Non-invasive pressure-flow testing
The penile cuff method, in which flow is interrupted to estimate isovolumetric bladder pressure, shows
promising data, with good test repeatability [121] and interobserver agreement [122]. A nomogram has also
been derived [123] whilst a method in which flow is not interrupted is also under investigation [124].
The data generated with the external condom method [125] correlates with invasive PFS in a high
proportion of patients [126]. Resistive index [127] and prostatic urethral angle [128] have also been proposed,
but are still experimental.

4.13.4 The diagnostic performance of non-invasive tests in diagnosing bladder outlet obstruction in
men with LUTS compared with pressure-flow studies
The diagnostic performance of non-invasive tests in diagnosing BOO in men with LUTS compared with PFS
has been investigated in a SR [129]. A total of 42 studies were included is this review. The majority were
prospective cohort studies, and the diagnostic accuracy of the following non-invasive tests were assessed:
penile cuff test; uroflowmetry; detrusor/bladder wall thickness; bladder weight; external condom catheter
method; IPP; Doppler US; prostate volume/height; and near-infrared spectroscopy. Overall, although the
majority of studies have a low risk of bias, data regarding the diagnostic accuracy of these non-invasive tests
is limited by the heterogeneity of the studies in terms of the threshold values used to define BOO, the different
urodynamic definitions of BOO used across different studies and the small number of studies for each test. It
was found that specificity, sensitivity, PPV and NPV of the non-invasive tests were highly variable. Therefore,
even though several tests have shown promising results regarding non-invasive diagnosis of BOO, invasive
urodynamics remains the modality of choice.

Summary of evidence LE
Data regarding the diagnostic accuracy of non-invasive tests is limited by the heterogeneity of the 1a
studies as well as the small number of studies for each test.
Specificity, sensitivity, PPV and NPV of the non-invasive tests were highly variable. 1a

Recommendation Strength rating


Do not offer non-invasive tests as an alternative to pressure-flow studies for diagnosing Strong
bladder outlet obstruction in men.

14 MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021
Figure 2: Assessment algorithm of LUTS in men aged 40 years or older
Readers are strongly recommended to read the full text that highlights the current position of each test in detail.

Male LUTS Manage according to


EAU mLUTS
treatment algorithm

History (+ sexual function) No


Symptom score questionnaire
Urinalysis
Physical examination
PSA (if diagnosis of PCa will change the
management – discuss with patient) Bothersome
Measurement of PVR symptoms

Yes
Abnormal DRE Significant PVR
Suspicion of
neurological disease
High PSA
Abnormal urinalysis FVC in cases of predominant
US of kidneys storage LUTS/nocturia
+/- Renal function US assessment of prostate
assessment Uroflowmetry

Evaluate according to
relevant guidelines or
clinical standard Benign conditions of
Medical treatment
according to treatment bladder and/or prostate
algorithm with baseline values
PLAN TREATMENT

Treat underlying
condition
(if any, otherwise Endoscopy (if test would alter the
return to initial choice of surgical modality)
assessment) Pressure flow studies (see text for
specific indications)

Surgical treatment according


to treatment algorithm

DRE = digital-rectal examination; FVC = frequency volume chart; LUTS = lower urinary tract symptoms;
PCa = prostate cancer; PSA = prostate specific antigen; PVR = post-void residual; US = ultrasound.

5. DISEASE MANAGEMENT
5.1 Conservative treatment
5.1.1 Watchful waiting (WW)
Many men with LUTS are not troubled enough by their symptoms to need drug treatment or surgical
intervention. All men with LUTS should be formally assessed prior to any allocation of treatment in order
to establish symptom severity and to differentiate between men with uncomplicated (the majority) and

MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021 15
complicated LUTS. Watchful waiting is a viable option for many men with non-bothersome LUTS as few will
progress to AUR and complications (e.g. renal insufficiency or stones) [130, 131], whilst others can remain
stable for years [132]. In one study, approximately 85% of men with mild LUTS were stable on WW at one year
[133].
A study comparing WW and transurethral resection of the prostate (TURP) in men with moderate
LUTS showed the surgical group had improved bladder function (flow rates and PVR volumes), especially
in those with high levels of bother; 36% of WW patients crossed over to surgery within five years, leaving
64% doing well in the WW group [134, 135]. Increasing symptom bother and PVR volumes are the strongest
predictors of WW failure. Men with mild-to-moderate uncomplicated LUTS who are not too troubled by their
symptoms are suitable for WW.

5.1.2 Behavioural and dietary modifications


It is customary for this type of management to include the following components:
• education (about the patient’s condition);
• reassurance (that cancer is not a cause of the urinary symptoms);
• periodic monitoring;
• lifestyle advice [132, 133, 136, 137] such as::
oo reduction of fluid intake at specific times aimed at reducing urinary frequency when most
inconvenient (e.g. at night or when going out in public);
oo avoidance/moderation of intake of caffeine or alcohol, which may have a diuretic and irritant effect,
thereby increasing fluid output and enhancing frequency, urgency and nocturia;
oo use of relaxed and double-voiding techniques;
oo urethral milking to prevent post-micturition dribble;
oo distraction techniques such as penile squeeze, breathing exercises, perineal pressure, and mental
tricks to take the mind off the bladder and toilet, to help control OAB symptoms;
oo bladder retraining that encourages men to hold on when they have urgency to increase their bladder
capacity and the time between voids;
oo reviewing the medication and optimising the time of administration or substituting drugs for others
that have fewer urinary effects (these recommendations apply especially to diuretics);
oo providing necessary assistance when there is impairment of dexterity, mobility, or mental state;
oo treatment of constipation.

There now exists evidence that self-management as part of WW reduces both symptoms and progression
[136, 137]. Men randomised to three self-care management sessions in addition to standard care had better
symptom improvement and QoL than men treated with standard care only, for up to a year [136].

5.1.3 Practical considerations


The components of self-care management have not been individually studied. The above components of
lifestyle advice have been derived from formal consensus methodology [138]. Further research in this area is
required.

Summary of evidence LE
Watchful waiting is usually a safe alternative for men who are less bothered by urinary difficulty or 1b
who wish to delay treatment. The treatment failure rate over a period of five years was 21%; 79% of
patients were clinically stable.
An additional study reported 81% of patients were clinically stable on WW after a mean follow-up of 2
seventeen months.
Men randomised to three self-management sessions in addition to standard care had better
symptom improvement and QoL than men treated with standard care alone at up to a year. Self-care 1b
management as part of WW reduces both symptoms and progression.

Recommendations Strength rating


Offer men with mild/moderate symptoms, minimally bothered by their symptoms, watchful Strong
waiting.
Offer men with LUTS lifestyle advice and self-care information prior to, or concurrent with, Strong
treatment.

16 MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021
5.2 Pharmacological treatment
5.2.1 α1-Adrenoceptor antagonists (α1-blockers)
Mechanism of action: α1-blockers aim to inhibit the effect of endogenously released noradrenaline on smooth
muscle cells in the prostate and thereby reduce prostate tone and BOO [139]. However, α1-blockers have little
effect on urodynamically determined bladder outlet resistance [140], and treatment-associated improvement of
LUTS correlates poorly with obstruction [141]. Thus, other mechanisms of action may also be relevant.
Alpha 1-adrenoceptors located outside the prostate (e.g. urinary bladder and/or spinal cord)
and α1-adrenoceptor subtypes (α1B- or α1D-adrenoceptors) may play a role as mediators of effects. Alpha
1-adrenoceptors in blood vessels, other non-prostatic smooth muscle cells, and the central nervous system
may mediate adverse events.
Currently available α1-blockers are: alfuzosin hydrochloride (alfuzosin); doxazosin mesylate
(doxazosin); silodosin; tamsulosin hydrochloride (tamsulosin); terazosin hydrochloride (terazosin); and
naftopidil. Alpha 1-blockers exist in different formulations. Although different formulations result in different
pharmacokinetic and tolerability profiles, the overall difference in clinical efficacy between the difference
formulations seems modest.

Efficacy: Indirect comparisons and limited direct comparisons between α1-blockers demonstrate that all
α1-blockers have a similar efficacy in appropriate doses [142]. Clinical effects take a few weeks to develop
fully, but significant efficacy over placebo can occur within hours to days [141].
Controlled studies show that α1-blockers typically reduce IPSS by approximately 30-40%
and increase Qmax by approximately 20-25%. However, considerable improvements also occurred in the
corresponding placebo arms [63, 143]. In open-label studies, an IPSS improvement of up to 50% and Qmax
increase of up to 40% were documented [63, 143]. A recent SR and meta-analysis suggested that Qmax
variation underestimates the real effect of α1-blockers on BPO, as small improvements in Qmax correspond to
relevant improvements in BOO index in PFS [144].
Alpha 1-blockers can reduce both storage and voiding LUTS. Prostate size does not affect
α1-blocker efficacy in studies with follow-up periods of less than one year, but α1-blockers do seem to be
more efficacious in patients with smaller prostates (< 40 mL) in longer-term studies [65, 145-148]. The efficacy
of α1-blockers is similar across age groups [143]. In addition, α1-blockers neither reduce prostate size nor
prevent AUR in long-term studies [146-148]; however, recent evidence suggests that the use of α1-blockers
(alfuzosin and tamsulosin) may improve resolution of AUR [149]. Nonetheless, IPSS reduction and Qmax
improvement during α1-blocker treatment appears to be maintained over at least four years.

Tolerability and safety: Tissue distribution, subtype selectivity, and pharmacokinetic profiles of certain
formulations may contribute to the tolerability profile of specific drugs. The most frequent adverse events of
α1-blockers are asthenia, dizziness and (orthostatic) hypotension. Vasodilating effects are most pronounced
with doxazosin and terazosin, and are less common with alfuzosin and tamsulosin [150]. Patients with
cardiovascular comorbidity and/or vaso-active co-medication may be susceptible to α1-blocker-induced
vasodilatation [151]. In contrast, the frequency of hypotension with the α1A-selective blocker silodosin is
comparable with placebo [152]. In a large retrospective cohort analysis of men aged > 66 years treated with
α1-blockers the risks of falling (odds ratio [OR] 1.14) and of sustaining a fracture (OR 1.16) was increased, most
likely as a result of induced hypotension [153].
An adverse ocular event termed intra-operative floppy iris syndrome (IFIS) was reported in 2005,
affecting cataract surgery [154]. A meta-analysis on IFIS after alfuzosin, doxazosin, tamsulosin or terazosin
exposure showed an increased risk for all α1-blockers [155]. However, the OR for IFIS was much higher for
tamsulosin. It appears prudent not to initiate α1-blocker treatment prior to scheduled cataract surgery, and the
ophthalmologist should be informed about α1-blocker use.
A SR concluded that α1-blockers do not adversely affect libido, have a small beneficial effect on
erectile function, but can cause abnormal ejaculation [156]. Originally, abnormal ejaculation was thought to be
retrograde, but more recent data demonstrate that it is due to a decrease or absence of seminal fluid during
ejaculation, with young age being an apparent risk factor. In a recent meta-analysis ejaculatory dysfunction
(EjD) was significantly more common with α1-blockers than with placebo (OR 5.88). In particular, EjD was
significantly more commonly related with tamsulosin or silodosin (OR 8.57 and 32.5) than placebo, while
both doxazosin and terazosin (OR 0.80 and 1.78) were associated with a low risk of EjD [157]. In the meta-
regression, the occurrence of EjD was independently associated with the improvement of urinary symptoms
and flow rate, suggesting that the more effective the α1-blocker is the greater the incidence of EjD.

MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021 17
Practical considerations: α1-blockers are usually considered the first-line drug treatment for male LUTS
because of their rapid onset of action, good efficacy, and low rate and severity of adverse events. However,
α1-blockers do not prevent occurrence of urinary retention or need for surgery. Ophthalmologists should be
informed about α1-blocker use prior to cataract surgery. Elderly patients treated with non-selective α1-blockers
should be informed about the risk of orthostatic hypotension. Sexually active patients treated with selective
α1-blockers should be counselled about the risk of EjD.

Summary of evidence LE
α1-blockers are effective in reducing urinary symptoms (IPSS) and increasing the peak urinary flow 1a
rate (Qmax) compared with placebo.
Alfuzosin, terazosin and doxazosin showed a statistically significant increased risk of developing 1a
vascular-related events compared with placebo.
Alfuzosin, doxazosin, tamsulosin or terazosin exposure has been associated with an increased risk of 1a
IFIS.
Ejaculatory dysfunction is significantly more common with α1-blockers than with placebo, particularly 1a
with more selective α1-blockers such as tamsulosin and silodosin.

Recommendation Strength rating


Offer α1-blockers to men with moderate-to-severe LUTS. Strong

5.2.2 5α-reductase inhibitors


Mechanism of action: Androgen effects on the prostate are mediated by dihydrotestosterone (DHT), which is
converted from testosterone by the enzyme 5α-reductase [158], which has two isoforms:
• 5α-reductase type 1: predominant expression and activity in the skin and liver.
• 5α-reductase type 2: predominant expression and activity in the prostate.

Two 5-ARIs are available for clinical use: dutasteride and finasteride. Finasteride inhibits only 5α-reductase
type 2, whereas dutasteride inhibits both 5α-reductase types (dual 5-ARI). The 5-ARIs induce apoptosis of
prostate epithelial cells [159] leading to prostate size reduction of about 18-28% and a decrease in circulating
PSA levels of about 50% after six to twelve months of treatment [160]. Mean prostate volume and PSA
reduction may be even more pronounced after long-term treatment. Continuous treatment reduces the serum
DHT concentration by approximately 70% with finasteride and 95% with dutasteride. However, prostate DHT
concentration is reduced to a similar level (85-90%) by both 5-ARIs.

Efficacy: Clinical effects relative to placebo are seen after treatment of at least six months. After two to four
years of treatment 5-ARIs improve IPSS by approximately 15-30%, decrease prostate volume by 18-28%, and
increase Qmax by 1.5-2.0 mL/s in patients with LUTS due to prostate enlargement [65, 147, 148, 161-167]. An
indirect comparison and one direct comparative trial (twelve months duration) indicate that dutasteride and
finasteride are equally effective in the treatment of LUTS [160, 168]. Symptom reduction depends on initial
prostate size.
Finasteride may not be more efficacious than placebo in patients with prostates < 40 mL [169].
However, dutasteride seems to reduce IPSS, prostate volume, and the risk of AUR, and to increase Qmax even
in patients with prostate volumes of between 30 and 40 mL [170, 171]. A long-term trial with dutasteride in
symptomatic men with prostate volumes > 30 mL and increased risk for disease progression showed that
dutasteride reduced LUTS at least as much as the α1-blocker tamsulosin [147, 167, 172]. The greater the
baseline prostate volume (or serum PSA level), the faster and more pronounced the symptomatic benefit of
dutasteride as compared to tamsulosin.
5α-reductase inhibitors, but not α1-blockers, reduce the long-term (> 1 year) risk of AUR or need
for surgery [65, 165, 173]. In the PLESS study, finasteride reduced the relative risk of AUR by 57% and need
for surgery by 55% (absolute risk reduction 4% and 7%, respectively) at four years, compared with placebo
[165]. In the MTOPS study, finasteride reduced the relative risk of AUR by 68% and need for surgery by
64% (absolute risk reduction 2% and 3%, respectively), also at four years [65]. A pooled analysis of three
randomised trials with two-year follow-up data, reported that treatment with finasteride decreased the relative
risk of AUR by 57%, and surgical intervention by 34% (absolute risk reduction 2% for both) in patients with
moderately symptomatic LUTS [174]. Dutasteride has also demonstrated efficacy in reducing the risks for AUR
and BPO-related surgery. Open-label trials have demonstrated relevant changes in urodynamic parameters
[175, 176]. Furthermore, finasteride might reduce blood loss during transurethral prostate surgery, probably due
to its effects on prostatic vascularisation [177, 178].

18 MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021
Tolerability and safety: The most common adverse events are reduced libido, erectile dysfunction (ED) and less
frequently, ejaculation disorders such as retrograde ejaculation, ejaculation failure, or decreased semen volume
[65, 148, 160, 179]. Gynaecomastia (with breast or nipple tenderness) develops in 1-2% of patients. Two
studies have suggested that treatment with 5-ARIs is associated with a higher incidence of high-grade cancers
although no causal relationship has been proven [180, 181]. There is a long-standing debate regarding potential
cardiovascular side effects of 5-ARIs, in particular dutasteride [182]. Population-based studies in Taiwan and
Ontario did not find an association between the use of 5-ARIs and increased cardiovascular side effects [182,
183].

Practical considerations: Treatment with 5-ARIs should be considered in men with moderate-to-severe LUTS
and an enlarged prostate (> 40 mL) and/or elevated PSA concentration (> 1.4-1.6 ng/mL). They can reduce the
risk of AUR and need for surgery. Due to the slow onset of action, they are not suitable for short-term use. Their
effect on PSA needs to be considered in relation to PCa screening.

Summary of evidence LE
After two to four years of treatment, 5-ARIs improve IPSS by approximately 15-30%, decrease 1b
prostate volume by 18-28%, and increase Qmax by 1.5-2.0 mL/s in patients with LUTS due to prostate
enlargement.
5α-reductase inhibitors can prevent disease progression with regard to AUR and the need for surgery. 1a
Due to 5-ARIs slow onset of action, they are suitable only for long-term treatment (years).
The most relevant adverse effects of 5-ARIs are related to sexual function, and include reduced libido, 1b
ED and less frequently, ejaculation disorders such as retrograde ejaculation, ejaculation failure, or
decreased semen volume.

Recommendations Strength rating


Use 5α-reductase inhibitors in men who have moderate-to-severe LUTS and an increased Strong
risk of disease progression (e.g. prostate volume > 40 mL).
Counsel patients about the slow onset of action of 5α-reductase inhibitors. Strong

5.2.3 Muscarinic receptor antagonists


Mechanism of action: The detrusor is innervated by parasympathetic nerves whose main neurotransmitter
is acetylcholine, which stimulates muscarinic receptors (M-cholinoreceptors) on the smooth muscle cells.
Muscarinic receptors are also present on other cell types, such as bladder urothelial cells and epithelial cells
of the salivary glands. Five muscarinic receptor subtypes (M1-M5) have been described, of which M2 and
M3 are predominant in the detrusor. The M2 subtype is more numerous, but the M3 subtype is functionally
more important in bladder contractions [184, 185]. Antimuscarinic effects might also be induced or modulated
through other cell types, such as the bladder urothelium or by the central nervous system [186, 187].
The following muscarinic receptor antagonists are licensed for treating OAB/storage symptoms:
darifenacin hydrobromide (darifenacin); fesoterodine fumarate (fesoterodine); oxybutynin hydrochloride
(oxybutynin); propiverine hydrochloride (propiverine); solifenacin succinate (solifenacin); tolterodine tartrate
(tolterodine); and trospium chloride. Transdermal preparations of oxybutynin have been formulated and
evaluated in clinical trials [188, 189].

Efficacy: Antimuscarinics were mainly tested in females in the past, as it was believed that LUTS in men were
caused by the prostate, so should be treated with prostate-specific drugs. However, there is no scientific
data for this assumption [190]. A sub-analysis of an open-label trial of OAB patients showed that age, but
not gender had an impact on urgency, frequency, or urgency incontinence [191]. In a pooled analysis, which
included a sub-analysis of male patients, fesoterodine 8 mg was superior to tolterodine extended release (ER) 4
mg for the improvement of severe urgency episodes/24 hours and the OAB-q Symptom Bother score at week
twelve, the urinary retention rate was around 2% [192].

The efficacy of antimuscarinics as single agents in men with OAB in the absence of BOO have been tested
[193-198]. Most trials lasted only twelve weeks. Four post hoc analyses of large RCTs on the treatment of
OAB in women and men without presumed BOO were performed focusing only on the men [190, 194, 199].
Tolterodine can significantly reduce urgency incontinence, daytime or 24-hour frequency and urgency-related
voiding whilst improving patient perception of treatment benefit. Solifenacin significantly improved mean
patient perception of bladder condition scores, mean OAB questionnaire scores, and overall perception of
bladder problems. Fesoterodine improved micturition frequency, urgency episodes, and urgency urinary
incontinence (UUI) episodes. In open-label trials with tolterodine, daytime frequency, nocturia, UUI, and

MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021 19
IPSS were significantly reduced compared with baseline values after 12-25 weeks [195, 198]. The TIMES
RCT reported that tolterodine ER monotherapy significantly improved UUI episodes per 24 hours compared
to placebo, at week twelve. Tolterodine ER did not significantly improve urgency, IPSS total or QoL score
compared with placebo. A significantly greater proportion of patients in the tolterodine ER plus tamsulosin
group reported treatment benefit compared with the other three treatment groups [197].
A further analysis showed that men with PSA levels of < 1.3 ng/mL (smaller prostates) might benefit
more from antimuscarinics [200]. Two other studies found a positive effect of antimuscarinics in patients with
OAB and concomitant BPO [198, 201]. In a small RCT propiverine improved frequency and urgency episodes
[201].

Tolerability and safety: Antimuscarinic drug trials generally show approximately 3-10% withdrawals, which
is similar to placebo. Drug-related adverse events include dry mouth (up to 16%), constipation (up to 4%),
micturition difficulties (up to 2%), nasopharyngitis (up to 3%), and dizziness (up to 5%).
Increased PVR in men without BOO is minimal and similar to placebo. Nevertheless, fesoterodine
8 mg showed higher PVRs (+20.2 mL) than placebo (-0.6 mL) or fesoterodine 4 mg (+9.6 mL) [195]. Incidence
of urinary retention in men without BOO was similar to placebo for tolterodine (0-1.3% vs. 0-1.4%). With
fesoterodine 8 mg, 5.3% had symptoms, which was higher than placebo or fesoterodine 4 mg (both 0.8%).
These symptoms appeared during the first two weeks of treatment and mainly affected men aged 66 years or
older.
Theoretically antimuscarinics might decrease bladder strength, and hence might be associated with
PVR urine or urinary retention. A twelve week safety study on men with mild-to-moderate BOO showed that
tolterodine increased the PVR (49 mL vs. 16 mL) but not AUR (3% in both arms) [202]. The urodynamic effects
included larger bladder volumes at first detrusor contraction, higher maximum cystometric capacity, and
decreased bladder contractility index, Qmax was unchanged. This trial indicated that short-term treatment with
antimuscarinics in men with BOO is safe [190].

Practical considerations: Not all antimuscarinics have been tested in elderly men, and long-term studies on
the efficacy of muscarinic receptor antagonists in men of any age with LUTS are not yet available. In addition,
only patients with low PVR volumes at baseline were included in the studies. These drugs should therefore be
prescribed with caution, and regular re-evaluation of IPSS and PVR urine is advised. Men should be advised to
discontinue medication if worsening voiding LUTS or urinary stream is noted after initiation of therapy.

Summary of evidence LE
Antimuscarinic monotherapy can significantly improve urgency, UUI, and increased daytime frequency. 2
Antimuscarinic monotherapy can be associated with increased PVR after therapy, but acute retention 2
is a rare event in men with a PVR volume of < 150 mL at baseline.

Recommendations Strength rating


Use muscarinic receptor antagonists in men with moderate-to-severe LUTS who mainly Strong
have bladder storage symptoms.
Do not use antimuscarinic overactive bladder medications in men with a post-void residual Weak
volume > 150 mL.

5.2.4 Beta-3 agonist


Mechanism of action: Beta-3 adrenoceptors are the predominant beta receptors expressed in the smooth
muscle cells of the detrusor and their stimulation is thought to induce detrusor relaxation. The mode of action
of beta-3 agonists is not fully elucidated [203].

Efficacy: Mirabegron 50 mg is the first clinically available beta-3 agonist with approval for use in adults with
OAB. Mirabegron has undergone extensive evaluation in RCTs conducted in Europe, Australia, North America
and Japan [204-208]. Mirabegron demonstrated significant efficacy in treating the symptoms of OAB, including
micturition frequency, urgency and UUI and also patient perception of treatment benefit. These studies had a
predominantly female study population. A meta-analysis of eight RCTS including 10,248 patients (27% male)
found that mirabegron treatment resulted in reduced frequency, urgency and UUI rates, as well as an improved
voided volume with a statistically significant improvement of nocturia as compared with both placebo and
tolterodine [209].
Mirabegron has been evaluated in male patients with OAB in the context of LUTS either associated
with or not associated with BPO confirmed by urodynamics [210]. Mirabegron 25 mg daily led to increased
satisfaction and improved QoL, but symptoms assessed by validated questionnaires (IPSS and OAB-SS), only

20 MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021
improved in non-obstructed patients. Mirabegron as an add-on therapy has been studied in OAB patients with
incontinence despite antimuscarinic therapy [211], again in a predominantly female study population. An Asian
study with a higher proportion of male subjects (approximately one third) reported superiority over placebo in
reducing frequency of micturition, but did not report the results separately for the genders [212].
In a study of more than 1,000 patients of whom approximately 30% were male, combination therapy
of mirabegron 25/50 mg and solifenacin 5/10 mg was associated with statistically significant improvements
in patient outcomes and health related QoL vs. solifenacin 5 mg and placebo; however, they did not separate
out the effects in men and women [213]. In another study, in which 28% patients were male, mirabegron
significantly improved patient reported perception of their condition and QoL whether or not patients were
incontinent [214]. A phase IV study, with a small proportion of male subjects, reported addition of mirabegron in
people with persisting urgency despite solifenacin in a Japanese population [215].
In an RCT evaluating add-on therapy with mirabegron for OAB symptoms remaining after treatment
with tamsulosin 0.2 mg daily in men with BPO, combination therapy was associated with greater improvements
in OAB symptom score, in the urinary urgency and daytime frequency part and storage subscore of the IPSS,
and in the QoL index compared to monotherapy with tamsulosin [216]. A prospective analysis of 50 elderly
men showed that mirabegron add-on therapy was effective for patients whose persistent LUTS and OAB
symptoms were not controlled with α1-blocker monotherapy, without causing negative effects on voiding
function [217].
An RCT compared the efficacy of mirabegron 50 mg or fesoterodine 4 mg add-on therapy to
silodosin in LUTS patients with persisting OAB symptoms [218]. At three months, fesoterodine add-on therapy
showed a significantly greater improvement then mirabegron add-on therapy in OAB symptom score-total
(-2.8 vs. -1.5, p = 0.004), IPSS-QoL (-1.5 vs. -1.1, p = 0.04), and OAB symptom score-urgency score (-1.5 vs.
-0.9, p = 0.008). Fesoterodine was also superior in alleviating detrusor overactivity (52.6% vs. 28.9%, p = 0.03).

Tolerability and safety: The most common treatment-related adverse events in the mirabegron groups were
hypertension, UTI, headache and nasopharyngitis [204-207]. Mirabegron is contraindicated in patients with
severe uncontrolled hypertension (systolic blood pressure ≥ 180 mmHg or diastolic blood pressure ≥ 110 mmHg,
or both). Blood pressure should be measured before starting treatment and monitored regularly during treatment.
A combination of thirteen clinical studies including 13,396 patients, 25% of whom were male, showed that
OAB treatments (anticholinergics or mirabegron) were not associated with an increased risk of hypertension or
cardiovascular events compared to placebo [219]. The proportion of patients with dry mouth and constipation
in the mirabegron groups was notably lower than reported in RCTs of other OAB agents or of the active control
tolterodine [204]. Evaluation of urodynamic parameters in men with combined BOO and OAB concluded that
mirabegron did not adversely affect voiding urodynamic parameters compared to placebo in terms of Qmax,
detrusor pressure at maximum flow and bladder contractility index [220]. The overall change in PVR with
mirabegron is small [220].
A small prospective study (mainly focused on males) has shown that mirabegron 25 mg is safe
in patients aged 80 years or more with multiple comorbidities [221]. A pooled analysis of three trials each of
twelve weeks and a one year trial showed, in patients aged > 65 years, a more favourable tolerability profile for
mirabegron than antimuscarinics [222]. In an eighteen week study of 3,527 patients (23% male), the incidence
of adverse events were higher in the combination (solifenacin 5 mg plus mirabegron 25 mg) group (40%) than
the mirabegron 25 mg alone group (32%). Events recorded as urinary retention were low (< 1%), but were
reported slightly more frequently in the combined group when compared with the monotherapy and placebo
groups. The PVR volume was slightly increased in the combined group compared with solifenacin 5 mg, and
the mirabegron monotherapy and placebo groups. Combined therapy with solifenacin 5 mg plus mirabegron 25
mg and solifenacin 5 mg plus mirabegron 50 mg provided improvements in efficacy generally consistent with
an additive effect [223].
In a retrospective analysis of persistence and adherence in 21,996 patients, of whom 30% were
male, the median time to discontinuation was significantly longer for mirabegron (169 days) compared to
tolterodine (56 days) and other antimuscarinics (30-78 days) (p < 0.0001). There was no statistical difference
between men and women [224]. Data on the safety of combination therapy at twelve months are awaited from
the SYNERGY II trial.

Practical considerations: Long-term studies on the efficacy and safety of mirabegron in men of any age with
LUTS are not yet available. Studies on the use of mirabegron in combination with other pharmacotherapeutic
agents for male LUTS are pending. However, pharmacokinetic interaction upon add-on of mirabegron or
tamsulosin to existing tamsulosin or mirabegron therapy does not cause clinically relevant changes in safety
profiles [225]. Available studies on mirabegron in combination with antimuscarinics in OAB patients had a
predominantly female study population, while further trials are still pending.

MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021 21
Summary of evidence LE
Mirabegron improves the symptoms of OAB, including micturition frequency, urgency and UUI. 2
Patients prescribed mirabegron remained on treatment longer than those prescribed antimuscarinics. 3

Recommendation Strength rating


Use beta-3 agonists in men with moderate-to-severe LUTS who mainly have bladder Weak
storage symptoms.

5.2.5 Phosphodiesterase 5 inhibitors


Mechanism of action: Phosphodiesterase 5 inhibitors (PDE5Is) increase intracellular cyclic guanosine
monophosphate, thus reducing smooth muscle tone of the detrusor, prostate and urethra. Nitric oxide and
PDE5Is might also alter reflex pathways in the spinal cord and neurotransmission in the urethra, prostate, or
bladder [226]. Moreover, chronic treatment with PDE5Is seems to increase blood perfusion and oxygenation
in the LUT [227]. Phosphodiesterase 5 inhibitors could also reduce chronic inflammation in the prostate and
bladder [228]. The exact mechanism of PDE5Is on LUTS remains unclear.
Although clinical trials of several selective oral PDE5Is have been conducted in men with LUTS, only
tadalafil (5 mg once daily) has been licensed for the treatment of male LUTS.

Efficacy: Several RCTs have demonstrated that PDE5Is reduce IPSS, storage and voiding LUTS, and improve
QoL. However, Qmax did not significantly differ from placebo in most trials.
A recent Cochrane review included a total of sixteen randomised trials that examined the effects
of PDE5Is compared to placebo and other standard of care drugs (α1-blockers and 5-ARIs) in men with LUTS
[229]. Phosphodiesterase 5 inhibitors led to a small reduction (mean difference (MD) 1.89 lower, 95% CI 2.27
lower to 1.50 lower) in IPSS compared to placebo. There was no difference between PDE5Is and α1-blockers
in IPSS. Most evidence was limited to short-term treatment up to twelve weeks and of moderate or low
certainty. In earlier [230] and more recent [231] meta-analysis, PDE5Is were also found to improve IPSS and
IIEF score, but not Qmax.
Tadalafil 5 mg reduces IPSS by 22-37% and improvement may be seen within a week of initiation
of treatment [232]. A three point or greater total IPSS improvement was observed in 60% of tadalafil treated
men within one week and in 80% within four weeks [233]. The maximum trial (open label) duration was 52
weeks [234]. A subgroup analysis of pooled data from four RCTs demonstrated a significant reduction in LUTS,
regardless of baseline severity, age, previous use of α-blockers or PDE5Is, total testosterone level or predicted
prostate volume [235]. In a recent post hoc analysis of pooled data from four RCTs, tadalafil was shown to also
be effective in men with cardiovascular risk factors/comorbidities, except for patients receiving more than one
antihypertensive medication. The use of diuretics may contribute to patients’ perception of a negated efficacy
[236]. Among sexually active men > 45 years with comorbid LUTS/BPH and ED, tadalafil improved both
conditions [235].
An integrated data analyses from four placebo controlled clinical studies showed that total IPSS
improvement was largely attributed to direct (92.5%, p < 0.001) vs. indirect (7.5%, p = 0.32) treatment effects
via IIEF-EF improvement [237]. Another analysis showed a small but significant increase in Qmax without any
effect on PVR [238]. An integrated analysis of RCTs showed that tadalfil was not superior to placebo for IPSS
improvement at twelve weeks in men ≥ 75 years (with varied effect size between studies), but was for men < 75
years [239]. An open label urodynamic study of 71 patients showed improvements in both voiding and storage
symptoms, confirmed by improvements in BOO index (61.3 to 47.1; p < 0.001), and resolution of DO in 15 (38%)
of 38 patients. Flow rate improved from 7.1 to 9.1 mL/s (p < 0.001) and mean IPSS from 18.2 to 13.4 [240].
A combination of PDE5Is and α-blockers has also been evaluated. A meta-analysis of five RCTs
(two studies with tadalafil 20 mg, two with sildenafil 25 mg, and one with vardenafil 20 mg), showed that
combination therapy significantly improved IPSS score (-1.8), IIEF score (+3.6) and Qmax (+1.5 mL/s) compared
with α-blockers alone [230]. A Cochrane review found similar findings [229]. The effects of tadalafil 5 mg
combined with finasteride 5 mg were assessed in a 26-week placebo-controlled RCT [241]. The combination
of tadalafil and finasteride provided an early improvement in urinary symptoms (p < 0.022 after 4, 12 and 26
weeks), with a significant improvement of storage and voiding symptoms and QoL. Combination therapy
was well tolerated and improved erectile function [241]. However, only tadalafil 5 mg has been licensed in the
context of LUTS management while data on combinations of PDE5Is and other LUTS medications is emerging.

Tolerability and safety: Reported adverse effects in RCTs comparing the effect of all PDE5Is vs. placebo in men
with LUTS include flushing, gastroesophageal reflux, headache, dyspepsia, back pain and nasal congestion
[230]. The discontinuation rate due to adverse effects for tadalafil was 2.0% [242] and did not differ by age,
LUTS severity, testosterone levels, or prostate volume in the pooled data analyses [235].

22 MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021
Phosphodiesterase 5 inhibitors are contraindicated in patients using nitrates, the potassium channel
opener nicorandil, or the α1-blockers doxazosin and terazosin. They are also contraindicated in patients
who have unstable angina pectoris, have had a recent myocardial infarction (< three months) or stroke (< six
months), myocardial insufficiency (New York Heart Association stage > 2), hypotension, poorly controlled blood
pressure, significant hepatic or renal insufficiency, or if anterior ischaemic optic neuropathy with sudden loss of
vision is known or was reported after previous use of PDE5Is.

Practical considerations: To date, only tadalafil 5 mg once daily has been officially licensed for the treatment of
male LUTS with or without ED. The meta-regression suggested that younger men with low body mass index
and more severe LUTS benefit the most from treatment with PDE5Is [230]. Long-term experience with tadalafil
in men with LUTS is limited to one trial with a one year follow-up [234]; therefore, conclusions about its efficacy
or tolerability greater than one year are not possible. There is limited information on reduction of prostate size
and no data on disease progression.

Summary of evidence LE
Phosphodiesterase 5 inhibitors improve IPSS and IIEF score, but not Qmax . 1a
A three point or greater total IPSS improvement was observed in 59.8% of tadalafil treated men within 1b
one week and in 79.3% within four weeks.

Recommendation Strength rating


Use phosphodiesterase type 5 inhibitors in men with moderate-to-severe LUTS with or Strong
without erectile dysfunction.

5.2.6 Plant extracts - phytotherapy


Potential mechanism of action: Herbal drug preparations are made of roots, seeds, pollen, bark, or fruits. There
are single plant preparations (mono-preparations) and preparations combining two or more plants in one pill
(combination preparations) [243].
Possible relevant compounds include phytosterols, ß-sitosterol, fatty acids, and lectins [243]. In
vitro, plant extracts can have anti-inflammatory, anti-androgenic and oestrogenic effects; decrease sexual
hormone binding globulin; inhibit aromatase, lipoxygenase, growth factor-stimulated proliferation of prostatic
cells, α-adrenoceptors, 5 α-reductase, muscarinic cholinoceptors, dihydropyridine receptors and vanilloid
receptors; and neutralise free radicals [243-245]. The in vivo effects of these compounds are uncertain, and the
precise mechanisms of plant extracts remain unclear.

Efficacy: The extracts of the same plant produced by different companies do not necessarily have the same
biological or clinical effects; therefore, the effects of one brand cannot be extrapolated to others [246]. In
addition, batches from the same producer may contain different concentrations of active ingredients [247].
A review of recent extraction techniques and their impact on the composition/biological activity of available
Serenoa repens based products showed that results from different clinical trials must be compared strictly
according to the same validated extraction technique and/or content of active compounds [248], as the
pharmacokinetic properties of the different preparations can vary significantly.
Heterogeneity and a limited regulatory framework characterise the current status of
phytotherapeutic agents. The European Medicines Agency (EMA) has developed the Committee on Herbal
Medicinal Products (HMPC). European Union (EU) herbal monographs contain the HMPC’s scientific opinion on
safety and efficacy data about herbal substances and their preparations intended for medicinal use. The HMPC
evaluates all available information, including non-clinical and clinical data, whilst also documenting long-
standing use and experience in the EU. European Union monographs are divided into two sections: a) Well
established use (marketing authorisation): when an active ingredient of a medicine has been used for more than
ten years and its efficacy and safety have been well established (including a review of the relevant literature);
and b) Traditional use (simplified registration): for herbal medicinal products which do not fulfil the requirements
for a marketing authorisation, but there is sufficient safety data and plausible efficacy on the basis of long-
standing use and experience. Table 1 lists the available EU monographs for herbal medicinal products.

MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021 23
Table 2: European Union monographs for herbal medicinal products

Herbal substance HMPC evaluation Therapeutic Date of monograph


Indication by HMPC
Serenoa repens, fructus (saw Well established use Symptomatic 14/01/2016
palmetto, fruit) treatment of BPH
Extraction solvent: hexane [1]
Serenoa repens, fructus (saw Traditional use LUTS related to BPH* 14/01/2016
palmetto, fruit)
Extraction solvent: ethanol [1]
Cucurbita pepo L, semen (pumpkin Traditional use LUTS related to BPH 25/03/2013
seed) or related to an OAB*
Preparation as defined in the
monograph [249]
Prunus africana (Hook f.) Kalkm., Traditional use LUTS related to BPH* 01/09/2017
cortex (pygeum africanum bark)
Preparation as defined in the
monograph [250]
Urtica dioica L., Urtica urens L., their Traditional use LUTS related to BPH* 05/11/2012
hybrids or their mixtures, radix
Preparation as defined in the
monograph [251]
Epilobium angustifolium L. and/or Traditional use LUTS related to BPH* 13/01/2016
Epilobium parviflorum Schreb., herba
(Willow herb)
Preparation as defined in the
monograph [252]
*After serious conditions have been excluded by a medical doctor.

Panel interpretation: Only hexane extracted Serenoa repens (HESr) has been recommended for well-
established use by the HMPC. Based on this a detailed scoping search covering the timeframe between the
search cut-off date of the EU monograph and April 2020 was conducted for HESr.

A large meta-analysis of 30 RCTs with 5,222 men and follow-up ranging from 4 to 60 weeks, demonstrated
no benefit of treatment with S. repens in comparison to placebo for the relief of LUTS [253]. It was concluded
that S. repens was not superior to placebo, finasteride, or tamsulosin with regard to IPSS improvement, Qmax,
or prostate size reduction; however, the similar improvement in IPSS or Qmax compared with finasteride or
tamsulosin could be interpreted as treatment equivalence. Importantly, in the meta-analysis all different brands
of S. repens were included regardless or not of the presence of HESr as the main ingredient in the extract.
Another SR focused on data from twelve RCTs on the efficacy and safety of HESr [254]. It was
concluded that HESr was superior to placebo in terms of improvement of nocturia and Qmax in patients with
enlarged prostates. Improvement in LUTS was similar to tamsulosin and short-term use of finasteride. An
updated SR analysed fifteen RCTs and also included twelve observational studies. It confirmed the results of
the previous SR on the efficacy of HESr [255]. Compared with placebo, HESr was associated with 0.64 (95%
CI 0.98 to 0.31) fewer voids/night and an additional mean increase in Qmax of 2.75 mL/s (95% CI 0.57 to 4.93),
both were significant. When compared with α-blockers, HESr showed similar improvements in IPSS (WMD
0.57, 95% CI 0.27 to 1.42) and a comparable increase in Qmax when compared to tamsulosin (WMD 0.02, 95%
CI 0.71 to 0.66). Efficacy assessed using IPSS was similar after six months of treatment between HESr and
5-ARIs. Analysis of all available published data for HESr showed a mean significant improvement in IPSS from
baseline of 5.73 points (95% CI 6.91 to 4.54) [255].
A network meta-analysis tried to compare the clinical efficacy of S. repens (HESr and non-HESr)
against placebo and α1-blockers in men with LUTS. Interestingly, only two RCTs on HESr were included in
the analysis. It was found that S. repens achieved no clinically meaningful improvement against placebo or
α1-blockers in short-term follow-up. However, S. repens showed a clinical benefit after a prolonged period of
treatment, and HESr demonstrated a greater improvement than non-HESr in terms of IPSS [256].

With respect to safety and tolerability, data from the SRs showed that HESr had a favourable safety profile with
gastrointestinal disorders being the most frequent adverse effects (mean incidence 3.8%) while HESr had very
limited impact on sexual function.

24 MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021
A cross-sectional study compared the combination of HESr with silodosin to silodosin monotherapy
in patients treated for at least twelve months (mean duration 13.5 months) [257]. It was reported that 69.9%
of the combination therapy patients achieved the predefined clinically meaningful improvement (improvement
more than three points in baseline IPSS) compared to 30.1% of patients treated only with silodosin. In addition,
a greater than 25% improvement in IPSS was found in 68.8% and 31.2% of the patients in the combination
and the monotherapy groups, respectively. These data suggest that combination of a α1-blocker with HESr
may result in greater clinically meaningful improvements in LUTS compared to α1-blocker monotherapy [257].

Practical considerations: Available RCTs do not use the same endpoints (e.g. IPSS). More studies on the use
of HESr in combination with other pharmacotherapeutic agents for male LUTS are pending. There is a need to
define the subpopulation of patients who will benefit most from therapy with HESr.

Summary of evidence LE
HESr improves Qmax and results in fewer voids/night [0.64 (95% CI 0.98 to 0.31)] compared to 2
placebo.
HESr has a very limited negative impact on sexual function. 2

Recommendations Strength rating


Offer hexane extracted Serenoa repens to men with LUTS who want to avoid any potential Weak
adverse events especially related to sexual function.
Inform the patient that the magnitude of efficacy may be modest. Strong

5.2.7 Combination therapies


5.2.7.1 α1-blockers + 5α-reductase inhibitors
Mechanism of action: Combination therapy consists of an α1-blocker (Section 5.2.1) together with a 5-ARI
(Section 5.2.2). The α1-blocker exhibits clinical effects within hours or days, whereas the 5-ARI needs several
months to develop full clinical efficacy. Finasteride has been tested in clinical trials with alfuzosin, terazosin,
doxazosin or terazosin, and dutasteride with tamsulosin.

Efficacy: Several studies have investigated the efficacy of combination therapy against an α1-blocker,
5-ARI or placebo alone. Initial studies with follow-up periods of six to twelve months demonstrated that the
α1-blocker was superior to finasteride in symptom reduction, whereas combination therapy of both agents
was not superior to α1-blocker monotherapy [162, 163, 258]. In studies with a placebo arm, the α1-blocker
was consistently more effective than placebo, but finasteride was not. Data at one year in the MTOPS study
showed similar results [65].
Long-term data (four years) from the MTOPS and CombAT studies showed that combination
treatment is superior to monotherapy for symptoms and Qmax, and superior to α1-blocker alone in reducing the
risk of AUR or need for surgery [65, 147, 148].
The CombAT study demonstrated that combination treatment is superior to either monotherapy
regarding symptoms and flow rate starting from month nine, and superior to α1-blocker for AUR and the
need for surgery after eight months [148]. Thus, the differences in MTOPS may reflect different inclusion and
exclusion criteria and baseline patient characteristics.
Discontinuation of the α1-blocker after six to nine months of combination therapy was investigated
by an RCT and an open-label multicentre trial [259, 260]. The first trial evaluated the combination of tamsulosin
with dutasteride and the impact of tamsulosin discontinuation after six months [259], with almost three quarters
of patients reporting no worsening of symptoms. However, patients with severe symptoms (IPSS > 20) at
baseline may benefit from longer combination therapy.
A more recent trial evaluated the symptomatic outcome of finasteride monotherapy at three and
nine months after discontinuation of nine-month combination therapy [260]. Lower urinary tract symptom
improvement after combination therapy was sustained at three months (IPSS difference 1.24) and nine months
(IPSS difference 0.4). The limitations of the studies include the short duration of the studies and the short
follow-up period after discontinuation.

In both the MTOPS and CombAT studies, combination therapy was superior to monotherapy in preventing
clinical progression as defined by an IPSS increase of at least four points, AUR, UTI, incontinence, or an
increase in creatinine > 50%. The MTOPS study found that the risk of long-term clinical progression (primarily
due to increasing IPSS) was reduced by 66% with combined therapy vs. placebo and to a greater extent than
with either finasteride or doxazosin monotherapy (34% and 39%, respectively) [65]. In addition, finasteride
(alone or in combination), but not doxazosin alone, significantly reduced both the risks of AUR and the need

MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021 25
for BPO-related surgery over the four-year study. In the CombAT study, combination therapy reduced the
relative risks of AUR by 68%, BPO-related surgery by 71%, and symptom deterioration by 41% compared with
tamsulosin, after four years [261]. To prevent one case of urinary retention and/or surgical treatment thirteen
patients need to be treated for four years with dutasteride and tamsulosin combination therapy compared to
tamsulosin monotherapy while the absolute risk reduction (risk difference) was 7.7%.

The CONDUCT study compared efficacy and safety of a fixed-dose combination of dutasteride and tamsulosin
to a WW approach with the potential initiation of tamsulosin (step-up approach) in a two year RCT with a total
of 742 patients. In both arms detailed lifestyle advice was given. This fixed-dose combination resulted in a
rapid and sustained improvement in men with moderate LUTS at risk of disease progression, the difference
in IPSS at 24 months was 5.4 in the active arm and 3.6 in the placebo arm (p < 0.001) [262]. Furthermore,
tamsulosin plus dutasteride significantly reduced the relative risk of clinical progression (mainly characterised
as a worsening in symptoms) by 43.1% when compared with WW, with an absolute risk reduction of 11.3%
(number needed to treat [NNT] = 9).
The influence of baseline variables on changes in IPSS after combination therapy with dutasteride
plus tamsulosin or either monotherapy was tested based on the four year results of the CombAT study.
Combination therapy provided consistent improvement of LUTS over tamsulosin across all analysed baseline
variables at 48 months [263].
More recently, a combination of the 5-ARI finasteride and tadalafil 5 mg was tested in a large scale
RCT against finasteride monotherapy. This study supports the concept of this novel combination therapy and is
described in more detail in the chapter on PDE5Is [241].

Tolerability and safety: Adverse events for both drug classes have been reported with combination treatment
[65, 147, 148]. The adverse events observed during combination treatment were typical of α1-blockers and
5-ARIs. The frequency of adverse events was significantly higher for combination therapy. The MTOPS study
demonstrated that the incidence of treatment related adverse events is higher during the first year of combined
treatment between doxazosin and finasteride [264]. A meta-analysis measuring the impact of medical
treatments for LUTS/BPH on ejaculatory function, reported that combination therapy with α1-blockers and
5-ARIs resulted in a three-fold increased risk of EjD as compared with each of the monotherapies [157].

Practical considerations: Compared with α1-blockers or 5-ARI monotherapy, combination therapy results in
a greater improvement in LUTS and increase in Qmax and is superior in prevention of disease progression.
However, combination therapy is also associated with a higher rate of adverse events. Combination therapy
should therefore be prescribed primarily in men who have moderate-to-severe LUTS and are at risk of disease
progression (higher prostate volume, higher PSA concentration, advanced age, higher PVR, lower Qmax, etc.).
Combination therapy should only be used when long-term treatment (more than twelve months) is intended
and patients should be informed about this. Discontinuation of the α1-blocker after six months might be
considered in men with moderate LUTS.

Summary of evidence LE
Long-term data (four years) from the MTOPS and CombAT studies showed that combination treatment 1b
is superior to monotherapy for symptoms and Qmax, and superior to α1-blocker alone in reducing the
risk of AUR or need for surgery.
The MTOPS study found that the risk of long-term clinical progression (primarily due to increasing 1b
IPSS) was reduced by 66% with combined therapy vs. placebo and to a greater extent than with either
finasteride or doxazosin monotherapy.
The CombAT study found that combination therapy reduced the relative risks of AUR by 68%, BPH- 1b
related surgery by 71%, and symptom deterioration by 41% compared with tamsulosin, after four years.
Adverse events of both drug classes are seen with combined treatment using α1-blockers and 5-ARIs. 1b

Recommendation Strength rating


Offer combination treatment with an α1-blocker and a 5α-reductase inhibitor to men with Strong
moderate-to-severe LUTS and an increased risk of disease progression (e.g. prostate
volume > 40 mL).

5.2.7.2 α1-blockers + muscarinic receptor antagonists


Mechanism of action: Combination treatment consists of an α1-blocker together with an antimuscarinic aiming
to antagonise both α1-adrenoceptors and muscarinic receptors. The possible combinations have not all been
tested in clinical trials yet.

26 MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021
Efficacy: Several RCTs and prospective studies investigated combination therapy, lasting four to twelve weeks,
either as an initial treatment in men with OAB and presumed BPO or as a sequential treatment for storage
symptoms persisting while on an α1-blocker [197, 261, 265-271]. One trial used the α1-blocker naftopidil (not
registered in most European countries) with and without antimuscarinics [272]. A high proportion of men with
voiding and storage LUTS need to add anticholinergics after α1-blocker monotherapy, particularly those with
longer duration of symptoms at presentation, and men with storage symptoms and a small prostate volume
[273].
Combination treatment is more efficacious in reducing urgency, UUI, voiding frequency, nocturia,
or IPSS compared with α1-blockers or placebo alone, and improves QoL [197, 274]. Symptom improvement
is higher regardless of PSA concentration with combination therapy, whereas tolterodine alone improved
symptoms mainly in men with a serum PSA of < 1.3 ng/mL [200].
Persistent LUTS during α1-blocker treatment can be reduced by the additional use of an
antimuscarinic, [261, 265, 271, 275, 276]. Two SRs of the efficacy and safety of antimuscarinics in men
suggested that combination treatment provides significant benefit [277, 278]. In a meta-analysis of sixteen
studies with 3,548 patients with BPH/OAB, initial combination treatment of an α1-blocker with anticholinergic
medication improvement storage symptoms and QoL compared to α1-blocker monotherapy without causing
significant deterioration of voiding function [279]. There was no difference in total IPSS and Qmax between the
two groups.
Effectiveness of therapy is evident primarily in those men with moderate-to-severe storage LUTS
[280]. Long term use of combination therapy has been reported in patients receiving treatment for up to one
year, showing symptomatic response is maintained, with a low incidence of AUR [281]. In men with moderate-
to-severe storage symptoms, voiding symptoms and PVR < 150 mL, the reduction in symptoms using
combination therapy is associated with patient-relevant improvements in health related quality of life compared
with placebo and α1-blocker monotherapy [282].

Tolerability and safety: Adverse events of both drug classes are seen with combined treatment using
α1-blockers and antimuscarinics. The most common side-effect is dry mouth. Some side-effects (e.g. dry
mouth or ejaculation failure) may show increased incidence which cannot simply be explained by summing
the incidence with the drugs used separately. Increased PVR may be seen, but is usually not clinically
significant, and risk of AUR is low up to one year of treatment [277, 283]. Antimuscarinics do not cause evident
deterioration in maximum flow rate used in conjunction with an α1-blocker in men with OAB symptoms [274,
284].
A recent RCT investigated safety in terms of maximum detrusor pressure and Qmax for solifenacin
(6 mg or 9 mg) with tamsulosin in men with LUTS and BOO compared with placebo [285]. The combination
therapy was not inferior to placebo for the primary urodynamic variables; Qmax was increased vs. placebo [285].

Practical considerations: Class effects are likely to underlie efficacy and QoL using an α1-blocker and
antimuscarinic. Trials used mainly storage symptom endpoints, were of short duration, and included only men
with low PVR volumes at baseline. Therefore, measuring PVR is recommended during combination treatment.

Summary of evidence LE
Combination treatment with α1-blockers and antimuscarinics is effective for improving LUTS-related 2
QoL impairment.
Combination treatment with α1-blockers and antimuscarinics is more effective for reducing urgency, 2
UUI, voiding frequency, nocturia, or IPSS compared with α1-blockers or placebo alone.
Adverse events of both drug classes are seen with combined treatment using α1-blockers and 1
antimuscarinics.
There is a low risk of AUR using α1-blockers and antimuscarinics in men known to have a PVR urine 2
volume of < 150 mL.

Recommendations Strength rating


Use combination treatment of a α1-blocker with a muscarinic receptor antagonist in Strong
patients with moderate-to-severe LUTS if relief of storage symptoms has been insufficient
with monotherapy with either drug.
Do not prescribe combination treatment in men with a post-void residual volume > 150 mL. Weak

Note: A
 ll patients should be counselled about pharmacological treatment related adverse events in
order to select the most appropriate treatment for each individual patient.

MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021 27
5.3 Surgical treatment
Surgical treatment is one of the cornerstones of LUTS/BPO management. Based on its ubiquitous availability,
as well as its efficacy, M-TURP has long been considered as the reference technique for the surgical
management of LUTS/BPO. However, in recent years various techniques have been developed with the aim
of providing a safe and effective alternative to M-TURP. Previously, the surgical section of the Guidelines was
based on technology rather than surgical approach; additionally, most of the studies are restricted by prostate
size, which is also reflected in the present Guidelines. Notably, only a small fraction of RCTs are performed
in patients with a prostate volume > 80 mL; therefore, high-level evidence for larger prostates is limited. As
the clinical reality is primarily reflected by surgical approach and not necessarily by a specific technology, the
chapter on surgical management has been restructured. It is now divided into the following five sections:

1. Resection;
2. Enucleation;
3. Vaporisation;
4. Alternative ablative techniques; and
5. Non-ablative techniques.

Based on Panel consensus, timeframes defining short-, mid- and long-term follow-up of patients submitted to
surgical treatments are 12, 36 and over 36 months, respectively. The durability of a technique is reflected by the
re-operation rate during follow-up, the failure to wean patients off medication as well as the initiation of novel
LUTS medication after surgery. However, for the majority of techniques only the re-operation rate is reported
and clinicians should inform patients that long-term surgical RCTs are often lacking. Some patients value
sexual function and perceived higher safety than maximum efficacy; therefore, it is not surprising that some
patients consciously choose an alternative ablative or non-ablative technique despite the knowledge that it
might not be their definitive treatment. In contrast, many urologists are critical about these ‘lesser’ procedures
due to their inferior relief of BOO.
Recommendations on new devices or interventions will only be included in the Guidelines once
supported by a minimum level of evidence. To clarify this the Panel have published their position on certainty
of evidence (CoE) [286]. In summary, a device or technology is only included once supported by RCTs looking
at both efficacy and safety, with adequate follow-up, and secondary studies to confirm the reproducibility
and generalisability of the first pivotal studies [286]. Otherwise, there is a danger that a single pivotal study
can be overexploited by device manufacturers. Studies that are needed include (1) proof of concept, (2) RCTs
on efficacy and safety, as well as (3) cohort studies with a broad range of inclusion and exclusion criteria
to confirm both reproducibility and generalisability of the benefits and harms [286]. The panel assesses the
quality of all RCTs and if they do not meet the standard required the intervention will continue to have no
recommendation i.e. a RCT does not guarantee inclusion in the Guidelines.
In addition, the Guidelines continues to include techniques under investigation. These are devices
or technologies that have shown promising results in initial studies; however, they do not meet the
aforementioned criteria yet to provide a CoE which allows the Panel to regard these devices or technologies
as recommended alternatives. To account for evolving evidence, recommendations for some techniques under
investigation have been made; however, these techniques remain under investigation until further studies
provide the recommended CoE.

5.3.1 Resection of the prostate


5.3.1.1 Monopolar and bipolar transurethral resection of the prostate
Mechanism of action: Transurethral resection of the prostate (TURP) is performed using two techniques:
monopolar TURP (M-TURP) and bipolar TURP (B-TURP). Monopolar transurethral resection of the prostate
removes tissue from the transition zone of the gland. Bipolar TURP addresses a major limitation of M-TURP by
allowing performance in normal saline. Prostatic tissue removal is identical to M-TURP. Contrary to M-TURP,
in B-TURP systems, the energy does not travel through the body to reach a skin pad. Bipolar circuitry is
completed locally; energy is confined between an active (resection loop) and a passive pole situated on the
resectoscope tip (“true” bipolar systems) or the sheath (“quasi” bipolar systems). The various bipolar devices
available differ in the way in which current flow is delivered [287, 288].

Efficacy: In a meta-analysis of 20 RCTs with a maximum follow-up of five years, M-TURP resulted in a
substantial mean Qmax improvement (+162%), a significant reduction in IPSS (-70%), QoL score (-69%),
and PVR (-77%) [289]. Monopolar-TURP delivers durable outcomes as shown by studies with a follow-up of
8-22 years. There are no similar data on durability for any other surgical treatment for BPO [290]. One study
with a mean follow-up of thirteen years reported a significant and sustained decrease in most symptoms
and improvement in urodynamic parameters. Failures were associated with DUA rather than re-development

28 MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021
of BPO [102]. A second prostatic operation, usually re-TURP, has been reported at a constant annual rate of
approximately 1-2%. A review analysing 29 RCTs found a retreatment rate of 2.6% after a mean follow-up of
sixteen months [291]. Data from an Austrian nationwide study of two cohorts totalling 41,059 men submitted to
M-TURP showed that the overall retreatment rates (re-TURP, urethrotomy and bladder neck incision) remained
unchanged during the last decade (0.9%, 3.7%, 9.5% and 12.7% at three months, one year, five years, and
eight years, respectively), and that the respective incidence of re-TURP was 0.8%, 2.4%, 6.1% and 8.3%,
respectively [292, 293].

Bipolar TURP is the most widely and thoroughly investigated alternative to M-TURP. Results from 56 RCTs
have been reported [294], of which around half have been pooled in RCT-based meta-analyses [289, 295-
299]. Early pooled results concluded that no clinically relevant differences exist in short-term efficacy (IPSS,
QoL score and Qmax) [296]. Subsequent meta-analyses supported these conclusions though trial quality was
generally poor [289, 295, 297-299]. Data from RCTs with mid- to long-term follow-up (up to 60 months) showed
no differences in efficacy parameters [300-308].
A meta-analysis was conducted to evaluate the quasi-bipolar transurethral resection in saline
(TURis, Olympus Medical) system vs. M-TURP. Ten unique RCTs (1,870 patients) were included and it was
concluded that TURis was of equivalent efficacy to M-TURP [309].

Tolerability and safety: Peri-operative mortality and morbidity of M-TURP have decreased over time, but
morbidity remains considerable (0.1% and 11.1%, respectively) [310]. Data from an Austrian nationwide study
of two cohorts totalling 41,059 men submitted to M-TURP showed a 20% reduction in mortality rate over time,
to 0.1% at 30 days and 0.5% at 90 days [292, 293].
The risk of TUR-syndrome decreased to < 1.1% [291, 311]. Data from 10,654 M-TURPs reported
bleeding requiring transfusion in 2.9% [310]. Short- to mid-term complications reported in an analysis of RCTs
using M-TURP as a comparator were: bleeding requiring transfusion 2% (0-9%), TUR-syndrome 0.8% (0-5%),
AUR 4.5% (0-13.3%), clot retention 4.9% (0-39%), and UTI 4.1% (0-22%) [289]. Long-term complications of
M-TURP comprise urinary incontinence, urinary retention and UTIs, bladder neck contracture (BNC), urethral
stricture, retrograde ejaculation and ED [291].

Early pooled results concluded that no differences exist in short-term urethral stricture/BNC rates, but B-TURP
is preferable to M-TURP due to a more favourable peri-operative safety profile (elimination of TUR-syndrome;
lower clot retention/blood transfusion rates; shorter irrigation, catheterisation, and possibly hospitalisation
times) [296]. Subsequent meta-analyses supported these conclusions [289, 295, 297-299]; however, trial
quality was relatively poor and limited follow-up might cause under-reporting of late complications, such as
urethral stricture/BNC [296]. An RCT based meta-analysis has shown that TURis reduces the risk of TUR-
syndrome and the need for blood transfusion compared to M-TURP [299]. It was concluded that TURis is
associated with improved peri-operative safety, eliminating the risk of TUR syndrome, reducing the risk of blood
transfusion/clot retention and hospital stay. No significant difference was detected in urethral stricture rates.
Data from the vast majority of individual RCTs with mid- to long-term follow-up (up to 60 months),
showed no differences between M-TURP and B-TURP in urethral stricture/BNC rates [300-308], in accordance
with all published meta-analyses. However, two individual RCTs have shown opposing results [307, 312]. A
significantly higher stricture (urethral stricture + BNC) rate was detected in the B-TURP arm performed with
a “quasi” bipolar system (TURis, Olympus Medical) in patients with a prostate volume > 70 mL at 36 months
follow-up [307]. In addition, a significantly higher BNC, but not urethral stricture, rate was detected in the
B-TURP arm performed with a “true” bipolar system (Gyrus PK SuperPulse, Olympus Medical) in 137 patients
followed up to twelve months [312].
Randomised controlled trials using the erectile function domain of the IIEF (IIEF-ED) and the
ejaculatory domain of the male sexual-health questionnaire (Ej-MSHQ) showed that M-TURP and B-TURP
have a similar effect on erectile and ejaculatory function [313, 314]. Comparative evaluations of the effects on
overall sexual function, quantified with IIEF-15, showed no differences between B-TURP and M-TURP at twelve
months follow-up (erection, orgasmic function, sexual desire, intercourse satisfaction, overall satisfaction) [314,
315].

Practical considerations: Monopolar-TURP is an effective treatment for moderate-to-severe LUTS secondary to


BPO. The choice should be based primarily on prostate volume (30-80 mL suitable for M-TURP). No studies on
the optimal cut-off value exist, but the complication rates increase with prostate size [310]. The upper limit for
M-TURP is suggested as 80 mL (based on Panel expert opinion, under the assumption that this limit depends
on the surgeon’s experience, choice of resectoscope size and resection speed), as surgical duration increases,
there is a significant increase in the rate of complications and the procedure is safest when performed in under
90 minutes [316].

MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021 29
Bipolar TURP in patients with moderate-to-severe LUTS secondary to BPO, has similar efficacy to
M-TURP, but lower peri-operative morbidity. The duration of improvements with B-TURP were documented in
a number of RCTs with mid-term follow-up. Long-term results (up to five years) for B-TURP showed that safety
and efficacy are comparable to M-TURP [300-308]. The choice of B-TURP should be based on equipment
availability, surgeon’s experience, and patient’s preference.

Summary of evidence LE
Bipolar- or monopolar-TURP is the current standard surgical procedure for men with prostate sizes of 1a
30-80 mL and bothersome moderate-to-severe LUTS secondary of BPO.
Bipolar-TURP achieves short-, mid- and long-term results comparable with M-TURP, but B-TURP has 1a
a more favourable peri-operative safety profile.

Recommendation Strength rating


Offer bipolar- or monopolar-transurethral resection of the prostate to surgically treat Strong
moderate-to-severe LUTS in men with prostate size of 30-80 mL.

5.3.1.2 Holmium laser resection of the prostate


With the advent of holmium laser enucleation of the prostate (section 5.3.2.3) and the fact that no relevant
publications on holmium laser resection of the prostate (HoLRP) have been published since 2004, HoLRP of
the prostate does not play a role in contemporary treatment algorithms.

5.3.1.3 Thulium:yttrium-aluminium-garnet laser (Tm:YAG) vaporesection of the prostate


Mechanism of action: In the thulium lasers, a wavelength between 1,940 nm (thulium fiber laser) and 2,013 nm
(Tm:YAG) is emitted in continuous wave mode. The laser is primarily used in front-fire applications [317].
Different applications such as vaporesection (ThuVARP) have been published [318].

Efficacy: Several meta-analyses with pooled data from both RCTs (generally low quality), and non-RCTs have
evaluated ThuVARP vs. M-TURP [319-321], and B-TURP [322-324]. The largest meta-analyses included nine
RCTs and seven non-RCTs and reported no clinically relevant differences in efficacy (IPSS, QoL score and
Qmax) between ThuVARP and M-TURP or B-TURP at twelve months [323]. Data from one RCT with long-term
follow-up showed no difference in efficacy and re-operation rates between ThuVARP and M-TURP (2.1%
vs. 4.1%, respectively) [325]. A prospective multicentre study on ThuVARP, including 2,216 patients, showed
durable post-operative improvement in IPSS, QoL, Qmax, and PVR for the entire eight years of follow-up [326].

Tolerability and safety: In a number of meta-analyses longer operation times, shorter catheterisation/
hospitalisation times and less blood loss without significant differences in transfusion rates or in any other
short-term complication rates have been reported for ThuVARP compared to TURP [319-324]. A significantly
higher transfusion rate was reported after M-TURP in two meta-analyses [321, 323]. However, overall RCT
quality was relatively low with limited follow-up potentially accounting for under-reporting of late complications,
such as urethral stricture/BNC [323]. Data from three RCTs with mid- to long-term follow-up (18 to 48 months)
showed no differences in late complication rates between ThuVARP and TURP (BNC: 0.0%-2.1% vs. 0.0%-
4.1%; stricture: 0.0%-2.2% vs. 0.0%-2.2%, respectively) [325, 327, 328].
Haemoglobin drop was significantly higher in the bridging group in a retrospectively analysed case
series of 103 patients who underwent ThuVARP and received either low molecular weight heparin bridging or
continued antiplatelet/anticoagulant therapy [329].

Practical considerations: As a limited number of RCTs with mid- to long-term follow-up support the efficacy of
ThuVARP, there is a need for ongoing investigation of the technique.

Summary of evidence LE
Laser vaporesection of the prostate using Tm:YAG laser (ThuVARP) has longer operation times and 1a
shorter catheterisation/hospitalisation times compared to TURP with no clinically relevant differences
in short-term efficacy and safety. Mid- to long-term results on efficacy and safety compared to TURP
are very limited.

Recommendation Strength rating


Offer laser resection of the prostate using Tm:YAG laser (ThuVARP) as an alternative to Weak
TURP.

30 MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021
5.3.1.4 Transurethral incision of the prostate
Mechanism of action: Transurethral incision of the prostate (TUIP) involves incising the bladder outlet without
tissue removal. Transurethral incision of the prostate is conventionally performed with Collins knife using
monopolar electrocautery; however, alternative energy sources such as holmium laser may be used [330]. This
technique may replace M-TURP in selected cases, especially in prostate sizes < 30 mL without a middle lobe.

Efficacy: An RCT comparing conventional TUIP vs. TUIP using holmium laser in prostates ≤ 30 mL with a
follow-up of twelve months, found both procedures to be equally effective in relieving BOO with similarly
low re-operation rates [330]. A meta-analysis of ten RCTs found similar LUTS improvements and lower but
significant improvements in Qmax for TUIP [331]. In this meta-analysis, an upper limit of prostate size was
reported as an entry criterion for eight studies with five < 30 mL and three < 60 mL. A meta-analysis of six trials
showed that re-operation was more common after TUIP (18.4%) than after M-TURP (7.2%) [331].

Tolerability and safety: An RCT comparing conventional TUIP vs. TUIP using holmium laser reported both
procedures to be safe with low complication rates; however, the operation time and retrograde ejaculation rate
was significantly lower in the conventional TUIP arm [330]. No cases of TUR-syndrome have been recorded
after TUIP. The risk of bleeding after TUIP is small [331].

Practical considerations: Transurethral incision of the prostate is an effective treatment for moderate-to-severe
LUTS secondary to BPO. The choice between M-TURP and TUIP should be based primarily on prostate
volume (30 mL TUIP) [331].

Summary of evidence LE
Transurethral incision of the prostate shows similar efficacy and safety to M-TURP for treating 1a
moderate-to-severe LUTS secondary to BPO in men with prostates < 30 mL.
No case of TUR-syndrome has been recorded, the risk of bleeding requiring transfusion is negligible 1a
and retrograde ejaculation rate is significantly lower after TUIP, but the re-operation rate is higher
compared to M-TURP.
The choice between TUIP and TURP should be based primarily on prostate volume (< 30 mL and 4
30-80 mL suitable for TUIP and TURP, respectively).

Recommendation Strength rating


Offer transurethral incision of the prostate to surgically treat moderate-to-severe LUTS in Strong
men with prostate size < 30 mL, without a middle lobe.

5.3.2 Enucleation of the prostate


5.3.2.1 Open prostatectomy
Mechanism of action: Open prostatectomy is the oldest surgical treatment for moderate-to-severe LUTS
secondary to BPO. Obstructive adenomas are enucleated using the index finger, approaching from within
the bladder (Freyer procedure) or through the anterior prostatic capsule (Millin procedure). It is used for
substantially enlarged glands (> 80-100 mL).

Efficacy: Open prostatectomy reduces LUTS by 63-86% (12.5-23.3 IPSS points), improves QoL score by
60-87%, increases mean Qmax by up to 375% (+16.5-20.2 mL/s), and reduces PVR by 86-98%. Efficacy is
maintained for up to six years [332-337]. Data from an Austrian nationwide study of 1,286 men submitted to
OP showed that the endourological re-intervention rates after primary OP were 0.9%, 3.0%, 6.0%, and 8.8%,
at three months, one year, five years, and eight years, respectively. The respective incidence of re-TURP was
0.5%, 1.8%, 3.7% and 4.3%, respectively [9].
Two meta-analyses [338, 339] evaluated the overall efficacy of OP performed via a transvesical
approach vs. two transurethral enucleation techniques for treating patients with large glands, namely bipolar
transurethral enucleation of the prostate (B-TUEP) and holmium laser enucleation of the prostate (HoLEP).
The larger study included nine RCTs involving 758 patients [339]. Five RCTs compared OP with B-TUEP [337,
340-343] and four RCTs compared OP with HoLEP [332, 333, 344, 345]. At three, six, twelve and 24-months
follow-up there were no significant differences in Qmax [339]. Post-void residual, PSA, IPSS and QoL score
showed no significant differences at one, three, six and twelve months follow-up [339]. Randomised controlled
trials indicate that OP is as effective as HoLEP for improving micturition in large prostates [332, 333], with
similar improvement regarding Qmax, IPSS score and re-operation rates after five years [332].

MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021 31
Tolerability and safety: Open prostatectomy mortality has decreased significantly during the past two decades
(< 0.25%) [336]. Data from an Austrian nationwide study totalling 1,286 men submitted to OP showed mortality
rates of 0.2% at 30 days and 0.4% at 90 days [293]. The estimated transfusion rate is about 7-14% [332,
335, 336, 338]. Long-term complications include transient urinary incontinence (up to 10%), BNC and urethral
stricture (about 6%) [332-334, 338, 346].
Two meta-analyses evaluated the overall safety of OP performed via a transvesical approach vs.
B-TUEP and HoLEP [338, 339]. Operation time did not differ significantly between OP and B-TUEP, but was
significantly shorter for OP compared to HoLEP. Catheterisation and hospitalisation time were significantly
longer for OP, which was also associated with more blood transfusions. There were no significant differences
regarding other complications. There was no significant difference in IIEF-5 at three, six, twelve and 24-months
follow-up.

Practical considerations: Open prostatectomy is the most invasive surgical method, but it is an effective
and durable procedure for the treatment of LUTS/BPO. In the absence of an endourological armamentarium
including a holmium laser or a bipolar system and with appropriate patient consent, OP is a reasonable surgical
treatment of choice for men with prostates > 80 mL.

Summary of evidence LE
Open prostatectomy is an effective and durable procedure for the treatment of LUTS/BPO, but it is the 1b
most invasive surgical method.
Open prostatectomy shows similar short- and mid-term efficacy to B-TUEP and HoLEP for treating 1a
moderate-to-severe LUTS secondary to BPO in patients with large prostates.
Open prostatectomy has a less favourable peri-operative safety profile compared to B-TUEP and 1a
HoLEP.
The long-term functional results of OP are comparable to HoLEP. 1b

Recommendation Strength rating


Offer open prostatectomy in the absence of bipolar transurethral enucleation of the prostate Strong
and holmium laser enucleation of the prostate to treat moderate-to-severe LUTS in men
with prostate size > 80 mL.

5.3.2.2 Bipolar transurethral enucleation of the prostate (B-TUEP)


Mechanism of action: Following the principles of bipolar technology (section 5.3.1.1), the obstructive adenoma
is enucleated endoscopically by the transurethral approach. Bipolar transurethral enucleation evolved from
plasmakinetic (PK) B-TURP and was introduced by Gyrus ACMI. The technique, also referred to as PK
enucleation of the prostate (PKEP), utilises a bipolar high-frequency generator and a variety of detaching
instruments, for this true bipolar system, including a point source in the form of a axipolar cystoscope
electrode suitable for enucleation [347] or a resectoscope tip/resection loop [348, 349]. More recently, a novel
form of B-TUEP has been described, bipolar plasma enucleation of the prostate (BPEP), stemming from
B-TURP (TURis, Olympus Medical), that utilises a bipolar high frequency generator and a variety of detaching
instruments including a mushroom- or button-like vapo-electrode [343, 350] and a Plasmasect enucleation
electrode [351] for this quasi-bipolar system. Bipolar transurethral enucleation of the prostate is followed by
either morcellation [343, 347] or resection [348-350, 352-354] of the enucleated adenoma.

Efficacy: One RCT evaluating PKEP vs. M-TURP in 204 patients with mean prostate volume < 80 mL reported
a significant improvement in IPSS, QoL, and Qmax, with urodynamically proven de-obstruction favouring PKEP
at 36 months follow-up [349]. The RCT concluded that the mid-term clinical efficacy of PKEP was comparable
to M-TURP [349]. A meta-analysis evaluating data from five RCTs including 666 patients submitted to B-TUEP
(PKEP or BPEP) vs. B-TURP, reported similar efficacy at twelve months in terms of IPSS, QoL and Qmax
[355]. Two RCTs evaluated the mid-term efficacy of PKEP vs. B-TURP at 36 months [348, 353] and one RCT
evaluated long-term efficacy at 60 months [354]. Efficacy was significantly better for PKEP in patients with large
prostates at 36, 48 and 60 months [348, 354]. Comparative data on efficacy for B-TUEP vs. OP and the various
forms of laser enucleation are presented in section 5.3.2.1 to 5.3.2.5, respectively.

Tolerability and safety: An RCT evaluating PKEP vs. M-TURP in patients with mean prostate volume < 80 mL
and 36 month follow-up reported that PKEP is superior to M-TURP in terms of haemoglobin drop, irrigation,
catheterisation and hospitalisation time [349]. No significant differences between the arms were reported in
operation time, blood transfusion rates, sexual function or any other reported complications (TUR-syndrome,
clot retention, incontinence, retrograde ejaculation, urethral structures/BNC) [349]. A meta-analysis including

32 MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021
five RCTs evaluating B-TUEP vs. B-TURP reported: similar operation, catheterisation and hospitalisation
times; lower acute urine retention rates; significantly reduced haemoglobin drop and blood transfusion rates;
no difference in erectile function; and no difference in all other reported complication rates including urethral
stricture/BNC rates for B-TUEP at 24 months follow-up [355]. No difference in urethral stricture/BNC rates was
reported at 60 months follow-up [354]. Comparative data on efficacy for B-TUEP vs. OP and the various forms
of laser enucleation are presented in section 5.3.2.1 to 5.3.2.5, respectively.

Summary of evidence LE
Bipolar transurethral (plasmakinetic) enucleation of the prostate shows favourable mid- to long-term 1b
efficacy compared to TURP.
Bipolar transurethral (plasmakinetic) enucleation of the prostate has a favourable peri-operative safety 1b
profile and demonstrates similar mid- to long-term safety compared to TURP.

Recommendation Strength rating


Offer bipolar transurethral (plasmakinetic) enucleation of the prostate to men with Weak
moderate-to-severe LUTS as an alternative to transurethral resection of the prostate.

5.3.2.3 Holmium laser enucleation of the prostate


Mechanism of action: The holmium:yttrium-aluminium garnet (Ho:YAG) laser (wavelength 2,140 nm) is a pulsed
solid-state laser that is absorbed by water and water-containing tissues. Tissue coagulation and necrosis are
limited to 3-4 mm, which is enough to obtain adequate haemostasis [356].

Efficacy: An initial meta-analysis reported no significant differences in short-term efficacy (Qmax) and
re-intervention rates (4.3% vs. 8.8%) between HoLEP and M-TURP [357]; however, subsequent meta-analyses
reported favourable short-term efficacy (Qmax and IPSS) for HoLEP [289, 320, 355, 358]. Two meta-analyses
evaluating HoLEP vs. B-TURP showed no significant differences in short-term efficacy (IPSS, QoL score and
Qmax) [355, 359]. One RCT comparing HoLEP with M-TURP in a small number of patients with mean prostate
volume < 80 mL and a seven year follow-up found that the functional long term results were comparable
[360]. Another RCT comparing HoLEP with B-TURP in patients with prostate volume < 80 mL reported no
significant difference in IPSS, QoL score and Qmax at 24 months [361]. Long-term (72 months) improvement in
IPSS and Qmax was better for HoLEP, but there were no clinically relevant differences between the arms [362].
Comparative efficacy data for HoLEP vs. OP is presented in section 5.3.2.1. One small RCT evaluating HoLEP
vs. PKEP in patients with mean prostate volume < 80 mL reported similar improvements in IPSS and Qmax as
well as similar re-operation rates at twelve months follow-up [347].

Tolerability and safety: Meta-analyses found that HoLEP has longer operation times, shorter catheterisation and
hospitalisation times, reduced blood loss, fewer blood transfusions and no significant differences in urethral
strictures (2.6% vs. 4.4%) and stress urinary incontinence (1.5% vs. 1.5%) rates compared to M-TURP [320,
355, 357, 358, 363]. Two meta-analyses evaluated HoLEP vs. B-TURP [355, 359]. Zhang et al., reported longer
operation times for HoLEP, but no significant differences in hospitalisation time or complication rates whilst
Qian et al., reported no significant differences in operation and catheterisation times or short-term complication
rates [355, 359]. A RCT comparing HoLEP with B-TURP in patients with prostate volume < 80 mL reported
longer operation time, shorter catheterisation and hospitalisation times and a lower risk for haemorrhage for
HoLEP with no significant differences in blood transfusion rates or other complication rates at 24 months
[361]. Comparative data on safety of HoLEP vs. OP are presented in section 5.3.2.1. One small RCT evaluating
HoLEP vs. PKEP in patients with mean prostate volume < 80 mL reported significantly shorter operation
times for HoLEP, but similar catheterisation and hospitalisation times and complication rates at twelve months
follow-up [347].
Holmium laser enucleation of the prostate has been safely performed in patients using anticoagulant
and/or antiplatelet medications [364, 365]. However, current limitations include: a lack of RCTs; limited data
on short- and mid-term complications and bridging therapy; data presentation does not allow for separate
interpretation of either of the two substantially different topics of antiplatelet and anticoagulant therapy.
The impact on erectile function and retrograde ejaculation is comparable between HoLEP and TURP
[366, 367]. Erectile function did not decrease from baseline in either group; three quarters of sexually active
patients had retrograde ejaculation after HoLEP. Data have shown that ejaculation and orgasm perception are
the two most impacted domains after HoLEP [368]. Attempts to maintain ejaculatory function with HoLEP have
been reported to be successful in up to 46.2% of patients [369].

MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021 33
Practical considerations: Holmium laser enucleation of the prostate requires experience and relevant
endoscopic skills. The experience of the surgeon is the most important factor affecting the overall occurrence
of complications [370, 371]. Mentorship programmes are advised to improve surgical performance from both
an institutional and personal learning curve perspective [372-374].

Summary of evidence LE
Laser enucleation of the prostate using Ho:YAG laser (HoLEP) demonstrates similar mid- to long-term 1b
efficacy when compared to TURP.
Laser enucleation of the prostate using Ho:YAG laser (HoLEP) demonstrates similar short-term safety 1a
when compared to TURP.
Laser enucleation of the prostate using Ho:YAG laser (HoLEP) demonstrates longer operation times, 1a
but a more favourable peri-operative profile when compared to TURP.

Recommendation Strength rating


Offer laser enucleation of the prostate using Ho:YAG laser (HoLEP) to men with moderate- Strong
to-severe LUTS as an alternative to transurethral resection of the prostate or open
prostatectomy.

5.3.2.4 Thulium:yttrium-aluminium-garnet laser (Tm:YAG) enucleation of the prostate


Mechanism of action: The Tm:YAG laser has been described in section 5.3.1.3. Enucleation using the Tm:YAG
laser includes ThuVEP (vapoenucleation i.e. excising technique) and ThuLEP (blunt enucleation).

Efficacy: A meta-analysis evaluating ThuLEP vs. M-TURP and B-TURP reported no clinically relevant
differences in Qmax, IPSS and QoL score [355]. An RCT with five years follow-up comparing ThuLEP with
B-TURP found no difference between the two procedures for Qmax, IPSS, PVR, and QoL [375]. A meta-analysis
[376] evaluating ThuLEP vs. HoLEP showed no clinically relevant differences in IPSS, QoL score and Qmax at
twelve months in accordance with one RCT showing similar results at eighteen months [377]. Furthermore,
ThuLEP and PKEP were compared in one RCT with twelve months follow-up the outcome of which showed no
difference with regard to efficacy [378].
There are mainly prospective case studies on ThuVEP showing a significant improvement in IPSS,
Qmax, and PVR after treatment [379-382].

Tolerability and safety: A meta-analysis evaluating ThuLEP vs. M-TURP and B-TURP reported a longer
operation time and a shorter hospitalisation time for ThuLEP compared to M-TURP and B-TURP, respectively
with no difference in complication rates [355]. A meta-analysis [376] evaluating ThuLEP vs. HoLEP showed
a significant difference is enucleation time favouring ThuLEP, but no significant differences in operation,
catheterisation and hospitalisation times short-term complication rates, in accordance with one RCT showing
similar results, including no urethral and bladder neck strictures, at eighteen months [377]. ThuLEP and PKEP
were compared in one RCT with twelve months follow-up [378]. No significant difference in complication rates
was detected, but haemoglobin level decrease and catheterisation time was significantly lower for ThuLEP.
In comparative studies ThuVEP show high intra-operative safety [383], as well as in case series in
patients with large prostates [379] and anticoagulation or bleeding disorders [380, 384]. A study focusing on
post-operative complications after ThuVEP reported adverse events in 31% of cases, with 6.6% complications
greater then Clavien grade 2 [385]. One case control study on ThuVEP with 48-month follow-up reported long-
term durability of voiding improvements and overall re-operation rates of 2.4% [384]. Two studies addressed
the impact of ThuVEP on sexual function, demonstrating no effect on erectile function with increased
prevalence of retrograde ejaculation post-operatively [386, 387].

Practical considerations: ThuLEP seems to offer similar efficacy and safety when compared to TURP, bipolar
enucleation and HoLEP; whereas, ThuVEP is not supported by RCTs. Based on the limited number of RCTs
and lack of long-term follow-up data there is a need for ongoing investigation of these techniques.

Summary of evidence LE
Enucleation of the prostate using the Tm:YAG laser demonstrates similar efficacy when compared to 1b
M-TURP/bipolar transurethral (plasmakinetic) enucleation, HoLEP and B-TURP in the short-, mid-, and
long-term, respectively.
Enucleation of the prostate using the Tm:YAG laser (ThuLEP) demonstrates similar safety compared 1b
to TURP/bipolar transurethral (plasmakinetic) enucleation, and HoLEP in the short- and mid-term,
respectively.

34 MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021
Vapoenucleation of the prostate using a Tm:YAG laser (ThuVEP) seems to be safe in patients with 2b
large prostates and those receiving anticoagulant or antiplatelet therapy.

Recommendations Strength rating


Offer enucleation of the prostate using the Tm:YAG laser (ThuLEP, ThuVEP) to men with Weak
moderate-to-severe LUTS as an alternative to transurethral resection of the prostate,
holmium laser enucleation or bipolar transurethral (plasmakinetic) enucleation.
Offer Tm:YAG laser enucleation of the prostate to patients receiving anticoagulant or Weak
antiplatelet therapy.

5.3.2.5 Diode laser enucleation of the prostate


Mechanism of action: For prostate surgery, diode lasers with a wavelength of 940, 980, 1,318, and 1,470 nm
(depending on the semiconductor used) are marketed for vaporisation and enucleation. Only a few have been
evaluated in clinical trials [385].

Efficacy: One small RCT comparing 1,318 nm diode laser enucleation of the prostate (DiLEP) with B-TURP in
patients with mean prostate volume < 80 mL reported no significant differences in IPSS, QoL score, Qmax and
PVR at six months follow-up [388]. Another RCT comparing 1,470 nm DiLEP with B-TURP in 157 patients with
mean prostate volume < 80 mL also reported no significant differences in IPSS, QoL score, Qmax and PVR at
twelve months follow-up [389]. In addition, three RCTs comparing 980 nm DiLEP with PKEP in patients with
mean prostate volume < 80 mL [390, 391] and > 80 mL [392] reported no significant differences in IPSS, QoL
score, Qmax and PVR at twelve months follow-up.

Tolerability and safety: One small RCT comparing 1,318 nm DiLEP with B-TURP in patients with mean prostate
volume < 80 mL and six months follow-up reported a significantly longer operation time for DiLEP, but shorter
catheterisation and hospitalisation times, as well as less blood loss (without differences in blood transfusion
rates) [388]. No differences in complication rates were reported between the two arms [388]. Another RCT
comparing 1,470 nm DiLEP with B-TURP in 157 patients with mean prostate volume < 80 mL and twelve
months follow-up reported significantly shorter operation, catheterisation and hospitalisation times with less
blood loss (without differences in blood transfusion rates) for DiLEP, with no differences in complication rates
between the two arms [389]. Three RCTs comparing 980 nm DiLEP with PKEP in patients with mean prostate
volume < 80 mL [390, 391] and > 80 mL [392] and twelve months follow-up reported conflicting peri-operative
outcomes: operation time (no difference between arms [390], significanly shorter for DiLEP [391] or sgnificanly
longer for DiLEP [392]); catheterisation time (no difference between the two arms [390], significanly shorter for
DiLEP [391, 392]); hospitalisation time (no difference between arms [390, 391], significanly shorter for DiLEP
[392]); blood loss (no difference in haemoglobin drop between arms [390], significantly lower haemoglobin drop
for DiLEP [391, 392]). All trials reported no differences in blood transfusion rates and complication rates [390-
392].

Practical considerations: Diode laser enucleation seems to offer similar efficacy and safety when compared
to either B-TURP or bipolar transurethral (plasmakinetic) enucleation. Based on the limited number, mainly
low-quality RCTs, and controversial data on the retreatment rate, results for diode laser enucleation should be
evaluated in further higher quality RCTs.

Summary of evidence LE
Laser enucleation of the prostate using the 1,318 nm or 1,470 laser showed comparable short-term 1b
efficacy and safety to B-TURP. Peri-operative parameters like blood loss, catheterisation time and
hospital stay are in favour of diode enucleation.
Laser enucleation of the prostate using the 980 nm laser showed comparable short-term efficacy and 1b
safety to bipolar transurethral (plasmakinetic) enucleation.

Recommendation Strength rating


Offer 120-W 980 nm, 1,318 nm or 1,470 nm diode laser enucleation of the prostate to Weak
men with moderate-to-severe LUTS as a comparable alternative to bipolar transurethral
(plasmakinetic) enucleation or bipolar transurethral resection of the prostate.

MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021 35
5.3.2.6 Enucleation techniques under investigation
5.3.2.6.1 Minimal invasive simple prostatectomy
Mechanism of action: The term minimal invasive simple prostatectomy (MISP) includes laparoscopic simple
prostatectomy (LSP) and robot-assisted simple prostatectomy (RASP). The technique for LSP was first
described in 2002 [393], while the first RASP was reported in 2008 [394]. Both LSP and RASP are performed
using different personalised techniques, based on the transcapsular (Millin) or transvesical (Freyer) techniques
of OP. An extraperitoneal approach is mostly used for LSP, while a transperitoneal approach is mostly used for
RASP.

Efficacy: A SR and meta-analysis showed that in 27 observational studies including 764 patients, the mean
increase in Qmax was 14.3 mL/s, and the mean improvement in IPSS was 17.2 [395]. Mean duration of
operation was 141 minutes and the mean intra-operative blood loss was 284 mL. One hundred and four
patients (13.6%) developed a surgical complication. In comparative studies to OP, length of hospital stay,
length of catheter use and estimated blood loss were significantly lower in the MISP group, while the duration
of operation was longer than in OP. There were no differences in improvements in Qmax, IPSS and peri-
operative complications between both procedures.
Two recent retrospective series on RASP were not included in the meta-analysis which confirm
these findings [396, 397]. The largest retrospective series reports 1,330 consecutive cases including 487
robotic (36.6%) and 843 laparoscopic (63.4%) simple prostatectomy cases. The authors confirm that both
techniques can be safely and effectively done in selected centres [396].

Tolerability and safety: In the largest series, the post-operative complication rate was 10.6% (7.1% for LSP and
16.6% for RASP), most of the complications being of low grade. The most common complications in the RASP
series were haematuria requiring irrigation, UTI and AUR; in the LSP series, the most common complications
were UTI, ileus and AUR. In the most recent, largest comparative analysis of robotic vs. OP for large-volume
prostates, a propensity score-matched analysis was performed with five covariates. Robotic compared with OP
demonstrated a significant shorter average length of hospital stay, but longer mean operative time. The robotic
approach was also associated with a lower estimated blood loss. Improvements in maximal flow rate, IPSS,
QoL, PVR and post-operative PSA levels were similar before and after surgery for both groups. There was no
difference in complications between the groups [398].

Practical considerations: Minimal invasive simple prostatectomy seems comparable to OP in terms of


efficacy and safety, providing similar improvements in Qmax and IPSS [395]. However, most studies are of a
retrospective nature. High-quality studies are needed to compare the efficacy, safety, and hospitalisation times
of MISP and both OP and endoscopic methods. Long-term outcomes, learning curve and cost of MISP should
also be evaluated.

Summary of evidence LE
Minimal invasive simple prostatectomy is feasible in men with prostate sizes > 80 mL needing surgical 1a
treatment; however, RCTs are needed.

5.3.2.6.2 532 nm (‘Greenlight’) laser enucleation of the prostate


Mechanism of action: The Potassium-Titanyl-Phosphate (KTP) and the lithium triborate (LBO) lasers work at
a wavelength of 532 nm. Laser energy is absorbed by haemoglobin, but not by water. Vaporisation leads to
immediate removal of prostatic tissue. Three “Greenlight” lasers exist, which differ not only in maximum power
output, but more significantly in fibre design and the associated energy tissue interaction of each. The standard
Greenlight device today is the 180-W XPS laser, but the majority of evidence is published with the former 80-W
KTP or 120-W HPS (LBO) laser systems.
Two approaches for KTP/LBO laser based enucleation technique exist [399]. GreenLEP is an
anatomical enucleation technique following the principle of blunt dissection of the adenoma with the sheath
and laser energy for incision as described for ThuLEP [400]. En bloc GreenLEP preparation has been
popularised with the same approach. A variation of the most commonly applied GreenLEP technique, with
tissue morcellation, is the in situ vaporisation of apically enucleated tissue, also referred to as anatomic
vaporisation-incision technique [400, 401]. Lithium triborate/Greenlight vapoenucleation on the other hand uses
vapoincising laser energy to cut out the tissue [402]. To date no RCTs evaluating enucleation using the KTP/
LBO laser exist [403].

36 MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021
5.3.3 Vaporisation of the prostate
5.3.3.1 Bipolar transurethral vaporisation of the prostate
Mechanism of action: Bipolar transurethral vaporisation of the prostate (B-TUVP) was introduced in the late
1990s (“PK” B-TUVP). The technique was derived from PK B-TURP and utilised a bipolar electrode and a high-
frequency generator to create a plasma effect able to vaporise prostatic tissue [404]. With minimal direct tissue
contact (near-contact; hovering technique) and heat production the bipolar electrode produces a constant
plasma field (thin layer of highly ionized particles; plasma corona), allowing it to glide over the tissue and
vaporise a limited layer of prostate cells without affecting the underlying tissue whilst achieving haemostasis,
leaving behind a TURP-like cavity [405]. A distinct difference between B-TUVP and its ancestor (monopolar
TUVP), is that B-TUVP displays thinner (< 2 mm) coagulation zones [406], compared to the disproportionate
extent of those created by the former (up to 10 mm) [407], that potentially lead to mostly irritative side-effects
and stress urinary incontinence [406, 408, 409].

Efficacy: Bipolar-TUVP has been evaluated as a TURP alternative for treating moderate-to-severe LUTS in
thirteen RCTs to date, including a total of 1,244 men with a prostate size of < 80 mL [303, 410-421]. Early
RCTs evaluated the PK B-TUVP system [410-414]; however, during the last decade, only the “plasma” B-TUVP
system with the “mushroom- or button-like” electrode (Olympus, Medical) has been evaluated [303, 415-421].
Results have been pooled in three RCT-based meta-analyses [289, 422, 423], and a narrative synthesis has
been produced in two SRs [289, 424]. The follow-up in most RCTs is twelve months [410-413, 415-417, 419,
421]. The longest follow-up is 36 months in a small RCT (n = 40) and eighteen months in a subsequent RCT
n = 340); evaluating PK [414] and plasma B-TUVP [303], respectively.
Early pooled results concluded that no significant differences exist in short-term efficacy (IPSS,
QoL score, Qmax and PVR) between PK B-TUVP and TURP [289]. However, the promising initial efficacy profile
of the former may be compromised by inferior clinical outcomes (IPSS, Qmax) at mid-term. Larger RCTs with
longer follow-up are necessary to draw definite conclusions [289, 414]. A SR of seven RCTs comparing PK
and plasma B-TUVP with TURP concluded that functional outcomes of B-TUVP and TURP do not differ [424].
The poor quality of the included RCTs and the fact that most data was derived from a single institution was
highlighted [424]. A similar SR of eight RCTs comparing both B-TUVP techniques with TURP concluded that
not enough consistent data suitable for a meta-analysis exists; that main functional results are contradictory;
and that heterogeneity of RCTs, non-standardised techniques and methodological limitations do not permit firm
conclusions [289]. A meta-analysis of six RCTs specifically evaluating plasma B-TUVP vs. TURP, concluded
that both techniques result in a similar improvement of LUTS [423].

Tolerability and safety: Early pooled results concluded that no statistically significant differences exist
collectively for intra-operative and short-term complications between PK B-TUVP and TURP, but peri-operative
complications are significantly fewer after B-TUVP [289]. However, the results of a statistical analysis comparing
pooled specific complication rates were not directly reported in this meta-analysis [289]. Mid-term safety results
(urethral stricture, ED, and retrograde ejaculation) have also been reported to be similar [414], but larger RCTs
with longer follow-up are necessary to draw definite conclusions [289, 414]. A SR of seven RCTs comparing
PK and plasma B-TUVP with TURP concluded that most RCTs suggest a better haemostatic efficiency for
B-TUVP, resulting in shorter catheterisation (42.5 vs. 77.5 hours) and hospitalisation times (3.1 vs. 4.4 days)
[424]. A similar SR of eight RCTs comparing both B-TUVP techniques with TURP concluded that not enough
consistent data suitable for a meta-analysis exists and that heterogeneity of RCTs, non-standardised techniques
and methodological limitations do not permit firm conclusions [289]. A meta-analysis of six RCTs specifically
evaluating plasma B-TUVP vs. TURP, concluded that no significant differences exist between the techniques
in overall complication and transfusion rates [423]. However, a statistically significant difference was detected
collectively in major complication rates (Clavien 3, 4; including urethral stricture, severe bleeding necessitating
re-operation and urinary incontinence) and in the duration of catheterisation favouring plasma B-TUVP.

Practical considerations: Bipolar-TUVP and TURP have similar short-term efficacy. Plasmakinetic B-TUVP has a
favourable peri-operative profile, similar mid-term safety, but inferior mid-term efficacy compared to TURP. Plasma
B-TUVP has a lower short-term major morbidity compared to TURP. Randomised controlled trials of higher quality,
multicentre RCTs, and longer follow-up periods are needed to evaluate B-TUVP in comparison to TURP.

Summary of evidence LE
Bipolar-TUVP and TURP have similar short-term efficacy. 1a
Plasmakinetic B-TUVP has a favourable peri-operative profile, similar mid-term safety, but inferior 1a
mid-term efficacy compared to TURP.
Plasma B-TUVP has a lower short-term major morbidity rate compared to TURP. 1a

MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021 37
Recommendation Strength rating
Offer bipolar transurethral vaporisation of the prostate as an alternative to monopolar Weak
transurethral resection of the prostate to surgically treat moderate-to-severe LUTS in men
with a prostate volume of 30-80 mL.

5.3.3.2 532 nm (‘Greenlight’) laser vaporisation of the prostate


Mechanism of action: The Potassium-Titanyl-Phosphate (KTP) and the lithium triborate (LBO) lasers have been
described in section 5.3.2.6.2.

Efficacy: A meta-analysis of the nine available RCTs comparing photoselective vaporisation of the prostate
(PVP) using the 80-W and 120-W lasers with TURP was performed in 2012 [425]. No differences were found in
Qmax and IPSS between 80-W PVP and TURP, but only three RCTs provided sufficient twelve month data to be
included in the meta-analysis [426-428]. Another meta-analysis from 2016, of four RCTs including 559 patients,
on the 120-W laser, demonstrated no significant difference in functional and symptomatic parameters at 6-,
12-, and 24-month follow-up when compared to TURP [429].
The only available RCT for the 180-W laser reported non-inferiority to TURP in terms of IPSS, Qmax,
PVR volume, prostate volume reduction, PSA decrease and QoL questionnaires. Efficacy outcomes were
similar to TURP with stable results at 24 months follow-up [430].
The longest RCT comparing the 120-W HPS laser with TURP had a follow-up of 36 months and
showed a comparable improvement in IPSS, Qmax, and PVR [431]. Comparable improvements in IPSS, QoL,
Qmax, or urodynamic parameters were reported from two RCTs with a maximum follow-up of 24 months [427,
432]. A SR and meta-analysis of eleven RCTs comparing M-TURP with the 80-W KTP or 120-W HPS system
found no significant difference with respect to IPSS and Qmax improvement [433].
One RCT comparing HoLEP to PVP, in patients with prostates > 60 mL, showed comparable
symptom improvement, but significantly higher flow rates and lower PVR volume after HoLEP at short-term
follow-up; in addition, PVP showed a 22% conversion rate to TURP [434].

Tolerability and safety: A meta-analysis of RCTs comparing the 80-W and 120-W lasers with TURP showed a
significantly longer operating time, but shorter catheterisation time and length of hospital stay after PVP [289].
Blood transfusions and clot retention were less with PVP. No difference was noted in post-operative urinary
retention, infection, meatal stenosis, urethral stricture, or bladder neck stenosis [289]. In a meta-analysis
including trials with the 120-W laser, patients in the PVP group demonstrated significantly lower transfusion
rates, shorter catheterisation time and shorter duration of hospital stay compared to TURP. Re-operation rates
and operation time were in favour of TURP. No significant differences were demonstrated for treatment for
urethral stricture, BNC, incidence of incontinence and infection [429]. A SR and meta-analysis of eleven RCTs
comparing M-TURP with the 80-W KTP or 120-W HPS system found that PVP was superior to M-TURP with
regard to transfusion rate, clot retention, catheterisation and hospitalisation time [433].
180-W Greenlight laser prostatectomy is non-inferior to TURP in terms of peri-operative
complications. Re-operation free survival during 24 months follow-up was comparable between the TURP-arm
and the 180-W XPS laser-arm [430]. In an RCT comparing the 120-W HPS laser with TURP, with a follow-up of
36 months, the re-operation rate was significantly higher after PVP (11% vs. 1.8%; p = 0.04) [431].
Based mostly on case series the 80-,120- and 180-W Greenlight laser appears to be safe in
high-risk patients undergoing anticoagulation treatment [435-438]; however, patients under anticoagulation
therapy were either excluded from or represented a very small sample in currently available RCTs. In one
study, anticoagulant patients had significantly higher rates of bladder irrigation (17.2%) compared with
those not taking anticoagulants (5.4%) [438]. In contrast, another retrospective study focusing on the 180-W
LBO laser did not find any significant differences between patients receiving or not receiving anticoagulants
[439]. A retrospective study of a mixed cohort of patients, treated with 80-W KTP PVP and 120-W LBO HPS,
revealed that delayed gross haematuria was common in patients (33.8%) during an average follow-up of 33
months [440]. A retrospective review of a database of patients undergoing 180-W PVP, without interruption of
anticoagulation therapy, had a 30.5% rate of peri-operative adverse events with a significant occurrence of
high grade Clavien Dindo events [441].
Safety in patients with urinary retention, impaired detrusor contractility, elderly patients or prostates
> 80 mL was shown in various prospective short-term non-randomised trials. No RCT including prostates
> 100 mL has been reported; therefore, comparison of retreatment rates between prostate volumes of different
sizes is not possible [442-444].
An RCT with twelve months follow-up reported a retrograde ejaculation rate of 49.9% following PVP
with an 80-W laser vs. 56.7% for TURP, there was no impact on erectile function in either arm of the trial [445].
Additional studies have also reported no difference between OP/TURP and Greenlight PVP for erectile function
[446, 447]. However, IIEF-5 scores were significantly decreased at 6-, 12-, and 24- months in patients with pre-
operative IIEF-5 greater than 19 [448].

38 MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021
Practical considerations: The 180-W XPS represents the current standard of generators for PVP; however, the
number and quality of supporting publications are low, especially for large glands (> 100 mL), with no long-term
follow up.

Summary of evidence LE
Laser vaporisation of the prostate using the 80-W KTP and the 120-W LBO laser (PVP) demonstrated 1a
higher intra-operative safety with regard to haemostatic properties when compared to TURP. Peri-
operative parameters such as catheterisation time and hospital stay are in favour of PVP, whereas
operation time and risk of re-operation are in favour of TURP. Short-term results for the 80-W KTP
laser and mid-term results for the 120-W LBO laser were comparable to TURP.
Laser vaporisation of the prostate using the 180-W LBO laser (PVP) demonstrated higher intra- 1b
operative safety with regard to haemostatic properties when compared to TURP. Peri-operative
parameters such as catheterisation time and hospital stay were in favour of PVP, whereas operation
time was in favour of TURP. Short- to mid-term results are comparable to TURP.
Laser vaporisation of the prostate using the 80-W KTP and 120-W LBO lasers seems to be safe for the 2
treatment of patients receiving antiplatelet or anticoagulant therapy.
Laser vaporisation of the prostate using the 180-W LBO laser seems to be safe for the treatment of 3
patients receiving antiplatelet or anticoagulant therapy; however, the level of available evidence is low.

Recommendations Strength rating


Offer 80-W 532-nm Potassium-Titanyl-Phosphate (KTP) laser vaporisation of the prostate Strong
to men with moderate-to-severe LUTS with a prostate volume of 30-80 mL as an alternative
to transurethral resection of the prostate (TURP).
Offer 120-W 532-nm Lithium Borat (LBO) laser vaporisation of the prostate to men with Strong
moderate-to-severe LUTS with a prostate volume of 30-80 mL as an alternative to TURP.
Offer 180-W 532-nm LBO laser vaporisation of the prostate to men with moderate-to- Strong
severe LUTS with a prostate volume of 30-80 mL as an alternative to TURP.
Offer laser vaporisation of the prostate using 80-W KTP, 120- or 180-W LBO lasers for the Weak
treatment of patients receiving antiplatelet or anticoagulant therapy with a prostate volume
< 80 mL.

5.3.3.3 Vaporisation techniques under investigation


5.3.3.3.1 Diode laser vaporisation of the prostate
Mechanism of action: For prostate surgery, diode lasers with a wavelength of 980 nm are marketed for
vaporisation; however, only a few have been evaluated in clinical trials [317].

Efficacy: Two RCTs for 120-W 980 nm diode laser vaporisation vs. M-TURP are available [449, 450]. The first
RCT with 24 month follow-up reported equal symptomatic and clinical parameters at one and six months.
However, at 12- and 24-months the results were significantly in favour of TURP, repeat TURP was more
frequent in the diode laser group [449]. The second RCT reported equivocal results for both interventions at
3-month follow-up [450].

Tolerability and safety: Published studies on 980 nm diode laser vaporisation indicate high haemostatic
potential, although anticoagulants or platelet aggregation inhibitors were taken in 24% and 52% of patients,
respectively [451, 452]. In a number of studies, a high rate of post-operative dysuria was reported [449, 451-
453]. In an RCT reflecting on peri-operative and post-operative complications no significant differences were
demonstrated for clot retention, re-catheterisation, UUI and UTI [449]. Moreover, for late complications no
significant differences could be demonstrated for re-operation rate, urethral stricture, bladder neck sclerosis,
de novo sexual dysfunction and mean time of dysuria [449].
Fibre modifications can potentially reduce surgical time [454]. Early publications on diode
vaporisation reported high re-operation rates (8-33%) and persisting stress urinary incontinence (9.1%) [449,
451-453].

Practical considerations: Diode laser vaporisation leads to similar improvements in clinical and symptomatic
parameters during short-term follow-up and provides good haemostatic properties. Based on the limited
number, mainly low quality RCTs, and controversial data on the retreatment rate, results for diode laser
vaporisation should be evaluated in further higher quality RCTs.

MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021 39
Summary of evidence LE
Laser vaporisation of the prostate using the 120-W 980 nm diode laser demonstrated high intra- 1b
operative safety with regard to haemostatic properties when compared to TURP. Peri-operative
parameters like catheterisation time and hospital stay were in favour of diode lasers. Evidence is
limited by the number and quality of the available studies.
In a number of studies severe post-operative complications such as severe storage symptoms and 3
persisting incontinence occurred with laser vaporisation of the prostate using the 120-W 980 nm diode
laser.
Laser vaporisation using the 120-W 980 nm diode laser seems to be safe with regard to haemostasis 3
in patients receiving anticoagulant therapy.

5.3.4 Alternative ablative techniques


5.3.4.1 Aquablation – image guided robotic waterjet ablation: AquaBeam
Mechanism of action: AquaBeam uses the principle of hydro-dissection to ablate prostatic parenchyma while
sparing collagenous structures like blood vessels and the surgical capsule. A targeted high velocity saline
stream ablates prostatic tissue without the generation of thermal energy under real-time transrectal ultrasound
guidance; therefore, it is truly a robotic operation. After completion of ablation haemostasis is performed with a
Foley balloon catheter on light traction or diathermy or low-powered laser, if necessary [455].

Efficacy: In a double-blind, multicentre, prospective RCT 181 patients were randomised to TURP or
Aquablation [456, 457]. Mean total operative time was similar for Aquablation and TURP (33 vs. 36 minutes),
but resection time was significantly lower for Aquablation (4 vs. 27 minutes). At six months patients treated with
Aquablation and TURP experienced large IPSS improvements (-16.9 and -15.1, respectively). The study non-
inferiority hypothesis was satisfied. Larger prostates (50-80 mL) demonstrated a more pronounced benefit. At
one year follow-up, mean IPSS reduction was 15.1 in the Aquablation group and 15.1 in the TURP group with a
mean percent reduction in IPSS score of 67% in both groups. Ninety three percent and 86.7% of patients had
improvements of at least five points from baseline, respectively. No significant difference in improvement of
IPSS, QoL, Qmax and reduction of PVR was reported between the groups. One TURP subject (1.5%) and three
Aquablation subjects (2.6%) underwent re-TURP within one year of the study procedure [458]. In the American
cohort of this study (90 patients) one year data showed less anejaculation in patients treated with aquablation
(9%) compared with TURP (45%) in sexually active subjects [457].
A subgroup analysis of the WATER study [456] reported that in men with larger more complex
prostates Aquablation was associated with both superior symptom improvement and a better safety profile
with less post-operative anejaculation [459].
In a cohort study of 101 men with a prostate volume between 80-150 mL mean IPSS improved
from 23.2 at baseline to 5.9 at six months. Improvement in IPSS, QoL, Qmax and reduction of PVR were also
significant at six months. No secondary procedures for tissue removal occurred as of the six months [460].
Another RCT comparing Aquablation with TURP performed urodynamic studies on 66 patients at six
months follow-up and reported significant changes in pdetQmax (reductions of 35 and 34 cm H20, respectively)
and large improvements in BOO Index in both groups [461].

Tolerability and safety: An RCT has shown that fewer men in the Aquablation group had a persistent Clavien-
Dindo grade 1 or 2 or higher adverse event compared to TURP (26% vs. 42%) at three months following
treatment. Among sexually active men the rate of anejaculation was lower in those treated with Aquablation
compared to TURP (10% vs. 36%, respectively). There were no procedure-related adverse events after six
months [458]. In patients with a prostate volume between 80-150 mL, bleeding related events were observed
in fourteen patients (13.9%) of which eight (7.9%) occurred prior to discharge and six (5.9%) occurred within
one month of discharge. Blood transfusions were required in eight (7.9%) patients, return to the theatre for
fulguration in three (3.0%) patients, and both transfusion and fulguration in two patients (2.0%). Ejaculatory
dysfunction occurred in 19% of sexually active men [460].

Practical considerations: During short-term follow-up, Aquablation provides non-inferior functional outcomes
compared to TURP in patients with LUTS and a prostate volume between 30-80 mL. Longer term follow-up is
necessary to assess the clinical value of Aquablation.

Summary of evidence LE
Aquablation appears to be as effective as TURP both subjectively and objectively; however, there are 1b
still some concerns about the best methods of achieving post-treatment haemostasis.

40 MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021
Recommendations Strength rating
Offer Aquablation* to patients with moderate-to-severe LUTS and a prostate volume of Weak
30-80 mL as an alternative to TURP.
Inform patients about the risk of bleeding and the lack of long-term follow-up data. Strong
*Approach remains under investigation

5.3.4.2 Prostatic artery embolisation


Mechanism of action: Prostatic artery embolisation (PAE) can be performed as a day case procedure under
local anaesthesia with access through the femoral or radial arteries. Digital subtraction angiography displays
arterial anatomy and the appropriate prostatic arterial supply is selectively embolised to effect stasis in treated
prostatic vessels. Different techniques have been used for PAE. Atherosclerosis, excessive tortuosity of the
arterial supply and the presence of adverse collaterals are anatomical obstacles for the technical approach.
Cone beam computed tomography and contrast enhanced MR angiography can help identify prostatic arteries
and prevent off-target embolisation particularly in patients with challenging anatomical configurations [462, 463].

Efficacy: Two RCTs were conducted for direct comparison of PAE with TURP [464, 465]. Both studies observed
significant treatment outcomes for both procedures as compared to baseline values, but TURP was superior
when considering urodynamic parameters such as Qmax and PVR. Improvement of LUTS as determined by
IPSS and QoL was slightly more pronounced after TURP and reduction of prostate volume was significantly
more efficient after TURP than PAE.
Another RCT comparing PAE with TURP in 99 patients showed a mean reduction in IPSS from
baseline to twelve weeks of −9.23 points after PAE and −10.77 points after TURP. At twelve weeks, PAE
was less effective than TURP regarding improvements in Qmax (5.19 mL/s vs. 15.34 mL/s), PVR (−86.36 mL
vs. −199.98 mL), prostate volume (−12.17 mL vs. −30.27 mL), and significant de-obstructive effectiveness
according to pressure flow studies (56% vs. 93% shift towards less obstructive category). For secondary
outcomes, compared with PAE, procedural time was shorter for TURP, but in PAE patients, bladder catheter
indwelling time, and duration of hospital stay were significantly shorter [462].
Another SR and meta-analysis including the three above mentioned RCTs and two non-randomised
comparative studies (n = 708 patients), showed that TURP achieved a significantly higher mean post-
operative difference for IPSS and IPSS-QoL, 3.80 and 0.73 points, respectively compared to PAE [466]. All of
the functional outcomes assessed were significantly superior after TURP: 3.62 mL/s for Qmax, 11.51 mL for
prostate volume, 11.86 mL for PVR, and 1.02 ng/mL for PSA [466].

Tolerability and safety: In a SR of comparative studies PAE resulted in significantly more adverse events than
TURP/OP (41.6% vs. 30.4%). The frequency of AUR after the procedures was significantly higher in the PAE
group (9.4% vs. 2.0%) [467].
Another RCT however, reported fewer adverse events occurred after PAE than after TURP (36 vs.
70 events; p = 0.003). For secondary outcomes, PAE showed favourable results in terms of blood loss [462]. A
SR and meta-analysis of four studies (506 patients) comparing PAE and TURP found no significant difference in
the post-operative complication rate between TURP and PAE [468]. A SR of 708 patients whilst confirming that
PAE was not as effective as TURP, it did report fewer side effects than established surgical procedures [466].
Post-operative erectile function measured by IIEF-5 was in favour of PAE with mean difference in change of
2.56 points. Concerns still exist about non-target embolisation, reported in earlier studies [469]; however, more
recent studies report less incidents [466, 470].

Practical considerations: A multidisciplinary team approach of urologists and radiologists is mandatory and
patient selection should be done by urologists and interventional radiologists together. The investigation of
patients with LUTS to indicate suitability for invasive techniques should be performed by urologists only. This
technically demanding procedure should only be done by an interventional radiologist with specific mentored
training and expertise in PAE [471]. Patients with larger prostates (> 80 mL) may have the most to gain from
PAE. The selection of LUTS patients who will benefit from PAE still needs to be better defined [466]. Further
data with medium- and long-term follow-up are still required and comparison with other minimally invasive
techniques would be valuable. However, current evidence of safety and efficacy of PAE appears adequate
to support the use of this procedure for men with moderate-to-severe LUTS provided proper arrangements
for consent and audit are in place; therefore, a recommendation has been given, but PAE remains under
investigation.

MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021 41
Summary of evidence LE
Prostatic artery embolisation is less effective than TURP at improving symptoms and urodynamic 1a
parameters such as flow rate.
Procedural time is longer for PAE compared to TURP, but blood loss, catheterisation and 1b
hospitalisation time are in favour of PAE.

Recommendations Strength rating


Offer prostatic artery embolisation (PAE)* to men with moderate-to-severe LUTS who Weak
wish to consider minimally invasive treatment options and accept less optimal outcomes
compared with transurethral resection of the prostate.
Perform PAE only in units where the work up and follow-up is performed by urologists Strong
working collaboratively with trained interventional radiologists for the identification of PAE
suitable patients.
*Approach remains under investigation

5.3.4.3 Alternative ablative techniques under investigation


5.3.4.3.1 Convective water vapour energy (WAVE) ablation: The Rezum system
Mechanism of action: The Rezum system uses radiofrequency power to create thermal energy in the form of
water vapour, which in turn deposits the stored thermal energy when the steam phase shifts to liquid upon cell
contact. The steam disperses through the tissue interstices and releases stored thermal energy onto prostatic
tissue effecting cell necrosis. The procedure can be performed in an office based setting. Usually, one to three
injections are needed for each lateral lobe and one to two injections may be delivered into the median lobe.

Efficacy: In a multicentre, randomised, controlled study 197 men were enrolled and randomised in a 2:1 ratio
to treatment with water vapour energy ablation or sham treatment [472]. At three months relief of symptoms,
measured by a change in IPSS and Qmax were significantly improved and maintained compared to the sham
arm, although only the active treatment arm was followed up to twelve months. No relevant impact was
observed on PVR. Quality of life outcome was significantly improved with a meaningful treatment response
of 52% at twelve months. Further validated objective outcome measures such as BPH impact index (BPHII),
Overactive Bladder Questionnaire Short Form for OAB bother, and impact on QoL and International Continence
Society Male Item Short Form Survey for male incontinence demonstrated significant amelioration of
symptoms at three months follow-up with sustained efficacy throughout the study period of twelve months.
The reported two year results in the Rezum cohort arm of the same study and the recently reported four year
results confirmed durability of the positive clinical outcome after convective water vapour energy ablation [473,
474]. Surgical retreatment rate was 4.4% over four years [474].

Tolerability and safety: Safety profile was favourable with adverse events documented to be mild-to-moderate
and resolving rapidly. Preservation of erectile and ejaculatory function after convective water vapour thermal
therapy was demonstrated utilising validated outcome instruments such as IIEF and Male Sexual Health
Questionnaire-Ejaculation Disorder Questionnaire [472].

Practical considerations: Randomised controlled trials against a reference technique are needed to confirm the
first promising clinical results and to evaluate mid- and long-term efficacy and safety of water vapour energy
treatment.

5.3.5 Non-ablative techniques


5.3.5.1 Prostatic urethral lift
Mechanism of action: The prostatic urethral lift (PUL) represents a novel minimally invasive approach under
local or general anaesthesia. Encroaching lateral lobes are compressed by small permanent suture-based
implants delivered under cystoscopic guidance (Urolift®) resulting in an opening of the prostatic urethra leaving a
continuous anterior channel through the prostatic fossa extending from the bladder neck to the verumontanum.

Efficacy: In general, PUL achieves a significant improvement in IPSS (-39% to -52%), Qmax (+32% to +59%)
and QoL (-48% to -53%) [475-480]. Prostatic urethral lift was initially evaluated vs. sham in a multicentre
study with one [477] three [481] and five [482] years follow-up. The primary endpoint was met at three months
with a 50% reduction in IPSS. In addition, Qmax increased significantly from 8.1 to 12.4 mL/s compared to
baseline at three months and this result was confirmed at twelve months. The difference in clinical response
for Qmax between both groups was of statistical significance. A relevant benefit with regard to PVR was
not demonstrated compared to baseline or sham. At three years, average improvements from baseline
were significant for total IPSS, QoL, Qmax and individual IPSS symptoms. There was no de novo, sustained

42 MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021
ejaculatory or erectile dysfunction events and all sexual function assessments showed average stability or
improvement after PUL. Improvements in IPSS, QoL, BPHII and Qmax were durable throughout the five years
with improvement rates of 36%, 50%, 52%, and 44%, respectively. The re-treatment rate was 13.6% over five
years. Adverse events were mild to moderate and transient. Sexual function was stable over five years with no
de novo, sustained erectile, or ejaculatory dysfunction.
Another RCT of 80 patients was conducted in three European countries, comparing PUL to
TURP. At twelve months, IPSS improvement was -11.4 for PUL and -15.4 for TURP. There was no retrograde
ejaculation among PUL patients with 40% in the TURP patients. Surgical recovery was measured using a
validated instrument and confirmed that recovery from PUL is more rapid and more extensive in the first three
to six months [483]. However, TURP resulted in much greater improvements in Qmax after twelve months
compared to PUL. At 24 months, significant improvements in IPSS, IPSS QoL, BPHII, and Qmax were observed
in both arms. Change in IPSS and Qmax in the TURP arm were superior to the PUL arm [484]. Improvements in
QoL and BPHII score were not statistically different between the study arms. Prostatic urethral lift resulted in
superior quality of recovery and ejaculatory function preservation. Ejaculatory function and bother scores did
not change significantly in either treatment arm.
In a meta-analysis of retrospective and prospective trials, pooled estimates showed an overall
improvement following PUL, including IPSS, Qmax, and QoL [480]. Sexual function was preserved with a small
improvement estimated at twelve months.

Tolerability and safety: The most common complications reported post-operatively included haematuria
(16-63%), dysuria (25-58%), pelvic pain (5-17.9%), urgency (7.1-10%), transient incontinence (3.6-16%), and
UTI (2.9-11%) [477, 480-482]. Most symptoms were mild-to-moderate in severity and resolved within two to
four weeks after the procedure.
Prostatic urethral lift seems to have no significant impact on sexual function. Evaluation of sexual
function as measured by IIEF-5, Male Sexual Health Questionnaire-Ejaculatory Dysfunction, and Male Sexual
Health Questionnaire-Bother in patients undergoing PUL showed that erectile and ejaculatory function were
preserved [475-479].

Practical considerations: There are only limited data on treating patients with an obstructed/protruding middle
lobe [485]. It appears that they can be effectively treated with a variation in the standard technique, but further
data are needed [485]. The effectiveness in large prostate glands has not been shown yet. Long-term studies
are needed to evaluate the duration of the effect in comparison to other techniques.

Summary of evidence LE
Prostatic urethral lift improves IPSS, Qmax and QoL; however, these improvements are inferior to TURP 1b
at 24 months.
Prostatic urethral lift has a low incidence of sexual side effects. 1b
Patients should be informed that long-term effects including the risk of retreatment have not been 4
evaluated.

Recommendation Strength rating


Offer Prostatic urethral lift (Urolift®) to men with LUTS interested in preserving ejaculatory Strong
function, with prostates < 70 mL and no middle lobe.

5.3.5.2 Intra-prostatic injections


Mechanism of action: Various substances have been injected directly into the prostate in order to improve
LUTS, these include Botulinum toxin-A (BoNT-A), fexapotide triflutate (NX-1207) and PRX302. The primary
mechanism of action of BoNT-A is through the inhibition of neurotransmitter release from cholinergic neurons
[486]. The detailed mechanisms of action for the injectables NX-1207 and PRX302 are not completely
understood, but experimental data associates apoptosis-induced atrophy of the prostate with both drugs [486].

Efficacy: Results from clinical trials have shown only modest clinical benefits, that do not seem to be superior
to placebo, for BoNT-A [487, 488]. A recent SR and meta-analysis showed no differences in efficacy compared
with placebo and concluded that there is no evidence of clinical benefits in medical practice [489]. The positive
results from Phase II-studies have not be confirmed in Phase III-trials for PRX302 [490, 491]. NX-1207 was
evaluated in two multicentre placebo controlled double-blind randomised parallel group trials including a total
of 995 patients with a mean follow-up of 3.6 years, IPSS change from baseline was significantly higher and
AUR rate was significantly reduced in the treatment arm. The authors concluded that NX-1207 is an effective
transrectal injectable for long-term treatment for LUTS and that treated patients have reduced need for further
intervention [492].

MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021 43
Safety: Studies including safety assessments have reported only a few mild and self-limiting adverse events
for all injectable drugs [486]. A recent SR and meta-analysis showed low incident rates of procedure-related
adverse events [489]. Two multicentre placebo controlled double-blind randomised parallel group trials with
long-term follow-up evaluating NX-1207 detected no significant safety differences between the study arms
[492].

Practical considerations: Although experimental evidence for compounds such as BoNT-A and PRX302 were
promising for their transition to clinical use positive results from Phase II-studies have not be confirmed in
Phase III-trials. Randomised controlled trials against a reference technique are needed to confirm the first
promising clinical results of NX-1207.

Summary of evidence LE
Results from clinical trials have shown no clinical benefits for BoNT-A compared to placebo for the 1a
management of LUTS due to BPO.
Results from clinical trials have shown clinical benefits for NX-1207 compared to placebo for the 1b
management of LUTS due to BPO.

Recommendation Strength rating


Do not offer intraprostatic Botulinum toxin-A injection treatment to patients with male Strong
LUTS.

5.3.5.3 Non-ablative techniques under investigation


5.3.5.3.1 (i)TIND
Mechanism of action: The iTIND is a device designed to remodel the bladder neck and the prostatic urethra
and is composed of three elongated struts and an anchoring leaflet, all made of nitinol. Under direct
visualisation the iTIND is deployed inside the prostate in expanded configuration. The intended mode of action
is to compress obstructive tissue by the expanded device, thereby exerting radial force leading to ischaemic
necrosis in defined areas of interest. The iTIND is left in position for five days. The resulting incisions may be
similar to a Turner Warwick incision. In an outpatient setting the device is removed by standard urethroscopy.

Efficacy: A single-arm, prospective study of 32 patients with a follow-up of three years was conducted to
evaluate feasibility and safety of the procedure [493]. The change from baseline in IPSS, QoL score and Qmax
was significant at every follow-up time point [494].

Tolerability and safety: The device has been reported to be well tolerated by all patients. Four early
complications (12.5%) were recorded, including one case of urinary retention (3.1%), one case of transient
incontinence due to device displacement (3.1%), and two cases of infection (6.2%). No further complications
were recorded during the 36-month follow-up period.

Practical considerations: Randomised controlled trials comparing iTIND to a reference technique are ongoing.

5.4 Patient selection


The choice of treatment depends on the assessed findings of patient evaluation, ability of the treatment to
change the findings, treatment preferences of the individual patient, and the expectations to be met in terms of
speed of onset, efficacy, side effects, QoL, and disease progression.
Behavioural modifications, with or without medical treatments, are usually the first choice of therapy.
Figure 3 provides a flow chart illustrating treatment choice according to evidence-based medicine and patient
profiles. Surgical treatment is usually required when patients have experienced recurrent or refractory urinary
retention, overflow incontinence, recurrent UTIs, bladder stones or diverticula, treatment-resistant macroscopic
haematuria due to BPH/BPE, or dilatation of the upper urinary tract due to BPO, with or without renal
insufficiency (absolute operation indications, need for surgery).
Additionally, surgery is usually needed when patients have not obtained adequate relief from
LUTS or PVR using conservative or medical treatments (relative operation indications). The choice of
surgical technique depends on prostate size, comorbidities of the patient, ability to have anaesthesia,
patients’ preferences, willingness to accept surgery-associated specific side-effects, availability of the surgical
armamentarium, and experience of the surgeon with these surgical techniques. An algorithm for surgical
approaches according to evidence-based medicine and the patient’s profile is provided in Figure 4.

44 MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021
Figure 3: Treatment algorithm of male LUTS using medical and/or conservative treatment options.
Treatment decisions depend on results assessed during initial evaluation.
Note that patients’ preferences may result in different treatment decisions.

Male LUTS
(without indications for surgery)

Bothersome
no symptoms? yes

Nocturnal
no polyuria yes
predominant

no Storage symptoms yes


predominant?

Prostate
no volume yes
> 40 mL?

Education + lifestyle Long-term


advice with or without no treatment?
α1-blocker/PDE5I

yes
Residual
storage
symptoms

Watchful Add muscarinic Education + Education + Education +


waiting receptor lifestyle advice lifestyle advice lifestyle advice
with or without antagonist/beta with or without with or without with or without
education + -3 agonist 5α-reductase muscarinic vasopressin
lifestyle advice inhibitor ± α1- receptor analogue
blocker/PDE5I antagonist/beta
-3 agonist

PDE5I = phosphodiesterase type 5 inhibitors.

MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021 45
Figure 4: Treatment algorithm of bothersome LUTS refractory to conservative/medical treatment or
in cases of absolute operation indications. The flowchart is stratified by the patient’s ability to
have anaesthesia, cardiovascular risk, and prostate size.

Male LUTS
with absolute indicaons for surgery or non-responders to medical treatment or those who do not
want medical treatment but request acve treatment

High-risk
no yes
paents?

Can have
yes surgery under no
anaesthesia?

Can stop
yes no
ancoagulaon/
anplatelet therapy

Prostate
< 30 mL volume > 80 mL

30 – 80 mL

TUIP (1) TURP (1) Open Laser PU liˆ


TURP Laser prostatectomy (1) vaporisaon (1)
enucleaon HoLEP (1) Laser
Bipolar Bipolar enucleaon
enucleaon enucleaon (1)
Laser Laser
vaporisaon vaporisaon
PU liˆ Thulium
enucleaon
TURP

(1) Current standard/first choice. The alternative treatments are presented in alphabetical order.
Laser vaporisation includes GreenLight, thulium, and diode laser vaporisation. Laser enucleation includes
holmium and thulium laser enucleation.
HoLEP = holmium laser enucleation; TUIP = transurethral incision of the prostate; TURP = transurethral
resection of the prostate and PU = prostatic urethral.

46 MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021
5.5 Management of Nocturia in men with lower urinary tract symptoms
The following section reports a SR of therapy for the management of nocturia in men with LUTS. It also
emphasises the need to consider the wide range of possible causes of nocturia [495].
Nocturia is defined as the complaint of waking at night to void [6]. It reflects the relationship
between the amount of urine produced while asleep, and the ability of the bladder to store the urine received.
Nocturia can occur as part of lower urinary tract dysfunction (LUTD), such as OAB and chronic pelvic pain
syndrome. Nocturia can also occur in association with other forms of LUTD, such as BOO, but here it is
debated whether the link is one of causation or simply the co-existence of two common conditions. Crucially,
nocturia may have behavioural, sleep disturbance (primary or secondary) or systemic causes unrelated to
LUTD (Table 2). Differing causes often co-exist and each has to be considered in all cases. Only where LUTD is
contributory should nocturia be termed a LUTS.

Table 2: Categories of nocturia


CATEGORY Disproportionate urine production Low volume of each void
(at all times, or during sleep) (at all times, or overnight)
Behavioural Inappropriate fluid intake “Bladder awareness” due to
secondary sleep disturbance
Systemic Water, salt and metabolite output
Sleep disorder Variable water and salt output “Bladder awareness” due to
primary sleep disturbance
LUTD Impaired storage function and
increased filling sensation

5.5.1 Diagnostic assessment


Evaluation is outlined in Figure 5;
1. Evaluate for LUTD according to the relevant guidelines. The severity and bother of individual LUTS
should be identified with a symptom score, supplemented by directed questioning if needed. A validated
bladder diary is mandatory.
2. Review whether behavioural factors affecting fluid balance and sleep are contributing.
3. Review of medical history and medications, including directed evaluation for key conditions, such as
renal failure, diabetes mellitus, cardiac failure, and obstructive sleep apnoea. If systemic factors or sleep
disorders are potentially important, consider involving appropriate medical expertise (see Figure 6). This
is appropriate where a known condition is suboptimally managed, or symptoms and signs suggest an
undiagnosed condition.

5.5.2 Medical conditions and sleep disorders Shared Care Pathway


Causative categories for nocturia comprise [496]:
1. bladder storage problems;
2. 24-hour polyuria (> 40 mL/kg urine output over a 24-hour period);
3. nocturnal polyuria (NP; nocturnal output exceeding 20% of 24-hour urine output in the young, or 33% of
urine output in people > 65 [6]);
4. sleep disorders;
5. mixed aetiology.
Potentially relevant systemic conditions are those which impair physiological fluid balance, including influences
on: levels of free water, salt, other solutes and plasma oncotic pressure; endocrine regulation e.g. by
antidiuretic hormone; natriuretic peptides; cardiovascular and autonomic control; renal function; neurological
regulation, e.g. circadian regulation of the pineal gland, and renal innervation. As nocturia is commonly referred
to the specialty without full insight into cause, the urologist must review the likely mechanisms underlying
a presentation with nocturia and instigate review by relevant specialties accordingly. Thus, the managing
urologist needs to evaluate nocturia patients in a context where additional medical expertise is available (Table
3). They should not proceed along any LUTD management pathway unless a causative link with LUTD is
justifiably suspected, and systemic or sleep abnormalities have been considered.
In patients with non-bothersome nocturia, the medical evaluation (history and physical examination)
should consider the possibility of early stages of systemic disease, and whether there is possibility of earlier
diagnosis or therapy adjustment.
Some important potentially treatable non-urological causes of nocturia include; obstructive sleep
aponea, congestive cardiac failure, poorly controlled diabetes mellitus and medications (e.g. diuretics, or
lithium).

MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021 47
Figure 5. Evaluation of Nocturia in non-neurogenic Male LUTS.

• History (+ sexual function)


• Symptom Score Questionnaire
• Physical Examination
• Urinalysis Bothersome Nocturia
• PSA (if diagnosis of PCa will change
the management – discuss with patient)
• Measurement of PVR
yes no

Significant PVR

• US assessment of prostate
• Uroflowmetry
• US of kidneys +/- renal
function assessment
• FVC with predominant storage LUTS

• Abnormal DRE, high PSA Mixed features


• Haematuria Polyuria/ LUTS
• Chronic pelvic pain NP
Evaluate according to
relevant guidelines or
clinical standard
Medical Conditions/Sleep Nocturia with LUTS in
disorders Care Pathway benign LUT conditions

Treat underlying condition Behavioural and drug NP Behavioural and drug


or sleep disorder treatment LUTS treatment
Offer shared care LUTS Algorithm

Interventional LUTS
treatment
(Indirect MoA for
nocturia)

Assessment must establish whether the patient has polyuria, LUTS, a sleep disorder or a combination. Therapy
may be driven by the bother it causes, but non-bothersome nocturia may warrant assessment with a frequency
volume chart (indicated by the dotted line) depending on history and clinical examination since potential
presence of a serious underlying medical condition must be considered.
FVC = frequency volume chart; DRE = digital rectal examination; NP = nocturnal polyuria; MoA = mechanism of
action; PVR = post-void residual; PSA = prostate-specific antigen; US = ultrasound.

48 MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021
Table 3: Shared care pathway for nocturia, highlighting the need to manage potentially complex patients using
relevant expertise for the causative factors.

UROLOGICAL CONTRIBUTION SHARED CARE MEDICAL CONTRIBUTION


Diagnosis of LUTD Diagnosis of conditions causing NP
• Urological/LUTS evaluation • Evaluate patient’s known
• Nocturia symptom scores conditions
• Bladder diary • Screening for sleep disorders
• Screening for potential causes
of polyuria*
Conservative management Conservative Management
Behavioural therapy management • Initiation of therapy for new
• Fluid/sleep habits advice • Antidiuretic diagnosis
• Drugs for storage LUTS • Diuretics • Optimised therapy of known
• Drugs for voiding LUTS • Drugs to aid conditions
• ISC/catherisation sleep
• Increased exercise * Potential causes of polyuria
NEPHROLOGICAL DISEASE
• Leg elevation • Tubular dysfunction
• Weight loss • Global renal dysfunction
CARDIOVASCULAR DISEASE
• Cardiac disease
Interventional therapy • Vascular disease
• Therapy of refractory ENDOCRINE DISEASE
storage LUTS • Diabetes insipidus/mellitus
• Hormones affecting diuresis/natriuresis
• Therapy of refractory NEUROLOGICAL DISEASE
voiding LUTS • Pituitary and renal innervation
• Autonomic dysfunction
RESPIRATORY DISEASE
• Obstructive sleep apnoea
BIOCHEMICAL
• Altered blood oncotic pressure

5.5.3 Treatment for Nocturia


5.5.3.1 Antidiuretic therapy
The antidiuretic hormone arginine vasopressin (AVP) plays a key role in body water homeostasis and control
of urine production by binding to V2 receptors in the renal collecting ducts. Arginine vasopressin increases
water re-absorption and urinary osmolality, so decreasing water excretion and total urine volume. Arginine
vasopressin also has V1 receptor mediated vasoconstrictive/hypertensive effects and a very short serum half-
life, which makes the hormone unsuitable for treating nocturia/nocturnal polyuria.
Desmopressin is a synthetic analogue of AVP with high V2 receptor affinity and no relevant
V1 receptor affinity. It has been investigated for treating nocturia [497], with specific doses, titrated
dosing, differing formulations, and options for route of administration. Most studies have short follow-up.
Global interpretation of existing studies is difficult due to the limitations, imprecision, heterogeneity and
inconsistencies of the studies.

A SR of randomised or quasi-randomised trials in men with nocturia found that desmopressin may decrease
the number of nocturnal voids by -0.46 compared to placebo over short-term follow-up (up to three months);
over intermediate-term follow-up (three to twelve months) there was a change of -0.85 in nocturnal voids in a
substantial number of participants without increase in major adverse events [498].
Another SR of comparative trials of men with nocturia as the primary presentation and LUTS
including nocturia or nocturnal polyuria found that antidiuretic therapy using dose titration was more effective
than placebo in relation to nocturnal voiding frequency and duration of undisturbed sleep [495]. Adverse events
include headache, hyponatremia, insomnia, dry mouth, hypertension, abdominal pain, peripheral edema, and
nausea. Three studies evaluating titrated-dose desmopressin in which men were included, reported seven
serious adverse events in 530 patients (1.3%), with one death. There were seventeen cases of hyponatraemia
(3.2%) and seven of hypertension (1.3%). Headache was reported in 53 (10%) and nausea in fifteen (2.8%)
[495]. Hyponatremia is the most important concern, especially in patients > 65 years of age, with potential life
threatening consequences. Baseline values of sodium over 130 mmol/L have been used as inclusion criteria
in some research protocols. Assessment of sodium levels must be undertaken at baseline, after initiation of
treatment or dose titration and during treatment. Desmopressin is not recommended in high-risk groups [495].

MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021 49
Desmopressin oral disintegrating tablets (ODT) have been studied separately in the sex-specific pivotal trials
CS41 and CS40 in patients with nocturia [499, 500]. Almost 87% of included patients had nocturnal polyuria
and approximately 48% of the patients were > 65 years. The co-primary endpoints in both trials were change
in number of nocturia episodes per night from baseline and at least a 33% decrease in the mean number
of nocturnal voids from baseline during three months of treatment. The mean change in nocturia episodes
from baseline was greater with desmopressin ODT compared to placebo (difference: women = -0.3 [95% CI:
-0.5, -0.1]; men = -0.4 [95% CI: -0.6, -0.2]). The 33% responder rate was also greater with desmopressin ODT
compared to placebo (women: 78% vs. 62%; men: 67% vs. 50%).
Analysis of three published placebo-controlled trials of desmopressin ODT for nocturia showed
that clinically significant hyponatraemia was more frequent in patients aged ≥ 65 years than in those aged
< 65 years in all dosage groups, including those receiving the minimum effective dose for desmopressin (11%
of men aged ≥ 65 years vs. 0% of men aged < 65 years receiving 50 mcg; 4% of women ≥ 65 aged years vs.
2% of women aged < 65 years receiving 25 mcg). Severe hyponatraemia, defined as ≤ 125 mmol/L serum
sodium, was rare, affecting 22/1,431 (2%) patients overall [501].
Low dose desmopressin (ODT) has been approved in Europe, Canada and Australia for the
treatment of nocturia with ≥ 2 episodes in gender-specific low doses 50 mcg for men and 25 mcg for women;
however, it initially failed to receive FDA approval, with the FDA citing uncertain benefit relative to risks as the
reason. Following resubmission to the FDA in June 2018 desmopressin acetate sublingual tablet, 50 mcg for
men and 25 mcg for women, was approved for the treatment of nocturia due to nocturnal polyuria in adults
who awaken at least two times per night to void with a boxed warning for hyponatremia.

Desmopressin acetate nasal spray is a new low-dose formulation of desmopressin and differs from other types
of desmopressin formulation due to its bioavailability and route of administration. Desmopressin acetate nasal
spray has been investigated in two RCTs including men and women with nocturia (over two episodes per night)
and a mean age of 66 years. The average benefit of treatment relative to placebo was statistically significant
but low, -0.3 and -0.2 for the 1.5 mcg and 0.75 mcg doses of desmopressin acetate, respectively. The number
of patients with a reduction of more than 50% of nocturia episodes was 48.5% and 37.9%, respectively
compared with 30% in the placebo group [502]. The reported adverse event rate of the studies was rather
low and the risk of hyponatremia was 1.2% and 0.9% for desmopressin acetate 1.5 mcg and 0.75 mcg,
respectively. Desmopressin acetate nasal spray was approved by the FDA in 2017 for the treatment of nocturia
due to nocturnal polyuria, but it is not available in Europe.

Practical considerations
A complete medical assessment should be made, to exclude potentially non-urological underlying causes,
e.g. sleep apnea, before prescribing desmopressin in men with nocturia due to nocturnal polyuria. The optimal
dose differs between patients, in men < 65 years desmopressin treatment should be initiated at a low dose
(0.1 mg/day) and may be gradually increased up to a dosage of 0.4 mg/day every week until maximum efficacy
is reached. Desmopressin is taken once daily before sleeping. Patients should avoid drinking fluids at least one
hour before and for eight hours after dosing. Low dose desmopressin may be prescribed in patients > 65 years.
In men ≥ 65 years or older, low dose desmopressin should not be used if the serum sodium concentration
is below normal: all patients should be monitored for hyponatremia. Urologists should be cautious when
prescribing low-dose desmopressin in patients under-represented in trials (e.g. patients > 75 years) who may
have an increased risk of hyponatremia.

5.5.3.2 Medications to treat LUTD


Where LUTD is diagnosed and considered causative of nocturia, relevant medications for storage (and
voiding) LUTS may be considered. Applicable medications include; selective α1-adrenergic antagonists
[503], antimuscarinics [504-506], 5-ARIs [507] and PDE5Is [508]. However, effect size of these medications
is generally small, or not significantly different from placebo when used to treat nocturia [495]. Data on OAB
medications (antimuscarinics, beta-3 agonist) generally had a female-predominant population. No studies
specifically addressing the impact of OAB medications on nocturia in men were identified [495]. Benefits with
combination therapies were not consistently observed.

5.5.3.3 Other medications


Agents to promote sleep [509], diuretics [510], non-steroidal anti-inflammatory agents (NSAIDs) [511] and
phytotherapy [512] were reported as being associated with response or QoL improvement [495]. Effect size of
these medications in nocturia is generally small, or not significantly different from placebo. Larger responses
have been reported for some medications, but larger scale confirmatory RCTs are lacking. Agents to promote
sleep do not appear to reduce nocturnal voiding frequency, but may help patients return to sleep.

50 MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021
Summary of evidence LE
No clinical trial of pathophysiology-directed primary therapy has been undertaken. 4
No robust clinical trial of behavioural therapy as primary intervention has been undertaken. 4
Antidiuretic therapy reduces nocturnal voiding frequency in men with baseline severity of two or more 1b
voids per night.
There is an increased risk of hyponatremia in patients 65 years of age or older under antidiuretic 1b
therapy.
Antidiuretic therapy increases duration of undisturbed sleep. 1b
Alpha 1-blocker use is associated with improvements in undisturbed sleep duration and nocturnal 2
voiding frequency, which are generally of only marginal clinical significance.
Antimuscarinic medications can reduce night-time urinary urgency severity, but the reduction in overall 2
nocturia frequency is small or non-significant.
Antimuscarinic medications are associated with higher incidence of dry mouth compared with placebo. 2
5α-reductase inhibitors reduce nocturia severity in men with baseline nocturia severity of two or more 2
voids per night.
A trial of timed diuretic therapy may be offered to men with nocturia due to nocturnal polyuria. 1b
Screening for hyponatremia should be undertaken at baseline and during treatment.

Recommendations Strength rating


Treat underlying causes of nocturia, including behavioural, systemic condition(s), sleep Weak
disorders, lower urinary tract dysfunction, or a combination of factors.
Discuss behavioural changes with the patient to reduce nocturnal urine volume and Weak
episodes of nocturia, and improve sleep quality.
Offer desmopressin to decrease nocturia due to nocturnal polyuria in men < 65 years of age. Weak
Offer low dose desmopressin for men > 65 years of age with nocturia at least twice per Weak
night due to nocturnal polyuria.
Screen for hyponatremia at baseline, day three and day seven, one month after initiating Strong
therapy and periodically during treatment. Measure serum sodium more frequently in
patients > 65 years of age and in patients at increased risk of hyponatremia.
Discuss with the patient the potential clinical benefit relative to the associated risks from Strong
the use of desmopressin, especially in men > 65 years of age.
Offer α1-adrenergic antagonists for treating nocturia in men who have nocturia associated Weak
with LUTS.
Offer antimuscarinic drugs for treating nocturia in men who have nocturia associated with Weak
overactive bladder.
Offer 5α-reductase inhibitors for treating nocturia in men who have nocturia associated with Weak
LUTS and an enlarged prostate (> 40 mL).
Do not offer phosphodiesterase type 5 inhibitors for the treatment of nocturia. Weak

6. FOLLOW-UP
6.1 Watchful waiting (behavioural)
Patients who elect to pursue a WW policy should be reviewed at six months and then annually, provided there
is no deterioration of symptoms or development of absolute indications for surgical treatment. The following
are recommended at follow-up visits: history, IPSS, uroflowmetry, and PVR volume.

6.2 Medical treatment


Patients receiving α1-blockers, muscarinic receptor antagonists, beta-3 agonists, PDE5Is or the combination
of α1-blockers and 5-ARIs or muscarinic receptor antagonists should be reviewed four to six weeks after
drug initiation to determine the treatment response. If patients gain symptomatic relief in the absence of
troublesome adverse events, drug therapy may be continued. Patients should be reviewed at six months
and then annually, provided there is no deterioration of symptoms or development of absolute indications
for surgical treatment. The following are recommended at follow-up visits: history, IPSS, uroflowmetry, and
PVR volume. Frequency volume charts or bladder diaries should be used to assess response to treatment for
predominant storage symptoms or nocturnal polyuria.

MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021 51
Patients receiving 5-ARIs should be reviewed after twelve weeks and six months to determine their
response and adverse events. The following are recommended at follow-up visits: history, IPSS, uroflowmetry
and PVR volume. Men taking 5-ARIs should be followed up regularly using serial PSA testing if life expectancy
is greater than ten years and if a diagnosis of PCa could alter management. A new baseline PSA should be
determined at six months, and any confirmed increase in PSA while on 5-ARIs should be evaluated.
In patients receiving desmopressin, serum sodium concentration should be measured at day three
and seven, one month after initiating therapy and periodically during treatment. If serum sodium concentration
has remained normal during periodic screening follow-up screening can be carried out every three months
subsequently. However, serum sodium concentration should be monitored more frequently in patients ≥ 65
years of age and in patients at increased risk of hyponatremia. The following tests are recommended at follow-
up visits: serum-sodium concentration and FVC. The follow-up sequence should be restarted after dose
escalation.

6.3 Surgical treatment


After prostate surgery, patients should be reviewed four to six weeks after catheter removal to evaluate
treatment response and adverse events. If patients have symptomatic relief and are without adverse events,
no further re-assessment is necessary. The following tests are recommended at follow-up visit after four to six
weeks: IPSS, uroflowmetry and PVR volume.

Summary of evidence LE
Follow-up for all conservative, medical, or operative treatment modalities is based on empirical data or 4
theoretical considerations, but not on evidence-based studies.

Recommendations Strength rating


Follow-up all patients who receive conservative, medical or surgical management. Weak
Define follow-up intervals and examinations according to the specific treatment. Weak

7. REFERENCES
1. Committee on Herbal Medicinal Products. European Union herbal monograph on Serenoa repens
(W. Bartram) Small, fructus. EMA/HMPC/280079/2013, 2015.
2. Guyatt, G.H., et al. GRADE: an emerging consensus on rating quality of evidence and strength of
recommendations. BMJ, 2008. 336: 924.
https://www.ncbi.nlm.nih.gov/pubmed/18436948
3. Guyatt, G.H., et al. What is “quality of evidence” and why is it important to clinicians? BMJ, 2008.
336: 995.
https://www.ncbi.nlm.nih.gov/pubmed/18456631
4. Phillips B, et al. Oxford Centre for Evidence-based Medicine Levels of Evidence. Updated by
Jeremy Howick March 2009. 1998.
https://www.cebm.net/2009/06/oxford-centre-evidence-based-medicine-levels-evidence-march-2009/
5. Guyatt, G.H., et al. Going from evidence to recommendations. BMJ, 2008. 336: 1049.
https://www.ncbi.nlm.nih.gov/pubmed/18467413
6. Abrams, P., et al. The standardisation of terminology of lower urinary tract function: report from the
Standardisation Sub-committee of the International Continence Society. Neurourol Urodyn, 2002.
21: 167.
https://www.ncbi.nlm.nih.gov/pubmed/11857671
7. Martin, S.A., et al. Prevalence and factors associated with uncomplicated storage and voiding lower
urinary tract symptoms in community-dwelling Australian men. World J Urol, 2011. 29: 179.
https://www.ncbi.nlm.nih.gov/pubmed/20963421
8. Société Internationale d’Urologie (SIU), Lower Urinary Tract Symptoms (LUTS): An International
Consultation on Male LUTS., C. Chapple & P. Abrams, Editors. 2013.
9. Kupelian, V., et al. Prevalence of lower urinary tract symptoms and effect on quality of life in a
racially and ethnically diverse random sample: the Boston Area Community Health (BACH) Survey.
Arch Intern Med, 2006. 166: 2381.
https://www.ncbi.nlm.nih.gov/pubmed/17130393

52 MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021
10. Agarwal, A., et al. What is the most bothersome lower urinary tract symptom? Individual- and
population-level perspectives for both men and women. Eur Urol, 2014. 65: 1211.
https://www.ncbi.nlm.nih.gov/pubmed/24486308
11. De Ridder, D., et al. Urgency and other lower urinary tract symptoms in men aged >/= 40 years: a
Belgian epidemiological survey using the ICIQ-MLUTS questionnaire. Int J Clin Pract, 2015. 69: 358.
https://www.ncbi.nlm.nih.gov/pubmed/25648652
12. Taub, D.A., et al. The economics of benign prostatic hyperplasia and lower urinary tract symptoms
in the United States. Curr Urol Rep, 2006. 7: 272.
https://www.ncbi.nlm.nih.gov/pubmed/16930498
13. Gacci, M., et al. Metabolic syndrome and benign prostatic enlargement: a systematic review and
meta-analysis. BJU Int, 2015. 115: 24.
https://www.ncbi.nlm.nih.gov/pubmed/24602293
14. Gacci, M., et al. Male Lower Urinary Tract Symptoms and Cardiovascular Events: A Systematic
Review and Meta-analysis. Eur Urol, 2016. 70: 788.
https://www.ncbi.nlm.nih.gov/pubmed/27451136
15. Kogan, M.I., et al. Epidemiology and impact of urinary incontinence, overactive bladder, and other
lower urinary tract symptoms: results of the EPIC survey in Russia, Czech Republic, and Turkey.
Curr Med Res Opin, 2014. 30: 2119.
https://www.ncbi.nlm.nih.gov/pubmed/24932562
16. Chapple, C.R., et al. Lower urinary tract symptoms revisited: a broader clinical perspective. Eur Urol,
2008. 54: 563.
https://www.ncbi.nlm.nih.gov/pubmed/18423969
17. Ficarra, V., et al. The role of inflammation in lower urinary tract symptoms (LUTS) due to benign
prostatic hyperplasia (BPH) and its potential impact on medical therapy. Curr Urol Rep, 2014. 15: 463.
https://www.ncbi.nlm.nih.gov/pubmed/25312251
18. He, Q., et al. Metabolic syndrome, inflammation and lower urinary tract symptoms: possible
translational links. Prostate Cancer Prostatic Dis, 2016. 19: 7.
https://www.ncbi.nlm.nih.gov/pubmed/26391088
19. Drake, M.J. Do we need a new definition of the overactive bladder syndrome? ICI-RS 2013.
Neurourol Urodyn, 2014. 33: 622.
https://www.ncbi.nlm.nih.gov/pubmed/24838519
20. Chapple, C.R., et al. Terminology report from the International Continence Society (ICS) Working
Group on Underactive Bladder (UAB). Neurourol Urodyn, 2018. 37: 2928.
https://www.ncbi.nlm.nih.gov/pubmed/30203560
21. Novara, G., et al. Critical Review of Guidelines for BPH Diagnosis and Treatment Strategy. Eur Urol
Suppl 2006. 4: 418.
https://www.eu-openscience.europeanurology.com/action/showPdf?pii
=S1569-9056%2806%2900012-1
22. McVary, K.T., et al. Update on AUA guideline on the management of benign prostatic hyperplasia.
J Urol, 2011. 185: 1793.
https://www.ncbi.nlm.nih.gov/pubmed/21420124
23. Bosch, J., et al. Etiology, Patient Assessment and Predicting Outcome from Therapy. International
Consultation on Urological Diseases Male LUTS Guideline 2013. 2013 (in press).
https://snucm.elsevierpure.com/en/publications/lower-urinary-tract-symptoms-in-men-male-luts-
etiology-patient-as
24. Martin, R.M., et al. Lower urinary tract symptoms and risk of prostate cancer: the HUNT 2 Cohort,
Norway. Int J Cancer, 2008. 123: 1924.
https://www.ncbi.nlm.nih.gov/pubmed/18661522
25. Young, J.M., et al. Are men with lower urinary tract symptoms at increased risk of prostate cancer?
A systematic review and critique of the available evidence. BJU Int, 2000. 85: 1037.
https://www.ncbi.nlm.nih.gov/pubmed/10848691
26. De Nunzio, C., et al. Erectile Dysfunction and Lower Urinary Tract Symptoms. Eur Urol Focus, 2017.
3: 352.
https://www.ncbi.nlm.nih.gov/pubmed/29191671
27. Barqawi, A.B., et al. Methods of developing UWIN, the modified American Urological Association
symptom score. J Urol, 2011. 186: 940.
https://www.ncbi.nlm.nih.gov/pubmed/21791346

MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021 53
28. Barry, M.J., et al. The American Urological Association symptom index for benign prostatic
hyperplasia. The Measurement Committee of the American Urological Association. J Urol, 1992.
148: 1549.
https://www.ncbi.nlm.nih.gov/pubmed/1279218
29. Donovan, J.L., et al. Scoring the short form ICSmaleSF questionnaire. International Continence
Society. J Urol, 2000. 164: 1948.
https://www.ncbi.nlm.nih.gov/pubmed/11061889
30. Epstein, R.S., et al. Validation of a new quality of life questionnaire for benign prostatic hyperplasia.
J Clin Epidemiol, 1992. 45: 1431.
https://www.ncbi.nlm.nih.gov/pubmed/1281223
31. Homma, Y., et al. Symptom assessment tool for overactive bladder syndrome--overactive bladder
symptom score. Urology, 2006. 68: 318.
https://www.ncbi.nlm.nih.gov/pubmed/16904444
32. Schou, J., et al. The value of a new symptom score (DAN-PSS) in diagnosing uro-dynamic
infravesical obstruction in BPH. Scand J Urol Nephrol, 1993. 27: 489.
https://www.ncbi.nlm.nih.gov/pubmed/7512747
33. Homma, Y., et al. Core Lower Urinary Tract Symptom score (CLSS) questionnaire: a reliable tool in
the overall assessment of lower urinary tract symptoms. Int J Urol, 2008. 15: 816.
https://www.ncbi.nlm.nih.gov/pubmed/18657204
34. D’Silva, K.A., et al. Does this man with lower urinary tract symptoms have bladder outlet
obstruction?: The Rational Clinical Examination: a systematic review. JAMA, 2014. 312: 535.
https://www.ncbi.nlm.nih.gov/pubmed/25096693
35. Bryan, N.P., et al. Frequency volume charts in the assessment and evaluation of treatment: how
should we use them? Eur Urol, 2004. 46: 636.
https://www.ncbi.nlm.nih.gov/pubmed/15474275
36. Gisolf, K.W., et al. Analysis and reliability of data from 24-hour frequency-volume charts in men with
lower urinary tract symptoms due to benign prostatic hyperplasia. Eur Urol, 2000. 38: 45.
https://www.ncbi.nlm.nih.gov/pubmed/10859441
37. Cornu, J.N., et al. A contemporary assessment of nocturia: definition, epidemiology,
pathophysiology, and management--a systematic review and meta-analysis. Eur Urol, 2012.
62: 877.
https://www.ncbi.nlm.nih.gov/pubmed/22840350
38. Weiss, J.P. Nocturia: “do the math”. J Urol, 2006. 175: S16.
https://www.ncbi.nlm.nih.gov/pubmed/16458734
39. Weiss, J.P., et al. Nocturia Think Tank: focus on nocturnal polyuria: ICI-RS 2011. Neurourol Urodyn,
2012. 31: 330.
https://www.ncbi.nlm.nih.gov/pubmed/22415907
40. Vaughan, C.P., et al. Military exposure and urinary incontinence among American men. J Urol, 2014.
191: 125.
https://www.ncbi.nlm.nih.gov/pubmed/23871759
41. Bright, E., et al. Urinary diaries: evidence for the development and validation of diary content,
format, and duration. Neurourol Urodyn, 2011. 30: 348.
https://www.ncbi.nlm.nih.gov/pubmed/21284023
42. Yap, T.L., et al. A systematic review of the reliability of frequency-volume charts in urological
research and its implications for the optimum chart duration. BJU Int, 2007. 99: 9.
https://www.ncbi.nlm.nih.gov/pubmed/16956355
43. Bright, E., et al. Developing and validating the International Consultation on Incontinence
Questionnaire bladder diary. Eur Urol, 2014. 66: 294.
https://www.ncbi.nlm.nih.gov/pubmed/24647230
44. Weissfeld, J.L., et al. Quality control of cancer screening examination procedures in the Prostate,
Lung, Colorectal and Ovarian (PLCO) Cancer Screening Trial. Control Clin Trials, 2000. 21: 390s.
https://www.ncbi.nlm.nih.gov/pubmed/11189690
45. Roehrborn, C.G. Accurate determination of prostate size via digital rectal examination and
transrectal ultrasound. Urology, 1998. 51: 19.
https://www.ncbi.nlm.nih.gov/pubmed/9586592
46. Roehrborn, C.G., et al. Interexaminer reliability and validity of a three-dimensional model to assess
prostate volume by digital rectal examination. Urology, 2001. 57: 1087.
https://www.ncbi.nlm.nih.gov/pubmed/11377314

54 MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021
47. Bosch, J.L., et al. Validity of digital rectal examination and serum prostate specific antigen in the
estimation of prostate volume in community-based men aged 50 to 78 years: the Krimpen Study.
Eur Urol, 2004. 46: 753.
https://www.ncbi.nlm.nih.gov/pubmed/15548443
48. Babjuk, M., et al. EAU Guidelines on Non-muscle-invasive Bladder Cancer Edn. presented at EAU
Annual Congress, Milan, 2021.
https://uroweb.org/guideline/non-muscle-invasive-bladder-cancer/
49. Bonkat, G., et al. EAU Guidelines on Urological Infections Edn. presented at EAU Annual Congress,
Milan, 2021.
https://uroweb.org/guideline/urological-infections/
50. Palou, J., et al. ICUD-EAU International Consultation on Bladder Cancer 2012: Urothelial carcinoma
of the prostate. Eur Urol, 2013. 63: 81.
https://www.ncbi.nlm.nih.gov/pubmed/22938869
51. Roupret, M., et al. EAU Guidelines on Upper Urinary Tract Urothelial Cell Carcinoma Edn. presented
at EAU Annual Congress, Milan, 2021.
https://uroweb.org/guideline/upper-urinary-tract-urothelial-cell-carcinoma/
52. Roehrborn, C.G., et al. Guidelines for the diagnosis and treatment of benign prostatic hyperplasia:
a comparative, international overview. Urology, 2001. 58: 642.
https://www.ncbi.nlm.nih.gov/pubmed/11711329
53. Abrams, P., et al. Evaluation and treatment of lower urinary tract symptoms in older men. J Urol,
2013. 189: S93.
https://www.ncbi.nlm.nih.gov/pubmed/23234640
54. European urinalysis guidelines. Scand J Clin Lab Invest Suppl, 2000. 231: 1.
https://www.ncbi.nlm.nih.gov/pubmed/12647764
55. Khasriya, R., et al. The inadequacy of urinary dipstick and microscopy as surrogate markers of
urinary tract infection in urological outpatients with lower urinary tract symptoms without acute
frequency and dysuria. J Urol, 2010. 183: 1843.
https://www.ncbi.nlm.nih.gov/pubmed/20303096
56. Roehrborn, C.G., et al. Serum prostate-specific antigen as a predictor of prostate volume in men
with benign prostatic hyperplasia. Urology, 1999. 53: 581.
https://www.ncbi.nlm.nih.gov/pubmed/10096388
57. Bohnen, A.M., et al. Serum prostate-specific antigen as a predictor of prostate volume in the
community: the Krimpen study. Eur Urol, 2007. 51: 1645.
https://www.ncbi.nlm.nih.gov/pubmed/17320271
58. Kayikci, A., et al. Free prostate-specific antigen is a better tool than total prostate-specific antigen at
predicting prostate volume in patients with lower urinary tract symptoms. Urology, 2012. 80: 1088.
https://www.ncbi.nlm.nih.gov/pubmed/23107399
59. Morote, J., et al. Prediction of prostate volume based on total and free serum prostate-specific
antigen: is it reliable? Eur Urol, 2000. 38: 91.
https://www.ncbi.nlm.nih.gov/pubmed/10859448
60. Mottet, N., et al. EAU Guidelines on Prostate Cancer Edn. presented at EAU Annual Congress,
Milan, 2021.
https://uroweb.org/guideline/prostate-cancer/
61. Roehrborn, C.G., et al. Serum prostate specific antigen is a strong predictor of future prostate
growth in men with benign prostatic hyperplasia. PROSCAR long-term efficacy and safety study.
J Urol, 2000. 163: 13.
https://www.ncbi.nlm.nih.gov/pubmed/10604304
62. Roehrborn, C.G., et al. Serum prostate-specific antigen and prostate volume predict long-term
changes in symptoms and flow rate: results of a four-year, randomized trial comparing finasteride
versus placebo. PLESS Study Group. Urology, 1999. 54: 662.
https://www.ncbi.nlm.nih.gov/pubmed/10510925
63. Djavan, B., et al. Longitudinal study of men with mild symptoms of bladder outlet obstruction
treated with watchful waiting for four years. Urology, 2004. 64: 1144.
https://www.ncbi.nlm.nih.gov/pubmed/15596187
64. Patel, D.N., et al. PSA predicts development of incident lower urinary tract symptoms: Results from
the REDUCE study. Prostate Cancer Prostatic Dis, 2018. 21: 238.
https://www.ncbi.nlm.nih.gov/pubmed/29795141
65. McConnell, J.D., et al. The long-term effect of doxazosin, finasteride, and combination therapy on
the clinical progression of benign prostatic hyperplasia. N Engl J Med, 2003. 349: 2387.
https://www.ncbi.nlm.nih.gov/pubmed/14681504

MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021 55
66. Roehrborn, C.G. Alfuzosin 10 mg once daily prevents overall clinical progression of benign prostatic
hyperplasia but not acute urinary retention: results of a 2-year placebo-controlled study. BJU Int,
2006. 97: 734.
https://www.ncbi.nlm.nih.gov/pubmed/16536764
67. Jacobsen, S.J., et al. Treatment for benign prostatic hyperplasia among community dwelling men:
the Olmsted County study of urinary symptoms and health status. J Urol, 1999. 162: 1301.
https://www.ncbi.nlm.nih.gov/pubmed/10492184
68. Lim, K.B., et al. Comparison of intravesical prostatic protrusion, prostate volume and serum
prostatic-specific antigen in the evaluation of bladder outlet obstruction. Int J Urol, 2006. 13: 1509.
https://www.ncbi.nlm.nih.gov/pubmed/17118026
69. Meigs, J.B., et al. Risk factors for clinical benign prostatic hyperplasia in a community-based
population of healthy aging men. J Clin Epidemiol, 2001. 54: 935.
https://www.ncbi.nlm.nih.gov/pubmed/11520654
70. Gerber, G.S., et al. Serum creatinine measurements in men with lower urinary tract symptoms
secondary to benign prostatic hyperplasia. Urology, 1997. 49: 697.
https://www.ncbi.nlm.nih.gov/pubmed/9145973
71. Oelke, M., et al. Can we identify men who will have complications from benign prostatic obstruction
(BPO)? ICI-RS 2011. Neurourol Urodyn, 2012. 31: 322.
https://www.ncbi.nlm.nih.gov/pubmed/22415947
72. Comiter, C.V., et al. Urodynamic risk factors for renal dysfunction in men with obstructive and
nonobstructive voiding dysfunction. J Urol, 1997. 158: 181.
https://www.ncbi.nlm.nih.gov/pubmed/9186351
73. Koch, W.F., et al. The outcome of renal ultrasound in the assessment of 556 consecutive patients
with benign prostatic hyperplasia. J Urol, 1996. 155: 186.
https://www.ncbi.nlm.nih.gov/pubmed/7490828
74. Rule, A.D., et al. The association between benign prostatic hyperplasia and chronic kidney disease
in community-dwelling men. Kidney Int, 2005. 67: 2376.
https://www.ncbi.nlm.nih.gov/pubmed/15882282
75. Hong, S.K., et al. Chronic kidney disease among men with lower urinary tract symptoms due to
benign prostatic hyperplasia. BJU Int, 2010. 105: 1424.
https://www.ncbi.nlm.nih.gov/pubmed/19874305
76. Lee, J.H., et al. Relationship of estimated glomerular filtration rate with lower urinary tract
symptoms/benign prostatic hyperplasia measures in middle-aged men with moderate to severe
lower urinary tract symptoms. Urology, 2013. 82: 1381.
https://www.ncbi.nlm.nih.gov/pubmed/24063940
77. Mebust, W.K., et al. Transurethral prostatectomy: immediate and postoperative complications.
A cooperative study of 13 participating institutions evaluating 3,885 patients. J Urol, 1989. 141: 243.
https://www.ncbi.nlm.nih.gov/pubmed/2643719
78. Rule, A.D., et al. Longitudinal changes in post-void residual and voided volume among community
dwelling men. J Urol, 2005. 174: 1317.
https://www.ncbi.nlm.nih.gov/pubmed/16145411
79. Sullivan, M.P., et al. Detrusor contractility and compliance characteristics in adult male patients with
obstructive and nonobstructive voiding dysfunction. J Urol, 1996. 155: 1995.
https://www.ncbi.nlm.nih.gov/pubmed/8618307
80. Oelke, M., et al. Diagnostic accuracy of noninvasive tests to evaluate bladder outlet obstruction in
men: detrusor wall thickness, uroflowmetry, postvoid residual urine, and prostate volume. Eur Urol,
2007. 52: 827.
https://www.ncbi.nlm.nih.gov/pubmed/17207910
81. Emberton, M. Definition of at-risk patients: dynamic variables. BJU Int, 2006. 97 Suppl 2: 12.
https://www.ncbi.nlm.nih.gov/pubmed/16507047
82. Mochtar, C.A., et al. Post-void residual urine volume is not a good predictor of the need for invasive
therapy among patients with benign prostatic hyperplasia. J Urol, 2006. 175: 213.
https://www.ncbi.nlm.nih.gov/pubmed/16406914
83. Jorgensen, J.B., et al. Age-related variation in urinary flow variables and flow curve patterns in
elderly males. Br J Urol, 1992. 69: 265.
https://www.ncbi.nlm.nih.gov/pubmed/1373664
84. Kranse, R., et al. Causes for variability in repeated pressure-flow measurements. Urology, 2003.
61: 930.
https://www.ncbi.nlm.nih.gov/pubmed/12736007

56 MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021
85. Reynard, J.M., et al. The ICS-’BPH’ Study: uroflowmetry, lower urinary tract symptoms and bladder
outlet obstruction. Br J Urol, 1998. 82: 619.
https://www.ncbi.nlm.nih.gov/pubmed/9839573
86. Idzenga, T., et al. Accuracy of maximum flow rate for diagnosing bladder outlet obstruction can be
estimated from the ICS nomogram. Neurourol Urodyn, 2008. 27: 97.
https://www.ncbi.nlm.nih.gov/pubmed/17600368
87. Siroky, M.B., et al. The flow rate nomogram: I. Development. J Urol, 1979. 122: 665.
https://www.ncbi.nlm.nih.gov/pubmed/159366
88. Siroky, M.B., et al. The flow rate nomogram: II. Clinical correlation. J Urol, 1980. 123: 208.
https://www.ncbi.nlm.nih.gov/pubmed/7354519
89. Grossfeld, G.D., et al. Benign prostatic hyperplasia: clinical overview and value of diagnostic
imaging. Radiol Clin North Am, 2000. 38: 31.
https://www.ncbi.nlm.nih.gov/pubmed/10664665
90. Thorpe, A., et al. Benign prostatic hyperplasia. Lancet, 2003. 361: 1359.
https://www.ncbi.nlm.nih.gov/pubmed/12711484
91. Wilkinson, A.G., et al. Is pre-operative imaging of the urinary tract worthwhile in the assessment of
prostatism? Br J Urol, 1992. 70: 53.
https://www.ncbi.nlm.nih.gov/pubmed/1379105
92. Loch, A.C., et al. Technical and anatomical essentials for transrectal ultrasound of the prostate.
World J Urol, 2007. 25: 361.
https://www.ncbi.nlm.nih.gov/pubmed/17701043
93. Stravodimos, K.G., et al. TRUS versus transabdominal ultrasound as a predictor of enucleated
adenoma weight in patients with BPH: a tool for standard preoperative work-up? Int Urol Nephrol,
2009. 41: 767.
https://www.ncbi.nlm.nih.gov/pubmed/19350408
94. Shoukry, I., et al. Role of uroflowmetry in the assessment of lower urinary tract obstruction in adult
males. Br J Urol, 1975. 47: 559.
https://www.ncbi.nlm.nih.gov/pubmed/1191927
95. Anikwe, R.M. Correlations between clinical findings and urinary flow rate in benign prostatic
hypertrophy. Int Surg, 1976. 61: 392.
https://www.ncbi.nlm.nih.gov/pubmed/61184
96. el Din, K.E., et al. The correlation between bladder outlet obstruction and lower urinary tract
symptoms as measured by the international prostate symptom score. J Urol, 1996. 156: 1020.
https://www.ncbi.nlm.nih.gov/pubmed/8709300
97. Oelke, M., et al. Age and bladder outlet obstruction are independently associated with detrusor
overactivity in patients with benign prostatic hyperplasia. Eur Urol, 2008. 54: 419.
https://www.ncbi.nlm.nih.gov/pubmed/18325657
98. Oh, M.M., et al. Is there a correlation between the presence of idiopathic detrusor overactivity and
the degree of bladder outlet obstruction? Urology, 2011. 77: 167.
https://www.ncbi.nlm.nih.gov/pubmed/20934743
99. Jeong, S.J., et al. Prevalence and Clinical Features of Detrusor Underactivity among Elderly with Lower
Urinary Tract Symptoms: A Comparison between Men and Women. Korean J Urol, 2012. 53: 342.
https://www.ncbi.nlm.nih.gov/pubmed/22670194
100. Thomas, A.W., et al. The natural history of lower urinary tract dysfunction in men: the influence of
detrusor underactivity on the outcome after transurethral resection of the prostate with a minimum
10-year urodynamic follow-up. BJU Int, 2004. 93: 745.
https://www.ncbi.nlm.nih.gov/pubmed/15049984
101. Al-Hayek, S., et al. Natural history of detrusor contractility--minimum ten-year urodynamic follow-up in
men with bladder outlet obstruction and those with detrusor. Scand J Urol Nephrol Suppl, 2004: 101.
https://www.ncbi.nlm.nih.gov/pubmed/15545204
102. Thomas, A.W., et al. The natural history of lower urinary tract dysfunction in men: minimum 10-year
urodynamic followup of transurethral resection of prostate for bladder outlet obstruction. J Urol,
2005. 174: 1887.
https://www.ncbi.nlm.nih.gov/pubmed/16217330
103. North Bristol NHS Trust. Urodynamics for Prostate Surgery Trial; Randomised Evaluation of
Assessment Methods (UPSTREAM). ClinicalTrials.gov, 2019. NCT02193451.
https://clinicaltrials.gov/ct2/show/NCT02193451

MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021 57
104. Lewis, A.L., et al. Clinical and Patient-reported Outcome Measures in Men Referred for
Consideration of Surgery to Treat Lower Urinary Tract Symptoms: Baseline Results and Diagnostic
Findings of the Urodynamics for Prostate Surgery Trial; Randomised Evaluation of Assessment
Methods (UPSTREAM). Eur Urol Focus, 2019. 5: 340.
https://www.ncbi.nlm.nih.gov/pubmed/31047905
105. Clement, K.D., et al. Invasive urodynamic studies for the management of lower urinary tract
symptoms (LUTS) in men with voiding dysfunction. Cochrane Database Syst Rev, 2015: CD011179.
https://www.ncbi.nlm.nih.gov/pubmed/25918922
106. Kim, M., et al. Effect of urodynamic preoperative detrusor overactivity on the outcomes of
transurethral surgery in patients with male bladder outlet obstruction: a systematic review and meta-
analysis. World J Urol, 2019. 37: 529.
https://www.ncbi.nlm.nih.gov/pubmed/30006907
107. Blok, B., et al. EAU Guidelines on Neuro-urology Edn. presented at the EAU Annual Congress,
Milan, 2021.
https://uroweb.org/guideline/neuro-urology/
108. Kojima, M., et al. Correlation of presumed circle area ratio with infravesical obstruction in men with
lower urinary tract symptoms. Urology, 1997. 50: 548.
https://www.ncbi.nlm.nih.gov/pubmed/9338730
109. Chia, S.J., et al. Correlation of intravesical prostatic protrusion with bladder outlet obstruction. BJU
Int, 2003. 91: 371.
https://www.ncbi.nlm.nih.gov/pubmed/12603417
110. Keqin, Z., et al. Clinical significance of intravesical prostatic protrusion in patients with benign
prostatic enlargement. Urology, 2007. 70: 1096.
https://www.ncbi.nlm.nih.gov/pubmed/18158025
111. Mariappan, P., et al. Intravesical prostatic protrusion is better than prostate volume in predicting the
outcome of trial without catheter in white men presenting with acute urinary retention: a prospective
clinical study. J Urol, 2007. 178: 573.
https://www.ncbi.nlm.nih.gov/pubmed/17570437
112. Tan, Y.H., et al. Intravesical prostatic protrusion predicts the outcome of a trial without catheter
following acute urine retention. J Urol, 2003. 170: 2339.
https://www.ncbi.nlm.nih.gov/pubmed/14634410
113. Arnolds, M., et al. Positioning invasive versus noninvasive urodynamics in the assessment of
bladder outlet obstruction. Curr Opin Urol, 2009. 19: 55.
https://www.ncbi.nlm.nih.gov/pubmed/19057217
114. Manieri, C., et al. The diagnosis of bladder outlet obstruction in men by ultrasound measurement of
bladder wall thickness. J Urol, 1998. 159: 761.
https://www.ncbi.nlm.nih.gov/pubmed/9474143
115. Kessler, T.M., et al. Ultrasound assessment of detrusor thickness in men-can it predict bladder
outlet obstruction and replace pressure flow study? J Urol, 2006. 175: 2170.
https://www.ncbi.nlm.nih.gov/pubmed/16697831
116. Blatt, A.H., et al. Ultrasound measurement of bladder wall thickness in the assessment of voiding
dysfunction. J Urol, 2008. 179: 2275.
https://www.ncbi.nlm.nih.gov/pubmed/18423703
117. Oelke, M. International Consultation on Incontinence-Research Society (ICI-RS) report on non-
invasive urodynamics: the need of standardization of ultrasound bladder and detrusor wall thickness
measurements to quantify bladder wall hypertrophy. Neurourol Urodyn, 2010. 29: 634.
https://www.ncbi.nlm.nih.gov/pubmed/20432327
118. Kojima, M., et al. Ultrasonic estimation of bladder weight as a measure of bladder hypertrophy in
men with infravesical obstruction: a preliminary report. Urology, 1996. 47: 942.
https://www.ncbi.nlm.nih.gov/pubmed/8677600
119. Kojima, M., et al. Noninvasive quantitative estimation of infravesical obstruction using ultrasonic
measurement of bladder weight. J Urol, 1997. 157: 476.
https://www.ncbi.nlm.nih.gov/pubmed/8996337
120. Akino, H., et al. Ultrasound-estimated bladder weight predicts risk of surgery for benign prostatic
hyperplasia in men using alpha-adrenoceptor blocker for LUTS. Urology, 2008. 72: 817.
https://www.ncbi.nlm.nih.gov/pubmed/18597835
121. McIntosh, S.L., et al. Noninvasive assessment of bladder contractility in men. J Urol, 2004. 172: 1394.
https://www.ncbi.nlm.nih.gov/pubmed/15371853

58 MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021
122. Drinnan, M.J., et al. Inter-observer agreement in the estimation of bladder pressure using a penile
cuff. Neurourol Urodyn, 2003. 22: 296.
https://www.ncbi.nlm.nih.gov/pubmed/12808703
123. Griffiths, C.J., et al. A nomogram to classify men with lower urinary tract symptoms using urine flow
and noninvasive measurement of bladder pressure. J Urol, 2005. 174: 1323.
https://www.ncbi.nlm.nih.gov/pubmed/16145412
124. Clarkson, B., et al. Continuous non-invasive measurement of bladder voiding pressure using an
experimental constant low-flow test. Neurourol Urodyn, 2012. 31: 557.
https://www.ncbi.nlm.nih.gov/pubmed/22190105
125. Van Mastrigt, R., et al. Towards a noninvasive urodynamic diagnosis of infravesical obstruction. BJU
Int, 1999. 84: 195.
https://www.ncbi.nlm.nih.gov/pubmed/10444152
126. Pel, J.J., et al. Development of a non-invasive strategy to classify bladder outlet obstruction in male
patients with LUTS. Neurourol Urodyn, 2002. 21: 117.
https://www.ncbi.nlm.nih.gov/pubmed/11857664
127. Shinbo, H., et al. Application of ultrasonography and the resistive index for evaluating bladder outlet
obstruction in patients with benign prostatic hyperplasia. Curr Urol Rep, 2011. 12: 255.
https://www.ncbi.nlm.nih.gov/pubmed/21475953
128. Ku, J.H., et al. Correlation between prostatic urethral angle and bladder outlet obstruction index in
patients with lower urinary tract symptoms. Urology, 2010. 75: 1467.
https://www.ncbi.nlm.nih.gov/pubmed/19962734
129. Malde, S., et al. Systematic Review of the Performance of Noninvasive Tests in Diagnosing Bladder
Outlet Obstruction in Men with Lower Urinary Tract Symptoms. Eur Urol, 2016.
https://www.ncbi.nlm.nih.gov/pubmed/27687821
130. Ball, A.J., et al. The natural history of untreated “prostatism”. Br J Urol, 1981. 53: 613.
https://www.ncbi.nlm.nih.gov/pubmed/6172172
131. Kirby, R.S. The natural history of benign prostatic hyperplasia: what have we learned in the last
decade? Urology, 2000. 56: 3.
https://www.ncbi.nlm.nih.gov/pubmed/11074195
132. Isaacs, J.T. Importance of the natural history of benign prostatic hyperplasia in the evaluation of
pharmacologic intervention. Prostate Suppl, 1990. 3: 1.
https://www.ncbi.nlm.nih.gov/pubmed/1689166
133. Netto, N.R., Jr., et al. Evaluation of patients with bladder outlet obstruction and mild international
prostate symptom score followed up by watchful waiting. Urology, 1999. 53: 314.
https://www.ncbi.nlm.nih.gov/pubmed/9933046
134. Flanigan, R.C., et al. 5-year outcome of surgical resection and watchful waiting for men with
moderately symptomatic benign prostatic hyperplasia: a Department of Veterans Affairs cooperative
study. J Urol, 1998. 160: 12.
https://www.ncbi.nlm.nih.gov/pubmed/9628595
135. Wasson, J.H., et al. A comparison of transurethral surgery with watchful waiting for moderate
symptoms of benign prostatic hyperplasia. The Veterans Affairs Cooperative Study Group on
Transurethral Resection of the Prostate. N Engl J Med, 1995. 332: 75.
https://www.ncbi.nlm.nih.gov/pubmed/7527493
136. Brown, C.T., et al. Self management for men with lower urinary tract symptoms: randomised
controlled trial. Bmj, 2007. 334: 25.
https://www.ncbi.nlm.nih.gov/pubmed/17118949
137. Yap, T.L., et al. The impact of self-management of lower urinary tract symptoms on frequency-
volume chart measures. BJU Int, 2009. 104: 1104.
https://www.ncbi.nlm.nih.gov/pubmed/19485993
138. Brown, C.T., et al. Defining the components of a self-management programme for men with
uncomplicated lower urinary tract symptoms: a consensus approach. Eur Urol, 2004. 46: 254.
https://www.ncbi.nlm.nih.gov/pubmed/15245822
139. Michel, M.C., et al. Alpha1-, alpha2- and beta-adrenoceptors in the urinary bladder, urethra and
prostate. Br J Pharmacol, 2006. 147 Suppl 2: S88.
https://www.ncbi.nlm.nih.gov/pubmed/16465187
140. Kortmann, B.B., et al. Urodynamic effects of alpha-adrenoceptor blockers: a review of clinical trials.
Urology, 2003. 62: 1.
https://www.ncbi.nlm.nih.gov/pubmed/12837408

MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021 59
141. Barendrecht, M.M., et al. Do alpha1-adrenoceptor antagonists improve lower urinary tract
symptoms by reducing bladder outlet resistance? Neurourol Urodyn, 2008. 27: 226.
https://www.ncbi.nlm.nih.gov/pubmed/17638312
142. Djavan, B., et al. State of the art on the efficacy and tolerability of alpha1-adrenoceptor antagonists
in patients with lower urinary tract symptoms suggestive of benign prostatic hyperplasia. Urology,
2004. 64: 1081.
https://www.ncbi.nlm.nih.gov/pubmed/15596173
143. Michel, M.C., et al. Comparison of tamsulosin efficacy in subgroups of patients with lower urinary
tract symptoms. Prostate Cancer Prostatic Dis, 1998. 1: 332.
https://www.ncbi.nlm.nih.gov/pubmed/12496876
144. Fusco, F., et al. alpha1-Blockers Improve Benign Prostatic Obstruction in Men with Lower Urinary
Tract Symptoms: A Systematic Review and Meta-analysis of Urodynamic Studies. Eur Urol, 2016.
69: 1091.
https://www.ncbi.nlm.nih.gov/pubmed/26831507
145. Boyle, P., et al. Meta-analysis of randomized trials of terazosin in the treatment of benign prostatic
hyperplasia. Urology, 2001. 58: 717.
https://www.ncbi.nlm.nih.gov/pubmed/11711348
146. Roehrborn, C.G. Three months’ treatment with the alpha1-blocker alfuzosin does not affect total or
transition zone volume of the prostate. Prostate Cancer Prostatic Dis, 2006. 9: 121.
https://www.ncbi.nlm.nih.gov/pubmed/16304557
147. Roehrborn, C.G., et al. The effects of dutasteride, tamsulosin and combination therapy on lower
urinary tract symptoms in men with benign prostatic hyperplasia and prostatic enlargement: 2-year
results from the CombAT study. J Urol, 2008. 179: 616.
https://www.ncbi.nlm.nih.gov/pubmed/18082216
148. Roehrborn, C.G., et al. The effects of combination therapy with dutasteride and tamsulosin on
clinical outcomes in men with symptomatic benign prostatic hyperplasia: 4-year results from the
CombAT study. Eur Urol, 2010. 57: 123.
https://www.ncbi.nlm.nih.gov/pubmed/19825505
149. Karavitakis, M., et al. Management of Urinary Retention in Patients with Benign Prostatic
Obstruction: A Systematic Review and Meta-analysis. Eur Urol, 2019.
https://www.ncbi.nlm.nih.gov/pubmed/30773327
150. Nickel, J.C., et al. A meta-analysis of the vascular-related safety profile and efficacy of alpha-
adrenergic blockers for symptoms related to benign prostatic hyperplasia. Int J Clin Pract, 2008.
62: 1547.
https://www.ncbi.nlm.nih.gov/pubmed/18822025
151. Barendrecht, M.M., et al. Treatment of lower urinary tract symptoms suggestive of benign prostatic
hyperplasia: the cardiovascular system. BJU Int, 2005. 95 Suppl 4: 19.
https://www.ncbi.nlm.nih.gov/pubmed/15871732
152. Chapple, C.R., et al. Silodosin therapy for lower urinary tract symptoms in men with suspected
benign prostatic hyperplasia: results of an international, randomized, double-blind, placebo- and
active-controlled clinical trial performed in Europe. Eur Urol, 2011. 59: 342.
https://www.ncbi.nlm.nih.gov/pubmed/21109344
153. Welk, B., et al. The risk of fall and fracture with the initiation of a prostate-selective alpha antagonist:
a population based cohort study. BMJ, 2015. 351: h5398.
https://www.ncbi.nlm.nih.gov/pubmed/26502947
154. Chang, D.F., et al. Intraoperative floppy iris syndrome associated with tamsulosin. J Cataract Refract
Surg, 2005. 31: 664.
https://www.ncbi.nlm.nih.gov/pubmed/15899440
155. Chatziralli, I.P., et al. Risk factors for intraoperative floppy iris syndrome: a meta-analysis.
Ophthalmology, 2011. 118: 730.
https://www.ncbi.nlm.nih.gov/pubmed/21168223
156. van Dijk, M.M., et al. Effects of alpha(1)-adrenoceptor antagonists on male sexual function. Drugs,
2006. 66: 287.
https://www.ncbi.nlm.nih.gov/pubmed/16526818
157. Gacci, M., et al. Impact of medical treatments for male lower urinary tract symptoms due to benign
prostatic hyperplasia on ejaculatory function: a systematic review and meta-analysis. J Sex Med,
2014. 11: 1554.
https://www.ncbi.nlm.nih.gov/pubmed/24708055

60 MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021
158. Andriole, G., et al. Dihydrotestosterone and the prostate: the scientific rationale for 5alpha-
reductase inhibitors in the treatment of benign prostatic hyperplasia. J Urol, 2004. 172: 1399.
https://www.ncbi.nlm.nih.gov/pubmed/15371854
159. Rittmaster, R.S., et al. Evidence for atrophy and apoptosis in the prostates of men given finasteride.
J Clin Endocrinol Metab, 1996. 81: 814.
https://www.ncbi.nlm.nih.gov/pubmed/8636309
160. Naslund, M.J., et al. A review of the clinical efficacy and safety of 5alpha-reductase inhibitors for the
enlarged prostate. Clin Ther, 2007. 29: 17.
https://www.ncbi.nlm.nih.gov/pubmed/17379044
161. Andersen, J.T., et al. Can finasteride reverse the progress of benign prostatic hyperplasia? A
two-year placebo-controlled study. The Scandinavian BPH Study Group. Urology, 1995. 46: 631.
https://www.ncbi.nlm.nih.gov/pubmed/7495111
162. Kirby, R.S., et al. Efficacy and tolerability of doxazosin and finasteride, alone or in combination, in
treatment of symptomatic benign prostatic hyperplasia: the Prospective European Doxazosin and
Combination Therapy (PREDICT) trial. Urology, 2003. 61: 119.
https://www.ncbi.nlm.nih.gov/pubmed/12559281
163. Lepor, H., et al. The efficacy of terazosin, finasteride, or both in benign prostatic hyperplasia.
Veterans Affairs Cooperative Studies Benign Prostatic Hyperplasia Study Group. N Engl J Med,
1996. 335: 533.
https://www.ncbi.nlm.nih.gov/pubmed/8684407
164. Marberger, M.J. Long-term effects of finasteride in patients with benign prostatic hyperplasia: a
double-blind, placebo-controlled, multicenter study. PROWESS Study Group. Urology, 1998. 51: 677.
https://www.ncbi.nlm.nih.gov/pubmed/9610579
165. McConnell, J.D., et al. The effect of finasteride on the risk of acute urinary retention and the need
for surgical treatment among men with benign prostatic hyperplasia. Finasteride Long-Term Efficacy
and Safety Study Group. N Engl J Med, 1998. 338: 557.
https://www.ncbi.nlm.nih.gov/pubmed/9475762
166. Nickel, J.C., et al. Efficacy and safety of finasteride therapy for benign prostatic hyperplasia: results
of a 2-year randomized controlled trial (the PROSPECT study). PROscar Safety Plus Efficacy
Canadian Two year Study. CMAJ, 1996. 155: 1251.
https://www.ncbi.nlm.nih.gov/pubmed/8911291
167. Roehrborn, C.G., et al. Efficacy and safety of a dual inhibitor of 5-alpha-reductase types 1 and 2
(dutasteride) in men with benign prostatic hyperplasia. Urology, 2002. 60: 434.
https://www.ncbi.nlm.nih.gov/pubmed/12350480
168. Nickel, J.C., et al. Comparison of dutasteride and finasteride for treating benign prostatic
hyperplasia: the Enlarged Prostate International Comparator Study (EPICS). BJU Int, 2011. 108: 388.
https://www.ncbi.nlm.nih.gov/pubmed/21631695
169. Boyle, P., et al. Prostate volume predicts outcome of treatment of benign prostatic hyperplasia with
finasteride: meta-analysis of randomized clinical trials. Urology, 1996. 48: 398.
https://www.ncbi.nlm.nih.gov/pubmed/8804493
170. Gittelman, M., et al. Dutasteride improves objective and subjective disease measures in men with
benign prostatic hyperplasia and modest or severe prostate enlargement. J Urol, 2006. 176: 1045.
https://www.ncbi.nlm.nih.gov/pubmed/16890688
171. Roehrborn, C.G., et al. Long-term sustained improvement in symptoms of benign prostatic
hyperplasia with the dual 5alpha-reductase inhibitor dutasteride: results of 4-year studies. BJU Int,
2005. 96: 572.
https://www.ncbi.nlm.nih.gov/pubmed/16104912
172. Roehrborn, C.G., et al. The influence of baseline parameters on changes in international prostate
symptom score with dutasteride, tamsulosin, and combination therapy among men with
symptomatic benign prostatic hyperplasia and an enlarged prostate: 2-year data from the CombAT
study. Eur Urol, 2009. 55: 461.
https://www.ncbi.nlm.nih.gov/pubmed/19013011
173. Roehrborn, C.G. BPH progression: concept and key learning from MTOPS, ALTESS, COMBAT, and
ALF-ONE. BJU Int, 2008. 101 Suppl 3: 17.
https://www.ncbi.nlm.nih.gov/pubmed/18307681
174. Andersen, J.T., et al. Finasteride significantly reduces acute urinary retention and need for surgery in
patients with symptomatic benign prostatic hyperplasia. Urology, 1997. 49: 839.
https://www.ncbi.nlm.nih.gov/pubmed/9187688

MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021 61
175. Kirby, R.S., et al. Long-term urodynamic effects of finasteride in benign prostatic hyperplasia: a pilot
study. Eur Urol, 1993. 24: 20.
https://www.ncbi.nlm.nih.gov/pubmed/7689971
176. Tammela, T.L., et al. Long-term effects of finasteride on invasive urodynamics and symptoms in the
treatment of patients with bladder outflow obstruction due to benign prostatic hyperplasia. J Urol,
1995. 154: 1466.
https://www.ncbi.nlm.nih.gov/pubmed/7544845
177. Donohue, J.F., et al. Transurethral prostate resection and bleeding: a randomized, placebo
controlled trial of role of finasteride for decreasing operative blood loss. J Urol, 2002. 168: 2024.
https://www.ncbi.nlm.nih.gov/pubmed/12394700
178. Khwaja, M.A., et al. The Effect of Two Weeks Preoperative Finasteride Therapy in Reducing Prostate
Vascularity. J Coll Physicians Surg Pak, 2016. 26: 213.
https://www.ncbi.nlm.nih.gov/pubmed/26975954
179. Corona, G., et al. Sexual dysfunction in subjects treated with inhibitors of 5alpha-reductase for
benign prostatic hyperplasia: a comprehensive review and meta-analysis. Andrology, 2017. 5: 671.
https://www.ncbi.nlm.nih.gov/pubmed/28453908
180. Andriole, G.L., et al. Effect of dutasteride on the risk of prostate cancer. N Engl J Med, 2010. 362: 1192.
https://www.ncbi.nlm.nih.gov/pubmed/20357281
181. Thompson, I.M., et al. The influence of finasteride on the development of prostate cancer. N Engl
J Med, 2003. 349: 215.
https://www.ncbi.nlm.nih.gov/pubmed/12824459
182. Hsieh, T.F., et al. Use of 5-alpha-reductase inhibitors did not increase the risk of cardiovascular
diseases in patients with benign prostate hyperplasia: a five-year follow-up study. PLoS One, 2015.
10: e0119694.
https://www.ncbi.nlm.nih.gov/pubmed/25803433
183. Skeldon, S.C., et al. The Cardiovascular Safety of Dutasteride. J Urol, 2017. 197: 1309.
https://www.ncbi.nlm.nih.gov/pubmed/27866006
184. Chess-Williams, R., et al. The minor population of M3-receptors mediate contraction of human
detrusor muscle in vitro. J Auton Pharmacol, 2001. 21: 243.
https://www.ncbi.nlm.nih.gov/pubmed/12123469
185. Matsui, M., et al. Multiple functional defects in peripheral autonomic organs in mice lacking muscarinic
acetylcholine receptor gene for the M3 subtype. Proc Natl Acad Sci U S A, 2000. 97: 9579.
https://www.ncbi.nlm.nih.gov/pubmed/10944224
186. Kono, M., et al. Central muscarinic receptor subtypes regulating voiding in rats. J Urol, 2006. 175: 353.
https://www.ncbi.nlm.nih.gov/pubmed/16406941
187. Wuest, M., et al. Effect of rilmakalim on detrusor contraction in the presence and absence of
urothelium. Naunyn Schmiedebergs Arch Pharmacol, 2005. 372: 203.
https://www.ncbi.nlm.nih.gov/pubmed/16283254
188. Goldfischer, E.R., et al. Efficacy and safety of oxybutynin topical gel 3% in patients with
urgency and/or mixed urinary incontinence: A randomized, double-blind, placebo-controlled study.
Neurourol Urodyn, 2015. 34: 37.
https://www.ncbi.nlm.nih.gov/pubmed/24133005
189. Baldwin, C.M., et al. Transdermal oxybutynin. Drugs, 2009. 69: 327.
https://www.ncbi.nlm.nih.gov/pubmed/19275276
190. Chapple, C.R., et al. A shifted paradigm for the further understanding, evaluation, and treatment of
lower urinary tract symptoms in men: focus on the bladder. Eur Urol, 2006. 49: 651.
https://www.ncbi.nlm.nih.gov/pubmed/16530611
191. Michel, M.C., et al. Does gender or age affect the efficacy and safety of tolterodine? J Urol, 2002.
168: 1027.
https://www.ncbi.nlm.nih.gov/pubmed/12187215
192. Chapple, C., et al. Fesoterodine clinical efficacy and safety for the treatment of overactive bladder in
relation to patient profiles: a systematic review. Curr Med Res Opin, 2015. 31: 1201.
https://www.ncbi.nlm.nih.gov/pubmed/25798911
193. Dmochowski, R., et al. Efficacy and tolerability of tolterodine extended release in male and female
patients with overactive bladder. Eur Urol, 2007. 51: 1054.
https://www.ncbi.nlm.nih.gov/pubmed/17097217
194. Herschorn, S., et al. Efficacy and tolerability of fesoterodine in men with overactive bladder: a
pooled analysis of 2 phase III studies. Urology, 2010. 75: 1149.
https://www.ncbi.nlm.nih.gov/pubmed/19914702

62 MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021
195. Hofner, K., et al. Safety and efficacy of tolterodine extended release in men with overactive bladder
symptoms and presumed non-obstructive benign prostatic hyperplasia. World J Urol, 2007. 25: 627.
https://www.ncbi.nlm.nih.gov/pubmed/17906864
196. Roehrborn, C.G., et al. Efficacy and tolerability of tolterodine extended-release in men with
overactive bladder and urgency urinary incontinence. BJU Int, 2006. 97: 1003.
https://www.ncbi.nlm.nih.gov/pubmed/16643482
197. Kaplan, S.A., et al. Tolterodine and tamsulosin for treatment of men with lower urinary tract
symptoms and overactive bladder: a randomized controlled trial. Jama, 2006. 296: 2319.
https://www.ncbi.nlm.nih.gov/pubmed/17105794
198. Kaplan, S.A., et al. Tolterodine extended release attenuates lower urinary tract symptoms in men
with benign prostatic hyperplasia. J Urol, 2005. 174: 2273.
https://www.ncbi.nlm.nih.gov/pubmed/16280803
199. Kaplan, S.A., et al. Solifenacin treatment in men with overactive bladder: effects on symptoms and
patient-reported outcomes. Aging Male, 2010. 13: 100.
https://www.ncbi.nlm.nih.gov/pubmed/20001469
200. Roehrborn, C.G., et al. Effects of serum PSA on efficacy of tolterodine extended release with or
without tamsulosin in men with LUTS, including OAB. Urology, 2008. 72: 1061.
https://www.ncbi.nlm.nih.gov/pubmed/18817961
201. Yokoyama, T., et al. Naftopidil and propiverine hydrochloride for treatment of male lower urinary
tract symptoms suggestive of benign prostatic hyperplasia and concomitant overactive bladder:
a prospective randomized controlled study. Scand J Urol Nephrol, 2009. 43: 307.
https://www.ncbi.nlm.nih.gov/pubmed/19396723
202. Abrams, P., et al. Safety and tolerability of tolterodine for the treatment of overactive bladder in men
with bladder outlet obstruction. J Urol, 2006. 175: 999.
https://www.ncbi.nlm.nih.gov/pubmed/16469601
203. Andersson, K.E. On the Site and Mechanism of Action of beta3-Adrenoceptor Agonists in the
Bladder. Int Neurourol J, 2017. 21: 6.
https://www.ncbi.nlm.nih.gov/pubmed/28361520
204. Chapple, C.R., et al. Randomized double-blind, active-controlled phase 3 study to assess 12-month
safety and efficacy of mirabegron, a beta(3)-adrenoceptor agonist, in overactive bladder. Eur Urol,
2013. 63: 296.
https://www.ncbi.nlm.nih.gov/pubmed/23195283
205. Herschorn, S., et al. A phase III, randomized, double-blind, parallel-group, placebo-controlled,
multicentre study to assess the efficacy and safety of the beta(3) adrenoceptor agonist, mirabegron,
in patients with symptoms of overactive bladder. Urology, 2013. 82: 313.
https://www.ncbi.nlm.nih.gov/pubmed/23769122
206. Khullar, V., et al. Efficacy and tolerability of mirabegron, a beta(3)-adrenoceptor agonist, in patients
with overactive bladder: results from a randomised European-Australian phase 3 trial. Eur Urol,
2013. 63: 283.
https://www.ncbi.nlm.nih.gov/pubmed/23182126
207. Nitti, V.W., et al. Results of a randomized phase III trial of mirabegron in patients with overactive
bladder. J Urol, 2013. 189: 1388.
https://www.ncbi.nlm.nih.gov/pubmed/23079373
208. Yamaguchi, O., et al. Efficacy and Safety of the Selective beta3 -Adrenoceptor Agonist Mirabegron
in Japanese Patients with Overactive Bladder: A Randomized, Double-Blind, Placebo-Controlled,
Dose-Finding Study. Low Urin Tract Symptoms, 2015. 7: 84.
https://www.ncbi.nlm.nih.gov/pubmed/26663687
209. Sebastianelli, A., et al. Systematic review and meta-analysis on the efficacy and tolerability
of mirabegron for the treatment of storage lower urinary tract symptoms/overactive bladder:
Comparison with placebo and tolterodine. Int J Urol, 2018. 25: 196.
https://www.ncbi.nlm.nih.gov/pubmed/29205506
210. Liao, C.H., et al. Mirabegron 25mg Monotherapy Is Safe but Less Effective in Male Patients With
Overactive Bladder and Bladder Outlet Obstruction. Urology, 2018. 117: 115.
https://www.ncbi.nlm.nih.gov/pubmed/29630956
211. Drake, M.J., et al. Efficacy and Safety of Mirabegron Add-on Therapy to Solifenacin in Incontinent
Overactive Bladder Patients with an Inadequate Response to Initial 4-Week Solifenacin
Monotherapy: A Randomised Double-blind Multicentre Phase 3B Study (BESIDE). Eur Urol, 2016.
70: 136.
https://www.ncbi.nlm.nih.gov/pubmed/26965560

MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021 63
212. Kuo, H.C., et al. Results of a randomized, double-blind, parallel-group, placebo- and active-
controlled, multicenter study of mirabegron, a beta3-adrenoceptor agonist, in patients with
overactive bladder in Asia. Neurourol Urodyn, 2015. 34: 685.
https://www.ncbi.nlm.nih.gov/pubmed/25130281
213. Abrams, P., et al. Combination treatment with mirabegron and solifenacin in patients with overactive
bladder: exploratory responder analyses of efficacy and evaluation of patient-reported outcomes
from a randomized, double-blind, factorial, dose-ranging, Phase II study (SYMPHONY). World
J Urol, 2017. 35: 827.
https://www.ncbi.nlm.nih.gov/pubmed/27514371
214. Khullar, V., et al. Patient-reported outcomes with the beta3 -adrenoceptor agonist mirabegron in a
phase III trial in patients with overactive bladder. Neurourol Urodyn, 2016. 35: 987.
https://www.ncbi.nlm.nih.gov/pubmed/26288118
215. Yamaguchi, O., et al. Safety and efficacy of mirabegron as ‘add-on’ therapy in patients with
overactive bladder treated with solifenacin: a post-marketing, open-label study in Japan (MILAI
study). BJU Int, 2015. 116: 612.
https://www.ncbi.nlm.nih.gov/pubmed/25639296
216. Ichihara, K., et al. A randomized controlled study of the efficacy of tamsulosin monotherapy and its
combination with mirabegron for overactive bladder induced by benign prostatic obstruction. J Urol,
2015. 193: 921.
https://www.ncbi.nlm.nih.gov/pubmed/25254938
217. Matsuo, T., et al. The efficacy of mirabegron additional therapy for lower urinary tract symptoms
after treatment with alpha1-adrenergic receptor blocker monotherapy: Prospective analysis of
elderly men. BMC Urol, 2016. 16: 45.
https://www.ncbi.nlm.nih.gov/pubmed/27473059
218. Matsukawa, Y., et al. Comparison in the efficacy of fesoterodine or mirabegron add-on therapy
to silodosin for patients with benign prostatic hyperplasia complicated by overactive bladder:
A randomized, prospective trial using urodynamic studies. Neurourol Urodyn, 2019.
https://www.ncbi.nlm.nih.gov/pubmed/30779375
219. White, W.B., et al. Cardiovascular safety of mirabegron: analysis of an integrated clinical trial
database of patients with overactive bladder syndrome. J Am Soc Hyperten, 2018. 12: 768.
https://www.ncbi.nlm.nih.gov/pubmed/30181042
220. Nitti, V.W., et al. Urodynamics and safety of the beta(3)-adrenoceptor agonist mirabegron in males
with lower urinary tract symptoms and bladder outlet obstruction. J Urol, 2013. 190: 1320.
https://www.ncbi.nlm.nih.gov/pubmed/23727415
221. Lee, Y.K., et al. Safety and therapeutic efficacy of mirabegron 25 mg in older patients with
overactive bladder and multiple comorbidities. Geriatr Gerontol Int, 2018. 18: 1330.
https://www.ncbi.nlm.nih.gov/pubmed/29931793
222. Wagg, A., et al. Oral pharmacotherapy for overactive bladder in older patients: mirabegron as a
potential alternative to antimuscarinics. Current Med Res Opin, 2016. 32: 621.
https://www.ncbi.nlm.nih.gov/pubmed/26828974
223. Herschorn, S., et al. Efficacy and safety of combinations of mirabegron and solifenacin compared
with monotherapy and placebo in patients with overactive bladder (SYNERGY study). BJU Int, 2017.
120: 562.
https://www.ncbi.nlm.nih.gov/pubmed/28418102
224. Chapple, C.R., et al. Persistence and Adherence with Mirabegron versus Antimuscarinic Agents in
Patients with Overactive Bladder: A Retrospective Observational Study in UK Clinical Practice. Eur
Urol, 2017. 72: 389.
https://www.ncbi.nlm.nih.gov/pubmed/28196724
225. Van Gelderen, M., et al. Absence of clinically relevant cardiovascular interaction upon add-on of
mirabegron or tamsulosin to an established tamsulosin or mirabegron treatment in healthy middle-
aged to elderly men. Int J Clin Pharmacol Ther, 2014. 52: 693.
https://www.ncbi.nlm.nih.gov/pubmed/24755125
226. Giuliano, F., et al. The mechanism of action of phosphodiesterase type 5 inhibitors in the treatment
of lower urinary tract symptoms related to benign prostatic hyperplasia. Eur Urol, 2013. 63: 506.
https://www.ncbi.nlm.nih.gov/pubmed/23018163
227. Morelli, A., et al. Phosphodiesterase type 5 expression in human and rat lower urinary tract tissues
and the effect of tadalafil on prostate gland oxygenation in spontaneously hypertensive rats. J Sex
Med, 2011. 8: 2746.
https://www.ncbi.nlm.nih.gov/pubmed/21812935

64 MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021
228. Vignozzi, L., et al. PDE5 inhibitors blunt inflammation in human BPH: a potential mechanism of
action for PDE5 inhibitors in LUTS. Prostate, 2013. 73: 1391.
https://www.ncbi.nlm.nih.gov/pubmed/23765639
229. Pattanaik, S., et al. Phosphodiesterase inhibitors for lower urinary tract symptoms consistent with
benign prostatic hyperplasia. Cochrane Database Syst Rev, 2018. 2018: CD010060.
https://www.ncbi.nlm.nih.gov/pubmed/30480763
230. Gacci, M., et al. A systematic review and meta-analysis on the use of phosphodiesterase
5 inhibitors alone or in combination with alpha-blockers for lower urinary tract symptoms due to
benign prostatic hyperplasia. Eur Urol, 2012. 61: 994.
https://www.ncbi.nlm.nih.gov/pubmed/22405510
231. Wang, Y., et al. Tadalafil 5 mg Once Daily Improves Lower Urinary Tract Symptoms and Erectile
Dysfunction: A Systematic Review and Meta-analysis. Low Urin Tract Symptoms, 2018. 10: 84.
https://www.ncbi.nlm.nih.gov/pubmed/29341503
232. Oelke, M., et al. Monotherapy with tadalafil or tamsulosin similarly improved lower urinary tract
symptoms suggestive of benign prostatic hyperplasia in an international, randomised, parallel,
placebo-controlled clinical trial. Eur Urol, 2012. 61: 917.
https://www.ncbi.nlm.nih.gov/pubmed/22297243
233. Oelke, M., et al. Time to onset of clinically meaningful improvement with tadalafil 5 mg once daily
for lower urinary tract symptoms secondary to benign prostatic hyperplasia: analysis of data pooled
from 4 pivotal, double-blind, placebo controlled studies. J Urol, 2015. 193: 1581.
https://www.ncbi.nlm.nih.gov/pubmed/25437533
234. Donatucci, C.F., et al. Tadalafil administered once daily for lower urinary tract symptoms secondary
to benign prostatic hyperplasia: a 1-year, open-label extension study. BJU Int, 2011. 107: 1110.
https://www.ncbi.nlm.nih.gov/pubmed/21244606
235. Porst, H., et al. Efficacy and safety of tadalafil 5 mg once daily for lower urinary tract symptoms
suggestive of benign prostatic hyperplasia: subgroup analyses of pooled data from 4 multinational,
randomized, placebo-controlled clinical studies. Urology, 2013. 82: 667.
https://www.ncbi.nlm.nih.gov/pubmed/23876588
236. Vlachopoulos, C., et al. Impact of cardiovascular risk factors and related comorbid conditions and
medical therapy reported at baseline on the treatment response to tadalafil 5 mg once-daily in men
with lower urinary tract symptoms associated with benign prostatic hyperplasia: an integrated analysis
of four randomised, double-blind, placebo-controlled, clinical trials. Int J Clin Pract, 2015. 69: 1496.
https://www.ncbi.nlm.nih.gov/pubmed/26299520
237. Brock, G.B., et al. Direct effects of tadalafil on lower urinary tract symptoms versus indirect effects
mediated through erectile dysfunction symptom improvement: integrated data analyses from
4 placebo controlled clinical studies. J Urol, 2014. 191: 405.
https://www.ncbi.nlm.nih.gov/pubmed/24096120
238. Roehrborn, C.G., et al. Effects of tadalafil once daily on maximum urinary flow rate in men with
lower urinary tract symptoms suggestive of benign prostatic hyperplasia. J Urol, 2014. 191: 1045.
https://www.ncbi.nlm.nih.gov/pubmed/24445278
239. Oelke, M., et al. Efficacy and safety of tadalafil 5 mg once daily in the treatment of lower urinary
tract symptoms associated with benign prostatic hyperplasia in men aged >=75 years: integrated
analyses of pooled data from multinational, randomized, placebo-controlled clinical studies. BJU
Int, 2017. 119: 793.
https://www.ncbi.nlm.nih.gov/pubmed/27988986
240. Matsukawa, Y., et al. Effects of tadalafil on storage and voiding function in patients with male lower
urinary tract symptoms suggestive of benign prostatic hyperplasia: A urodynamic-based study. Int
J Urol, 2018. 25: 246.
https://www.ncbi.nlm.nih.gov/pubmed/29164680
241. Casabe, A., et al. Efficacy and safety of the coadministration of tadalafil once daily with finasteride
for 6 months in men with lower urinary tract symptoms and prostatic enlargement secondary to
benign prostatic hyperplasia. J Urol, 2014. 191: 727.
https://www.ncbi.nlm.nih.gov/pubmed/24096118
242. Gacci, M., et al. The use of a single daily dose of tadalafil to treat signs and symptoms of benign
prostatic hyperplasia and erectile dysfunction. Res Rep Urol, 2013. 5: 99.
https://www.ncbi.nlm.nih.gov/pubmed/24400241
243. Madersbacher, S., et al. Plant extracts: sense or nonsense? Curr Opin Urol, 2008. 18: 16.
https://www.ncbi.nlm.nih.gov/pubmed/18090484

MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021 65
244. Buck, A.C. Is there a scientific basis for the therapeutic effects of serenoa repens in benign prostatic
hyperplasia? Mechanisms of action. J Urol, 2004. 172: 1792.
https://www.ncbi.nlm.nih.gov/pubmed/15540722
245. Levin, R.M., et al. A scientific basis for the therapeutic effects of Pygeum africanum and Serenoa
repens. Urol Res, 2000. 28: 201.
https://www.ncbi.nlm.nih.gov/pubmed/10929430
246. Habib, F.K., et al. Not all brands are created equal: a comparison of selected components of
different brands of Serenoa repens extract. Prostate Cancer Prostatic Dis, 2004. 7: 195.
https://www.ncbi.nlm.nih.gov/pubmed/15289814
247. Scaglione, F., et al. Comparison of the potency of different brands of Serenoa repens extract on
5alpha-reductase types I and II in prostatic co-cultured epithelial and fibroblast cells. Pharmacology,
2008. 82: 270.
https://www.ncbi.nlm.nih.gov/pubmed/18849646
248. De Monte, C., et al. Modern extraction techniques and their impact on the pharmacological profile of
Serenoa repens extracts for the treatment of lower urinary tract symptoms. BMC Urol, 2014. 14: 63.
https://www.ncbi.nlm.nih.gov/pubmed/25112532
249. Committee on Herbal Medicinal Products. Community herbal monograph on Cucurbita pepo L.,
semen. EMA/HMPC/136024/2010, 2012.
https://www.ema.europa.eu/en/documents/herbal-monograph/final-community-herbal-monograph-
cucurbita-pepo-l-semen_en.pdf
250. Committee on Herbal Medicinal Products. European Union herbal monograph on Prunus africana
(Hook f.) Kalkm., cortex. EMA/HMPC/680626/2013, 2016.
https://www.ema.europa.eu/en/documents/herbal-monograph/draft-european-union-herbal-
monograph-prunus-africana-hook-f-kalkm-cortex_en.pdf
251. Committee on Herbal Medicinal Products. Community herbal monograph on Urtica dioica L., Urtica
urens L., their hybrids or their mixtures, radix. EMA/HMPC/461160/2008, 2012.
https://www.ema.europa.eu/en/documents/herbal-monograph/final-community-herbal-monograph-
urtica-dioica-l-urtica-urens-l-their-hybrids-their-mixtures-radix_en.pdf
252. Committee on Herbal Medicinal Products. European Union herbal monograph on Epilobium
angustifolium L. and/or Epilobium parviflorum Schreb., herba. EMA/HMPC/712511/2014, 2015.
https://www.ema.europa.eu/en/documents/herbal-monograph/final-european-union-herbal-
monograph-epilobium-angustifolium-l/epilobium-parviflorum-schreb-herba_en.pdf
253. Tacklind, J., et al. Serenoa repens for benign prostatic hyperplasia. Cochrane Database Syst Rev,
2009: CD001423.
https://www.ncbi.nlm.nih.gov/pubmed/19370565
254. Novara, G., et al. Efficacy and Safety of Hexanic Lipidosterolic Extract of Serenoa repens (Permixon)
in the Treatment of Lower Urinary Tract Symptoms Due to Benign Prostatic Hyperplasia: Systematic
Review and Meta-analysis of Randomized Controlled Trials. Eur Urol Focus, 2016. 2: 553.
https://www.ncbi.nlm.nih.gov/pubmed/28723522
255. Vela-Navarrete, R., et al. Efficacy and safety of a hexanic extract of Serenoa repens (Permixon(R))
for the treatment of lower urinary tract symptoms associated with benign prostatic hyperplasia
(LUTS/BPH): systematic review and meta-analysis of randomised controlled trials and observational
studies. BJU Int, 2018. 122: 1049.
https://www.ncbi.nlm.nih.gov/pubmed/29694707
256. Russo, G.I., et al. Clinical Efficacy of Serenoa repens Versus Placebo Versus Alpha-blockers for the
Treatment of Lower Urinary Tract Symptoms/Benign Prostatic Enlargement: A Systematic Review
and Network Meta-analysis of Randomized Placebo-controlled Clinical Trials. Eur Urol Focus, 2020.
https://www.ncbi.nlm.nih.gov/pubmed/31952967
257. Boeri, L., et al. Clinically Meaningful Improvements in LUTS/BPH Severity in Men Treated with
Silodosin Plus Hexanic Extract of Serenoa Repens or Silodosin Alone. Sci Rep, 2017. 7: 15179.
https://www.ncbi.nlm.nih.gov/pubmed/29123161
258. Debruyne, F.M., et al. Sustained-release alfuzosin, finasteride and the combination of both in the
treatment of benign prostatic hyperplasia. European ALFIN Study Group. Eur Urol, 1998. 34: 169.
https://www.ncbi.nlm.nih.gov/pubmed/9732187
259. Barkin, J., et al. Alpha-blocker therapy can be withdrawn in the majority of men following initial
combination therapy with the dual 5alpha-reductase inhibitor dutasteride. Eur Urol, 2003. 44: 461.
https://www.ncbi.nlm.nih.gov/pubmed/14499682
260. Nickel, J.C., et al. Finasteride monotherapy maintains stable lower urinary tract symptoms in men with
benign prostatic hyperplasia following cessation of alpha blockers. Can Urol Assoc J, 2008. 2: 16.
https://www.ncbi.nlm.nih.gov/pubmed/18542722

66 MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021
261. Athanasopoulos, A., et al. Combination treatment with an alpha-blocker plus an anticholinergic for
bladder outlet obstruction: a prospective, randomized, controlled study. J Urol, 2003. 169: 2253.
https://www.ncbi.nlm.nih.gov/pubmed/12771763
262. Roehrborn, C.G., et al. Efficacy and safety of a fixed-dose combination of dutasteride and
tamsulosin treatment (Duodart((R)) ) compared with watchful waiting with initiation of tamsulosin
therapy if symptoms do not improve, both provided with lifestyle advice, in the management of
treatment-naive men with moderately symptomatic benign prostatic hyperplasia: 2-year CONDUCT
study results. BJU Int, 2015. 116: 450.
https://www.ncbi.nlm.nih.gov/pubmed/25565364
263. Roehrborn, C.G., et al. Influence of baseline variables on changes in International Prostate
Symptom Score after combined therapy with dutasteride plus tamsulosin or either monotherapy in
patients with benign prostatic hyperplasia and lower urinary tract symptoms: 4-year results of the
CombAT study. BJU Int, 2014. 113: 623.
https://www.ncbi.nlm.nih.gov/pubmed/24127818
264. Kaplan, S.A., et al. Time Course of Incident Adverse Experiences Associated with Doxazosin,
Finasteride and Combination Therapy in Men with Benign Prostatic Hyperplasia: The MTOPS Trial.
J Urol, 2016. 195: 1825.
https://www.ncbi.nlm.nih.gov/pubmed/26678956
265. Chapple, C., et al. Tolterodine treatment improves storage symptoms suggestive of overactive
bladder in men treated with alpha-blockers. Eur Urol, 2009. 56: 534.
https://www.ncbi.nlm.nih.gov/pubmed/19070418
266. Kaplan, S.A., et al. Safety and tolerability of solifenacin add-on therapy to alpha-blocker treated
men with residual urgency and frequency. J Urol, 2009. 182: 2825.
https://www.ncbi.nlm.nih.gov/pubmed/19837435
267. Lee, J.Y., et al. Comparison of doxazosin with or without tolterodine in men with symptomatic
bladder outlet obstruction and an overactive bladder. BJU Int, 2004. 94: 817.
https://www.ncbi.nlm.nih.gov/pubmed/15476515
268. Lee, K.S., et al. Combination treatment with propiverine hydrochloride plus doxazosin controlled
release gastrointestinal therapeutic system formulation for overactive bladder and coexisting benign
prostatic obstruction: a prospective, randomized, controlled multicenter study. J Urol, 2005. 174: 1334.
https://www.ncbi.nlm.nih.gov/pubmed/16145414
269. MacDiarmid, S.A., et al. Efficacy and safety of extended-release oxybutynin in combination with
tamsulosin for treatment of lower urinary tract symptoms in men: randomized, double-blind,
placebo-controlled study. Mayo Clin Proc, 2008. 83: 1002.
https://www.ncbi.nlm.nih.gov/pubmed/18775200
270. Saito, H., et al. A comparative study of the efficacy and safety of tamsulosin hydrochloride (Harnal
capsules) alone and in combination with propiverine hydrochloride (BUP-4 tablets) in patients with
prostatic hypertrophy associated with pollakisuria and/or urinary incontinence. Jpn J Urol Surg,
1999. 12: 525. [No abstract available].
271. Yang, Y., et al. Efficacy and safety of combined therapy with terazosin and tolteradine for patients
with lower urinary tract symptoms associated with benign prostatic hyperplasia: a prospective
study. Chin Med J (Engl), 2007. 120: 370.
https://www.ncbi.nlm.nih.gov/pubmed/17376305
272. Maruyama, O., et al. Naftopidil monotherapy vs naftopidil and an anticholinergic agent combined
therapy for storage symptoms associated with benign prostatic hyperplasia: A prospective
randomized controlled study. Int J Urol, 2006. 13: 1280.
https://www.ncbi.nlm.nih.gov/pubmed/17010005
273. Lee, H.N., et al. Rate and associated factors of solifenacin add-on after tamsulosin monotherapy in
men with voiding and storage lower urinary tract symptoms. Int J Clin Pract, 2015. 69: 444.
https://www.ncbi.nlm.nih.gov/pubmed/25363606
274. van Kerrebroeck, P., et al. Combination therapy with solifenacin and tamsulosin oral controlled
absorption system in a single tablet for lower urinary tract symptoms in men: efficacy and safety
results from the randomised controlled NEPTUNE trial. Eur Urol, 2013. 64: 1003.
https://www.ncbi.nlm.nih.gov/pubmed/23932438
275. Kaplan, S.A., et al. Add-on fesoterodine for residual storage symptoms suggestive of overactive
bladder in men receiving alpha-blocker treatment for lower urinary tract symptoms. BJU Int, 2012.
109: 1831.
https://www.ncbi.nlm.nih.gov/pubmed/21966995

MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021 67
276. Kim, T.H., et al. Comparison of the efficacy and safety of tolterodine 2 mg and 4 mg combined
with an alpha-blocker in men with lower urinary tract symptoms (LUTS) and overactive bladder: a
randomized controlled trial. BJU Int, 2016. 117: 307.
https://www.ncbi.nlm.nih.gov/pubmed/26305143
277. Athanasopoulos, A., et al. The role of antimuscarinics in the management of men with symptoms
of overactive bladder associated with concomitant bladder outlet obstruction: an update. Eur Urol,
2011. 60: 94.
https://www.ncbi.nlm.nih.gov/pubmed/21497434
278. Kaplan, S.A., et al. Antimuscarinics for treatment of storage lower urinary tract symptoms in men:
a systematic review. Int J Clin Pract, 2011. 65: 487.
https://www.ncbi.nlm.nih.gov/pubmed/21210910
279. Kim, H.J., et al. Efficacy and Safety of Initial Combination Treatment of an Alpha Blocker with
an Anticholinergic Medication in Benign Prostatic Hyperplasia Patients with Lower Urinary Tract
Symptoms: Updated Meta-Analysis. PLoS One, 2017. 12: e0169248.
https://www.ncbi.nlm.nih.gov/pubmed/28072862
280. Van Kerrebroeck, P., et al. Efficacy and safety of solifenacin plus tamsulosin OCAS in men with
voiding and storage lower urinary tract symptoms: results from a phase 2, dose-finding study
(SATURN). Eur Urol, 2013. 64: 398.
https://www.ncbi.nlm.nih.gov/pubmed/23537687
281. Drake, M.J., et al. Long-term safety and efficacy of single-tablet combinations of solifenacin and
tamsulosin oral controlled absorption system in men with storage and voiding lower urinary tract
symptoms: Results from the NEPTUNE study and NEPTUNE II open-label extension. Eur Urol, 2015.
67: 262.
https://www.ncbi.nlm.nih.gov/pubmed/25070148
282. Drake, M.J., et al. Responder and health-related quality of life analyses in men with lower urinary
tract symptoms treated with a fixed-dose combination of solifenacin and tamsulosin OCAS: results
from the NEPTUNE study. BJU Int, 2015.
https://www.ncbi.nlm.nih.gov/pubmed/25907003
283. Drake, M.J., et al. Incidence of urinary retention during treatment with single tablet combinations
of solifenacin+tamsulosin OCAS for up to 1 year in adult men with both storage and voiding LUTS:
A subanalysis of the NEPTUNE/NEPTUNE II randomized controlled studies. PLoS One, 2017.
12: e0170726.
https://www.ncbi.nlm.nih.gov/pubmed/28166296
284. Gong, M., et al. Tamsulosin combined with solifenacin versus tamsulosin monotherapy for male
lower urinary tract symptoms: a meta-analysis. Curr Med Res Opin, 2015. 31: 1781.
https://www.ncbi.nlm.nih.gov/pubmed/26211817
285. Kaplan, S.A., et al. Solifenacin plus tamsulosin combination treatment in men with lower urinary
tract symptoms and bladder outlet obstruction: a randomized controlled trial. Eur Urol, 2013. 63:
158.
https://www.ncbi.nlm.nih.gov/pubmed/22831853
286. Speakman, M.J., et al. What Is the Required Certainty of Evidence for the Implementation of Novel
Techniques for the Treatment of Benign Prostatic Obstruction? Eur Urol Focus, 2019. 5: 351.
https://www.ncbi.nlm.nih.gov/pubmed/31204291
287. Issa, M.M. Technological advances in transurethral resection of the prostate: bipolar versus
monopolar TURP. J Endourol, 2008. 22: 1587.
https://www.ncbi.nlm.nih.gov/pubmed/18721041
288. Rassweiler, J., et al. Bipolar transurethral resection of the prostate--technical modifications and early
clinical experience. Minim Invasive Ther Allied Technol, 2007. 16: 11.
https://www.ncbi.nlm.nih.gov/pubmed/17365673
289. Cornu, J.N., et al. A Systematic Review and Meta-analysis of Functional Outcomes and
Complications Following Transurethral Procedures for Lower Urinary Tract Symptoms Resulting
from Benign Prostatic Obstruction: An Update. Eur Urol, 2015. 67: 1066.
https://www.ncbi.nlm.nih.gov/pubmed/24972732
290. Reich, O., et al. Techniques and long-term results of surgical procedures for BPH. Eur Urol, 2006.
49: 970.
https://www.ncbi.nlm.nih.gov/pubmed/16481092
291. Madersbacher, S., et al. Is transurethral resection of the prostate still justified? BJU Int, 1999. 83: 227.
https://www.ncbi.nlm.nih.gov/pubmed/10233485

68 MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021
292. Madersbacher, S., et al. Reoperation, myocardial infarction and mortality after transurethral and
open prostatectomy: a nation-wide, long-term analysis of 23,123 cases. Eur Urol, 2005. 47: 499.
https://www.ncbi.nlm.nih.gov/pubmed/15774249
293. Eredics, K., et al. Reoperation Rates and Mortality After Transurethral and Open Prostatectomy in a
Long-term Nationwide Analysis: Have We Improved Over a Decade? Urology, 2018. 118: 152.
https://www.ncbi.nlm.nih.gov/pubmed/29733869
294. Alexander, C.E., et al. Bipolar versus monopolar transurethral resection of the prostate for lower
urinary tract symptoms secondary to benign prostatic obstruction. Cochrane Database Syst Rev,
2019. CD009629.
https://www.cochranelibrary.com/cdsr/doi/10.1002/14651858.CD009629.pub4/full
295. Burke, N., et al. Systematic review and meta-analysis of transurethral resection of the prostate
versus minimally invasive procedures for the treatment of benign prostatic obstruction. Urology,
2010. 75: 1015.
https://www.ncbi.nlm.nih.gov/pubmed/19854492
296. Mamoulakis, C., et al. Bipolar versus monopolar transurethral resection of the prostate: a systematic
review and meta-analysis of randomized controlled trials. Eur Urol, 2009. 56: 798.
https://www.ncbi.nlm.nih.gov/pubmed/19595501
297. Omar, M.I., et al. Systematic review and meta-analysis of the clinical effectiveness of bipolar
compared with monopolar transurethral resection of the prostate (TURP). BJU Int, 2014. 113: 24.
https://www.ncbi.nlm.nih.gov/pubmed/24053602
298. Inzunza, G., et al. Bipolar or monopolar transurethral resection for benign prostatic hyperplasia?
Medwave, 2018. 18: e7134.
https://www.ncbi.nlm.nih.gov/pubmed/29351269
299. Treharne, C., et al. Economic Value of the Transurethral Resection in Saline System for Treatment
of Benign Prostatic Hyperplasia in England and Wales: Systematic Review, Meta-analysis, and
Cost-Consequence Model. Eur Urol Focus, 2016.
https://www.ncbi.nlm.nih.gov/pubmed/28753756
300. Autorino, R., et al. Four-year outcome of a prospective randomised trial comparing bipolar
plasmakinetic and monopolar transurethral resection of the prostate. Eur Urol, 2009. 55: 922.
https://www.ncbi.nlm.nih.gov/pubmed/19185975
301. Chen, Q., et al. Bipolar transurethral resection in saline vs traditional monopolar resection of the
prostate: results of a randomized trial with a 2-year follow-up. BJU Int, 2010. 106: 1339.
https://www.ncbi.nlm.nih.gov/pubmed/20477825
302. Fagerstrom, T., et al. Complications and clinical outcome 18 months after bipolar and monopolar
transurethral resection of the prostate. J Endourol, 2011. 25: 1043.
https://www.ncbi.nlm.nih.gov/pubmed/21568691
303. Geavlete, B., et al. Bipolar plasma vaporization vs monopolar and bipolar TURP-A prospective,
randomized, long-term comparison. Urology, 2011. 78: 930.
https://www.ncbi.nlm.nih.gov/pubmed/21802121
304. Giulianelli, R., et al. Comparative randomized study on the efficaciousness of endoscopic bipolar
prostate resection versus monopolar resection technique. 3 year follow-up. Arch Ital Urol Androl,
2013. 85: 86.
https://www.ncbi.nlm.nih.gov/pubmed/23820656
305. Mamoulakis, C., et al. Midterm results from an international multicentre randomised controlled trial
comparing bipolar with monopolar transurethral resection of the prostate. Eur Urol, 2013. 63: 667.
https://www.ncbi.nlm.nih.gov/pubmed/23102675
306. Xie, C.Y., et al. Five-year follow-up results of a randomized controlled trial comparing bipolar
plasmakinetic and monopolar transurethral resection of the prostate. Yonsei Med J, 2012. 53: 734.
https://www.ncbi.nlm.nih.gov/pubmed/22665339
307. Komura, K., et al. Incidence of urethral stricture after bipolar transurethral resection of the prostate
using TURis: results from a randomised trial. BJU Int, 2015. 115: 644.
https://www.ncbi.nlm.nih.gov/pubmed/24909399
308. Kumar, N., et al. Prospective Randomized Comparison of Monopolar TURP, Bipolar TURP and
Photoselective Vaporization of the Prostate in Patients with Benign Prostatic Obstruction: 36
Months Outcome. Low Urin Tract Sympt, 2018. 10: 17.
https://www.ncbi.nlm.nih.gov/pubmed/27168018
309. National Institute for Health and Care Excellence. The TURis system for transurethral resection of
the prostate. NICE GUidelines, 2015.
https://www.nice.org.uk/guidance/mtg23

MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021 69
310. Reich, O., et al. Morbidity, mortality and early outcome of transurethral resection of the prostate: a
prospective multicenter evaluation of 10,654 patients. J Urol, 2008. 180: 246.
https://www.ncbi.nlm.nih.gov/pubmed/18499179
311. Rassweiler, J., et al. Complications of transurethral resection of the prostate (TURP)--incidence,
management, and prevention. Eur Urol, 2006. 50: 969.
https://www.ncbi.nlm.nih.gov/pubmed/16469429
312. Stucki, P., et al. Bipolar versus monopolar transurethral resection of the prostate: a prospective
randomized trial focusing on bleeding complications. J Urol, 2015. 193: 1371.
https://www.ncbi.nlm.nih.gov/pubmed/25464004
313. Akman, T., et al. Effects of bipolar and monopolar transurethral resection of the prostate on urinary
and erectile function: a prospective randomized comparative study. BJU Int, 2013. 111: 129.
https://www.ncbi.nlm.nih.gov/pubmed/22672229
314. El-Assmy, A., et al. Erectile and ejaculatory functions changes following bipolar versus monopolar
transurethral resection of the prostate: a prospective randomized study. Int Urol Nephrol, 2018.
50: 1569.
https://www.ncbi.nlm.nih.gov/pubmed/30083842
315. Mamoulakis, C., et al. Bipolar vs monopolar transurethral resection of the prostate: evaluation of
the impact on overall sexual function in an international randomized controlled trial setting. BJU Int,
2013. 112: 109.
https://www.ncbi.nlm.nih.gov/pubmed/23490008
316. Riedinger, C.B., et al. The impact of surgical duration on complications after transurethral resection
of the prostate: an analysis of NSQIP data. Prostate Cancer Prostatic Dis, 2019. 22: 303.
https://www.ncbi.nlm.nih.gov/pubmed/30385836
317. Bach, T., et al. Laser treatment of benign prostatic obstruction: basics and physical differences.
Eur Urol, 2012. 61: 317.
https://www.ncbi.nlm.nih.gov/pubmed/22033173
318. Xia, S.J., et al. Thulium laser versus standard transurethral resection of the prostate: a randomized
prospective trial. Eur Urol, 2008. 53: 382.
https://www.ncbi.nlm.nih.gov/pubmed/17566639
319. Jiang, H., et al. Safety and Efficacy of Thulium Laser Prostatectomy Versus Transurethral Resection
of Prostate for Treatment of Benign Prostate Hyperplasia: A Meta-Analysis. Low Urin Tract Sympt,
2016. 8: 165.
https://www.ncbi.nlm.nih.gov/pubmed/27619781
320. Zhang, X., et al. Different lasers in the treatment of benign prostatic hyperplasia: a network meta-
analysis. Sci Rep, 2016. 6: 23503.
https://www.ncbi.nlm.nih.gov/pubmed/27009501
321. Zhu, Y., et al. Thulium laser versus standard transurethral resection of the prostate for benign
prostatic obstruction: a systematic review and meta-analysis. World J Urol, 2015. 33: 509.
https://www.ncbi.nlm.nih.gov/pubmed/25298242
322. Zhao, C., et al. Thulium Laser Resection Versus Plasmakinetic Resection of Prostates in the
Treatment of Benign Prostate Hyperplasia: A Meta-Analysis. J Laparoen Adv Surg Tech Part A,
2016. 26: 789.
https://www.ncbi.nlm.nih.gov/pubmed/27500451
323. Deng, Z., et al. Thulium laser VapoResection of the prostate versus traditional transurethral resection
of the prostate or transurethral plasmakinetic resection of prostate for benign prostatic obstruction:
a systematic review and meta-analysis. World J Urol, 2018. 36: 1355.
https://www.ncbi.nlm.nih.gov/pubmed/29651642
324. Lan, Y., et al. Thulium (Tm:YAG) laser vaporesection of prostate and bipolar transurethral resection
of prostate in patients with benign prostate hyperplasia: a systematic review and meta-analysis.
Lasers Med Sci, 2018. 33: 1411.
https://www.ncbi.nlm.nih.gov/pubmed/29947009
325. Cui, D., et al. A randomized trial comparing thulium laser resection to standard transurethral
resection of the prostate for symptomatic benign prostatic hyperplasia: four-year follow-up results.
World J Urol, 2014. 32: 683.
https://www.ncbi.nlm.nih.gov/pubmed/23913094
326. Sun, F., et al. Long-term results of thulium laser resection of the prostate: a prospective study at
multiple centers. World J Urol, 2015. 33: 503.
https://www.ncbi.nlm.nih.gov/pubmed/25487702

70 MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021
327. Yang, Z., et al. Thulium laser enucleation versus plasmakinetic resection of the prostate: a
randomized prospective trial with 18-month follow-up. Urology, 2013. 81: 396.
https://www.ncbi.nlm.nih.gov/pubmed/23374815
328. Wei, H., et al. Thulium laser resection versus plasmakinetic resection of prostates larger than 80 ml.
World J Urol, 2014. 32: 1077.
https://www.ncbi.nlm.nih.gov/pubmed/24264126
329. Sener, T.E., et al. Thulium laser vaporesection of the prostate: Can we operate without interrupting
oral antiplatelet/ anticoagulant therapy? Invest Clin Urol, 2017. 58: 192.
https://www.ncbi.nlm.nih.gov/pubmed/28480345
330. Bansal, A., et al. Holmium Laser vs Monopolar Electrocautery Bladder Neck Incision for Prostates
Less Than 30 Grams: A Prospective Randomized Trial. Urology, 2016. 93: 158.
https://www.ncbi.nlm.nih.gov/pubmed/27058689
331. Lourenco, T., et al. The clinical effectiveness of transurethral incision of the prostate: a systematic
review of randomised controlled trials. World J Urol, 2010. 28: 23.
https://www.ncbi.nlm.nih.gov/pubmed/20033744
332. Kuntz, R.M., et al. Holmium laser enucleation of the prostate versus open prostatectomy for
prostates greater than 100 grams: 5-year follow-up results of a randomised clinical trial. Eur Urol,
2008. 53: 160.
https://www.ncbi.nlm.nih.gov/pubmed/17869409
333. Naspro, R., et al. Holmium laser enucleation of the prostate versus open prostatectomy for
prostates >70 g: 24-month follow-up. Eur Urol, 2006. 50: 563.
https://www.ncbi.nlm.nih.gov/pubmed/16713070
334. Skolarikos, A., et al. Eighteen-month results of a randomized prospective study comparing
transurethral photoselective vaporization with transvesical open enucleation for prostatic adenomas
greater than 80 cc. J Endourol, 2008. 22: 2333.
https://www.ncbi.nlm.nih.gov/pubmed/18837655
335. Varkarakis, I., et al. Long-term results of open transvesical prostatectomy from a contemporary
series of patients. Urology, 2004. 64: 306.
https://www.ncbi.nlm.nih.gov/pubmed/15302484
336. Gratzke, C., et al. Complications and early postoperative outcome after open prostatectomy in
patients with benign prostatic enlargement: results of a prospective multicenter study. J Urol, 2007.
177: 1419.
https://www.ncbi.nlm.nih.gov/pubmed/17382744
337. Chen, S., et al. Plasmakinetic enucleation of the prostate compared with open prostatectomy for
prostates larger than 100 grams: a randomized noninferiority controlled trial with long-term results at
6 years. Eur Urol, 2014. 66: 284.
https://www.ncbi.nlm.nih.gov/pubmed/24502959
338. Li, M., et al. Endoscopic enucleation versus open prostatectomy for treating large benign prostatic
hyperplasia: a meta-analysis of randomized controlled trials. PLoS One, 2015. 10: e0121265.
https://www.ncbi.nlm.nih.gov/pubmed/25826453
339. Lin, Y., et al. Transurethral enucleation of the prostate versus transvesical open prostatectomy for
large benign prostatic hyperplasia: a systematic review and meta-analysis of randomized controlled
trials. World J Urol, 2016. 34: 1207.
https://www.ncbi.nlm.nih.gov/pubmed/26699627
340. Ou, R., et al. Transurethral enucleation and resection of the prostate vs transvesical prostatectomy
for prostate volumes >80 mL: a prospective randomized study. BJU Int, 2013. 112: 239.
https://www.ncbi.nlm.nih.gov/pubmed/23795788
341. Rao, J.M., et al. Plasmakinetic enucleation of the prostate versus transvesical open prostatectomy
for benign prostatic hyperplasia >80 mL: 12-month follow-up results of a randomized clinical trial.
Urology, 2013. 82: 176.
https://www.ncbi.nlm.nih.gov/pubmed/23601443
342. Geavlete, B., et al. Bipolar vaporization, resection, and enucleation versus open prostatectomy:
optimal treatment alternatives in large prostate cases? J Endourol, 2015. 29: 323.
https://www.ncbi.nlm.nih.gov/pubmed/25111385
343. Geavlete, B., et al. Bipolar plasma enucleation of the prostate vs open prostatectomy in large benign
prostatic hyperplasia cases - a medium term, prospective, randomized comparison. BJU Int, 2013.
111: 793.
https://www.ncbi.nlm.nih.gov/pubmed/23469933

MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021 71
344. Salonia, A., et al. Holmium laser enucleation versus open prostatectomy for benign prostatic
hyperplasia: an inpatient cost analysis. Urology, 2006. 68: 302.
https://www.ncbi.nlm.nih.gov/pubmed/16904441
345. Zhang, Y., et al. [Transurethral holmium laser enucleation for prostate adenoma greater than 100 g].
Zhonghua Nan Ke Xue, 2007. 13: 1091.
https://www.ncbi.nlm.nih.gov/pubmed/18284057
346. Tubaro, A., et al. A prospective study of the safety and efficacy of suprapubic transvesical
prostatectomy in patients with benign prostatic hyperplasia. J Urol, 2001. 166: 172.
https://www.ncbi.nlm.nih.gov/pubmed/11435849
347. Neill, M.G., et al. Randomized trial comparing holmium laser enucleation of prostate with
plasmakinetic enucleation of prostate for treatment of benign prostatic hyperplasia. Urology, 2006.
68: 1020.
https://www.ncbi.nlm.nih.gov/pubmed/17095078
348. Li, K., et al. A Novel Modification of Transurethral Enucleation and Resection of the Prostate in
Patients With Prostate Glands Larger than 80 mL: Surgical Procedures and Clinical Outcomes.
Urology, 2018. 113: 153.
https://www.ncbi.nlm.nih.gov/pubmed/29203184
349. Zhao, Z., et al. A prospective, randomised trial comparing plasmakinetic enucleation to standard
transurethral resection of the prostate for symptomatic benign prostatic hyperplasia: three-year
follow-up results. Eur Urol, 2010. 58: 752.
https://www.ncbi.nlm.nih.gov/pubmed/20800340
350. Zhang, K., et al. Plasmakinetic Vapor Enucleation of the Prostate with Button Electrode versus
Plasmakinetic Resection of the Prostate for Benign Prostatic Enlargement >90 ml: Perioperative and
3-Month Follow-Up Results of a Prospective, Randomized Clinical Trial. Urol Int, 2015. 95: 260.
https://www.ncbi.nlm.nih.gov/pubmed/26044933
351. Wang, Z., et al. A prospective, randomised trial comparing transurethral enucleation with bipolar
system (TUEB) to monopolar resectoscope enucleation of the prostate for symptomatic benign
prostatic hyperplasia. Biomed Res, 2017. 28.
https://www.alliedacademies.org/articles/a-prospective-randomised-trial-comparing-transurethral-
enucleation-with-bipolar-system-tueb-to-monopolar-resectoscope-enucleation-.html
352. Ran, L., et al. Comparison of fluid absorption between transurethral enucleation and transurethral
resection for benign prostate hyperplasia. Urol Int, 2013. 91: 26.
https://www.ncbi.nlm.nih.gov/pubmed/23571450
353. Luo, Y.H., et al. Plasmakinetic enucleation of the prostate vs plasmakinetic resection of the prostate
for benign prostatic hyperplasia: comparison of outcomes according to prostate size in 310
patients. Urology, 2014. 84: 904.
https://www.ncbi.nlm.nih.gov/pubmed/25150180
354. Zhu, L., et al. Electrosurgical enucleation versus bipolar transurethral resection for prostates larger
than 70 ml: a prospective, randomized trial with 5-year followup. J Urol, 2013. 189: 1427.
https://www.ncbi.nlm.nih.gov/pubmed/23123549
355. Zhang, Y., et al. Efficacy and safety of enucleation vs. resection of prostate for treatment of benign
prostatic hyperplasia: a meta-analysis of randomized controlled trials. Prostate Cancer Prostatic Dis,
2019.
https://www.ncbi.nlm.nih.gov/pubmed/30816336
356. Gilling, P.J., et al. Combination holmium and Nd:YAG laser ablation of the prostate: initial clinical
experience. J Endourol, 1995. 9: 151.
https://www.ncbi.nlm.nih.gov/pubmed/7633476
357. Tan, A., et al. Meta-analysis of holmium laser enucleation versus transurethral resection of the
prostate for symptomatic prostatic obstruction. Br J Surg, 2007. 94: 1201.
https://www.ncbi.nlm.nih.gov/pubmed/17729384
358. Yin, L., et al. Holmium laser enucleation of the prostate versus transurethral resection of the
prostate: a systematic review and meta-analysis of randomized controlled trials. J Endourol, 2013.
27: 604.
https://www.ncbi.nlm.nih.gov/pubmed/23167266
359. Qian, X., et al. Functional outcomes and complications following B-TURP versus HoLEP for the
treatment of benign prostatic hyperplasia: a review of the literature and Meta-analysis. Aging Male,
2017. 20: 184.
https://www.ncbi.nlm.nih.gov/pubmed/28368238

72 MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021
360. Gilling, P.J., et al. Long-term results of a randomized trial comparing holmium laser enucleation of
the prostate and transurethral resection of the prostate: results at 7 years. BJU Int, 2012. 109: 408.
https://www.ncbi.nlm.nih.gov/pubmed/21883820
361. Chen, Y.B., et al. A prospective, randomized clinical trial comparing plasmakinetic resection of the
prostate with holmium laser enucleation of the prostate based on a 2-year followup. J Urol, 2013.
189: 217.
https://www.ncbi.nlm.nih.gov/pubmed/23174256
362. Gu, M., et al. Comparison of Holmium Laser Enucleation and Plasmakinetic Resection of Prostate:
A Randomized Trial with 72-Month Follow-Up. J Endourol, 2018. 32: 139.
https://www.ncbi.nlm.nih.gov/pubmed/29239228
363. Lourenco, T., et al. Alternative approaches to endoscopic ablation for benign enlargement of the
prostate: systematic review of randomised controlled trials. Bmj, 2008. 337: a449.
https://www.ncbi.nlm.nih.gov/pubmed/18595932
364. El Tayeb, M.M., et al. Holmium Laser Enucleation of the Prostate in Patients Requiring
Anticoagulation. J Endourol, 2016. 30: 805.
https://www.ncbi.nlm.nih.gov/pubmed/27065437
365. Sun, J., et al. Safety and feasibility study of holmium laser enucleation of the prostate (HOLEP) on
patients receiving dual antiplatelet therapy (DAPT). World J Urol, 2018. 36: 271.
https://www.ncbi.nlm.nih.gov/pubmed/29138929
366. Briganti, A., et al. Impact on sexual function of holmium laser enucleation versus transurethral
resection of the prostate: results of a prospective, 2-center, randomized trial. J Urol, 2006. 175: 1817.
https://www.ncbi.nlm.nih.gov/pubmed/16600770
367. Li, Z., et al. The impact of surgical treatments for lower urinary tract symptoms/benign prostatic
hyperplasia on male erectile function: A systematic review and network meta-analysis. Medicine
(Baltimore), 2016. 95: e3862.
https://www.ncbi.nlm.nih.gov/pubmed/27310968
368. Elshal, A.M., et al. Prospective controlled assessment of men’s sexual function changes following
Holmium laser enucleation of the prostate for treatment of benign prostate hyperplasia. Int Urol
Nephrol, 2017. 49: 1741.
https://www.ncbi.nlm.nih.gov/pubmed/28780626
369. Kim, M., et al. Pilot study of the clinical efficacy of ejaculatory hood sparing technique for
ejaculation preservation in Holmium laser enucleation of the prostate. Int J Impot Res, 2015. 27: 20.
https://www.ncbi.nlm.nih.gov/pubmed/25007827
370. Elzayat, E.A., et al. Holmium laser enucleation of the prostate (HoLEP): long-term results,
reoperation rate, and possible impact of the learning curve. Eur Urol, 2007. 52: 1465.
https://www.ncbi.nlm.nih.gov/pubmed/17498867
371. Du, C., et al. Holmium laser enucleation of the prostate: the safety, efficacy, and learning experience
in China. J Endourol, 2008. 22: 1031.
https://www.ncbi.nlm.nih.gov/pubmed/18377236
372. Robert, G., et al. Multicentre prospective evaluation of the learning curve of holmium laser
enucleation of the prostate (HoLEP). BJU Int, 2016. 117: 495.
https://www.ncbi.nlm.nih.gov/pubmed/25781490
373. Aho, T., et al. Description of a modular mentorship programme for holmium laser enucleation of the
prostate. World J Urol, 2015. 33: 497.
https://www.ncbi.nlm.nih.gov/pubmed/25271105
374. Enikeev, D., et al. A Randomized Trial Comparing The Learning Curve of 3 Endoscopic Enucleation
Techniques (HoLEP, ThuFLEP, and MEP) for BPH Using Mentoring Approach-Initial Results. Urology,
2018. 121: 51.
https://www.ncbi.nlm.nih.gov/pubmed/30053397
375. Yang, Z., et al. Comparison of thulium laser enucleation and plasmakinetic resection of the prostate
in a randomized prospective trial with 5-year follow-up. Lasers Med Sci, 2016. 31: 1797.
https://www.ncbi.nlm.nih.gov/pubmed/27677474
376. Xiao, K.W., et al. Enucleation of the prostate for benign prostatic hyperplasia thulium laser versus
holmium laser: a systematic review and meta-analysis. Lasers Med Sci, 2019.
https://www.ncbi.nlm.nih.gov/pubmed/30604345
377. Zhang, F., et al. Thulium laser versus holmium laser transurethral enucleation of the prostate:
18-month follow-up data of a single center. Urology, 2012. 79: 869.
https://www.ncbi.nlm.nih.gov/pubmed/22342411

MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021 73
378. Feng, L., et al. Thulium Laser Enucleation Versus Plasmakinetic Enucleation of the Prostate: A
Randomized Trial of a Single Center. J Endourol, 2016. 30: 665.
https://www.ncbi.nlm.nih.gov/pubmed/26886719
379. Bach, T., et al. Thulium:YAG vapoenucleation in large volume prostates. J Urol, 2011. 186: 2323.
https://www.ncbi.nlm.nih.gov/pubmed/22014812
380. Hauser, S., et al. Thulium laser (Revolix) vapoenucleation of the prostate is a safe procedure in
patients with an increased risk of hemorrhage. Urol Int, 2012. 88: 390.
https://www.ncbi.nlm.nih.gov/pubmed/22627127
381. Netsch, C., et al. Comparison of 120-200 W 2 mum thulium:yttrium-aluminum-garnet
vapoenucleation of the prostate. J Endourol, 2012. 26: 224.
https://www.ncbi.nlm.nih.gov/pubmed/22191688
382. Netsch, C., et al. 120-W 2-microm thulium:yttrium-aluminium-garnet vapoenucleation of the
prostate: 12-month follow-up. BJU Int, 2012. 110: 96.
https://www.ncbi.nlm.nih.gov/pubmed/22085294
383. Chang, C.H., et al. Vapoenucleation of the prostate using a high-power thulium laser: a one-year
follow-up study. BMC Urol, 2015. 15: 40.
https://www.ncbi.nlm.nih.gov/pubmed/25956819
384. Netsch, C., et al. Safety and effectiveness of Thulium VapoEnucleation of the prostate (ThuVEP) in
patients on anticoagulant therapy. World J Urol, 2014. 32: 165.
https://www.ncbi.nlm.nih.gov/pubmed/23657354
385. Gross, A.J., et al. Complications and early postoperative outcome in 1080 patients after thulium
vapoenucleation of the prostate: results at a single institution. Eur Urol, 2013. 63: 859.
https://www.ncbi.nlm.nih.gov/pubmed/23245687
386. Tiburtius, C., et al. Impact of thulium VapoEnucleation of the prostate on erectile function: a
prospective analysis of 72 patients at 12-month follow-up. Urology, 2014. 83: 175.
https://www.ncbi.nlm.nih.gov/pubmed/24103563
387. Wang, Y., et al. Impact of 120-W 2-mum continuous wave laser vapoenucleation of the prostate on
sexual function. Lasers Med Sci, 2014. 29: 689.
https://www.ncbi.nlm.nih.gov/pubmed/23828495
388. Lusuardi, L., et al. Safety and efficacy of Eraser laser enucleation of the prostate: preliminary report.
J Urol, 2011. 186: 1967.
https://www.ncbi.nlm.nih.gov/pubmed/21944122
389. Zhang, J., et al. 1470 nm Diode Laser Enucleation vs Plasmakinetic Resection of the Prostate for
Benign Prostatic Hyperplasia: A Randomized Study. J Endourol, 2019. 33: 211.
https://www.ncbi.nlm.nih.gov/pubmed/30489151
390. Zou, Z., et al. Dual-centre randomized-controlled trial comparing transurethral endoscopic
enucleation of the prostate using diode laser vs. bipolar plasmakinetic for the treatment of LUTS
secondary of benign prostate obstruction: 1-year follow-up results. World J Urol, 2018.
https://www.ncbi.nlm.nih.gov/pubmed/29459994
391. Xu, A., et al. A randomized trial comparing diode laser enucleation of the prostate with
plasmakinetic enucleation and resection of the prostate for the treatment of benign prostatic
hyperplasia. J Endourol, 2013. 27: 1254.
https://www.ncbi.nlm.nih.gov/pubmed/23879477
392. Wu, G., et al. A comparative study of diode laser and plasmakinetic in transurethral enucleation of
the prostate for treating large volume benign prostatic hyperplasia: a randomized clinical trial with
12-month follow-up. Lasers Med Sci, 2016. 31: 599.
https://www.ncbi.nlm.nih.gov/pubmed/26822403
393. Mariano, M.B., et al. Laparoscopic prostatectomy with vascular control for benign prostatic
hyperplasia. J Urol, 2002. 167: 2528.
https://www.ncbi.nlm.nih.gov/pubmed/11992078
394. Sotelo, R., et al. Robotic simple prostatectomy. J Urol, 2008. 179: 513.
https://www.ncbi.nlm.nih.gov/pubmed/18076926
395. Lucca, I., et al. Outcomes of minimally invasive simple prostatectomy for benign prostatic
hyperplasia: a systematic review and meta-analysis. World J Urol, 2015. 33: 563.
https://www.ncbi.nlm.nih.gov/pubmed/24879405
396. Autorino, R., et al. Perioperative Outcomes of Robotic and Laparoscopic Simple Prostatectomy:
A European-American Multi-institutional Analysis. Eur Urol, 2015. 68: 86.
https://www.ncbi.nlm.nih.gov/pubmed/25484140

74 MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021
397. Pokorny, M., et al. Robot-assisted Simple Prostatectomy for Treatment of Lower Urinary Tract
Symptoms Secondary to Benign Prostatic Enlargement: Surgical Technique and Outcomes in a
High-volume Robotic Centre. Eur Urol, 2015. 68: 451.
https://www.ncbi.nlm.nih.gov/pubmed/25887786
398. Sorokin, I., et al. Robot-Assisted Versus Open Simple Prostatectomy for Benign Prostatic
Hyperplasia in Large Glands: A Propensity Score-Matched Comparison of Perioperative and Short-
Term Outcomes. J Endourol, 2017. 31: 1164.
https://www.ncbi.nlm.nih.gov/pubmed/28854815
399. Stoddard, M.D., et al. Standardization of 532 nm Laser Terminology for Surgery in Benign Prostatic
Hyperplasia: A Systematic Review. J Endourol, 2020. 34: 121.
https://www.ncbi.nlm.nih.gov/pubmed/31880953
400. Gomez Sancha, F., et al. Common trend: move to enucleation-Is there a case for GreenLight
enucleation? Development and description of the technique. World J Urol, 2015. 33: 539.
https://www.ncbi.nlm.nih.gov/pubmed/24929643
401. Law, K.W., et al. Anatomic GreenLight laser vaporization-incision technique for benign prostatic
hyperplasia using the XPS LBO-180W system: How I do it. Can J Urol, 2019. 26: 9963.
https://www.ncbi.nlm.nih.gov/pubmed/31629449
402. Brunken, C., et al. Transurethral GreenLight laser enucleation of the prostate--a feasibility study.
J Endourol, 2011. 25: 1199.
https://www.ncbi.nlm.nih.gov/pubmed/21612434
403. Elshal, A.M., et al. Prospective Assessment of Learning Curve of Holmium Laser Enucleation of the
Prostate for Treatment of Benign Prostatic Hyperplasia Using a Multidimensional Approach. J Urol,
2017. 197: 1099.
https://www.ncbi.nlm.nih.gov/pubmed/27825972
404. Botto, H., et al. Electrovaporization of the prostate with the Gyrus device. J Endourol, 2001. 15: 313.
https://www.ncbi.nlm.nih.gov/pubmed/11339400
405. Bucuras, V., et al. Bipolar vaporization of the prostate: Is it ready for the primetime? Ther Adv Urol,
2011. 3: 257.
https://www.ncbi.nlm.nih.gov/pubmed/22164195
406. Reich, O., et al. Plasma Vaporisation of the Prostate: Initial Clinical Results. Eur Urol, 2010. 57: 693.
https://www.ncbi.nlm.nih.gov/pubmed/19482414
407. Reich, O., et al. In vitro comparison of transurethral vaporization of the prostate (TUVP), resection of
the prostate (TURP), and vaporization-resection of the prostate (TUVRP). Urol Res, 2002. 30: 15.
https://www.ncbi.nlm.nih.gov/pubmed/11942320
408. Gallucci, M., et al. Transurethral electrovaporization of the prostate vs. transurethral resection.
Results of a multicentric, randomized clinical study on 150 patients. Eur Urol, 1998. 33: 359.
https://www.ncbi.nlm.nih.gov/pubmed/9612677
409. Poulakis, V., et al. Transurethral electrovaporization vs transurethral resection for symptomatic
prostatic obstruction: a meta-analysis. BJU Int, 2004. 94: 89.
https://www.ncbi.nlm.nih.gov/pubmed/15217438
410. Dunsmuir, W.D., et al. Gyrus bipolar electrovaporization vs transurethral resection of the prostate:
a randomized prospective single-blind trial with 1 y follow-up. Prostate Cancer Prostatic Dis, 2003.
6: 182.
https://www.ncbi.nlm.nih.gov/pubmed/12806380
411. Fung, B.T., et al. Prospective randomized controlled trial comparing plasmakinetic vaporesection
and conventional transurethral resection of the prostate. Asian J Surg, 2005. 28: 24.
https://www.ncbi.nlm.nih.gov/pubmed/15691793
412. Karaman, M.I., et al. Comparison of transurethral vaporization using PlasmaKinetic energy and
transurethral resection of prostate: 1-year follow-up. J Endourol, 2005. 19: 734.
https://www.ncbi.nlm.nih.gov/pubmed/16053367
413. Hon, N.H., et al. A prospective, randomized trial comparing conventional transurethral prostate
resection with PlasmaKinetic vaporization of the prostate: physiological changes, early
complications and long-term followup. J Urol, 2006. 176: 205.
https://www.ncbi.nlm.nih.gov/pubmed/16753403
414. Kaya, C., et al. The long-term results of transurethral vaporization of the prostate using
plasmakinetic energy. BJU Int, 2007. 99: 845.
https://www.ncbi.nlm.nih.gov/pubmed/17378844

MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021 75
415. Geavlete, B., et al. Transurethral resection (TUR) in saline plasma vaporization of the prostate vs
standard TUR of the prostate: ‘the better choice’ in benign prostatic hyperplasia? BJU Int, 2010.
106: 1695.
https://www.ncbi.nlm.nih.gov/pubmed/20518763
416. Nuhoglu, B., et al. The role of bipolar transurethral vaporization in the management of benign
prostatic hyperplasia. Urol Int, 2011. 87: 400.
https://www.ncbi.nlm.nih.gov/pubmed/22086154
417. Zhang, S.Y., et al. Efficacy and safety of bipolar plasma vaporization of the prostate with “button-
type” electrode compared with transurethral resection of prostate for benign prostatic hyperplasia.
Chin Med J (Engl), 2012. 125: 3811.
https://www.ncbi.nlm.nih.gov/pubmed/23106879
418. Falahatkar, S., et al. Bipolar transurethral vaporization: a superior procedure in benign prostatic
hyperplasia: a prospective randomized comparison with bipolar TURP. Int Braz J Urol, 2014. 40: 346.
https://www.ncbi.nlm.nih.gov/pubmed/25010300
419. Geavlete, B., et al. Continuous vs conventional bipolar plasma vaporisation of the prostate and
standard monopolar resection: A prospective, randomised comparison of a new technological
advance. BJU Int, 2014. 113: 288.
https://www.ncbi.nlm.nih.gov/pubmed/24053794
420. Yip, S.K., et al. A randomized controlled trial comparing the efficacy of hybrid bipolar transurethral
vaporization and resection of the prostate with bipolar transurethral resection of the prostate.
J Endourol, 2011. 25: 1889.
https://www.ncbi.nlm.nih.gov/pubmed/21923418
421. Elsakka, A.M., et al. A prospective randomised controlled study comparing bipolar plasma
vaporisation of the prostate to monopolar transurethral resection of the prostate. Arab J Urol, 2016.
14: 280.
https://www.ncbi.nlm.nih.gov/pubmed/27900218
422. Lee, S.W., et al. Transurethral procedures for lower urinary tract symptoms resulting from benign
prostatic enlargement: A quality and meta-analysis. Int Neurourol J, 2013. 17: 59.
https://www.ncbi.nlm.nih.gov/pubmed/23869269
423. Wroclawski, M.L., et al. ‘Button type’ bipolar plasma vaporisation of the prostate compared with
standard transurethral resection: A systematic review and meta-analysis of short-term outcome
studies. BJU Int, 2016. 117: 662.
https://www.ncbi.nlm.nih.gov/pubmed/26299915
424. Robert, G., et al. Bipolar plasma vaporization of the prostate: ready to replace GreenLight? A
systematic review of randomized control trials. World J Urol, 2015. 33: 549.
https://www.ncbi.nlm.nih.gov/pubmed/25159871
425. Thangasamy, I.A., et al. Photoselective vaporisation of the prostate using 80-W and 120-W laser
versus transurethral resection of the prostate for benign prostatic hyperplasia: a systematic review
with meta-analysis from 2002 to 2012. Eur Urol, 2012. 62: 315.
https://www.ncbi.nlm.nih.gov/pubmed/22575913
426. Bouchier-Hayes, D.M., et al. A randomized trial of photoselective vaporization of the prostate using
the 80-W potassium-titanyl-phosphate laser vs transurethral prostatectomy, with a 1-year follow-up.
BJU Int, 2010. 105: 964.
https://www.ncbi.nlm.nih.gov/pubmed/19912196
427. Capitan, C., et al. GreenLight HPS 120-W laser vaporization versus transurethral resection of the
prostate for the treatment of lower urinary tract symptoms due to benign prostatic hyperplasia:
a randomized clinical trial with 2-year follow-up. Eur Urol, 2011. 60: 734.
https://www.ncbi.nlm.nih.gov/pubmed/21658839
428. Skolarikos, A., et al., 80W PVP versus TURP: results of a randomized prospective study at 12
months of follow-up., in Abstract presented at: American Urological Association annual meeting.
2008: Orlando, FL, USA.
429. Zhou, Y., et al. Greenlight high-performance system (HPS) 120-W laser vaporization versus
transurethral resection of the prostate for the treatment of benign prostatic hyperplasia: a meta-
analysis of the published results of randomized controlled trials. Lasers Med Sci, 2016. 31: 485.
https://www.ncbi.nlm.nih.gov/pubmed/26868032
430. Thomas, J.A., et al. A Multicenter Randomized Noninferiority Trial Comparing GreenLight-XPS Laser
Vaporization of the Prostate and Transurethral Resection of the Prostate for the Treatment of Benign
Prostatic Obstruction: Two-yr Outcomes of the GOLIATH Study. Eur Urol, 2016. 69: 94.
https://www.ncbi.nlm.nih.gov/pubmed/26283011

76 MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021
431. Al-Ansari, A., et al. GreenLight HPS 120-W laser vaporization versus transurethral resection of the
prostate for treatment of benign prostatic hyperplasia: a randomized clinical trial with midterm
follow-up. Eur Urol, 2010. 58: 349.
https://www.ncbi.nlm.nih.gov/pubmed/20605316
432. Pereira-Correia, J.A., et al. GreenLight HPS 120-W laser vaporization vs transurethral resection of
the prostate (<60 mL): a 2-year randomized double-blind prospective urodynamic investigation. BJU
Int, 2012. 110: 1184.
https://www.ncbi.nlm.nih.gov/pubmed/22257240
433. Kang, D.H., et al. A Systematic Review and Meta-Analysis of Functional Outcomes and
Complications Following the Photoselective Vaporization of the Prostate and Monopolar
Transurethral Resection of the Prostate. World J Men’s Health, 2016. 34: 110.
https://www.ncbi.nlm.nih.gov/pubmed/27574594
434. Elmansy, H., et al. Holmium laser enucleation versus photoselective vaporization for prostatic
adenoma greater than 60 ml: preliminary results of a prospective, randomized clinical trial. J Urol,
2012. 188: 216.
https://www.ncbi.nlm.nih.gov/pubmed/22591968
435. Chung, D.E., et al. Outcomes and complications after 532 nm laser prostatectomy in anticoagulated
patients with benign prostatic hyperplasia. J Urol, 2011. 186: 977.
https://www.ncbi.nlm.nih.gov/pubmed/21791350
436. Reich, O., et al. High power (80 W) potassium-titanyl-phosphate laser vaporization of the prostate in
66 high risk patients. J Urol, 2005. 173: 158.
https://www.ncbi.nlm.nih.gov/pubmed/15592063
437. Ruszat, R., et al. Safety and effectiveness of photoselective vaporization of the prostate (PVP) in
patients on ongoing oral anticoagulation. Eur Urol, 2007. 51: 1031.
https://www.ncbi.nlm.nih.gov/pubmed/16945475
438. Sandhu, J.S., et al. Photoselective laser vaporization prostatectomy in men receiving
anticoagulants. J Endourol, 2005. 19: 1196.
https://www.ncbi.nlm.nih.gov/pubmed/16359214
439. Lee, D.J., et al. Laser Vaporization of the Prostate With the 180-W XPS-Greenlight Laser in Patients
With Ongoing Platelet Aggregation Inhibition and Oral Anticoagulation. Urology, 2016. 91: 167.
https://www.ncbi.nlm.nih.gov/pubmed/26829717
440. Jackson, R.E., et al. Risk factors for delayed hematuria following photoselective vaporization of the
prostate. J Urol, 2013. 190: 903.
https://www.ncbi.nlm.nih.gov/pubmed/23538242
441. Knapp, G.L., et al. Perioperative adverse events in patients on continued anticoagulation undergoing
photoselective vaporisation of the prostate with the 180-W Greenlight lithium triborate laser. BJU Int,
2017. 119: 33.
https://www.ncbi.nlm.nih.gov/pubmed/28544292
442. Woo, H., et al. Outcome of GreenLight HPS 120-W laser therapy in specific patient populations:
those in retention, on anticoagulants, and with large prostates (>80 ml). Eur Urol Suppl 2008. 7: 378.
https://www.sciencedirect.com/science/article/abs/pii/S1569905608000274
443. Rajbabu, K., et al. Photoselective vaporization of the prostate with the potassium-titanyl-phosphate
laser in men with prostates of >100 mL. BJU Int, 2007. 100: 593.
https://www.ncbi.nlm.nih.gov/pubmed/17511771
444. Ruszat, R., et al. Photoselective vaporization of the prostate: subgroup analysis of men with
refractory urinary retention. Eur Urol, 2006. 50: 1040.
https://www.ncbi.nlm.nih.gov/pubmed/16481099
445. Horasanli, K., et al. Photoselective potassium titanyl phosphate (KTP) laser vaporization versus
transurethral resection of the prostate for prostates larger than 70 mL: a short-term prospective
randomized trial. Urology, 2008. 71: 247.
https://www.ncbi.nlm.nih.gov/pubmed/18308094
446. Alivizatos, G., et al. Transurethral photoselective vaporization versus transvesical open enucleation for
prostatic adenomas >80ml: 12-mo results of a randomized prospective study. Eur Urol, 2008. 54: 427.
https://www.ncbi.nlm.nih.gov/pubmed/18069117
447. Bouchier-Hayes, D.M., et al. KTP laser versus transurethral resection: early results of a randomized
trial. J Endourol, 2006. 20: 580.
https://www.ncbi.nlm.nih.gov/pubmed/16903819
448. Bruyere, F., et al. Influence of photoselective vaporization of the prostate on sexual function: results
of a prospective analysis of 149 patients with long-term follow-up. Eur Urol, 2010. 58: 207.
https://www.ncbi.nlm.nih.gov/pubmed/20466480

MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021 77
449. Razzaghi, M.R., et al. Diode laser (980 nm) vaporization in comparison with transurethral resection
of the prostate for benign prostatic hyperplasia: randomized clinical trial with 2-year follow-up.
Urology, 2014. 84: 526.
https://www.ncbi.nlm.nih.gov/pubmed/25168526
450. Cetinkaya, M., et al. 980-Nm Diode Laser Vaporization versus Transurethral Resection of the
Prostate for Benign Prostatic Hyperplasia: Randomized Controlled Study. Urol J, 2015. 12: 2355.
https://www.ncbi.nlm.nih.gov/pubmed/26571321
451. Chiang, P.H., et al. GreenLight HPS laser 120-W versus diode laser 200-W vaporization of the
prostate: comparative clinical experience. Lasers Surg Med, 2010. 42: 624.
https://www.ncbi.nlm.nih.gov/pubmed/20806388
452. Ruszat, R., et al. Prospective single-centre comparison of 120-W diode-pumped solid-state high-
intensity system laser vaporization of the prostate and 200-W high-intensive diode-laser ablation of
the prostate for treating benign prostatic hyperplasia. BJU Int, 2009. 104: 820.
https://www.ncbi.nlm.nih.gov/pubmed/19239441
453. Seitz, M., et al. The diode laser: a novel side-firing approach for laser vaporisation of the human
prostate--immediate efficacy and 1-year follow-up. Eur Urol, 2007. 52: 1717.
https://www.ncbi.nlm.nih.gov/pubmed/17628326
454. Shaker, H.S., et al. Quartz head contact laser fiber: a novel fiber for laser ablation of the prostate
using the 980 nm high power diode laser. J Urol, 2012. 187: 575.
https://www.ncbi.nlm.nih.gov/pubmed/22177175
455. MacRae, C., et al. How I do it: Aquablation of the prostate using the AQUABEAM system. Can
J Urol, 2016. 23: 8590.
https://www.ncbi.nlm.nih.gov/pubmed/27995858
456. Gilling, P., et al. WATER: A Double-Blind, Randomized, Controlled Trial of Aquablation vs
Transurethral Resection of the Prostate in Benign Prostatic Hyperplasia. J Urol, 2018. 199: 1252.
https://www.ncbi.nlm.nih.gov/pubmed/29360529
457. Kasivisvanathan, V., et al. Aquablation versus transurethral resection of the prostate: 1 year United
States - cohort outcomes. Can J Urol, 2018. 25: 9317.
https://www.ncbi.nlm.nih.gov/pubmed/29900819
458. Gilling, P.J., et al. Randomized Controlled Trial of Aquablation versus Transurethral Resection of the
Prostate in Benign Prostatic Hyperplasia: One-year Outcomes. Urology, 2019. 125: 169.
https://www.ncbi.nlm.nih.gov/pubmed/30552937
459. Plante, M., et al. Symptom relief and anejaculation after aquablation or transurethral resection of the
prostate: subgroup analysis from a blinded randomized trial. BJU Int, 2019. 123: 651.
https://www.ncbi.nlm.nih.gov/pubmed/29862630
460. Desai, M., et al. Aquablation for benign prostatic hyperplasia in large prostates (80-150 mL):
6-month results from the WATER II trial. BJU Int, 2019.
https://www.ncbi.nlm.nih.gov/pubmed/30734990
461. Pimentel, M.A., et al. Urodynamic Outcomes After Aquablation. Urology, 2019.
https://www.ncbi.nlm.nih.gov/pubmed/30721737
462. Abt, D., et al. Comparison of prostatic artery embolisation (PAE) versus transurethral resection of the
prostate (TURP) for benign prostatic hyperplasia: randomised, open label, non-inferiority trial. BMJ,
2018. 361: k2338.
https://www.ncbi.nlm.nih.gov/pubmed/29921613
463. Zhang, J.L., et al. Effectiveness of Contrast-enhanced MR Angiography for Visualization of the
Prostatic Artery prior to Prostatic Arterial Embolization. Radiology, 2019: 181524.
https://www.ncbi.nlm.nih.gov/pubmed/30806596
464. Gao, Y.A., et al. Benign prostatic hyperplasia: prostatic arterial embolization versus transurethral
resection of the prostate--a prospective, randomized, and controlled clinical trial. Radiology, 2014.
270: 920.
https://www.ncbi.nlm.nih.gov/pubmed/24475799
465. Carnevale, F.C., et al. Transurethral Resection of the Prostate (TURP) Versus Original and
PErFecTED Prostate Artery Embolization (PAE) Due to Benign Prostatic Hyperplasia (BPH):
Preliminary Results of a Single Center, Prospective, Urodynamic-Controlled Analysis. Cardiovasc
Intervent Radiol, 2016. 39: 44.
https://www.ncbi.nlm.nih.gov/pubmed/26506952
466. Zumstein, V., et al. Prostatic Artery Embolization versus Standard Surgical Treatment for Lower
Urinary Tract Symptoms Secondary to Benign Prostatic Hyperplasia: A Systematic Review and
Meta-analysis. Eur Urol Focus, 2018.
https://www.ncbi.nlm.nih.gov/pubmed/30292422

78 MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021
467. Shim, S.R., et al. Efficacy and Safety of Prostatic Arterial Embolization: Systematic Review with
Meta-Analysis and Meta-Regression. J Urol, 2017. 197: 465.
https://www.ncbi.nlm.nih.gov/pubmed/27592008
468. Jiang, Y.L., et al. Transurethral resection of the prostate versus prostatic artery embolization in the
treatment of benign prostatic hyperplasia: A meta-analysis. BMC Urol, 2019. 19: 11.
https://www.ncbi.nlm.nih.gov/pubmed/30691478
469. Moreira, A.M., et al. A Review of Adverse Events Related to Prostatic Artery Embolization for
Treatment of Bladder Outlet Obstruction Due to BPH. Cardiovasc Intervent Radiol, 2017. 40: 1490.
https://www.ncbi.nlm.nih.gov/pubmed/28795212
470. Ray, A.F., et al. Efficacy and safety of prostate artery embolization for benign prostatic hyperplasia:
an observational study and propensity-matched comparison with transurethral resection of the
prostate (the UK-ROPE study). BJU Int, 2018. 122: 270.
https://www.ncbi.nlm.nih.gov/pubmed/29645352
471. National Institute for Health and Care Excellence. Prostate artery embolisation for lower urinary tract
symptoms caused by benign prostatic hyperplasia. NICE GUidelines, 2018.
https://www.nice.org.uk/guidance/ipg611
472. McVary, K.T., et al. Erectile and Ejaculatory Function Preserved With Convective Water Vapor
Energy Treatment of Lower Urinary Tract Symptoms Secondary to Benign Prostatic Hyperplasia:
Randomized Controlled Study. J Sex Med, 2016. 13: 924.
https://www.ncbi.nlm.nih.gov/pubmed/27129767
473. Roehrborn, C.G., et al. Convective Thermal Therapy: Durable 2-Year Results of Randomized
Controlled and Prospective Crossover Studies for Treatment of Lower Urinary Tract Symptoms Due
to Benign Prostatic Hyperplasia. J Urol, 2017. 197: 1507.
https://www.ncbi.nlm.nih.gov/pubmed/27993667
474. McVary, K.T., et al. Rezum Water Vapor Thermal Therapy for Lower Urinary Tract Symptoms
Associated With Benign Prostatic Hyperplasia: 4-Year Results From Randomized Controlled Study.
Urology, 2019. 126: 171.
https://www.ncbi.nlm.nih.gov/pubmed/30677455
475. Chin, P.T., et al. Prostatic urethral lift: two-year results after treatment for lower urinary tract
symptoms secondary to benign prostatic hyperplasia. Urology, 2012. 79: 5.
https://www.ncbi.nlm.nih.gov/pubmed/22202539
476. McNicholas, T.A., et al. Minimally invasive prostatic urethral lift: surgical technique and multinational
experience. Eur Urol, 2013. 64: 292.
https://www.ncbi.nlm.nih.gov/pubmed/23357348
477. Roehrborn, C.G., et al. The prostatic urethral lift for the treatment of lower urinary tract symptoms
associated with prostate enlargement due to benign prostatic hyperplasia: the L.I.F.T. Study. J Urol,
2013. 190: 2161.
https://www.ncbi.nlm.nih.gov/pubmed/23764081
478. Woo, H.H., et al. Safety and feasibility of the prostatic urethral lift: a novel, minimally invasive
treatment for lower urinary tract symptoms (LUTS) secondary to benign prostatic hyperplasia (BPH).
BJU Int, 2011. 108: 82.
https://www.ncbi.nlm.nih.gov/pubmed/21554526
479. Woo, H.H., et al. Preservation of sexual function with the prostatic urethral lift: a novel treatment for
lower urinary tract symptoms secondary to benign prostatic hyperplasia. J Sex Med, 2012. 9: 568.
https://www.ncbi.nlm.nih.gov/pubmed/22172161
480. Perera, M., et al. Prostatic urethral lift improves urinary symptoms and flow while preserving sexual
function for men with benign prostatic hyperplasia: a systematic review and meta-analysis. Eur Urol,
2015. 67: 704.
https://www.ncbi.nlm.nih.gov/pubmed/25466940
481. Roehrborn, C.G., et al. Three year results of the prostatic urethral L.I.F.T. study. Can J Urol, 2015.
22: 7772.
https://www.ncbi.nlm.nih.gov/pubmed/26068624
482. Roehrborn, C.G., et al. Five year results of the prospective randomized controlled prostatic urethral
L.I.F.T. study. Can J Urol, 2017. 24: 8802.
https://www.ncbi.nlm.nih.gov/pubmed/28646935
483. Sonksen, J., et al. Prospective, Randomized, Multinational Study of Prostatic Urethral Lift Versus
Transurethral Resection of the Prostate: 12-month Results from the BPH6 Study. Eur Urol, 2015.
68: 643.
https://www.ncbi.nlm.nih.gov/pubmed/25937539

MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021 79
484. Gratzke, C., et al. Prostatic urethral lift vs transurethral resection of the prostate: 2-year results of
the BPH6 prospective, multicentre, randomized study. BJU Int, 2017. 119: 767.
https://www.ncbi.nlm.nih.gov/pubmed/27862831
485. Rukstalis, D., et al. Prostatic Urethral Lift (PUL) for obstructive median lobes: 12 month results of the
MedLift Study. Prostate Cancer Prostatic Dis, 2019. 22: 411.
https://www.ncbi.nlm.nih.gov/pubmed/30542055
486. Magistro, G., et al. New intraprostatic injectables and prostatic urethral lift for male LUTS. Nat Rev
Urol, 2015. 12: 461.
https://www.ncbi.nlm.nih.gov/pubmed/26195444
487. Marberger, M., et al. A randomized double-blind placebo-controlled phase 2 dose-ranging study of
onabotulinumtoxinA in men with benign prostatic hyperplasia. Eur Urol, 2013. 63: 496.
https://www.ncbi.nlm.nih.gov/pubmed/23098762
488. McVary, K.T., et al. A multicenter, randomized, double-blind, placebo controlled study of
onabotulinumtoxinA 200 U to treat lower urinary tract symptoms in men with benign prostatic
hyperplasia. J Urol, 2014. 192: 150.
https://www.ncbi.nlm.nih.gov/pubmed/24508634
489. Shim, S.R., et al. Efficacy and safety of botulinum toxin injection for benign prostatic hyperplasia: a
systematic review and meta-analysis. Int Urol Nephrol, 2016. 48: 19.
https://www.ncbi.nlm.nih.gov/pubmed/26560471
490. Elhilali, M.M., et al. Prospective, randomized, double-blind, vehicle controlled, multicenter phase
IIb clinical trial of the pore forming protein PRX302 for targeted treatment of symptomatic benign
prostatic hyperplasia. J Urol, 2013. 189: 1421.
https://www.ncbi.nlm.nih.gov/pubmed/23142202
491. Denmeade, S.R., et al. Phase 1 and 2 studies demonstrate the safety and efficacy of intraprostatic
injection of PRX302 for the targeted treatment of lower urinary tract symptoms secondary to benign
prostatic hyperplasia. Eur Urol, 2011. 59: 747.
https://www.ncbi.nlm.nih.gov/pubmed/21129846
492. Shore, N., et al. Fexapotide triflutate: results of long-term safety and efficacy trials of a novel
injectable therapy for symptomatic prostate enlargement. World J Urol, 2018. 36: 801.
https://www.ncbi.nlm.nih.gov/pubmed/29380128
493. Porpiglia, F., et al. Temporary implantable nitinol device (TIND): a novel, minimally invasive treatment
for relief of lower urinary tract symptoms (LUTS) related to benign prostatic hyperplasia (BPH):
feasibility, safety and functional results at 1 year of follow-up. BJU Int, 2015. 116: 278.
https://www.ncbi.nlm.nih.gov/pubmed/25382816
494. Porpiglia, F., et al. 3-Year follow-up of temporary implantable nitinol device implantation for the
treatment of benign prostatic obstruction. BJU Int, 2018. 122: 106.
https://www.ncbi.nlm.nih.gov/pubmed/29359881
495. Sakalis, V.I., et al. Medical Treatment of Nocturia in Men with Lower Urinary Tract Symptoms:
Systematic Review by the European Association of Urology Guidelines Panel for Male Lower Urinary
Tract Symptoms. Eur Urol, 2017. 72: 757.
https://www.ncbi.nlm.nih.gov/pubmed/28666669
496. Marshall, S.D., et al. Nocturia: Current Levels of Evidence and Recommendations From the
International Consultation on Male Lower Urinary Tract Symptoms. Urology, 2015.
https://www.ncbi.nlm.nih.gov/pubmed/25881866
497. Cannon, A., et al. Desmopressin in the treatment of nocturnal polyuria in the male. BJU Int, 1999.
84: 20.
https://www.ncbi.nlm.nih.gov/pubmed/10444118
498. Han, J., et al. Desmopressin for treating nocturia in men. Cochrane Database Syst Rev, 2017. 10:
CD012059.
https://www.ncbi.nlm.nih.gov/pubmed/29055129
499. Weiss, J.P., et al. Efficacy and safety of low dose desmopressin orally disintegrating tablet in men
with nocturia: results of a multicenter, randomized, double-blind, placebo controlled, parallel group
study. J Urol, 2013. 190: 965.
https://www.ncbi.nlm.nih.gov/pubmed/23454402
500. Sand, P.K., et al. Efficacy and safety of low dose desmopressin orally disintegrating tablet in women
with nocturia: results of a multicenter, randomized, double-blind, placebo controlled, parallel group
study. J Urol, 2013. 190: 958.
https://www.ncbi.nlm.nih.gov/pubmed/23454404

80 MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021
501. Juul, K.V., et al. Low-dose desmopressin combined with serum sodium monitoring can prevent
clinically significant hyponatraemia in patients treated for nocturia. BJU Int, 2017. 119: 776.
https://www.ncbi.nlm.nih.gov/pubmed/27862898
502. Cohn, J.A., et al. Desmopressin acetate nasal spray for adults with nocturia. Expert Rev Clin
Pharmacol, 2017. 10: 1281.
https://www.ncbi.nlm.nih.gov/pubmed/29048257
503. Djavan, B., et al. The impact of tamsulosin oral controlled absorption system (OCAS) on nocturia
and the quality of sleep: Preliminary results of a pilot study. Eur Urol Suppl, 2005. 4: 1119.
https://www.eu-openscience.europeanurology.com/article/S1569-9056(04)00127-7/fulltext
504. Yokoyama, O., et al. Efficacy of fesoterodine on nocturia and quality of sleep in Asian patients with
overactive bladder. Urology, 2014. 83: 750.
https://www.ncbi.nlm.nih.gov/pubmed/24518285
505. Yokoyama, O., et al. Efficacy of solifenacin on nocturia in Japanese patients with overactive bladder:
impact on sleep evaluated by bladder diary. J Urol, 2011. 186: 170.
https://www.ncbi.nlm.nih.gov/pubmed/21575976
506. Johnson, T.M., 2nd, et al. The effect of doxazosin, finasteride and combination therapy on nocturia
in men with benign prostatic hyperplasia. J Urol, 2007. 178: 2045.
https://www.ncbi.nlm.nih.gov/pubmed/17869295
507. Oelke, M., et al. Impact of dutasteride on nocturia in men with lower urinary tract symptoms
suggestive of benign prostatic hyperplasia (LUTS/BPH): a pooled analysis of three phase III studies.
World J Urol, 2014. 32: 1141.
https://www.ncbi.nlm.nih.gov/pubmed/24903347
508. Oelke, M., et al. Effects of tadalafil on nighttime voiding (nocturia) in men with lower urinary tract
symptoms suggestive of benign prostatic hyperplasia: a post hoc analysis of pooled data from four
randomized, placebo-controlled clinical studies. World J Urol, 2014. 32: 1127.
https://www.ncbi.nlm.nih.gov/pubmed/24504761
509. Drake, M.J., et al. Melatonin pharmacotherapy for nocturia in men with benign prostatic
enlargement. J Urol, 2004. 171: 1199.
https://www.ncbi.nlm.nih.gov/pubmed/14767300
510. Reynard, J.M., et al. A novel therapy for nocturnal polyuria: a double-blind randomized trial of
frusemide against placebo. Br J Urol, 1998. 81: 215.
https://www.ncbi.nlm.nih.gov/pubmed/9488061
511. Falahatkar, S., et al. Celecoxib for treatment of nocturia caused by benign prostatic hyperplasia: a
prospective, randomized, double-blind, placebo-controlled study. Urology, 2008. 72: 813.
https://www.ncbi.nlm.nih.gov/pubmed/18692876
512. Sigurdsson, S., et al. A parallel, randomized, double-blind, placebo-controlled study to investigate
the effect of SagaPro on nocturia in men. Scand J Urol, 2013. 47: 26.
https://www.ncbi.nlm.nih.gov/pubmed/23323790

8. CONFLICT OF INTEREST
All members of the EAU Non-neurogenic Male LUTS Guidelines Panel have provided disclosure statements
on all relationships that they have that might be perceived to be a potential source of a conflict of interest.
This information is publically accessible through the EAU website: http://www.uroweb.org/guidelines/. These
Guidelines were developed with the financial support of the EAU. No external sources of funding and support
have been involved. The EAU is a non-profit organisation, and funding is limited to administrative assistance
and travel and meeting expenses. No honoraria or other reimbursements have been provided.

9. CITATION INFORMATION
The format in which to cite the EAU Guidelines will vary depending on the style guide of the journal in which the
citation appears. Accordingly, the number of authors or whether, for instance, to include the publisher, location,
or an ISBN number may vary.

MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021 81
The compilation of the complete Guidelines should be referenced as:
EAU Guidelines. Edn. presented at the EAU Annual Congress Milan 2021. ISBN 978-94-92671-13-4.

If a publisher and/or location is required, include:


EAU Guidelines Office, Arnhem, the Netherlands. http://uroweb.org/guidelines/compilations-of-all-guidelines/

References to individual guidelines should be structured in the following way:


Contributors’ names. Title of resource. Publication type. ISBN. Publisher and publisher location, year.

82 MANAGEMENT OF NON-NEUROGENIC MALE LOWER URINARY TRACT SYMPTOMS (LUTS) - UPDATE MARCH 2021
EAU Guidelines on
Muscle-invasive
and Metastatic
Bladder Cancer
J.A. Witjes (Chair), H.M. Bruins, R. Cathomas, E. Compérat,
N.C. Cowan, J.A. Efstathiou, R. Fietkau, G. Gakis,
V. Hernández, A. Lorch, M.I. Milowsky, M.J. Ribal (Vice-chair),
G.N. Thalmann, A.G. van der Heijden, E. Veskimäe
Guidelines Associates: E. Linares Espinós, M. Rouanne,
Y. Neuzillet

© European Association of Urology 2021


TABLE OF CONTENTS PAGE
1. INTRODUCTION 6
1.1 Aims and scope 6
1.2 Panel composition 6
1.3 Available publications 6
1.4 Publication history and summary of changes 6
1.4.1 Publication history 6
1.4.2 Summary of changes 6

2. METHODS 9
2.1 Data identification 9
2.2 Peer-review 10
2.2.1 Lay review 10
2.3 Future goals 11

3. EPIDEMIOLOGY, AETIOLOGY AND PATHOLOGY 11


3.1 Epidemiology 11
3.2 Aetiology 11
3.2.1 Tobacco smoking 11
3.2.2 Occupational exposure to chemicals 11
3.2.3 Radiotherapy 11
3.2.4 Dietary factors 12
3.2.5 Metabolic disorders 12
3.2.6 Bladder schistosomiasis and chronic urinary tract infection 12
3.2.7 Gender 12
3.2.8 Genetic factors 13
3.2.9 Summary of evidence and guidelines for epidemiology and risk factors 13
3.3 Pathology 13
3.3.1 Handling of transurethral resection and cystectomy specimens 13
3.3.2 Pathology of muscle-invasive bladder cancer 14
3.3.3 Guidelines for the assessment of tumour specimens 14
3.3.4 EAU-ESMO consensus statements on the management of advanced- and
variant bladder cancer 15

4. STAGING AND CLASSIFICATION SYSTEMS 15


4.1 Pathological staging 15
4.2 Tumour, node, metastasis classification 15

5. DIAGNOSTIC EVALUATION 16
5.1 Primary diagnosis 16
5.1.1 Symptoms 16
5.1.2 Physical examination 16
5.1.3 Bladder imaging 16
5.1.4 Urinary cytology 16
5.1.5 Cystoscopy 16
5.1.6 Transurethral resection of invasive bladder tumours 17
5.1.7 Concomitant prostate cancer 17
5.1.8 Summary of evidence and guidelines for the primary assessment of presumably
invasive bladder tumours 17
5.1.9 EAU-ESMO consensus statements on the management of advanced- and
variant bladder cancer 18
5.2 Imaging for staging of MIBC 18
5.2.1 Local staging of MIBC 18
5.2.1.1 CT imaging for local staging of MIBC 19
5.2.2 Imaging of lymph nodes in MIBC 19
5.2.3 Upper urinary tract urothelial carcinoma 19
5.2.3.1 Computed tomography urography 19
5.2.3.2 Magnetic resonance urography 19
5.2.4 Distant metastases at sites other than lymph nodes 19

2 LIMITED UPDATE MARCH 2021


5.2.5 Future developments 19
5.2.6 Summary of evidence and guidelines for staging in muscle-invasive bladder
cancer 20
5.3 MIBC and health status 20
5.3.1 Evaluation of comorbidity, frailty and cognition 20
5.3.2 Comorbidity scales, anaesthetic risk classification and geriatric assessment 22
5.3.3 Summary of evidence and guidelines for comorbidity scales 23

6. MARKERS 24
6.1 Introduction 24
6.2 Prognostic markers 24
6.2.1 Histopathological and clinical markers 24
6.2.2 Molecular markers 24
6.2.2.1 Molecular subtypes based on the Cancer Genome Atlas cohort 24
6.3 Predictive markers 25
6.3.1 Clinical and histopathological markers 25
6.3.2 Molecular markers 25
6.4 Conclusion 26
6.5 Summary of evidence and recommendations for urothelial markers 26

7. DISEASE MANAGEMENT 26
7.1 Neoadjuvant therapy 26
7.1.1 Introduction 26
7.1.2 Role of cisplatin-based chemotherapy 26
7.1.2.1 Summary of available data 27
7.1.3 The role of imaging and predictive biomarkers 28
7.1.4 Role of neoadjuvant immunotherapy 28
7.1.5 Summary of evidence and guidelines for neoadjuvant therapy 28
7.2 Pre- and post-operative radiotherapy in muscle-invasive bladder cancer 29
7.2.1 Post-operative radiotherapy 29
7.2.2 Pre-operative radiotherapy 29
7.2.3 Summary of evidence and guidelines for pre- and post-operative radiotherapy 30
7.2.4 EAU-ESMO consensus statements on the management of advanced- and
variant bladder cancer 30
7.3 Radical surgery and urinary diversion 30
7.3.1 Removal of the tumour-bearing bladder 30
7.3.1.1 Introduction 30
7.3.1.2 Radical cystectomy: timing 30
7.3.2 Radical cystectomy: indications 31
7.3.3 Radical cystectomy: technique and extent 31
7.3.3.1 Radical cystectomy in men 31
7.3.3.1.1 Summary of evidence and recommendations for sexual-
preserving techniques in men 32
7.3.3.2 Radical cystectomy in women 32
7.3.3.2.1 Summary of evidence and recommendations for sexual-
preserving techniques in women 32
7.3.4 Lymphadenectomy: role and extent 33
7.3.5 Laparoscopic/robotic-assisted laparoscopic cystectomy 33
7.3.5.1 Laparoscopic radical cystectomy versus robot-assisted radical
cystectomy 34
7.3.5.2 Summary of evidence and guidelines for laparoscopic/robotic-
assisted laparoscopic cystectomy 35
7.3.6 Urinary diversion after radical cystectomy 35
7.3.6.1 Patient selection and preparations for surgery 35
7.3.6.2 Different types of urinary diversion 36
7.3.6.2.1 Uretero-cutaneostomy 36
7.3.6.2.2 Ileal conduit 37
7.3.6.2.3 Orthotopic neobladder 37
7.3.7 Morbidity and mortality 37
7.3.8 Survival 39

LIMITED UPDATE MARCH 2021 3


7.3.9 Impact of hospital and surgeon volume on treatment outcomes 39
7.3.10 Summary of evidence and guidelines for radical cystectomy and
urinary diversion 40
7.3.11 EAU-ESMO consensus statements on the management of advanced-
and variant bladder cancer 41
7.4 Unresectable tumours 42
7.4.1 Palliative cystectomy for muscle-invasive bladder carcinoma 42
7.4.1.1 Guidelines for unresectable tumours 42
7.4.1.2 EAU-ESMO consensus statements on the management of advanced-
and variant bladder cancer 42
7.4.2 Supportive care 42
7.4.2.1 Obstruction of the upper urinary tract 42
7.4.2.2 Bleeding and pain 42
7.5 Bladder-sparing treatments for localised disease 42
7.5.1 Transurethral resection of bladder tumour 42
7.5.1.1 Guideline for transurethral resection of bladder tumour 43
7.5.1.2 EAU-ESMO consensus statements on the management of advanced-
and variant bladder cancer 43
7.5.2 External beam radiotherapy 43
7.5.2.1 Summary of evidence and guideline for external beam radiotherapy 44
7.5.2.2 EAU-ESMO consensus statements on the management of advanced-
and variant bladder cancer 44
7.5.3 Chemotherapy 44
7.5.3.1 Summary of evidence and guideline for chemotherapy 44
7.5.4 Trimodality bladder-preserving treatment 45
7.5.4.1 Summary of evidence and guidelines for trimodality bladder-
preserving treatment 46
7.5.4.2 EAU-ESMO consensus statements on the management of advanced-
and variant bladder cancer 47
7.6 Adjuvant therapy 47
7.6.1 Role of adjuvant platinum-based chemotherapy 47
7.6.2 Role of adjuvant immunotherapy 48
7.6.3 Guidelines for adjuvant therapy 48
7.7 Metastatic disease 48
7.7.1 Introduction 48
7.7.1.1 Prognostic factors and treatment decisions 48
7.7.1.2 Comorbidity in metastatic disease 49
7.7.2 First-line systemic therapy for metastatic disease 49
7.7.2.1 Definitions: ‘Fit for cisplatin, fit for carboplatin, unfit for any platinum-
based chemotherapy’ 49
7.7.2.2 Chemotherapy in patients fit for cisplatin 49
7.7.2.3 Chemotherapy in patients fit for carboplatin (but unfit for cisplatin) 50
7.7.2.4 Integration of immunotherapy in the first-line treatment of patients fit
for platinum-based chemotherapy 50
7.7.2.4.1 Immunotherapy combination approaches 50
7.7.2.4.2 Use of single-agent immunotherapy 50
7.7.2.4.3 Switch maintenance with immunotherapy 51
7.7.2.5 Treatment of patients unfit for any platinum-based chemotherapy 51
7.7.2.6 Non-platinum combination chemotherapy 51
7.7.3 Second-line systemic therapy for metastatic disease 51
7.7.3.1 Second-line chemotherapy 51
7.7.3.2 Second-line immunotherapy for platinum-pre-treated patients 52
7.7.3.2.1 Side-effect profile of immunotherapy 52
7.7.4 Novel agents for second- or later-line therapy 52
7.7.5 Post-chemotherapy surgery and oligometastatic disease 53
7.7.6 Treatment of patients with bone metastases 53
7.7.7 Summary of evidence and guidelines for metastatic disease 54
7.8 Quality of life 55
7.8.1 Introduction 55
7.8.2 Neoadjuvant chemotherapy 55

4 LIMITED UPDATE MARCH 2021


7.8.3 Radical cystectomy and urinary diversion 56
7.8.4 Bladder sparing trimodality therapy 56
7.8.5 Non-curative or metastatic bladder cancer 56
7.8.6 Summary of evidence and recommendations for health-related quality of life 56

8. FOLLOW-UP 57
8.1 Follow-up in muscle invasive bladder cancer 57
8.2 Site of recurrence 57
8.2.1 Local recurrence 57
8.2.2 Distant recurrence 57
8.2.3 Urothelial recurrences 58
8.3 Time schedule for surveillance 58
8.4 Follow-up of functional outcomes and complications 58
8.5 Summary of evidence and recommendations for specific recurrence sites 59
8.6 EAU-ESMO consensus statements on the management of advanced- and variant
bladder cancer 59

9. REFERENCES 60

10. CONFLICT OF INTEREST 94

11. CITATION INFORMATION 94

LIMITED UPDATE MARCH 2021 5


1. INTRODUCTION
1.1 Aims and scope
The European Association of Urology (EAU) Guidelines Panel for Muscle-invasive and Metastatic Bladder
Cancer (MIBC) have prepared these guidelines to help urologists assess the evidence-based management of
MIBC and to incorporate guideline recommendations into their clinical practice.
Separate EAU guidelines documents are available addressing upper urinary tract (UUT) tumours [1],
non-muscle-invasive bladder cancer (TaT1 and carcinoma in situ) (NMIBC) [2], and primary urethral carcinomas [3].
It must be emphasised that clinical guidelines present the best evidence available to the experts
but following guideline recommendations will not necessarily result in the best outcome. Guidelines can never
replace clinical expertise when making treatment decisions for individual patients, but rather help to focus
decisions - also taking personal values and preferences/individual circumstances of patients into account.
Guidelines are not mandates and do not purport to be a legal standard of care.

1.2 Panel composition


The EAU Guidelines Panel consists of an international multidisciplinary group of clinicians, including urologists,
oncologists, a pathologist, a radiologist and radiotherapists. Section 5.3 -MIBC and health status, was
developed with the assistance of Dr. S. O’Hanlon, consultant geriatrician, International Society of Geriatric
Oncology (SIOG) representative and member of the EAU-EANM-ESTRO-ESUR-SIOG Prostate Cancer
Guidelines Panel. The MIBC Panel is most grateful for his support.
All experts involved in the production of this document have submitted potential conflict
of interest statements which can be viewed on the EAU website Uroweb: http://uroweb.org/guideline/
bladdercancermuscle-invasive-and-metastatic/?type=panel.

1.3 Available publications


A quick reference document (Pocket Guidelines) is available, both in print and as an app for iOS and Android
devices. These are abridged versions which may require consultation together with the full text version.
Several scientific publications are available (the most recent paper dating back to 2020 [4]), as are
a number of translations of all versions of the EAU MIBC Guidelines. All documents are accessible through the
EAU website: http://uroweb.org/guideline/bladder-cancer-muscle-invasive-and-metastatic/.

1.4 Publication history and summary of changes


1.4.1 Publication history
The EAU published its first guidelines on bladder cancer (BC) in 2000. This document covered both NMIBC and
MIBC. Since these conditions require different treatment strategies, it was decided to give each condition its
own guidelines, resulting in the first publication of the MIBC Guidelines in 2004. This 2021 document presents
a limited update of the 2020 version.

1.4.2 Summary of changes


New relevant references have been identified through a structured assessment of the literature and
incorporated in the various chapters of the 2021 EAU MIBC Guidelines resulting in new sections and added
and revised recommendations in:

• Section 3.3.3 Guidelines for the assessment of tumour specimens

Recommendation Strength rating


Record the sampling sites, as well as information on tumour size when providing Strong
specimens to the pathologist.

•  ection 5.1.8 Summary of evidence and guidelines for the primary assessment of presumably invasive
S
bladder tumours

Summary of evidence LE
In men, prostatic urethral biopsy includes resection from the bladder neck to the verumontanum 2b
(between the 5 and 7 o’clock position) using a resection loop. In case any abnormal-looking
areas in the prostatic urethra are present at this time, these need to be biopsied as well.

6 LIMITED UPDATE MARCH 2021


Recommendations Strength rating
In men with a negative prostatic urethral biopsy undergoing subsequent orthotopic Strong
neobladder construction, an intra-operative frozen section can be omitted.
In men with a prior positive transurethral prostatic biopsy, subsequent orthotopic Strong
neobladder construction should not be denied a priori, unless an intra-operative
frozen section of the distal urethral stump reveals malignancy at the level of urethral
dissection.

• Section 5.2.1 - Local staging of MIBC; inclusion of data on multiparametric MRI using the Vesical Imaging
Reporting and Data System (VI-RADS) scoring system. No new recommendation has been provided.

• Section 5.3 - MIBC and health status; this section has been updated, introducing the concept of frailty.

• Chapter 6 - Markers; this chapter has been significantly revised, presenting two new recommendations.

6.5 Summary of evidence and recommendations for urothelial markers

Summary of evidence LE
There is insufficient evidence to use TMB, molecular subtypes, immune or other gene -
expression signatures for the management of patients with urothelial cancer.

Recommendations Strength rating


Evaluate PD-L1 expression (by immunohistochemistry) to determine the potential Weak
for use of pembrolizumab or atezolizumab in previously untreated patients with
locally advanced or metastatic urothelial cancer who are unfit for cisplatin-based
chemotherapy.
Evaluate for FGFR2/3 genetic alterations for the potential use of erdafitinib in Weak
patients with locally advanced or metastatic urothelial carcinoma who have
progressed following platinum-containing chemotherapy (including within 12 months
of neoadjuvant or adjuvant platinum-containing chemotherapy).

•  .2 - Pre- and post-operative radiotherapy in muscle-invasive bladder cancer; this section was revised
7
and new data added, resulting in an additional recommendation.

7.2.3 Summary of evidence and guidelines for pre- and post-operative radiotherapy

Summary of evidence LE
Addition of adjuvant RT to chemotherapy is associated with an improvement in local relapse- 2a
free survival following cystectomy for locally advanced bladder cancer (pT3b─4, or node-
positive).

Recommendation Strength rating


Consider offering adjuvant radiation in addition to chemotherapy following radical Weak
cystectomy, based on pathologic risk (pT3b–4, or positive nodes, or positive margins).

7.3.10 Summary of evidence and guidelines for radical cystectomy and urinary diversion

Summary of evidence LE
Ensuring that patients are well informed about the various urinary diversion options prior to 3
making a decision may help prevent or reduce decision regret, independent of the method of
diversion selected.

• 7.5.4 Trimodality bladder-preserving treatment

• 7.7 Metastatic disease: data from a number of key trials has been included, in particular on
immunotherapy combinations in first- and later-line setting, resulting in a number of new
recommendations and a change to the treatment flowchart (Figure 7.2).

LIMITED UPDATE MARCH 2021 7


7.7.8 Summary of evidence and guidelines for metastatic disease

Summary of evidence LE
PD-1 inhibitor pembrolizumab has been approved for patients that have progressed during or
after previous platinum-based chemotherapy based on the results of a phase III trial.
PD-L1 inhibitors atezolizumab, nivolumab, durvalumab and avelumab have been FDA approved
for patients that have progressed during or after previous platinum-based chemotherapy based
on the results of a phase II trial.
PD-1 inhibitor pembrolizumab and PD-L1 inhibitor atezolizumab have been approved for
patients with advanced or metastatic UC unfit for cisplatinum-based first-line chemotherapy
and with overexpression of PD-L1 based on the results of single-arm phase II trials.
The combination of chemotherapy plus pembrolizumab or atezolizumab and the combination
of durvalumab and tremelimumab have not demonstrated an OS survival benefit compared to
platinum-based chemotherapy alone.
Switch maintenance with the PD-L1 inhibitor avelumab has demonstrated significant OS benefit
in patients achieving at least stable disease on first-line platinum-based chemotherapy.

Recommendations Strength rating


First-line treatment for platinum-fit patients
Use cisplatin-containing combination chemotherapy with GC or HD-MVAC. Strong
In patients unfit for cisplatin but fit for carboplatin use the combination of Strong
carboplatin and gemcitabine.
In patients achieving stable disease, or better, after first-line platinum-based Strong
chemotherapy use maintenance treatment with PD-L1 inhibitor avelumab.
First-line treatment in patients unfit for platinum-based chemotherapy
Consider checkpoint inhibitors pembrolizumab or atezolizumab. Weak
Second-line treatment
Offer checkpoint inhibitor pembrolizumab to patients progressing during, or after, Strong
platinum-based combination chemotherapy for metastatic disease. If this is not
possible, offer atezolizumab, nivolumab (EMA, FDA approved); avelumab or
durvalumab (FDA approved).
Further treatment after platinum- and immunotherapy
Offer treatment in clinical trials testing novel antibody drug conjugates (enfortumab Strong
vedotin, sacituzumab govitecan); or in case of patients with FGFR3 alterations,
FGFR tyrosine kinase inhibitors.
GC = gemcitabine plus cisplatin; FGFR = fibroblast growth factor receptor; HD-MVAC = high-dose intensity
methotrexate, vinblastine, adriamycin plus cisplatin.

8 LIMITED UPDATE MARCH 2021


Figure 7.2: Flow chart for the management of metastatic urothelial cancer*

PLATINUM-ELIGIBLE PLATINUM-INELIGIBLE
PS 2 and GFR < 60 mL/min
PS > 2; GFR < 30 mL/min

cisplatin carboplatin
PS 0-1 and PS 2 or GFR 30-60
GFR > 50-60 mL/min mL/min

cisplatin/gemcitabine carboplatin/gemcitabine
or DD-MVAC 4-6 cycles 4-6 cycles
PD-L1 + PD-L1 -

PD CR/PR/SD

PD Watchful
2nd line therapy Maintenance
waiting

• pembrolizumab
Switch Immunotherapy Best
(atezolizumab, avelumab,
maintenance: • atezolizumab supportive
durvalumab, nivolumab)
avelumab • pembrolizumab care
• Trials

LATER-LINE THERAPY

FGFR3 mutation UC progressing on UC patients refractory to platinum-based


platinum-based chemotherapy ± prior IO chemotherapy and IO

• erdafitinib (FDA) – in trial • enfortumab vedotin (FDA) – in trial


• chemotherapy: • chemotherapy:
o paclitaxel o paclitaxel
o Docetaxel o docetaxel
o vinflunine o vinflunine
• Trials • Trials

*Treatment within clinical trials is highly encouraged.


BSC = best supportive care; CR = complete response; DD-MVAC = dose dense methotrexate vinblastine
doxorubicin cisplatin; EV = enfortumab vedotin; FDA = US Food and Drug Administration;
FGFR = pan-fibroblast growth factor receptor tyrosine kinase inhibitor; GFR = glomerular filtration rate;
IO = immunotherapy; PR = partial response; PS = performance status; SD = stable disease.

2. METHODS
2.1 Data identification
For the 2021 MIBC Guidelines, new and relevant evidence has been identified, collated and appraised through a
structured assessment of the literature. A broad and comprehensive literature search, covering all sections of the
MIBC Guideline was performed. The search was limited to English language publications. Databases searched
included Medline, EMBASE and the Cochrane Libraries, covering a time frame between May 10th, 2019 and
May 14th, 2020. A total of 1,837 unique records were identified, retrieved and screened for relevance resulting
in 83 new publications having been included in the 2021 print. A detailed search strategy is available online:
http://uroweb.org/guideline/bladdercancer-muscle-invasive-andmetastatic/?type=appendices-publications.

LIMITED UPDATE MARCH 2021 9


For each recommendation within the guidelines there is an accompanying online strength rating form, the basis
of which is a modified GRADE methodology [5, 6] which addresses a number of key elements namely:
1. the overall quality of the evidence which exists for the recommendation, references used in
this text are grade according to a classification system modified from the Oxford Centre for
Evidence-Based Medicine Levels of Evidence [7];
2. the magnitude of the effect (individual or combined effects);
3. the certainty of the results (precision, consistency, heterogeneity and other statistical or
study related factors);
4. the balance between desirable and undesirable outcomes;
5. the impact of patient values and preferences on the intervention;
6. the certainty of those patient values and preferences.

These key elements are the basis which panels use to define the strength rating of each recommendation.
The strength of each recommendation is represented by the words ‘strong’ or ‘weak’ [6]. The strength of each
recommendation is determined by the balance between desirable and undesirable consequences of alternative
management strategies, the quality of the evidence (including certainty of estimates), and nature and variability
of patient values and preferences.
Additional information can be found in the general Methodology section of this print, and online at
the EAU website; http://www.uroweb.org/guideline/. A list of Associations endorsing the EAU Guidelines can
also be viewed online at the above address.

The results of a collaborative multi-stakeholder consensus project on the management of advanced and
variant bladder cancer have been incorporated in the 2020 MIBC Guidelines update [8, 9]. Only statements
which reached the a priori defined level of agreement - ≥ 70% agreement and ≤ 15% disagreement - across
all stakeholders involved in this consensus project are listed. The methodology is presented in detail in the
scientific publications. Since the publication of these consensus papers, emerging evidence prompted a
re-evaluation of these findings, resulting in the removal of a number of consensus statements.

2.2 Peer-review
The 2021 print of the MIBC guidelines was peer reviewed prior to publication.

2.2.1 Lay review


Post publication, the 2018 MIBC Guidelines were shared with seven patients treated for MIBC. Their comments
were requested, but not limited to:
• the overall tone of the guidelines content;
• any missing information;
• any information considered incorrect;
• any information which is not presented in a clear fashion;
• any text which is considered redundant and should be omitted;
• any text section that should be more detailed.

Common comments across reviewers:


• In general, the overall tone of the text was considered informational and instructive, but the language
used obviously targets medical professionals, which make certain parts of the text difficult to understand
for lay persons. The use of many abbreviations is considered an additional hindrance, as are the
methodological elements. In case the EAU are considering producing a lay version of this text, the
language needs to be adapted and clear instructions are to be provided.
• It is difficult for lay reviewers to comment on what may be omitted since, in their opinion, they lack the
expertise.
• Some sections, such as ‘Recurrent disease’ and ‘Markers’ denote areas where less evidence is available.
Consequently, the available data is less systematically presented which makes these sections more
difficult to understand.
• There is an interest whether screening for BC is a consideration.
• In particular ‘follow-up’, ‘quality of life’ and ‘survivorship aspects’ should be elaborated on; providing
additional information on what may be expected after treatment is considered very helpful for patients
and their families. Also lifestyle elements would be of relevance (healthy living, “what to do to
prevent cancer”). For this section, in particular, involvement of patients in the text development was
considered missing. Transparency about the process of patient involvement in guidelines development
was considered most relevant.

The MIBC Guidelines Panel is most grateful for the unique insights and guidance provided by the lay reviewers.

10 LIMITED UPDATE MARCH 2021


2.3 Future goals
The MIBC Panel will integrate two patient advocates in the course of 2021, to ensure that patient views will be
appropriately represented in the course of the production of these guidelines.

Topics considered for inclusion in the 2022 update of the MIBC Guidelines:
• development of a diagnostic pathway for the assessment of visible and non-visible haematuria;
• participation in developing strategies to ensure meaningful participation of patients in the development
and implementation of the MIBC Guidelines.

3. EPIDEMIOLOGY, AETIOLOGY AND


PATHOLOGY
3.1 Epidemiology
Bladder cancer is the 7th most commonly diagnosed cancer in males, whilst it drops to 10th position when both
genders are considered [10]. The worldwide age-standardised incidence rate (per 100,000 person/years) is 9.5
for men and 2.4 for women [10]. In the European Union, the age-standardised incidence rate is 20 for men and
4.6 for women [10]. In Europe, the highest age-standardised incidence rate has been reported in Belgium (31 in
men and 6.2 in women) and the lowest in Finland (18.1 in men and 4.3 in women) [10].
Worldwide, the BC age-standardised mortality rate (per 100,000 person/years) was 3.2 for men
vs. 0.9 for women in 2012 [10]. Bladder cancer incidence and mortality rates vary across countries due to
differences in risk factors, detection and diagnostic practices, and availability of treatments. The variations
are, however, also partly caused by the different methodologies used in the studies and the quality of data
collection [11, 12].
The incidence and mortality of BC has decreased in some registries, possibly reflecting the
decreased impact of causative agents [12, 13].
Approximately 75% of patients with BC present with disease confined to the mucosa (stage Ta,
carcinoma in situ [CIS]) or submucosa (stage T1). In younger patients (< 40 years) this percentage is even
higher [14]. Patients with TaT1 and CIS have a high prevalence due to long-term survival in many cases and
lower risk of cancer-specific mortality (CSM) compared to T2-4 tumours [10, 11].

3.2 Aetiology
3.2.1 Tobacco smoking
Tobacco smoking is the most well-established risk factor for BC, causing 50–65% of male cases and 20–30%
of female cases [15]. A causal relationship has been established between exposure to tobacco and cancer in
studies in which chance, bias and confounding can be discounted with reasonable confidence [16].
The incidence of BC is directly related to the duration of smoking and the number of cigarettes
smoked per day [17]. A meta-analysis looked at 216 observational studies on cigarette smoking and
cancer published between 1961 and 2003, and the pooled risk estimates for BC demonstrated a significant
association for both current and former smokers [18]. Recently, an increase in risk estimates for current
smokers relative to never smokers has been described suggesting this could be due to changes in cigarette
composition [15]. Starting to smoke at a younger age increased the risk of death from BC [19]. An immediate
decrease in the risk of BC was observed in those who stopped smoking. The reduction was about 40% within
one to four years of quitting smoking and 60% after 25 years of cessation [17]. Encouraging people to stop
smoking would result in the incidence of BC decreasing equally in men and women [15].

3.2.2 Occupational exposure to chemicals


Occupational exposure is the second most important risk factor for BC. Work-related cases accounted for
20–25% of all BC cases in several series and it is likely to occur in occupations in which dyes, rubbers, textiles,
paints, leathers, and chemicals are used [20]. The risk of BC due to occupational exposure to carcinogenic
aromatic amines is significantly greater after ten years or more of exposure; the mean latency period usually
exceeds 30 years [21, 22]. Population-based studies established the occupational attribution for BC in men to
be 7.1%, while no such attribution was discernible for women [11, 23].

3.2.3 Radiotherapy
Increased rates of secondary bladder malignancies have been reported after external-beam radiotherapy
(EBRT) for gynaecological malignancies, with relative risks (RR) of 2-4 [24]. In a population-based cohort study,

LIMITED UPDATE MARCH 2021 11


the standardised incidence ratios for BC developing after radical prostatectomy (RP), EBRT, brachytherapy,
and EBRT-brachytherapy were 0.99, 1.42, 1.10, and 1.39, respectively, in comparison with the general U.S.
population [25].
It has recently been proposed that patients who have received radiotherapy (RT) for prostate
cancer with modern modalities such as intensity-modulated radiotherapy (IMRT) may have lower rates of
in-field bladder- and rectal secondary malignancies [26]. Nevertheless, since longer follow-up data are not
yet available, and as BC requires a long period to develop, patients treated with radiation and with a long life
expectancy are at a higher risk of developing BC [26].

3.2.4 Dietary factors


Several dietary factors have been related to BC; however, the links remain controversial. The European
Prospective Investigation into Cancer and Nutrition (EPIC) study is an on-going multicentre cohort study designed
to examine the association between diet, lifestyle, environmental factors and cancer. They found no links between
BC and fluid intake, red meat, vegetable and fruit consumption and only recently an inverse association between
dietary intake of flavonoids and lignans and the risk of aggressive BC tumours has been described [27].

3.2.5 Metabolic disorders


In a large prospective study pooling six cohorts from Norway, Sweden, and Austria (The Metabolic syndrome
and Cancer project, Me-Can 2.0), metabolic aberrations, especially elevated blood pressure and triglycerides,
were associated with increased risks of BC among men, whereas high body mass index (BMI) was associated
with decreased BC risk. The associations between BMI, blood pressure and BC risk significantly differed
between men and women [28].

The association of diabetes mellitus (DM) with the risk of BC has been evaluated in numerous meta-analyses
with inconsistent results. When analysing specific subpopulations, DM was associated with BC or CSM risk
especially in men [29]. Thiazolidinediones (pioglitazone and rosiglitazone) are oral hypoglycaemic drugs used
for the management of type 2 DM. Their use and the association with BC is still a matter of debate. In a recent
meta-analysis of observational studies the summary results indicated that pioglitazone use was significantly
associated with an increased risk of BC which appears to be linked to higher dose and longer duration of
treatment [30]. The U.S. Food and Drug Administration (FDA) recommend that healthcare professionals should
not prescribe pioglitazone in patients with active BC. Several countries in Europe have removed this agent from
the market or included warnings for prescription. Moreover, the benefits of glycaemic control vs. unknown risks
for cancer recurrence with pioglitazone should be considered in patients with a prior history of BC.

3.2.6 Bladder schistosomiasis and chronic urinary tract infection


Bladder schistosomiasis (bilharzia) is the second most common parasitic infection after malaria, with about
600 million people exposed to infection in Africa, Asia, South America, and the Caribbean [31]. There is a well-
established relationship between schistosomiasis and urothelial carcinoma (UC) of the bladder, which can
progress to squamous cell carcinoma (SCC), however, better control of the disease is decreasing the incidence
of SCC of the bladder in endemic zones such as Egypt [32, 33].
Similarly, invasive SCC has been linked to the presence of chronic urinary tract infection (UTI)
distinct from schistosomiasis. A direct association between BC and UTIs has been observed in several case-
control studies, which have reported a two-fold increased risk of BC in patients with recurrent UTIs in some
series [34]. However, a recent meta-analysis found no statistical association when pooling data from the
most recent and highest quality studies which highlights the need for higher quality data to be able to draw
conclusions [35].
Similarly, urinary calculi and chronic irritation or inflammation of the urothelium have been described
as possible risk factors for BC. A meta-analysis of case-control and cohort studies suggests a positive
association between history of urinary calculi and BC [36].

3.2.7 Gender
Although men are more likely to develop BC than women, women present with more advanced disease and
have worse survival rates. A meta-analysis including nearly 28,000 patients shows that female gender was
associated with a worse survival outcome (hazard ratio [HR]: 1.20, 95% CI: 1.09–1.32) compared to male
gender after radical cystectomy (RC) [37]. This finding had already been presented in a descriptive nation-
wide analysis based on 27,773 Austrian patients. After their analysis the authors found that cancer-specific
survival (CSS) was identical for pT1-tumours in both sexes, while women had a worse CSS in both age cohorts
(< 70 years and ≥ 70 years) with higher tumour stages [38]. However, treatment patterns are unlikely to explain
the differences in overall survival (OS) [39]. In a population-based study from the Ontario Cancer Registry
analysing all patients with BC treated with cystectomy or radical RT between 1994 and 2008, no differences in

12 LIMITED UPDATE MARCH 2021


OS, mortality and outcomes were found between males and females following radical therapy [40]. The gender-
specific difference in survival for patients with BC was also analysed in the Norwegian population. Survival
was inferior for female patients but only within the first 2 years after diagnosis. This discrepancy was partly
attributed to a more severe T-stage in female patients at initial diagnoses [41].
A population-based study from the MarketScan databases suggests that a possible reason for
worse survival in the female population may be that women experienced longer delays in diagnosis than men,
as the differential diagnosis in women includes diseases that are more prevalent than BC [42]. Furthermore,
differences in the gender prevalence of BC may be due to other factors besides tobacco and chemical
exposure. In a large prospective cohort study, post-menopausal status was associated with an increase in
BC risk, even after adjustment for smoking status. This finding suggests that the differences in oestrogen
and androgen levels between men and women may be responsible for some of the difference in the gender
prevalence of BC [43-45]. Moreover, a recent population study assessing impact of hormones on BC suggests
that younger age at menopause (≤ 45 years) is associated with an increased risk of BC [46].

3.2.8 Genetic factors


There is growing evidence that genetic susceptibility factors and family association may influence the
incidence of BC. A recent population-based study of cancer risk in relatives and spouses of UC patients
showed an increased risk for first- and second-degree relatives, and suggests genetic or environmental
roots independent of smoking-related behaviour [47]. Shared environmental exposure was recognised as a
potentially confounding factor [48]. Recent studies detected genetic susceptibility with independent loci, which
are associated with BC risk [49].
Genome-wide association studies (GWAS) of BC identified several susceptibility loci associated
with BC risk [50, 51].

3.2.9 Summary of evidence and guidelines for epidemiology and risk factors

Summary of evidence LE
Worldwide, bladder cancer is the 10th most commonly diagnosed cancer. 2a
Several risk factors associated with bladder cancer diagnosis have been identified. 3
Active and passive tobacco smoking continues to be the main risk factor, while exposure-related 2a
incidence is decreasing.
The increased risk of developing bladder cancer in patients undergoing external-beam radiotherapy 3
(EBRT), brachytherapy, or a combination of EBRT and brachytherapy, must be considered during
patient follow-up. As bladder cancer requires time to develop, patients treated with radiation at a
young age are at the greatest risk and should be followed-up closely.

Recommendations Strength rating


Council patients to stop active and avoid passive smoking. Strong
Inform workers in potentially hazardous workplaces of the potential carcinogenic effects of Strong
a number of recognised substances, including duration of exposure and latency periods.
Protective measures are recommended.
Do not prescribe pioglitazone to patients with active bladder cancer or a history of bladder Strong
cancer.

3.3 Pathology
3.3.1 Handling of transurethral resection and cystectomy specimens
During transurethral resection (TUR), a specimen from the tumour and normal looking bladder wall should be
taken, if possible. Specimens should be taken from the superficial and deep areas of the tumour and sent
to the pathology laboratory separately, in case the outcome will impact on treatment decisions. If random
biopsies of the flat mucosa are taken, each biopsy specimen of the flat mucosa should also be submitted
separately [52]. The sampling sites must be recorded by the urologist; the pathologist report should include
location of tumour tissue in the cystectomy specimen. Anatomical tumour location is relevant for staging and
prognosis [53, 54].
In RC, bladder fixation must be carried out as soon as possible. The pathologist must open the
specimen from the urethra to the bladder dome and fix the specimen. In a female cystectomy specimen,
the length of the urethral segment removed en bloc with the specimen should be checked, preferably by the
urological surgeon [55].

LIMITED UPDATE MARCH 2021 13


Specimen handling should follow the general rules as published by a collaborative group of pathologists and
urologists [56, 57]. It must be stressed that it may be very difficult to confirm the presence of a neoplastic lesion
using gross examination of the cystectomy specimen after TUR or chemotherapy, so the entire retracted or
ulcerated area should be included.
It is compulsory to study the urethra, the ureters, the prostate in men and the radial margins [58]. In
urethra-sparing cystectomy; the level of urethral dissection, completeness of the prostate, specifically at the
apex (in men), and the inclusion of the entire bladder neck and amount of adjacent urethra, uterus and vaginal
vault (in women) have to be documented by the pathologist.

All lymph node (LN) specimens should be provided in their totality, in clearly labelled containers. In case of
doubt, or adipose differentiation of the LNs, the entire specimen is to be included. Lymph nodes should be
counted and measured on slides; capsular extension and percentage of LN invasion should be reported as
well as vascular embols [59, 60]. In case of metastatic spread in the perivesical fat without real LN structures
(capsule, subcapsular sinus), this localisation should nevertheless be considered as N+.
Potentially positive soft tissue margins should be inked by the pathologist for evaluation [61]. In rare
cases, fresh frozen sections may be helpful to determine treatment strategy [62].

3.3.2 Pathology of muscle-invasive bladder cancer


All MIBC cases are high-grade UCs. For this reason, no prognostic information can be provided by grading
MIBC [63]. However, identification of morphological subtypes is important for prognostic reasons and treatment
decisions [64-66].

The data presented in these guidelines are based on the 2004/2016 World Health Organization (WHO)
classifications [67, 68].

Currently the following differentiations are used [64, 69]:


1. urothelial carcinoma (more than 90% of all cases);
2. urothelial carcinomas with partial squamous and/or glandular or trophoblastic differentiation;
3. micropapillary urothelial carcinoma;
4. nested variant (including large nested variant) and microcystic urothelial carcinoma;
5. plasmacytoid, giant cell, signet ring, diffuse, undifferentiated;
6. lymphoepithelioma-like;
7. small-cell carcinomas;
8. sarcomatoid urothelial carcinomas;
9. neuroendocrine variant of urothelial carcinoma;
10. some urothelial carcinomas with other rare differentiations.

Outcomes may vary for divergent histologies, which need to be mentioned following international reporting
standards [64, 70].

3.3.3 Guidelines for the assessment of tumour specimens

Recommendations Strength rating


Record the depth of invasion (categories pT2a and pT2b, pT3a and pT3b or pT4a and pT4b). Strong
Record margins with special attention paid to the radial margin, prostate, ureter, urethra,
peritoneal fat, uterus and vaginal vault.
Record the total number of lymph nodes (LNs), the number of positive LNs and extranodal
spread.
Record lymphatic or blood vessel invasion.
Record the presence of carcinoma in situ.
Record the sampling sites as well as information on tumour size when providing specimens
to the pathologist.

14 LIMITED UPDATE MARCH 2021


3.3.4 EAU-ESMO consensus statements on the management of advanced- and variant bladder
cancer [8, 9]*

Consensus statement
Bladder urothelial carcinoma with small cell neuroendocrine variant should be treated with neoadjuvant
chemotherapy followed by consolidating local therapy.
Muscle-invasive pure squamous cell carcinoma of the bladder should be treated with primary radical
cystectomy and lymphadenectomy.
Muscle-invasive pure adenocarcinoma of the bladder should be treated with primary radical cystectomy and
lymphadenectomy.
Muscle-invasive small cell neuroendocrine variant of bladder urothelial carcinoma should not receive
preventive brain irradiation to avoid brain recurrence.
Differentiating between urachal and non-urachal subtypes of adenocarcinoma is essential when making
treatment decisions.
*Only statements which met the a priori consensus threshold across all three stakeholder groups are listed
(defined as ≥ 70% agreement and ≤ 15% disagreement, or vice versa).

4. STAGING AND CLASSIFICATION SYSTEMS


4.1 Pathological staging
For staging, the Tumour, Node, Metastasis (TNM) Classification (2017, 8th edition) is recommended [71]. Blood
and lymphatic vessel invasion have an independent prognostic significance [72, 73].

4.2 Tumour, node, metastasis classification


The TNM classification of malignant tumours is the method most widely used to classify the extent of cancer
spread [71] (Table 4.1).

Table 4.1: TNM Classification of urinary bladder cancer [71]

T - Primary Tumour
Tx Primary tumour cannot be assessed
T0 No evidence of primary tumour
Ta Non-invasive papillary carcinoma
Tis Carcinoma in situ: “flat tumour”
T1 Tumour invades subepithelial connective tissue
T2 Tumour invades muscle
T2a Tumour invades superficial muscle (inner half)
T2b Tumour invades deep muscle (outer half)
T3 Tumour invades perivesical tissue:
T3a microscopically
T3b macroscopically (extravesical mass)
T4 Tumour invades any of the following: prostate stroma, seminal vesicles, uterus, vagina, pelvic wall,
abdominal wall
T4a Tumour invades prostate stroma, seminal vesicles, uterus, or vagina
T4b Tumour invades pelvic wall or abdominal wall
N - Regional Lymph Nodes
Nx Regional lymph nodes cannot be assessed
N0 No regional lymph node metastasis
N1 Metastasis in a single lymph node in the true pelvis (hypogastric, obturator, external iliac, or presacral)
N2 Metastasis in multiple regional lymph nodes in the true pelvis (hypogastric, obturator, external iliac, or
presacral)
N3 Metastasis in a common iliac lymph node(s)

LIMITED UPDATE MARCH 2021 15


M - Distant Metastasis
M0 No distant metastasis
M1a Non-regional lymph nodes
M1b Other distant metastasis

Staging after neo-adjuvant chemotherapy (NAC) and radical cystectomy can be done, but must be mentioned
as ypTNM (International Collaboration on Cancer Reporting) [70]. ypT0N0 after NAC and cystectomy is
associated with good prognosis [74, 75].

5. DIAGNOSTIC EVALUATION
5.1 Primary diagnosis
5.1.1 Symptoms
Painless visible haematuria is the most common presenting complaint. Other presenting symptoms and clinical
signs include non-visible haematuria, urgency, dysuria, increased frequency, and in more advanced tumours,
pelvic pain and symptoms related to urinary tract obstruction.

5.1.2 Physical examination


Physical examination should include rectal and vaginal bimanual palpation. A palpable pelvic mass can be
found in patients with locally advanced tumours. In addition, bimanual examination under anaesthesia should
be carried out before and after TUR of the bladder (TURB) to assess whether there is a palpable mass or if the
tumour is fixed to the pelvic wall [76, 77]. However, considering the discrepancy between bimanual examination
and pT stage after cystectomy (11% clinical overstaging and 31% clinical understaging), some caution is
suggested with the interpretation of bimanual examination [78].

5.1.3 Bladder imaging


Patients with a bladder mass identified by any diagnostic imaging technique should undergo cystoscopy,
biopsy and/or resection for histopathological diagnosis and staging.
The high specificity of diagnostic imaging for detecting bladder cancer means that patients
with imaging positive for bladder cancer may avoid diagnostic flexible cystoscopy and go directly to rigid
cystoscopy and transurethral resection [79, 80].

5.1.4 Urinary cytology


Examination of voided urine or bladder washings for exfoliated cancer cells has high sensitivity in high-grade
tumours and is a useful indicator in cases of high-grade malignancy or CIS. However, positive urinary cytology
may originate from a urothelial tumour located anywhere in the urinary tract.
Evaluation of cytology specimens can be hampered by low cellular yield, UTIs, stones or intravesical
instillations, but for experienced readers, specificity exceeds 90% [81, 82]. However, negative cytology does
not exclude a tumour. There is no known urinary marker specific for the diagnosis of invasive BC [83].

A standardised reporting system, the ‘Paris System’ redefining urinary cytology diagnostic categories was
published in 2016 [84]:
• adequacy of urine specimens (Adequacy);
• negative for high-grade UC (Negative);
• atypical urothelial cells (AUC);
• suspicious for high-grade UC (Suspicious);
• high-grade UC (HGUC);
• low-grade urothelial neoplasia (LGUN).

5.1.5 Cystoscopy
Ultimately, the diagnosis of BC is made by cystoscopy and histological evaluation of resected tissue. An
(outpatient) flexible cystoscopy is recommended to obtain a complete image of the bladder. However, in
daily practice, If a bladder tumour has been visualised unequivocally by imaging studies such as computed
tomography (CT), magnetic resonance imaging (MRI), or ultrasound (US), diagnostic cystoscopy may be
omitted and the patient can proceed directly to TURB for histological diagnosis and resection. During the
procedure, a thorough investigation of the bladder with rigid cystoscopy under anaesthesia is mandatory in
order not to miss any tumours at the level of the bladder neck.

16 LIMITED UPDATE MARCH 2021


Currently, there is no evidence for the role of photodynamic diagnosis (PDD) in the standard diagnosis of
invasive BC.
A careful description of the cystoscopic findings is necessary. This should include documentation
of the site, size, number, and appearance (papillary or solid) of the tumours, as well as a description of any
mucosal abnormalities [85]. The use of a bladder diagram is recommended.
The use of PDD could be considered if a T1 high-grade tumour is present and to identify associated
CIS. Presence of CIS may lead to a modified treatment plan (see EAU Guidelines on Non-muscle-invasive
Bladder Cancer [2]). Photodynamic diagnosis is highly sensitive for the detection of CIS and in experienced
hands the rate of false-positive results may be similar to that with regular white-light cystoscopy [73, 86].

5.1.6 Transurethral resection of invasive bladder tumours


The goal of TURB is to enable histopathological diagnosis and staging, which requires the inclusion of bladder
muscle in the resection specimen.
In case MIBC is suspected, tumours need to be resected separately in parts, which include the
exophytic part of the tumour, the underlying bladder wall with the detrusor muscle, and the edges of the
resection area. At least the deeper part of the resection specimen must be referred to the pathologist in a
separate labelled container to enable making a correct diagnosis. In cases in which RT is considered and CIS is
to be excluded, PDD can be used [87].
The involvement of the prostatic urethra and ducts in men with bladder tumours has been
reported. The exact risk is not known, but it seems to be higher if the tumour is located on the trigone or
bladder neck, with concomitant bladder CIS, and in the case of multiple tumours [54, 88, 89]. Involvement
of the prostatic urethra can be determined either at the time of primary TURB or by frozen section during the
cystoprostatectomy procedure. A frozen section has a higher negative-predictive value and is more accurate
[90-92].

A negative urethral frozen section can reliably identify patients in whom urethrectomy should be avoided.
However, a positive pre-operative biopsy seems to have limited utility as these findings are not reliably
associated with final margin status [90, 93].
Diagnosis of a urethral tumour before cystectomy will result in a urethrectomy which could be a
contraindication for an orthotopic diversion. However, an orthotopic diversion should not be denied based
on positive pre-operative biopsy findings alone and frozen section should be part of the radical cystectomy
procedure, in particular in male patients [94, 95].

5.1.7 Concomitant prostate cancer


Prostate cancer is found in 21–50% of male patients undergoing RC for BC [96-99]. Incidentally discovered
clinically significant prostatic adenocarcinoma did not alter survival [98, 99]. Pathological reporting of
the specimens should follow the recommendations as presented in the EAU-EANM-ESTRO-ESUR-SIOG
Guidelines on Prostate Cancer [100].

5.1.8 Summary of evidence and guidelines for the primary assessment of presumably invasive
bladder tumours

Summary of evidence LE
Cystoscopy is necessary for the diagnosis of bladder cancer. 1
Urinary cytology has high sensitivity in high-grade tumours including carcinoma in situ. 2b
In men, prostatic urethral biopsy includes resection from the bladder neck to the verumontanum 2b
(between the 5 and 7 o’clock position) using a resection loop. In case any abnormal-looking areas in
the prostatic urethra are present at this time, these need to be biopsied as well.

Recommendations Strength rating


Describe all macroscopic features of the tumour (site, size, number and appearance) and Strong
mucosal abnormalities during cystoscopy. Use a bladder diagram.
Take a biopsy of the prostatic urethra in cases of bladder neck tumour, when bladder Strong
carcinoma in situ is present or suspected, when there is positive cytology without evidence
of tumour in the bladder, or when abnormalities of the prostatic urethra are visible.
In men with a negative prostatic urethral biopsy undergoing subsequent orthotopic Strong
neobladder construction an intra-operative frozen section can be omitted.

LIMITED UPDATE MARCH 2021 17


In men with a prior positive transurethral prostatic biopsy, subsequent orthotopic Strong
neobladder construction should not be denied a priori, unless an intra-operative frozen
section of the distal urethral stump reveals malignancy at the level of urethral dissection.
In women undergoing subsequent orthotopic neobladder construction, obtain procedural Strong
information (including histological evaluation) of the bladder neck and urethral margin, either
prior to, or at the time of cystectomy.
In the pathology report, specify the grade, depth of tumour invasion, and whether the Strong
lamina propria and muscle tissue are present in the specimen.
(For general information on the assessment of bladder tumours, see EAU Guidelines on Non-muscle-invasive
Bladder Cancer [2]).

5.1.9 EAU-ESMO consensus statements on the management of advanced- and variant bladder
cancer [8, 9]*

Consensus statement
Differentiating between urachal and non-urachal subtypes of adenocarcinoma is essential when making
treatment decisions.
*Only statements which met the a priori consensus threshold across all three stakeholder groups are listed
(defined as ≥ 70% agreement and ≤ 15% disagreement, or vice versa).

5.2 Imaging for staging of MIBC


In clinical practice, tumour stage and histopathological grade are used to guide treatment and determine
prognosis [69, 101, 102].
Computed tomography and MRI are the imaging techniques most commonly used for tumour
staging. If the correct choice of treatment is to be made, staging must be accurate.

The goal of imaging patients with bladder cancer is to determine:


• extent of local tumour invasion, ideally differentiating T1 from T2 tumours as the treatment is different;
• tumour spread to LNs;
• tumour spread to the upper UT and other distant organs (e.g., liver, lungs, bones, peritoneum, pleura, and
adrenal glands).

5.2.1 Local staging of MIBC


Multiparametric MRI (mpMRI) using the Vesical Imaging Reporting and Data System (VI-RADS) scoring system
may differentiate between muscle- and non-muscle-invasive primary BC with high diagnostic accuracy. A
recent meta-analysis found that the pooled sensitivity and specificity of mpMRI with VI-RADS acquisition
and scoring for predicting MIBC were 0.83 and 0.90, respectively. This diagnostic performance of VI-RADS is
similar to the diagnostic performance of bladder MRI in determining MIBC prior to the introduction of VI-RADS
based on a previous meta-analysis of 24 studies in which the pooled sensitivity and specificity were 0.92
(95% CI: 0.88–0.95) and 0.87 (95% CI: 0.78–0.93), respectively [103]. The Vesical Imaging Reporting and Data
System offers a standardised approach to both acquisition and reporting of mpMRI for BC. How mpMRI is best
used in clinical practice and which cut off levels to use for VI-RADS scoring is still to be determined [103].

Magnetic resonance imaging has superior soft tissue contrast resolution compared with CT. The accuracy of
MRI for primary tumour staging varies from 73% to 96% (mean 85%). A meta-analysis of 17 studies showed
a 91% sensitivity and 96% specificity for 3.0-T device MRI combined with diffusion-weighted imaging (DWI) to
differentiate ≤ T1 tumours from ≥ T2 tumours before surgery [104].
Both CT and MRI may be used for assessment of local invasion by T3b disease, or higher, but they
are unable to accurately diagnose microscopic invasion of perivesical fat (T2 vs. T3a) [105].

Magnetic resonance imaging may evaluate post-biopsy reaction because enhancement of the tumour occurs
earlier than that of the normal bladder wall due to neovascularisation [106, 107].

In 2006, a link was established between the use of gadolinium-based contrast agents and nephrogenic
systemic fibrosis (NSF), which may result in fatal or severely debilitating systemic fibrosis. Patients with
impaired renal function are at risk of developing NSF and non-ionic linear gadolinium-based contrast agents
should be avoided (gadodiamide, gadopentetate dimeglumine and gadoversetamide). A stable macrocyclic
contrast agent should be used (gadobutrol, gadoterate meglumine or gadoteridol). Contrast-enhanced CT
using iodinated contrast media can be considered as an alternative [108].

18 LIMITED UPDATE MARCH 2021


5.2.1.1 CT imaging for local staging of MIBC
The advantages of CT include high spatial resolution, shorter acquisition time, wider coverage in a single breath
hold, and lower susceptibility to variable patient factors. Computed tomography is unable to differentiate
between stages Ta to T3a tumours, but it is useful for detecting invasion into the perivesical fat (T3b) and
adjacent organs. The accuracy of CT in determining extravesical tumour extension varies from 55% to 92%
[109] and increases with more advanced disease [110].

5.2.2 Imaging of lymph nodes in MIBC


Assessment of LN metastases based solely on size is limited by the inability of both CT and MRI to identify
metastases in normal-sized or minimally-enlarged nodes. The sensitivity for detection of LN metastases is low
(48–87%). Specificity is also low because nodal enlargement may be due to benign disease. Overall, CT and
MRI show similar results in the detection of LN metastases in a variety of primary pelvic tumours [111-115].
Pelvic nodes > 8 mm and abdominal nodes > 10 mm in maximum short-axis diameter, detected by CT or MRI,
should be regarded as pathologically enlarged [116, 117].
Positron emission tomography (PET) combined with CT is increasingly being used in clinical
practice and its exact role continues to be evaluated [118].

5.2.3 Upper urinary tract urothelial carcinoma


5.2.3.1 Computed tomography urography
Computed tomography urography has the highest diagnostic accuracy of the available imaging techniques
[119]. The sensitivity of CT urography for UTUC is 0.67–1.0 and specificity is 0.93–0.99 [120].
Rapid acquisition of thin sections allows high-resolution isotropic images that can be viewed in
multiple planes to assist with diagnosis without loss of resolution. Epithelial “flat lesions” without mass effect or
urothelial thickening are generally not visible with CT.
The secondary sign of hydronephrosis is associated with advanced disease and poor oncological
outcome [121, 122]. The presence of enlarged LNs is highly predictive of metastases in UTUC [123].

5.2.3.2 Magnetic resonance urography


Magnetic resonance urography is indicated in patients who cannot undergo CT urography, usually when
radiation or iodinated contrast media are contraindicated [124]. The sensitivity of MR urography is 0.75 after
contrast injection for tumours < 2 cm [124]. The use of MR urography with gadolinium-based contrast media
should be limited in patients with severe renal impairment (< 30 mL/min creatinine clearance), due to the risk
of NSF. Computed tomography urography is generally preferred to MR urography for diagnosing and staging
UTUC.

5.2.4 Distant metastases at sites other than lymph nodes


Prior to any curative treatment, it is essential to evaluate the presence of distant metastases. Computed
tomography and MRI are the diagnostic techniques of choice to detect lung [125] and liver metastases [126],
respectively. Bone and brain metastases are rare at the time of presentation of invasive BC. A bone scan and
additional brain imaging are therefore not routinely indicated unless the patient has specific symptoms or signs
to suggest bone or brain metastases [127, 128]. Magnetic resonance imaging is more sensitive and specific for
diagnosing bone metastases than bone scintigraphy [129, 130].

5.2.5 Future developments


Evidence is accruing in the literature suggesting that 18F-fluorodeoxyglucose (FDG)-positron emission
tomography (PET)/CT might have potential clinical use for staging metastatic BC [131, 132], but there is no
consensus as yet. The results of further trials are awaited before a recommendation can be made. Recently,
the first study was published showing the superior feasibility of DWI over T2-weighted and dynamic contrast-
enhanced (DCE)-MRI in assessing the therapeutic response to induction chemotherapy against MIBC [133].
The high specificity of DWI indicates that it is useful for accurate prediction of a complete histopathological
response, allowing better patient selection for bladder-sparing protocols. Results from prospective studies are
awaited.

LIMITED UPDATE MARCH 2021 19


5.2.6 Summary of evidence and guidelines for staging in muscle-invasive bladder cancer

Summary of evidence LE
Imaging as part of staging in muscle-invasive bladder cancer (MIBC) provides information about 2b
prognosis and assists in selection of the most appropriate treatment.
There are currently insufficient data on the use of diffusion-weighted imaging (DWI) and
18F-fluorodeoxyglucose-positron emission tomography/computed tomography (FDG-PET/CT) in MIBC

to allow for a recommendation to be made.


The diagnosis of upper tract urothelial carcinoma depends on CT urography and ureteroscopy. 2

Recommendations Strength rating


In patients with confirmed muscle-invasive bladder cancer, use computed tomography (CT) Strong
of the chest, abdomen and pelvis for staging, including some form of CT urography with
designated phases for optimal urothelial evaluation.
Use magnetic resonance urography when CT urography is contraindicated for reasons Strong
related to contrast administration or radiation dose.
Use CT or magnetic resonance imaging (MRI) for staging locally advanced or metastatic Strong
disease in patients in whom radical treatment is considered.
Use CT to diagnose pulmonary metastases. Computed tomography and MRI are generally Strong
equivalent for diagnosing local disease and distant metastases in the abdomen.

5.3 MIBC and health status


Complications from RC may be directly related to pre-existing comorbidity as well as the surgical procedure,
bowel anastomosis, or urinary diversion. A significant body of literature has evaluated the usefulness of age
as a prognostic factor for RC, although chronological age is less important than frailty [134-136]. Frailty is a
syndrome of reduced ability to respond to stressors. Patients with frailty have a higher risk of mortality and
negative side effects of cancer treatment [137]. Controversy remains regarding age, RC and the type of urinary
diversion. Radical cystectomy is associated with the greatest risk reduction in disease-related and non-
disease-related death in patients aged < 80 years [138].
The largest retrospective study on RC in septuagenarians and octogenarians based on data from
the National Surgical Quality Improvement Program database (n = 1,710) showed no significant difference
for wound, cardiac, or pulmonary complications. However, the risk of mortality in octogenarians compared
to septuagenarians is higher (4.3% vs. 2.3%) [139]. Although some octogenarians successfully underwent a
neobladder procedure, most patients were treated with an ileal conduit diversion. It is important to evaluate
functioning and quality of life (QoL) of older patients using a standardised geriatric assessment, as well as
carrying out a standard medical evaluation [140].

Sarcopenia has been shown to be an independent predictor for OS and CSS in a large multicentre study
with patients undergoing RC for BC [141]. In order to predict CSM after RC in patients receiving neoadjuvant
chemotherapy (NAC), sarcopenia should be assessed after completing the chemotherapy [142]. Other
risk factors for morbidity include prior abdominal surgery, extravesical disease, and prior RT [143]. Female
gender, an increased BMI and lower pre-operative albumin levels are associated with a higher rate of
parastomal hernias [144]. Low pre-operative serum albumin is also associated with impaired wound healing,
gastrointestinal (GI) complications and a decrease of recurrence-free and OS after RC [145, 146]. Therefore, it
could be used as a prognostic biomarker for patients undergoing RC.

5.3.1 Evaluation of comorbidity, frailty and cognition


Rochon et al. have shown that evaluation of comorbidity provides a better indicator of life expectancy in MIBC
than patient age [147]. Evaluation of comorbidity helps to identify factors likely to interfere with, or have an
impact on, treatment and the evolution and prognosis of MIBC [148].
The value of assessing overall health before recommending and proceeding with surgery was
emphasised by Zietman et al., who have demonstrated an association between comorbidity and adverse
pathological and survival outcomes following RC [149]. Similar results were found for the impact of comorbidity
on cancer-specific and other-cause mortality in a population-based competing risk analysis of > 11,260
patients from the Surveillance, Epidemiology, and End Results (SEER) registries. Age carried the highest risk for
other-cause mortality but not for increased cancer-specific death, while the stage of locally advanced tumour
was the strongest predictor for decreased CSS [150].
Stratifying older patients according to frailty using a multidisciplinary approach will help select

20 LIMITED UPDATE MARCH 2021


patients most likely to benefit from radical surgery and to optimise treatment outcomes [151]. There are many
different screening tools available for frailty and local approaches can be used. Examples include the G8 and
the Clinical Frailty Scale (See Table 5.1 and Figure 5.1 below).
Cognitive impairment can be screened for using a tool such as the mini-COG (https://mini-cog.
com/), which consists of three-word recall and a clock-drawing test, and can be completed within 5 minutes.
A score of ≤ 3/5 indicates the need to refer the patient for full cognitive assessment. Patients with any form of
cognitive impairment (e.g., Alzheimer’s or vascular dementia) may need a capacity assessment of their ability
to make an informed decision, which is an important factor in health status assessment. Cognitive impairment
also predicts risk of delirium, which is important for patients undergoing surgery [152].

Table 5.1: G8 screening tool (adapted from [153])

Items Possible responses (score)


A Has food intake declined over the past 3 months 0 = severe decrease in food intake
due to loss of appetite, digestive problems, 1 = moderate decrease in food intake
chewing, or swallowing difficulties? 2 = no decrease in food intake
B Weight loss during the last 3 months? 0 = weight loss > 3 kg
1 = does not know
2 = weight loss between 1 and 3 kg
3 = no weight loss
C Mobility? 0 = bed or chair bound
1 = able to get out of bed/chair but does not go out
2 = goes out
D Neuropsychological problems? 0 = severe dementia or depression
1 = mild dementia
2 = no psychological problems
E BMI? (weight in kg)/(height in m2) 0 = BMI < 19
1 = BMI 19 to < 21
2 = BMI 21 to < 23
3 = BMI ≥ 23
F Takes more than three prescription drugs per day? 0 = yes
1 = no
G In comparison with other people of the same 0.0 = not as good
age, how does the patient consider his/her health 0.5 = does not know
status? 1.0 = as good
2.0 = better
H Age 0 = ≥ 85
1 = 80–85
2 = < 80
Total score 0–17

LIMITED UPDATE MARCH 2021 21


Figure 5.1: Clinical Frailty Scale©, Version 2.0* [154]

CLINICAL FRAILTY SCALE 6 LIVING


WITH
People who need help with all outside
activities and with keeping house.
MODERATE Inside, they often have problems with
stairs and need help with bathing and
1 VERY People who are robust, active, energetic
FIT and motivated. They tend to exercise
FRAILTY might need minimal assistance (cuing,
standby) with dressing.
regularly and are among the fittest for
their age.
7 LIVING
WITH
Completely dependent for personal
care, from whatever cause (physical or
2 FIT People who have no active disease
symptoms but are less fit than category
SEVERE
FRAILTY
cognitive). Even so, they seem stable
and not at high risk of dying (within ~6
1. Often, they exercise or are very active months).
occasionally, e.g., seasonally.
8 LIVING Completely dependent for personal care

3 MANAGING People whose medical problems are


WELL well controlled, even if occasionally
WITH VERY
SEVERE
and approaching end of life. Typically,
they could not recover even from a
symptomatic, but often are not FRAILTY minor illness.
regularly active beyond routine walking.
9 TERMINALLY Approaching the end of life. This
4 LIVING
WITH
Previously “vulnerable,” this category
marks early transition from complete
ILL category applies to people with a life
expectancy <6 months, who are not
VERY MILD independence. While not dependent on otherwise living with severe frailty.
(Many terminally ill people can still
FRAILTY others for daily help, often symptoms
exercise until very close to death.)
limit activities. A common complaint
is being “slowed up” and/or being tired
during the day. SCORING FRAILTY IN PEOPLE WITH DEMENTIA
The degree of frailty generally In moderate dementia, recent memory is
5 LIVING
WITH
People who often have more evident
slowing, and need help with high
corresponds to the degree of
dementia. Common symptoms in
very impaired, even though they seemingly
can remember their past life events well.
mild dementia include forgetting They can do personal care with prompting.
MILD order instrumental activities of daily
the details of a recent event, though In severe dementia, they cannot do
FRAILTY living (finances, transportation, heavy still remembering the event itself, personal care without help.
housework). Typically, mild frailty repeating the same question/story
In very severe dementia they are often
progressively impairs shopping and and social withdrawal.
bedfast. Many are virtually mute.
walking outside alone, meal preparation,
medications and begins to restrict light Clinical Frailty Scale ©2005–2020 Rockwood,
housework. Version 2.0 (EN). All rights reserved. For permission:
www.geriatricmedicineresearch.ca
Rockwood K et al. A global clinical measure of fitness
www.geriatricmedicineresearch.ca and frailty in elderly people. CMAJ 2005;173:489–495.

*Permission to reproduce the Clinical Frailty Scale© has been granted by the copyright holder.

5.3.2 Comorbidity scales, anaesthetic risk classification and geriatric assessment


A range of comorbidity scales has been developed [155], seven of which have been validated [156-162]. The
Charlson Comorbidity Index (CCI) ranges from 0 to 30 according to the importance of comorbidity described
at four levels and is calculated by healthcare practitioners based on patients’ medical records. The score has
been widely studied in patients with BC and found to be an independent prognostic factor for peri-operative
mortality [163, 164], overall mortality [165], and CSM [138, 166-168]. Only the age-adjusted version of the CCI
was correlated with both cancer-specific and other-cause mortality [169]. The age-adjusted CCI (Table 5.2) is
the most widely used comorbidity index in cancer for estimating long-term survival and is easily calculated
[170].

Health assessment of oncology patients must be supplemented by measuring their activity level. Extermann
et al. have shown that there is no correlation between morbidity and competitive activity level [171]. The
Eastern Cooperative Oncology Group (ECOG) performance status (PS) scores and Karnofsky index have been
validated to measure patient activity [172]. Performance score is correlated with patient OS after RC [167] and
palliative chemotherapy [173-175].

Patients who have screened positive for frailty or cognitive impairment benefit from an assessment by
a geriatrician. This allows identification of geriatric syndromes and any scope for optimisation. The most
complete protocol is the Comprehensive Geriatric Assessment (CGA) [176] which is useful in the care of cancer
patients [177]. In BC, the CGA has been used to adapt gemcitabine chemotherapy in previously untreated older
patients with advanced BC [178].

22 LIMITED UPDATE MARCH 2021


Table 5.2: Calculation of the Charlson Comorbidity Index

Number of points Conditions


1 50–60 years
Myocardial infarction
Heart failure
Peripheral vascular insufficiency
Cerebrovascular disease
Dementia
Chronic lung disease
Connective tissue disease
Ulcer disease
Mild liver disease
Diabetes
2 61–70 years
Hemiplegia
Moderate to severe kidney disease
Diabetes with organ damage
Tumours of all origins
3 71–80 years
Moderate to severe liver disease
4 81–z years
5 > 90 years
6 Metastatic solid tumours
AIDS

Interpretation:
1. Calculate Charlson Comorbidity Score or Index = i
a. Add comorbidity score to age score
b. Total denoted as ‘i’ in the Charlson Probability calculation (see below).
i = sum of comorbidity score to age score
2. Calculate Charlson Probability (10-year mortality = Y)
a. Calculate Y = 10(i x 0.9)
b. Calculate Z = 0.983Y (where Z is the 10-year survival)

5.3.3 Summary of evidence and guidelines for comorbidity scales

Summary of evidence LE
Chronological age is of limited relevance. 3
It is important to screen for frailty and cognitive impairment and provide a Comprehensive Geriatric 3
Assessment (CGA) where optimisation is needed.

Recommendations Strength rating


Base the decision on bladder-sparing treatment or radical cystectomy in older/frail patients Strong
with invasive bladder cancer on tumour stage and frailty.
Assess comorbidity by a validated score, such as the Charlson Comorbidity Index. Strong
The American Society of Anesthesiologists score should not be used in this setting (see
Section 5.3.2).

LIMITED UPDATE MARCH 2021 23


6. MARKERS
6.1 Introduction
Both patient and tumour characteristics guide treatment decisions and prognosis of patients with MIBC.

6.2 Prognostic markers


6.2.1 Histopathological and clinical markers
The most important histopathological prognostic variables after RC and LN dissection are tumour stage and
LN status [179]. In addition, other histopathological parameters of the RC specimen have been associated with
prognosis.
The value of lymphovascular invasion was reported in a systematic review and meta-analysis
including 78,000 patients from 65 studies treated with RC for BC [180]. Lymphovascular invasion was present
in 35% of the patients and correlated with a 1.5-fold higher risk of recurrence and CSM, independent of
pathological stage and peri-operative chemotherapy. This correlation was even stronger in those patients with
node-negative disease [181].
In a systematic review and meta-analysis including 23 studies and over 20,000 patients, the presence
of concomitant CIS in the RC specimen was associated with a higher odds ratio (OR) of ureteral involvement
(pooled OR: 4.51, 2.59–7.84). Concomitant CIS was not independently associated with OS, recurrence-free
survival (RFS) and DSS in all patients, but in patients with organ-confined disease concomitant CIS was
associated with worse RFS (pooled HR: 1.57, 1.12–2.21) and CSM (pooled HR: 1.51, 1.001–2.280) [181].

Tumour location has been associated with prognosis. Tumours located at the bladder neck or trigone of the
bladder appear to have an increased likelihood of nodal metastasis (OR: 1.83, 95% CI: 1.11–2.99) and have
been associated with decreased survival [179, 182-184].
Prostatic urethral involvement at the time of RC was also found to be associated with worse survival
outcomes. In a series of 995 patients, prostatic involvement was recorded in 31% of patients. The 5-year CSS
in patients with CIS of the prostatic urethra was 40%, whilst the prognosis of patients with UC invading the
prostatic stroma was worse with a 5-year CSS of only 12% [185].
Neutrophil-to-lymphocyte ratio (NLR) has emerged as a prognostic factor in UUT tumours [1] and
other non-urological malignancies. In a pooled analysis of 21 studies analysing the prognostic role of NLR in
BC, the authors correlated elevated pre-treatment NLR with OS, RFS and disease-free survival (DFS) in both
localised and metastatic disease [186]. In contrast, a secondary analysis of the Southwest Oncology Group
(SWOG) 8710 trial, a randomised phase III trial assessing cystectomy ± NAC in patients with MIBC, suggests
that NLR is neither a prognostic nor a predictive biomarker for OS in MIBCs [187].

In patients with LN-positive disease, the American Joint Committee on Cancer (AJCC)-TNM staging system
provides 3 subcategories. In addition, several other prognostic LN-related parameters have been reported.
These include, but are not limited to, the number of positive LNs, the number of LNs removed, LN density
(the ratio of positive LNs to the number of LNs removed) and extranodal extension. In a systematic review
and meta-analysis, it was reported that LN density was independently associated with OS (HR: 1.45, 95%,
CI: 1.11–1.90) [188]. It has been suggested that LN density outperforms the AJCC-TNM staging system for
LN-positive disease in terms of prognostic value [189]. However, in spite of these studies supporting the
use of LN density, LN density relies on the number of LNs removed which, in turn, is subject to surgical and
pathological factors. This makes the concept of LN density difficult to apply uniformly [190].
Two studies investigated whether any of the reported LN-related parameters may be superior to the
routinely used AJCC-TNM staging system [190, 191]. Whilst the conclusion was that the AJCC-TNM staging
system for LN status did not perform well, none of the other tested variables outperformed the AJCC system.

6.2.2 Molecular markers


6.2.2.1 Molecular subtypes based on the Cancer Genome Atlas cohort
The updated Cancer Genome Atlas (TCGA) reported on 412 MIBCs and identified two main groups; luminal
and basal-squamous - consisting of five mRNA expression-based molecular subtypes including luminal-
papillary, luminal-infiltrated, luminal; basal-squamous; and neuronal, a subtype associated with poor survival in
which the majority of tumours do not have small cell or neuroendocrine histology. Each subtype is associated
with distinct mutational profiles, histopathological features and prognostic and treatment implications [192].
The basal-squamous subtype is characterised by expression of basal keratin markers, immune
infiltrates and is felt to be chemo sensitive. The different luminal subtypes are characterised by fibroblast
growth factor receptor 3 (FGFR3) alterations (luminal-papillary), epithelial-mesenchymal transition (EMT)
markers (luminal-infiltrated) and may be associated with chemotherapy resistance [65, 66, 192, 193].

24 LIMITED UPDATE MARCH 2021


In 2019, a consensus on molecular subtype classification was reported [194]. The authors analysed 1,750
MIBC transcriptomic profiles from 18 datasets and identified six MIBC molecular classes that reconcile
all previously published classification schemes. The molecular subgroup classes include luminal papillary
(LumP), luminal non-specified (LumNS), luminal unstable (LumU), stroma-rich, basal/squamous (Ba/Sq), and
neuroendocrine-like (NE-like). Each class has distinct differentiation patterns, oncogenic mechanisms, tumour
micro-environments and histological and clinical associations. However, the authors stressed that consensus
was reached for biological rather than clinical classes. Therefore, at this time, the classification should be
considered as a research tool for retrospective and prospective studies until future studies establish how these
molecular subgroups can be used best in a clinical setting.
Molecular classification of MIBC is still evolving and treatment tailored to molecular subtype is
not a standard yet. A novel 12-gene signature derived from patients in the TCGA utilising published gene
signatures has been developed and externally validated to predict OS in MIBC [195]. Interestingly, a recently
published analysis of molecular subtyping in MIBC demonstrated that although molecular subtypes reflect the
heterogeneity of bladder tumours and are associated with tumour grade, clinical parameters outperformed
subtypes for predicting outcome [196]. In the coming years, new insights into BC carcinogenesis may change
our management of the disease and our ability to better predict outcomes.

6.3 Predictive markers


6.3.1 Clinical and histopathological markers
Based on retrospective data only, patients with secondary MIBC have a worse response to NAC compared
to patients with primary MIBC [197]. Pietzak et al. retrospectively analysed clinico-pathologic outcomes
comparing 245 patients with clinical T2–4a N0M0 primary MIBC and 43 patients with secondary MIBC
treated with NAC and RC. They found that patients with secondary MIBC had lower pathologic response
rates following NAC than those with primary MIBC (univariable 26% vs. 45%, multivariable OR: 0.4 [95% CI:
0.18–0.84, p = 0.02]). They also found that MIBC patients progressing after NAC had worse CSS as compared
to patients treated with cystectomy alone (p = 0.002).
Variant histologies and non-UC have also been linked to worse outcomes after NAC, but there is, as
yet, insufficient data to conclude that they can be considered as predictive markers [198].

6.3.2 Molecular markers


Several predictive biomarkers have been investigated such as serum vascular endothelial growth factor [199],
circulating tumour cells as well as defects in DNA damage repair (DDR) genes including ERCC2, ATM, RB1 and
FANCC that may predict response to cisplatin-based NAC [200, 201]. More recently, alterations in FGFR3 including
both mutations and gene fusions have been shown to be associated with response to FGFR inhibitors [202, 203].

More recent efforts have focused on markers for predicting response to immune checkpoint inhibition.
Programmed death-ligand 1 (PD-L1) expression by immunohistochemistry has been evaluated in several studies
with mixed results which may in part be related to the use of different antibodies and various scoring systems
evaluating different compartments i.e., tumour cells, immune cells, or both. The major limitation of PD-L1
staining relates to the significant proportion of PD-L1-negative patients that respond to immune checkpoint
blockade. For example, in the IMvigor 210 phase II study of atezolizumab in patients with advanced/metastatic
UC who progressed after platinum-based chemotherapy, responses were seen in 18% of patients with low/
no PD-L1 expression [204]. At present, the only indication for PD-L1 testing relates to the use of immune
checkpoint inhibitors as monotherapy in patients with locally advanced or metastatic UC unfit for cisplatin-
containing chemotherapy who have not received prior therapy. In this setting, pembrolizumab or atezolizumab
should only be used in patients unfit for cisplatin-containing chemotherapy whose tumours overexpress PD-L1
(i.e., Combined Positive Score [CPS] ≥ 10 using the 22C3 assay for pembrolizumab and tumour-infiltrating
immune cells [IC] covering ≥ 5% of the tumour area using the SP142 assay for atezolizumab) [205].

Urothelial cancer is associated with a high tumour mutational burden (TMB) [206]. Both predicted neoantigen
burden and tumour mutational burden have been associated with response to immune checkpoint blockade
in several malignancies. High TMB has been associated with response to immune checkpoint inhibitors
in metastatic BC [204, 207]. Conflicting results have been seen in studies evaluating immune checkpoint
inhibitors in the neoadjuvant setting with the Pembrolizumab as Neoadjuvant Therapy Before Radical
Cystectomy in Patients With Muscle-Invasive Urothelial Bladder Carcinoma (PURE)-01 study utilising
pembrolizumab demonstrating an association of high TMB with response while there was no association with
atezolizumab in the Phase II study investigating the safety and efficacy of neoadjuvant atezolizumab in MIBC
(ABACUS) [208, 209].

LIMITED UPDATE MARCH 2021 25


Other markers that have been evaluated in predicting response to immune checkpoint inhibitors include
molecular subtypes as discussed earlier, CD8 expression by immunohistochemistry and other immune gene
cell signatures. Recent work has focused on the importance of stroma including the role of TGFß in predicting
response to immune checkpoint blockade [210, 211]. Although promising, there are currently no validated
predictive molecular markers that are routinely used in clinical practice. Further validation studies are awaited.

6.4 Conclusion
The updated Cancer Genome Atlas and other efforts have refined our understanding of the molecular
underpinnings of bladder cancer biology. Molecular subtypes, immune gene signatures as well as stromal
signatures may ultimately have an important role in predicting response to immunotherapy. Although PD-L1
expression by immunohistochemistry and TMB have demonstrated predictive value in certain settings,
additional studies are needed. Prospectively validated prognostic and predictive molecular biomarkers will
present valuable adjuncts to clinical and pathological data, but large phase III randomised controlled trials
(RCTs) with long-term follow-up will be needed to clarify the many questions remaining.

6.5 Summary of evidence and recommendations for urothelial markers

Summary of evidence LE
There is insufficient evidence to use TMB, molecular subtypes, immune or other gene expression -
signatures for the management of patients with urothelial cancer.

Recommendations Strength rating


Evaluate PD-L1 expression (by immunohistochemistry) to determine the potential for use of Weak
pembrolizumab or atezolizumab in previously untreated patients with locally advanced or
metastatic urothelial cancer who are unfit for cisplatin-based chemotherapy.
Evaluate for FGFR2/3 genetic alterations for the potential use of erdafitinib in patients Weak
with locally advanced or metastatic urothelial carcinoma who have progressed following
platinum-containing chemotherapy (including within 12 months of neoadjuvant or adjuvant
platinum-containing chemotherapy).

7. DISEASE MANAGEMENT
7.1 Neoadjuvant therapy
7.1.1 Introduction
The standard treatment for patients with urothelial MIBC and MIBC with variant histologies is RC. However,
RC only provides 5-year survival in about 50% of patients [212-216]. To improve these results in patients with
cN0M0 disease, cisplatin-based NAC has been used since the 1980s [212-218].

7.1.2 Role of cisplatin-based chemotherapy


There are theoretical advantages and disadvantages of administering chemotherapy before planned definitive
surgery to patients with resectable muscle-invasive UC of the bladder and cN0M0 disease:
• Chemotherapy is delivered at the earliest time-point, when the burden of micrometastatic disease is
expected to be low.
• Potential reflection of in-vivo chemosensitivity.
• Tolerability of chemotherapy and patient compliance are expected to be better pre-cystectomy.
• Patients may respond to NAC and have a favourable pathological response as determined mainly by
achieving ypT0, ≤ ypT1, ypN0 and negative surgical margins.
• Delayed cystectomy might compromise the outcome in patients not sensitive to chemotherapy [219-
221]. However, there are no prospective trials indicating that delayed surgery due to NAC has a negative
impact on survival. In the recently reported French Genito-Urinary Group and the French Association of
Urology (GETUG/AFU) V05 Randomized Phase III VESPER trial, comparing gemcitabine/cisplatin (GC) vs.
high-dose-intensity methotrexate, vinblastine, doxorubicine and cisplatin (HD-MVAC) in the peri-operative
setting, approximately 90% of patients proceeded to surgery after neoadjuvant dose-dense MVAC
(ddMVAC) or GC (median delay of surgery was 48 days for GC and 51 days for ddMVAC) [222].

26 LIMITED UPDATE MARCH 2021


• Neoadjuvant chemotherapy does not seem to affect the outcome of surgical morbidity. In one
randomised trial the same distribution of grade 3–4 post-operative complications was seen in both
treatment arms [223]. In the combined Nordic trials (n = 620), NAC did not have a major adverse effect
on the percentage of performable cystectomies. The cystectomy frequency was 86% in the experimental
arm and 87% in the control arm with 71% of patients receiving all three chemotherapy cycles [224].
• Clinical staging using bimanual palpation, CT or MRI may result in over- and understaging and have a
staging accuracy of only 70% [78]. Overtreatment is a possible negative consequence.
• Gender may have an impact on chemotherapeutic response and oncologic outcomes [225, 226].
• Neoadjuvant chemotherapy should only be used in patients eligible for cisplatin-combination
chemotherapy; other combinations (or monotherapies) are inferior in metastatic BC and have not been
fully tested in a neoadjuvant setting [223, 227-235].

7.1.2.1 Summary of available data


Several randomised phase III trials addressed the potential survival benefit of NAC administration [223, 227-
232, 236-240]. The main differences in trial designs were the type of chemotherapy (i.e., single-agent cisplatin
or combination chemotherapy) and the number of cycles provided. Patients had to be fit for cisplatin. Since
these studies differed considerably for patient numbers, patient characteristics (e.g., clinical T-stages included)
and the type of definitive treatment offered (cystectomy and/or RT), pooling of results was not possible.
Three meta-analyses were undertaken to establish if NAC prolongs survival [233-235]. In a
meta-analysis, published in 2005 [235] with updated patient data from 11 randomised trials (n = 3,005),
a significant survival benefit was shown in favour of NAC. The most recent meta-analysis included four
additional randomised trials, and used the updated results from the Nordic I, Nordic II, and BA06 30894 trials
including data from 427 new patients and updated information from 1,596 patients. The results of this analysis
confirmed the previously published data and showed an 8% absolute improvement in survival at five years
with a number needed-to-treat of 12.5 [241]. Only cisplatin-combination chemotherapy with at least one
additional chemotherapeutic agent resulted in a meaningful therapeutic benefit [233, 235]; the regimens tested
were methotrexate, vinblastine, adriamycin (epirubicin) plus cisplatin (MVA(E)C), cisplatin, methotrexate plus
vinblastine (CMV), cisplatin plus methotrexate (CM), cisplatin plus adriamycin and cisplatin plus 5-fluorouracil
(5-FU) [242].
The updated analysis of a large phase III RCT [227] with a median follow-up of eight years
confirmed previous results and provided additional findings:
• 16% reduction in mortality risk;
• improvement in 10-year survival from 30% to 36% with neoadjuvant CMV;
• benefit with regard to distant metastases;
• the addition of neoadjuvant CMV provided no benefit for locoregional control and locoregional DFS,
independent of the definitive treatment.

More modern chemotherapeutic regimens such as GC have shown similar pT0/pT1 rates as methotrexate,
vinblastine, adriamycin plus cisplatin in retrospective series and pooled data analyses [242-245]. Modified
ddMVAC was tested in two small single-arm phase II studies demonstrating high rates of pathologic complete
remission [246, 247]. Moreover, a large cross-sectional analysis showed higher rates of down-staging and
pathological complete response for ddMVAC [248]. The recently reported results from the GETUG/AFU V05
VESPER randomised trial of ddMVAC vs. gemcitabine and cisplatin as peri-operative chemotherapy in patients
with MIBC demonstrated similar pathologic response rates (ypT0N0) in patients treated with neoadjuvant
ddMVAC or GC of 42% and 36%, respectively (p = 0.2). An organ-confined status (< ypT3pN0) was obtained
in 154 (77%) and 124 (63%) patients, respectively (p = 0.001). Dose-dense MVAC was associated with more
severe asthenia and gastrointestinal side effects than GC. Although a higher local control rate (complete
pathological response, tumour down-staging, or organ confined) was observed in the ddMVAC arm (p = 0.021),
the primary endpoint of PFS at three years is not yet mature [222]. Another dose-dense regimen using cisplatin/
gemcitabine was reported in two small phase II trials [249, 250]. While pathological response rates (< pT2)
in the range of 45%–57% were achieved, one trial had to be closed prematurely due to high rates of severe
vascular events [249]. This approach is therefore not recommended outside of clinical trials.

As an alternative to the standard dose of cisplatin-based NAC with 70 mg/m2 on day 1, split-dose
modifications regimens are often used with 35 mg/m2 on days 1+8 or days 1+2. In a retrospective analysis
the standard schedule was compared to a split-dose schedule in terms of complete and partial pathological
response. A lower number of complete and partial response rates was seen in the split-dose group, but these
results were not statistically significant [251].

LIMITED UPDATE MARCH 2021 27


There seem to be differences in the outcomes of patients treated with NAC for primary or secondary MIBC.
However, in the absence of prospective data, patients with secondary MIBC should be treated similarly to
those presenting with primary MIBC [197].

It is unclear, if patients with non-UC histology will also benefit from NAC. A retrospective analysis demonstrated
that patients with neuroendocrine tumours had improved OS and lower rates of non-organ-confined disease
when receiving NAC. In case of micropapillary differentiation, sarcomatoid differentiation and adenocarcinoma,
lower rates of non-organ confined disease were found, but no statistically significant impact on OS. Patients
with SCC did not benefit from NAC [252].

A retrospective analysis assessed the use of NAC in MIBC based on data from the U.S. National Cancer
Database [253]. Only 19% of all patients received NAC before RC (1,619 of 8,732 patients) and no clear
survival advantage for NAC following propensity score adjustment was found despite efforts to include patients
based on SWOG 8710 study criteria [223]. These results have to be interpreted with caution, especially
since no information was available for the type of NAC applied. Such analyses emphasise the importance of
pragmatically designed studies that reflect real-life practice.

7.1.3 The role of imaging and predictive biomarkers


Data from small imaging studies aiming to identify responders in patients treated with NAC suggest that
response after two cycles of treatment is predictive of outcome. Although mpMRI has the advantage of better
resolution of the bladder wall tissue planes as compared to CT, it is not ready yet for standard patient care.
However, bladder mpMRI may be useful to inform on tumour stage after TURB and response to NAC [254]. So
far PET/CT, MRI or DCE-MRI cannot accurately assess treatment response [255-258]. To identify progression
during NAC imaging is being used in many centres notwithstanding the lack of supporting evidence.
For responders to NAC, especially in those with a complete response (pT0 N0), treatment has a
major positive impact on OS [259, 260]. Therefore, reliable predictive markers to identify patients most likely
to benefit from chemotherapy are needed. Molecular tumour profiling might guide the use of NAC in the future
but, as yet, this is not applicable in routine practice [261-263] (see Chapter 6 - Markers).

7.1.4 Role of neoadjuvant immunotherapy


Inhibition of programmed death receptor-1 (PD-1)/PD-L1 checkpoint has demonstrated significant benefit in
patients with unresectable and metastatic BC in the second-line setting and in platinum-ineligible PD-L1+
patients as first-line treatment using different agents. Checkpoint inhibitors are increasingly tested also in
the neoadjuvant setting; either as monotherapy or in combination with chemotherapy or CTLA-4 checkpoint
inhibition. Data from two phase II trials have been presented with encouraging results [208, 209]. The results
of the phase II trial using the PD-1 inhibitor pembrolizumab reported a complete pathological remission (pT0)
in 42% and pathological response (< pT2) in 54% of patients, whereas in the single-arm phase II trial with
atezolizumab a pathologic complete response rate of 31% was reported. However, immunotherapy is not yet
approved in the neoadjuvant setting.

7.1.5 Summary of evidence and guidelines for neoadjuvant therapy

Summary of evidence LE
Neoadjuvant cisplatin-containing combination chemotherapy improves overall survival (OS) (5–8% at 1a
five years).
Neoadjuvant treatment has a major impact on OS in patients who achieve ypT0 or at least ypT2. 2
Currently immunotherapy with checkpoint inhibitors as monotherapy, or in different combinations, is -
being tested in phase II and III trials. Initial results are promising.
There are still no tools available to select patients who have a higher probability of benefitting from -
NAC. In the future, genetic markers in a personalised medicine setting might facilitate the selection of
patients for NAC and differentiate responders from non-responders.
Neoadjuvant chemotherapy has its limitations regarding patient selection, current development of 3
surgical techniques and current chemotherapy combinations.

28 LIMITED UPDATE MARCH 2021


Recommendations Strength rating
Offer neoadjuvant chemotherapy (NAC) for T2–T4a, cN0 M0 bladder cancer. In this case, Strong
always use cisplatin-based combination therapy.
Do not offer NAC to patients who are ineligible for cisplatin-based combination Strong
chemotherapy.
Only offer neoadjuvant immunotherapy to patients within a clinical trial setting. Strong

7.2 Pre- and post-operative radiotherapy in muscle-invasive bladder cancer


7.2.1 Post-operative radiotherapy
Given the high rates of local-regional failure after RC in patients with locally advanced (pT3–4) BC, estimated
at~30%, as well as the high risk of distant failure and poor survival for these patients, there is an interest
in adjuvant therapies that address both the risk of local and distant disease. Data on adjuvant RT after RC
are limited and further prospective studies are needed, but a more recent phase II trial compared adjuvant
sequential chemotherapy and radiation vs. adjuvant chemotherapy alone in 120 patients with locally advanced
disease and negative margins after RC (with one or more risk factors: ≥ pT3b, grade 3, or node-positive), in a
study population with 53% UC and 47% SCC. Addition of adjuvant RT to chemotherapy alone was associated
with a statistically significant improvement in local relapse-free survival (at 2 years 96% vs. 69% favouring the
addition of RT). Disease-free survival and OS also favoured the addition of RT, but those differences were not
statistically significant and the study was not powered for those endpoints. Late-grade ≥ 3 gastrointestinal
toxicity in the chemoradiation arm was low (7% of patients) [264].

A 2019 systematic review evaluating the efficacy of adjuvant radiation for BC or UTUC found no clear benefit of
adjuvant radiation following radical surgery (e.g., cystectomy), although the combination of adjuvant radiation
with chemotherapy may be beneficial in locally advanced disease [265].

While there are no conclusive data demonstrating improvements in OS it is reasonable to consider adjuvant
radiation in patients with pT3/pT4 pN0–2 urothelial BC following RC, although this approach has been
evaluated in only a limited number of studies. Radiation fields should encompass areas at risk for harbouring
residual microscopic disease based on pathologic findings at surgery and may include cystectomy bed and
pelvic LNs. Doses in the range of 45 to 50.4 Gy may be considered. For patients who have not had prior NAC,
it may be reasonable to sandwich adjuvant radiation between cycles of adjuvant chemotherapy. The safety and
efficacy of concurrent radiosensitising chemotherapy in the adjuvant setting needs further study.

7.2.2 Pre-operative radiotherapy


To date, six RCTs have been published investigating pre-operative RT, although all are from several decades
ago. In the largest trial, pre-operative RT at a dose of 45 Gy was used in patients with muscle-invasive
tumours resulting in a significant increase in pathological complete response (9% to 34%) in favour of pre-
operative RT, which was also a prognostic factor for survival [266]. The OS data were difficult to interpret since
chemotherapy was used in a subset of patients only and more than 50% of patients (241/475) did not receive
the planned treatment and were excluded from the final analyses. Two smaller studies using a dose of 20 Gy
showed only a small survival advantage in ≥ T3 tumours [267, 268]. Two other small trials confirmed down-
staging after pre-operative RT [269, 270].
A meta-analysis of five RCTs showed a difference in 5-year survival (OR: 0.71, 95% CI: 0.48–1.06)
in favour of pre-operative RT [271]. However, the meta-analysis was potentially biased by data from the largest
trial in which patients were not given the planned treatment. When the largest trial was excluded from the
analysis, the OR became 0.94 (95% CI: 0.57–1.55), which was not significant.
A more recent RCT, comparing pre-operative vs. post-operative RT and RC (n = 100), showed
comparable OS, DFS and complication rates [272]. Approximately half of these patients had UC, while the
other half had SCC.

In general, such older data is limited in being able to provide a robust evidence base for modern guideline
recommendations.

LIMITED UPDATE MARCH 2021 29


7.2.3 Summary of evidence and guidelines for pre- and post-operative radiotherapy

Summary of evidence LE
No contemporary data exists to support that pre-operative RT for operable MIBC increases survival. 2a
Pre-operative RT for operable MIBC, using a dose of 45–50 Gy in fractions of 1.8–2 Gy, results in 2
down-staging after 4 to 6 weeks.
Limited high-quality evidence supports the use of pre-operative RT to decrease local recurrence of 3
MIBC after radical cystectomy.
Addition of adjuvant RT to chemotherapy is associated with an improvement in local relapse-free 2a
survival following cystectomy for locally-advanced bladder cancer (pT3b–4, or node-positive).

Recommendations Strength rating


Do not offer pre-operative radiotherapy (RT) for operable muscle-invasive bladder cancer Strong
since it will only result in down-staging, but will not improve survival.
Do not offer pre-operative RT when subsequent radical cystectomy (RC) with urinary Strong
diversion is planned.
Consider offering adjuvant radiation in addition to chemotherapy following RC, based on Weak
pathologic risk (pT3b–4 or positive nodes or positive margins).

7.2.4 EAU-ESMO consensus statements on the management of advanced- and variant bladder
cancer [8, 9]*

Consensus statement
Candidates for curative treatment, such as cystectomy or bladder preservation, should be clinically assessed
by at least an oncologist, a urologist and a neutral HCP such as a specialist nurse.
When assessing patient eligibility for bladder preservation, the likelihood of successful debulking surgery
should be taken into consideration (optimal debulking).
*Only statements which met the a priori consensus threshold across all three stakeholder groups are listed
(defined as ≥ 70% agreement and ≤ 15% disagreement, or vice versa).
HCP = healthcare professional.

7.3 Radical surgery and urinary diversion


7.3.1 Removal of the tumour-bearing bladder
7.3.1.1 Introduction
Radical cystectomy is the standard treatment for localised MIBC in most Western countries [212, 273].
Increased recognition of the central patient role as a healthcare consumer and a greater focus on patients’
QoL contributed to an increasing trend of utilising bladder-preserving treatment modalities, such as radio-
and/or chemotherapy (see Section 7.5). Performance status and life expectancy influence the choice of
primary management as well as the type of urinary diversion with RC being reserved for patients with a longer
life expectancy without concomitant disease and a better PS. The value of assessing overall health before
proceeding with surgery was emphasised in a multivariate analysis [138]. The analysis found an association
between comorbidity and adverse pathological- and survival outcomes following RC [138].
Performance status and comorbidity have a different impact on treatment outcomes and must be
evaluated independently [171].

7.3.1.2 Radical cystectomy: timing


A population-based study utilising the U.S. SEER database analysed patients who underwent a RC between
1992 and 2001 and the authors concluded that a delay of > 12 weeks has a negative impact on outcome
and should be avoided [274]. Moreover, the SEER analysis did not show any significant utilisation and timing
differences between men and women. A 2020 meta-analysis including 19 studies confirmed these findings,
showing that a longer delay of RC after diagnosis (> 3 months) was found to have a detrimental effect on OS
(HR: 134, 95% CI: 1.18–1.53). The authors highlight the lack of standardisation how delays were defined in
the included studies which prohibited defining a clear cut-off time, although most studies used a cut-off of
< 3 months [275]. Overall conclusion was that BC patients scheduled for RC should be treated without delays
to maximise survival.

30 LIMITED UPDATE MARCH 2021


7.3.2 Radical cystectomy: indications
Traditionally, RC was recommended in patients with T2–T4a, N0–Nx, M0 disease [273]. Other indications
include recurrent high-risk non-muscle-invasive tumours, BCG-refractory, BCG-relapsing and BCG-
unresponsive NMIBC (see EAU Guidelines on Non-muscle-invasive Bladder Cancer [2]), as well as extensive
papillary disease that cannot be controlled with TURB and intravesical therapy alone.
Salvage cystectomy is indicated in non-responders to conservative therapy, recurrence after
bladder-sparing treatment, and non-UC. It is also used as a purely palliative intervention, including for fistula
formation, pain and recurrent visible haematuria (see Section 7.4.1 - Palliative cystectomy).

7.3.3 Radical cystectomy: technique and extent


Different approaches have been described to improve voiding and sexual function in patients undergoing RC
for BC. No consensus exists regarding which approach preserves function best. Concern remains regarding
the impact of “sparing-techniques” on oncological outcomes.
To determine the effect of sexual function-preserving cystectomy (SPC) on functional and
oncological outcomes the Panel undertook two systematic reviews addressing sparing techniques in men and
women [276, 277].

In men, standard RC includes removal of the bladder, prostate, seminal vesicles, distal ureters, and regional
LNs. In women, standard RC includes removal of the bladder, the entire urethra and adjacent vagina, uterus,
distal ureters, and regional LNs [278].

7.3.3.1 Radical cystectomy in men


Four main types of sexual-preserving techniques have been described:
1. Prostate sparing cystectomy: part of or the whole prostate is preserved including seminal vesicles, vas
deferens and neurovascular bundles.
2. Capsule sparing cystectomy: the capsule or peripheral part of the prostate is preserved with adenoma
(including prostatic urethra) removed by TURP or en bloc with the bladder. Seminal vesicles, vas deferens
and neurovascular bundles are also preserved.
3. Seminal sparing cystectomy: seminal vesicles, vas deferens and neurovascular bundles are preserved.
4. Nerve-sparing cystectomy: the neurovascular bundles are the only tissue left in place.

Twelve studies recruiting a total of 1,098 patients were identified, including nine comparative studies [279-289]
and three single-arm case series [290-292]. In the majority of cases, an open surgical approach was used and
the urinary diversion of choice was an orthotopic neobladder. Median follow-up was longer than three years in
nine studies, with three studies presenting results with a median follow-up longer than five years.

The majority of the studies included patients who were potent pre-operatively with organ-confined disease
without tumour in the bladder neck and/or prostatic urethra. Prostate cancer was ruled out in all of the SPC
techniques, except in nerve-sparing cystectomy.
Oncological outcomes did not differ between groups in any of the comparative studies that
measured local recurrence, metastatic recurrence, DSS and OS, at a median follow-up of three to five years.
Local recurrence after SPC was commonly defined as any UC recurrence below the iliac bifurcation within the
pelvic soft tissue and ranged from 1.2–61.1% vs. 16–55% in the control group. Metastatic recurrence ranged
from 0–33.3%.
For techniques preserving prostatic tissue (prostate- or capsule-sparing), rates of incidental prostate
cancer in the intervention group ranged from 0–15%. In no case was incidental prostate cancer with ISUP
grade ≥ 4 reported.
Post-operative potency was significantly better in patients who underwent any type of sexual-
preserving technique compared to conventional RC (p < 0.05), ranging from 80–90%, 50–100% and 29–78%
for prostate-, capsule- or nerve-sparing techniques, respectively. Data did not show superiority of any sexual-
preserving technique.
Urinary continence, defined as the use of “no pads” in the majority of studies, ranged from
88–100% (day-time continence) and from 31–96% (night-time continence) in the prostate-sparing cystectomy
patients. No major impact was shown with regard to continence rates for any of the three approaches.

The evidence base suggests that these procedures may yield better sexual outcomes than standard RC
without compromising oncological outcomes. However, the overall quality of the evidence was moderate, and
hence if a sexual-preserving technique is offered, patients must be carefully selected, counselled and closely
monitored.

LIMITED UPDATE MARCH 2021 31


7.3.3.1.1 Summary of evidence and recommendations for sexual-preserving techniques in men

Summary of evidence LE
The majority of patients motivated to preserve their sexual function will benefit from sexual-preserving 2a
techniques.
None of the sexual-preserving techniques (prostate/capsule/seminal/nerve-sparing) have shown to be 3
superior, and no particular technique can be recommended.

Recommendations Strength rating


Do not offer sexual-preserving radical cystectomy to men as standard therapy for muscle- Strong
invasive bladder cancer.
Offer sexual-preserving techniques to men motivated to preserve their sexual function since Strong
the majority will benefit.
Select patients based on: Strong
• organ-confined disease;
• absence of any kind of tumour at the level of the prostate, prostatic urethra or bladder
neck.

7.3.3.2 Radical cystectomy in women


Pelvic floor disorders, sexual and voiding dysfunction in female patients are prevalent after RC [293]. As part
of the pre-operative evaluation a gynaecological history should be obtained and patients should be counselled
on the potential negative impact of RC on sexual function and/or vaginal prolapse. Most importantly, a history
of cervical cancer screening, abnormal vaginal bleeding and a family history of breast and/or ovarian cancer
should be recorded, as well as ruling out possible pelvic organ prolapse. Equally important is screening for
sexual and urinary function and prolapse post-operatively. Better imaging modalities, increased knowledge of
the function of the pelvic structures and improved surgical techniques have enabled less destructive methods
for treating high-risk BC.
Pelvic organ-preserving techniques involve preserving the neurovascular bundle, vagina, uterus,
ovaries or variations of any of the stated techniques. From an oncological point of view, concomitant
malignancy in gynaecological organs is rare and local recurrences reported after RC are infrequent [294,
295]. In premenopausal women, by preserving ovaries, hormonal homeostasis will be preserved, decreasing
risk of cognitive impairment, cardiovascular diseases and loss of bone density. In case of an increased risk
of hereditary breast or ovarian cancer (i.e., BRCA1/2 mutation carriers or patients with Lynch syndrome),
salpingooophorectomy should be advised after childbearing and to all women over 40 years of age [296]. On
the other hand, preservation of the uterus and vagina will provide the necessary support for the neobladder,
thereby reducing the risk of urinary retention. It also helps to avoid post-operative prolapse as removal of
the uterus predisposes to an anterior or posterior vaginal prolapse. In case of an already existing prolapse
of the uterus, either isolated or combined with a vaginal prolapse, removing the uterus will be beneficial. It
is noteworthy that by resecting the vaginal wall, the vagina shortens which could potentially impair sexual
satisfaction and function.
Based on retrospective low quality data only, a systematic review evaluating the advantages and
disadvantages of sexual-function preserving RC and orthotopic neobladder in female patients concluded that
in well-selected patients, sparing female reproductive organs during RC appears to be oncologically safe and
provides improved functional outcomes [277].
Pelvic organ-preserving RC could be considered also in elderly and fragile patients having
abdominal diversions. By reducing excision range, it might be beneficial from the point of reduced operating
time, estimated blood loss and quicker bowel recovery [297].

7.3.3.2.1 Summary of evidence and recommendations for sexual-preserving techniques in women

Summary of evidence LE
Data regarding pelvic organ-preserving RC for female patients remain immature. 3

Recommendations Strength rating


Do not offer pelvic organ-preserving radical cystectomy to women as standard therapy for Strong
muscle-invasive bladder cancer.

32 LIMITED UPDATE MARCH 2021


Offer sexual organ-preserving techniques to women motivated to preserve their sexual Weak
function since the majority will benefit.
Select patients based on: Strong
• absence of tumour in the area to be preserved to avoid positive soft tissue margins;
• absence of pT4 urothelial carcinoma.

7.3.4 Lymphadenectomy: role and extent


Controversies in evaluating the clinical significance of lymphadenectomy (LND) are related to two main aspects
of nodal dissection: therapeutic procedure and/or staging instrument.

Two important autopsy studies have been performed for RC so far. The first study showed that in 215 patients
with MIBC and nodal dissemination, the frequency of metastasis was 92% in regional (perivesical or pelvic),
72% in retroperitoneal, and 35% in abdominal LNs. There was also a significant correlation between nodal
metastases and concomitant distant metastases (p < 0.0001). Approximately 47% of the patients had both
nodal metastases and distant dissemination and only 12% of the patients had nodal dissemination as the sole
metastatic manifestation [298].
The second autopsy study focused on the nodal yield when super-extended pelvic LND was
performed. Substantial inter-individual differences were found with counts ranging from 10 to 53 nodes [299].
These findings demonstrate the limited utility of node count as a surrogate for extent of dissection.
Regional LNs have been shown to consist of all pelvic LNs below the bifurcation of the aorta [300-
304]. Mapping studies also found that skipping lesions at locations above the bifurcation of the aorta without
more distally located LN metastases is rare [304, 305].

The optimal extent of LND has not been established to date. Standard LND in BC patients involves removal of
nodal tissue cranially up to the common iliac bifurcation, with the ureter being the medial border, and including
the internal iliac, presacral, obturator fossa and external iliac nodes [306]. Extended LND includes all LNs in
the region of the aortic bifurcation, and presacral and common iliac vessels medial to the crossing ureters. The
lateral borders are the genitofemoral nerves, caudally the circumflex iliac vein, the lacunar ligament and the LN
of Cloquet, as well as the area described for standard LND [306-310]. A super-extended LND extends cranially
to the level of the inferior mesenteric artery [311, 312].

In order to assess how and if cancer outcome is influenced by the extent of LND in patients with clinical
N0M0 MIBC, a systematic review of the literature was undertaken [313]. Out of 1,692 abstracts retrieved and
assessed, nineteen studies fulfilled the review criteria [306-310, 312, 314-326]. All five studies comparing LND
vs. no LND reported a better oncological outcome for the LND group. Seven out of twelve studies comparing
(super)extended with limited or standard LND reported a beneficial outcome for (super)extended LND in at
least a subset of patients which is in concordance with the findings of several other meta-analyses [327, 328].
No difference in outcome was reported between extended and super-extended LND in the two high-volume-
centre studies identified [312, 324]. The LEA trial, a prospective phase III RCT, including 401 patients with a
median follow-up of 43 months recently reported [329]. Extended LND failed to show a significant advantage
(the trial was designed to show an absolute improvement of 15% in 5-year RFS by extended LND) over limited
LND in RFS, CSS, and OS. Results from another large RCT on the therapeutic impact of the extent of LND are
expected shortly.

It has been suggested that PFS as well as OS might be correlated with the number of LNs removed during
surgery. Although there are no data from RCTs on the minimum number of LNs that should be removed,
survival rates increase with the number of dissected LNs [330]. In retrospective studies removal of at least
ten LNs has been postulated as sufficient for evaluation of LN status, as well as being beneficial for OS [331].
Submitting separate nodal packets instead of en bloc has shown significant increased total LN yield, but did
not result in an increased number of positive LNs, making LN density an inaccurate prognosticator [332]. In
conclusion, extended LND might have a therapeutic benefit compared to less extensive LND, but due to study
bias no firm conclusions can be drawn [145, 313, 333].

7.3.5 Laparoscopic/robotic-assisted laparoscopic cystectomy


A number of recent systematic reviews comparing open RC (ORC) and robot-assisted RC (RARC) reach similar
conclusions; RARC has an approximately one-day shorter length of hospital stay (LOS) and less blood loss, but
a longer operative time. Complication rates seem similar for both approaches but all published reviews suffer
from low quality data.
In minimally-invasive cystectomy, with increasing age, LOS is markedly shorter; up to 2.56 days in
patients over 80 years old [334].

LIMITED UPDATE MARCH 2021 33


Although the low level of evidence of the studies included in these reviews remains a major limitation, a recent
Cochrane review incorporating data from all five published RCTs corroborates most findings [335]. Time to
recurrence, positive surgical margin rates, grade 3–5 complications and QoL were comparable for RARC and
ORC, whilst transfusion rate was likely lower after RARC. For other endpoints outcomes were uncertain due to
study limitations.

The Pasadena Consensus Panel (a group of experts on RC, LND and urinary reconstruction) reached similar
conclusions [336]. Additionally, they reported that RARC was associated with increased costs, although
compared to laparoscopic RC (LRC) there are ergonomic advantages for the surgeon. For both techniques,
surgeons’ experience and institutional volume strongly predicted outcome. According to the literature,
proficiency is reached after 20–250 cases. However, after statistical modelling, the Pasadena Consensus Panel
suggested 30 cases but they also concluded that challenging patients (high BMI, post chemotherapy or RT,
pelvic surgery, T4 or bulky tumours or positive nodes) should be performed by experienced robotic surgeons
only. Safety of RC after RT was confirmed by a small retrospective study (n = 46) [337]. In experienced hands
the percentage of 90-day (major) complications after robotic cystectomy was independent of previous RT. A
recent population-based study addressed benign uretero-enteric strictures, which appeared 5% higher after
robotic - as compared to open surgery (15% vs. 9.5% at 12 months, p = 0.01) [338].

Positive surgical margins, as a surrogate for oncological outcome, are comparable between RARC and ORC,
although with low certainty [335]. Recurrence-free survival, CSS and OS have been documented as similar
in all RCTs including the largest RAZOR (Robot-assisted radical cystectomy versus open radical cystectomy
in patients with bladder cancer) trial (n = 302) [339]. Age over 70, poor PS and major complications were
significant predictors of 36-month PFS whilst stage and positive margins were significant predictors of
recurrence, PFS and OS. The surgical approach was not a significant predictor of any outcome. A larger
(n = 595) single-centre study with a median follow-up of over five years also found comparable recurrence
and survival data, including atypical recurrences (defined as one or a combination of the following: port-
site metastasis or peritoneal carcinomatosis) [340]. However, recently, port-site metastases and atypical
recurrences were reviewed by Mantica et al. [341]. Based on 31 studies and 6,720 evaluable patients, 105
patients (1.63%) were identified with an atypical recurrence, of which 63 (60%) were peritoneal carcinomatosis
and 11 (10.5%) port-site metastases. The authors acknowledge, however, that these results may be linked to
publication bias and retrospective study design of the included studies. Wei et al. detected residual cancer
cells in pelvic washing specimens during or after, but not before, RARC in approximately half of the patients
(9/17), which was associated with aggressive variant histology and cancer recurrence. These findings need
confirmation in larger studies [342].
The largest RCT to date, the RAZOR trial, supports all of the above findings showing RARC to be
non-inferior to ORC in terms of 2-year PFS (72.3% vs. 71.6%), adverse events (67% vs. 69%) and QoL [343].

Most reviewed series, including the RAZOR trial, offer extracorporeal reconstruction. Hussein et al.
retrospectively compared extracorporeal reconstruction (n = 1,031) to intracorporeal reconstruction (n = 1,094);
the latter was associated with a shorter operative time and fewer blood transfusions but more high-grade
complications, which, again, decreased over time [344]. A recent retrospective report from a high-volume
centre found less (major) complications after intracorporeal reconstruction (n = 301) as compared to
extracorporeal reconstruction (n = 375) and open RC (n = 272) [345]. It is important to note that, although an
intracorporeal neobladder is a very complex robotic procedure [346], the choice for neobladder or cutaneous
diversion should not depend on the surgical approach.

7.3.5.1 Laparoscopic radical cystectomy versus robot-assisted radical cystectomy


For LRC a review including sixteen studies came to similar conclusions as described for RARC [346]. As
compared to ORC, LRC had a significantly longer operative time, fewer overall complications, less blood
transfusions and analgesic use, less blood loss and a shorter LOS. However, the review was limited by the
inherent limitations of the included studies. Although this review also showed better oncological outcomes,
these appeared comparable to ORC series in a large LRC multicentre study [347].
The CORAL study was a small single-centre RCT comparing open (n = 20) vs. robotic (n = 20)
vs. laparoscopic (n = 19) RC [348, 349]. The 30-day complication rate was significantly higher in the open
arm (70%) compared to the laparoscopic arm (26%). There was no difference between the 90-day Clavien
complication rates in the three study arms. Limitations of this study include the small sample size, three
different although experienced surgeons, and cross over between arms.

34 LIMITED UPDATE MARCH 2021


7.3.5.2 Summary of evidence and guidelines for laparoscopic/robotic-assisted laparoscopic cystectomy

Summary of evidence LE
Robot-assisted radical cystectomy has longer operative time (1–1.5 hours) and major costs, but 1
shorter length of hospital stay (1–1.5 days) and less blood loss compared to ORC.
Robotic cystectomy and open cystectomy may result in similar rates of (major) complications. 2
Most endpoints, if reported, including intermediate-term oncological endpoint and QoL, are not 2
different between RARC and ORC.
Surgeons experience and institutional volume are considered the key factor for outcome of both 2
RARC and ORC, not the technique.

Recommendations Strength rating


Inform the patient of the advantages and disadvantages of open radical cystectomy (ORC) Strong
and robot-assisted radical cystectomy (RARC) to allow selection of the proper procedure.
Select experienced centres, not specific techniques, both for RARC and ORC. Strong

7.3.6 Urinary diversion after radical cystectomy


From an anatomical standpoint, three alternatives are currently used after cystectomy:
• abdominal diversion, such as an uretero-cutaneostomy, ileal or colonic conduit, and various forms of a
continent pouch (infrequently used);
• urethral diversion, which includes various forms of gastrointestinal pouches attached to the urethra as a
continent, orthotopic urinary diversion (neobladder, orthotopic bladder substitution);
• rectosigmoid diversions, such as uretero-(ileo-)rectostomy (infrequently used).

Different types of segments of the intestinal tract have been used to reconstruct the urinary tract, including the
stomach, ileum, colon and appendix [350].
Several studies have compared certain aspects of health-related quality of life (HRQoL) such as
sexual function, urinary continence and body image in patient cohorts with different types of urinary diversion
[351]. However, further research evaluating the impact of pre-operative tumour stage, functional- and socio-
economic status, and time interval to primary surgery are needed.

7.3.6.1 Patient selection and preparations for surgery


In consultation with the patient, both an orthotopic neobladder and ileal conduit should be considered in case
reconstructive surgery exposes the patient to excessive risk (as determined by comorbidity and age).
Ensuring that patients make a well-informed decision about the type of urinary diversion is
associated with less decision regret post-operatively, independent of the method selected [352].
Diagnosis of an invasive urethral tumour prior to cystectomy leads to urethrectomy which could be a
contraindication for a neobladder reconstruction. If indicated; in males, in case of CIS and extension of tumour
in the prostatic urethra, urethral frozen section has to be performed on the cystoprostatectomy specimen just
under the verumontanum and on the inferior limits of the bladder neck; in females a urethral frozen section has
to be taken just below the bladder neck.

Non-muscle-invasive BC in prostatic urethra or bladder neck biopsies does not necessarily preclude orthotopic
neobladder substitution, provided that patients undergo regular follow-up cystoscopy and urinary cytology [353].
In the presence of positive LNs, orthotopic neobladder can nevertheless be considered in case of
N1 involvement (metastasis in a single node in the true pelvis) but not in N2 or N3 tumours [354].
Oncological results after orthotopic neobladder substitution or conduit diversion are similar in terms
of local or distant metastasis recurrence, but secondary urethral tumours seem less common in patients with a
neobladder compared to those with conduits or continent cutaneous diversions [355].

For cystectomy, general preparations are necessary as for any other major pelvic and abdominal surgery. If the
urinary diversion is constructed from gastrointestinal segments, the length or size of the respective segments
and their pathophysiology when storing urine must be considered [356]. Despite the necessary interruption
and re-anastomosis of the bowel, formal bowel preparation may not be necessary [357]. Bowel recovery
time can be reduced by the use of early mobilisation and early oralisation, gastrointestinal stimulation with
metoclopramide and chewing gum [358]. Patients treated according to the “fast tract”/ERAS (Early Recovery
After Surgery) protocol have shown to score better on the emotional and physical functioning scores and suffer
less from wound healing disorders, fever and thrombosis [359].

LIMITED UPDATE MARCH 2021 35


A cornerstone of the ERAS protocol is post-operative pain management, which involves significantly
reducing the use of opioids; offering opioids mainly as breakthrough pain medication. Instead of patient-
controlled analgesia and epidural opioids, most patients receive high-dose acetaminophen and/or ketorolac,
starting intra-operatively. Patients on ERAS experience more pain as compared to patients on a traditional
protocol (Visual Analogue Scale 3.1 vs. 1.1, p < 0.001), but post-operative ileus decreased from 22% to 7.3%
(p = 0.003) [360].
A multicentre randomised placebo-controlled trial showed that patients receiving alvimopan, a
peripherally acting μ-opioid receptor antagonist, had quicker bowel recovery compared to patients receiving
placebo [361]. However, this drug is, as yet, not approved in Europe.
Venous thromboembolism (VTE) prophylaxis may be implemented as part of an ERAS protocol. A
single-centre non-randomised study showed a significant lower 30-day VTE incidence rate in patients treated for
28 days with enoxaparin compared to patients without prophylaxis [362]. Data from the Ontario Cancer Registry
including 4,205 cystectomy patients of whom 1,084 received NAC showed that VTE rates are higher in patients
treated with NAC as compared to patients treated with cystectomy only (12% vs. 8%, p = 0.002) [363, 364].

Patients undergoing continent urinary diversion must be motivated to learn about their diversion and to be
manually skilful in manipulating their diversion. Contraindications to more complex forms of urinary diversion
include:
• debilitating neurological and psychiatric illnesses;
• limited life expectancy;
• impaired liver or renal function;
• urothelial carcinoma positive surgical margins.

Relative contraindications specific for an orthotopic neobladder are high-dose pre-operative RT, complex
urethral stricture disease and severe urethral sphincter-related incontinence [365].

7.3.6.2 Different types of urinary diversion


Radical cystectomy and urinary diversion are the two steps of one operation. However, the literature uniformly
reports complications of RC while ignoring the fact that most complications are diversion related [366]. Age
alone is not a criterion for offering continent diversion [365, 367]. Comorbidity, cardiac- and pulmonary function
and cognitive function are all important factors that should be considered, along with the patient’s social
support and preference.
Age > 80 years is often considered to be the threshold after which neobladder reconstruction
is not recommended. However, there is no exact age for a strict contraindication. In most large series from
experienced centres, the rate of orthotopic bladder substitution after cystectomy for bladder tumour is up
to 80% in men and 50% in women [368-371]. Nevertheless, no RCTs comparing conduit diversion with
neobladder or continent cutaneous diversion have been performed.
A retrospective study including 1,383 patients showed that the risk of a decline in estimated
glomerular filtration rate (eGFR) did not significantly differ after ileal conduit vs. neobladder in patients with pre-
operative chronic kidney disease 2 (eGFR 60–89 mL/min/1.73 m2) or 3a (eGFR 45–59 mL/min/1.73 m2) [372].
Only age and anastomotic strictures were found to be associated with a decline in eGFR.

7.3.6.2.1 Uretero-cutaneostomy
Ureteral diversion to the abdominal wall is the simplest form of cutaneous diversion. Operating time,
complication rate, stay at intensive care and length of hospital stay are lower in patients treated with uretero-
cutaneostomy as compared to ileal conduit [373]. Therefore, in frail patients and/or in those with a solitary
kidney who need a supravesical diversion, uretero-cutaneostomy is the preferred procedure [374, 375]. Quality
of life, which was assessed using the Bladder Cancer Index (BCI), showed equal urinary bother and function
for patients treated with ileal conduit and uretero-cutaneostomy [373]. However, others have demonstrated
that in carefully selected elderly patients, all other forms of wet and dry urinary diversions, including orthotopic
bladder substitutions, are possible [376].

Technically, in case patients have both kidneys; either one ureter, to which the other shorter one is attached
end-to-side, is connected to the skin (trans-uretero-cutaneostomy) or both ureters are directly anastomosed
to the abdominal wall creating a stoma. Due to the smaller diameter of the ureters, stoma stenosis has been
observed more frequently for this technique as compared to using small or large bowel to create an intestinal
stoma [374].
In a retrospective multicentre study peri-operative morbidity was evaluated for urinary diversion
using bowel as compared to uretero-cutaneostomy. Patients selected for a uretero-cutaneostomy were older
and had a higher ASA score, while their mean Charlson score was lower (4.2 vs. 5.6, p < 0.001) [377].

36 LIMITED UPDATE MARCH 2021


Despite the limited comparative data available, it must be taken into consideration that older
data and clinical experience suggest ureter stenosis at the skin level and ascending UTI are more frequent
complications in uretero-cutaneostomy compared to an ileal conduit diversion. In a retrospective study
comparing various forms of intestinal diversion, ileal conduits had fewer late complications than continent
abdominal pouches or orthotopic neobladders [378].

7.3.6.2.2 Ileal conduit


The ileal conduit is still an established option with well-known/predictable results. However, up to 48% of
patients develop early complications including UTIs, pyelonephritis, ureteroileal leakage and stenosis [378].
The main complications in long-term follow-up studies are stomal complications in up to 24% of patients and
functional and/or morphological changes of the UUT in up to 30% [378-380]. An increase in complications
was seen with longer follow-up in the Berne series of 131 patients who were followed for a minimum of five
years (median follow-up 98 months) [381]; the rate of complications increased from 45% at five years to 94%
in those surviving > 15 years. In the latter group, 50% of patients developed UUT changes and 38% developed
urolithiasis.

7.3.6.2.3 Orthotopic neobladder


An orthotopic bladder substitution to the urethra is now commonly used both in men and women.
Contemporary reports document the safety and long-term reliability of this procedure. In several large centres,
this has become the diversion of choice for most patients undergoing cystectomy [213, 273, 365]. However, in
elderly patients (> 80 years) it is rarely performed even in high-volume expert centres [382, 383].
The terminal ileum is the gastrointestinal segment most often used for bladder substitution.
There is less experience with the ascending colon, including the caecum, and the sigmoid [273]. Emptying
of the reservoir anastomosed to the urethra requires abdominal straining, intestinal peristalsis, and sphincter
relaxation. Early and late morbidity in up to 22% of patients is reported [384, 385]. In two studies of 1,054
and 1,300 patients [365, 386], long-term complications included diurnal (8–10%) and nocturnal (20–30%)
incontinence, uretero-intestinal stenosis (3–18%), metabolic disorders, and vitamin B12 deficiency. A study
comparing cancer control and patterns of disease recurrence in patients with neobladder and ileal conduit
showed no difference in CSS between the two groups when adjusting for pathological stage [387]. Urethral
recurrence in neobladder patients seems rare (1.5–7% in both male and female patients) [365, 388]. These
results indicate that neobladder in male and female patients does not compromise the oncological outcome of
cystectomy. It remains debatable whether patient’s QoL for neobladder is better compared to non-continent
urinary diversion [389, 390].

Continent cutaneous urinary diversion (a low-pressure detubularised ileal reservoir for self-catheterisation)
and uretero-rectosigmoidostomy are rarely used techniques nowadays, due to their high complication rates,
including stomal stenosis, incontinence in the continent cutaneous diversion, UUT infections and stone
formation in case of uretero-rectosigmoidostomy [391].

Various forms of UUT reflux protection, including a simple isoperistaltic tunnel, ileal intussusception, tapered
ileal prolongation implanted subserosally, and direct (sub)mucosal or subserosal ureteral implantation, have
been described [385, 392]. According to the long-term results, the UUT is protected sufficiently by either
method.

A detailed investigation of the bladder neck prior to RC is important for women who are scheduled for an
orthotopic bladder substitute [393]. In women undergoing RC the rate of concomitant urethral malignancy has
been reported to range from 12–16% [394]. Localisation of the primary tumour at the bladder neck correlated
strongly with concomitant urethral malignancy. In addition, tumour involving the bladder neck and urethra
tended to be associated with a higher risk of advanced stage and nodal involvement [395].

Currently, it is not possible to recommend a particular type of urinary diversion. However, most institutions
prefer ileal orthotopic neobladders and ileal conduits based on clinical experience [396, 397]. In selected
patients, such as patients with a single kidney, uretero-cutaneostomy is surgically the least burdensome type of
diversion. Recommendations related to RC and urinary diversions are listed in Section 7.3.10.

7.3.7 Morbidity and mortality


In three long-term studies and one population-based cohort study, the peri-operative mortality was reported
as 1.2–3.2% at 30 days and 2.3–8.0% at 90 days [213, 366, 368, 398, 399]. In a large single-centre series
early complications (within three months of surgery) were seen in 58% of patients [366]. Late morbidity was
usually linked to the type of urinary diversion (see also above) [369, 400]. Early morbidity associated with RC

LIMITED UPDATE MARCH 2021 37


for NMIBC (at high risk for disease progression) is similar and no less than that associated with muscle-invasive
tumours [401]. In general, lower morbidity and (peri-operative) mortality have been observed by surgeons and
in hospitals with a higher case load and therefore more experience [398, 402-406].

Table 7.6: Management of neobladder morbidity (30-64%) [407]

CLAVIEN Morbidity Management


System
Grade I Any deviation from the normal Immediate complications:
post-operative course without Post-operative ileus Nasogastric intubation (usually
the need for pharmacological removed at day 1)
treatment or surgical, endoscopic Chewing gum
and radiological interventions. Avoid fluid excess and
hypovolemia (provoke
Allowed therapeutic regimens splanchnic hypoperfusion)
are: drugs such as antiemetics, Post-operative nausea Antiemetic agent
antipyretics, analgesics, and vomiting (decrease opioids)
diuretics and electrolytes and Nasogastric intubation
physiotherapy. Urinary infection Antibiotics, no ureteral catheter
removal
This grade also includes wound Check the 3 drainages (ureters
infections opened at the bedside. and neobladdder)
Ureteral catheter Inject 5 cc saline in the
obstruction ureteral catheter to resolve the
obstruction
Increase volume infusion to
increase diuresis
Intra-abdominal urine Check drainages and watchful
leakage (anastomosis waiting
leakage)
Anaemia well tolerated Martial treatment
(give iron supplement)
Late complications:
Non compressive Watchful waiting
lymphocele
Mucus cork Cough
Indwelling catheter to remove
the obstruction
Incontinence Urine analysis (infection),
echography (post-void residual)
Physiotherapy
Retention Drainage and self-
catheterisation education
Grade II Requiring pharmacological Anaemia badly tolerated Transfusion1,2
treatment with drugs other or if myocardial
than those allowed for grade I cardiopathy history
complications. Blood transfusions Pulmonary embolism Heparinotherapy3
and total parenteral nutrition are Pyelonephritis Antibiotics and check kidney
also included. drainage
(nephrostomy if necessary)
Confusion or neurological Neuroleptics and avoid opioids
disorder
Grade III Requiring surgical, endoscopic or Ureteral catheter Indwelling leader to raise the
radiological intervention accidentally dislodged ureteral catheter
Anastomosis stenosis Renal drainage (ureteral
(7%) catheter or nephrostomy)
Ureteral reflux No treatment if asymptomatic

38 LIMITED UPDATE MARCH 2021


III-a Intervention not under general Compressive lymphocele Transcutaneous drainage or
anaesthesia intra-operative marsupialisation
(cf grade III)
III-b Intervention under general Ileal anastomosis leakage Ileostomy, as soon as possible
anaesthesia Evisceration Surgery in emergency
Compressive lymphocele Surgery (marsupialisation)
Grade IV Life-threatening complication Rectal necrosis Colostomy
(including central nervous Neobladder rupture Nephrostomy and indwelling
system complications: brain catheter/surgery for repairing
haemorrhage, ischaemic stroke, neobladder
subarachnoid bleeding, but Severe sepsis Antibiotics and check all the
excluding transient ischaemic urinary drainages and CT scan
attacks) requiring intensive care/ in emergency
intensive care unit management.
IV-a Single organ dysfunction Non-obstructive renal Bicarbonate/aetiology treatment
(including dialysis) failure
IV-b Multi-organ dysfunction Obstructive pyelonephritis Nephrostomy and antibiotics
and septicaemia
Grade V Death of a patient
Suffix ‘d’ If the patient suffers from a complication at the time of discharge, the suffix “d” (for ‘disability’)
is added to the respective grade of complication. This label indicates the need for a follow-up to
fully evaluate the complication.

1  systematic review showed that peri-operative blood transfusion (PBT) in patients who undergo RC
A
correlates with increased overall mortality, CSM and cancer recurrence. The authors hypothesised that this
may be caused by the suggested immunosuppressive effect of PBT. The foreign antigens in transfused blood
induce immune suppression, which may lead to tumour cell spread, tumour growth and reduced survival in
already immunosuppressed cancer patients. As other possible causes for this finding increased post-operative
infections and blood incompatibility were mentioned [408]. Buchner and co-workers showed similar results
in a retrospective study. The 5-year CSS decreased in cases where intra-operative blood transfusion (CSS
decreased from 67% to 48%) or post-operative blood transfusion (CSS decreased from 63% to 48%) were
given [409].

2 Intra-operative tranexamin acid infusion reduces peri-operative blood transfusion rates from 57.7% to 31.1%.
There was no increase seen in peri-operative VTE [410].

3  ammond and co-workers reviewed 20,762 cases of VTE after major surgery and found cystectomy patients
H
to have the second highest rate of VTE among all cancers studied [411]. These patients benefit from 30
days low-molecular-weight heparin prophylaxis. Subsequently, it was demonstrated that BMI > 30 and non-
urothelial BCs are independently associated with VTE after cystectomy. In these patients extended (90 days)
heparin prophylaxis should be considered [412].

7.3.8 Survival
According to a multi-institutional database of 888 consecutive patients undergoing RC for BC, the 5-year
RFS rate was 58% and CSS was 66% [413]. External validation of post-operative nomograms for BC-specific
mortality showed similar results, with bladder-CSS of 62% [414].
Recurrence-free survival and OS in a large single-centre study of 1,054 patients was 68% and 66%
at five years and 60% and 43%, at ten years, respectively [212]. However, the 5-year RFS in node-positive
patients who underwent cystectomy was considerably less at 34–43% [415, 416]. In a surgery-only study, the
5-year RFS was 76% in patients with pT1 tumours, 74% for pT2, 52% for pT3, and 36% for pT4 [212].
A trend analysis based on 148,315 BC patients identified in the SEER database between 1973 and
2009 showed increased stage-specific 5-year survival rates for all stages, except for metastatic disease [417].

7.3.9 Impact of hospital and surgeon volume on treatment outcomes


Recently, a systemic review was performed to assess the impact of hospital and/or surgeon volume on peri-
operative outcomes of RC [418]. In total, 40 studies including over 560,000 patients were included. All studies
were retrospective cohort studies.

LIMITED UPDATE MARCH 2021 39


Twenty-two studies reported on hospital volume only, six studies on surgeon volume only and twelve studies
reported on both. The results of this systematic review suggests that a higher hospital volume is likely
associated with lower in-hospital, 30-day and 90-day mortality rates. Also, higher volume hospitals are likely
to have lower positive surgical margins, higher LND and neobladder rates and lower complication rates. For
surgeon volume, less evidence is available and it seems that outcome after RC is mainly hospital-driven. In
spite of the lower quality, the available evidence suggests that performing more than 10 RCs per year per
hospital reduces 30- and 90-day mortality. Performing more than 20 RCs per hospital per year might even
further reduce these mortality rates.

7.3.10 Summary of evidence and guidelines for radical cystectomy and urinary diversion

Summary of evidence LE
Ensuring that patients are well informed about the various urinary diversion options prior to making a 3
decision may help prevent or reduce decision regret, independent of the method of diversion selected.
Higher hospital volume likely improves quality of care and reduction in peri-operative mortality and 3
morbidity.
Radical cystectomy includes removal of regional lymph nodes. 3
There are data to support that extended lymph node dissection (LND) (vs. standard or limited LND) 3
improves survival after RC.
Radical cystectomy in both sexes must not include removal of the entire urethra in all cases, which 3
may then serve as the outlet for an orthotopic bladder substitution. The terminal ileum and colon are
the intestinal segments of choice for urinary diversion.
The type of urinary diversion does not affect oncological outcome. 3
The use of extended venous thromboembolism (VTE) prophylaxis significantly decreases the incidence 3
of VTE after RC.
In patients aged > 80 years with MIBC, cystectomy is an option. 3
Surgical outcome is influenced by comorbidity, age, previous treatment for bladder cancer or other 2
pelvic diseases, surgeon and hospital volumes of cystectomy, and type of urinary diversion.
Surgical complications of cystectomy and urinary diversion should be reported using a uniform 2
grading system. Currently, the best-adapted grading system for cystectomy is the Clavien grading
system.
No conclusive evidence exists as to the optimal extent of LND. 2a

Recommendations Strength rating


Do not delay radical cystectomy (RC) for > 3 months as it increases the risk of progression Strong
and cancer-specific mortality, unless the patient receives neo-adjuvant chemotherapy.
Perform at least 10, and preferably > 20, RCs per hospital/per year. Strong
Before RC, fully inform the patient about the benefits and potential risks of all possible Strong
alternatives. The final decision should be based on a balanced discussion between the
patient and the surgeon.
Do not offer an orthotopic bladder substitute diversion to patients who have a tumour in the Strong
urethra or at the level of urethral dissection.
Pre-operative bowel preparation is not mandatory. “Fast track” measurements may reduce Strong
the time to bowel recovery.
Offer pharmacological prophylaxis, such as low-molecular-weight heparin to RC patients, Strong
starting the first day post-surgery, for a period of 4 weeks.
Offer RC to patients with T2–T4a, N0M0 disease or high-risk non-muscle-invasive bladder Strong
cancer.
Perform a lymph node dissection as an integral part of RC. Strong
Do not preserve the urethra if margins are positive. Strong

40 LIMITED UPDATE MARCH 2021


7.3.11 EAU-ESMO consensus statements on the management of advanced- and variant bladder
cancer [8, 9]*

Consensus statement
Candidates for curative treatment, such as cystectomy or bladder preservation, should be clinically assessed
by at least an oncologist, a urologist, a radiation oncologist (in case adjuvant radiotherapy or bladder
preservation is considered) and a neutral healthcare professional such as a specialist nurse.
Muscle-invasive pure squamous cell carcinoma of the bladder should be treated with primary radical
cystectomy and lymphadenectomy.
Muscle-invasive pure adenocarcinoma of the bladder should be treated with primary radical cystectomy and
lymphadenectomy.
T1 high-grade bladder urothelial cancer with micropapillary histology (established after complete TURBT and/
or re-TURBT) should be treated with immediate radical cystectomy and lymphadenectomy.
*Only statements which met the a priori consensus threshold across all three stakeholder groups are listed
(defined as ≥ 70% agreement and ≤ 15% disagreement, or vice versa).
TURBT = transurethral resection of bladder tumour.

Figure 7.1: Flow chart for the management of T2–T4a N0M0 urothelial bladder cancer

Diagnosis 1 – Males: biopsy apical prostatic urethra


or frozen section during surgery if
• Cystoscopy and tumour resection
indicated (see below)
• Evaluation of urethra1
• CT imaging of abdomen, chest, UUT
– Females : biopsy of proximal urethra or
• MRI can be used for local staging
frozen section during surgery if indicated

Findings cT2N0M0 selected patients


– Mult modality bladder-sparing therapy
• cT2-4N0M0
can be considered for T2 tumours
(Note: alternative, not the standard option )

Neo-adjuvant therapy
• Chemotherapy
Recommended in cisplatin-fit patients
(5-8% survival benefit)
• Radiotherapy
Not recommended
• Immunotherapy
Experimental, only in clinical trial setting

Radical cystectomy
• Know general aspects of surgery
- Preparation
- Surgical technique
- Integrated node dissection
- Urinary diversion
- Timing of surgery
• A higher case load improves outcome

Adjuvant chemotherapy
• Consider in high-risk patients only if no
neo-adjuvant therapy was given

CT = computed tomography; MRI = magnetic resonance imaging; UUT = upper urinary tract.

LIMITED UPDATE MARCH 2021 41


7.4 Unresectable tumours
7.4.1 Palliative cystectomy for muscle-invasive bladder carcinoma
Locally advanced tumours (T4b, invading the pelvic or abdominal wall) may be accompanied by several
debilitating symptoms, including bleeding, pain, dysuria and urinary obstruction. These patients are candidates
for palliative treatments, such as palliative RT. Palliative cystectomy with urinary diversion carries the greatest
morbidity and should be considered for symptom relief only if there are no other options [419-421].

Locally advanced MIBC can be associated with ureteral obstruction due to a combination of mechanical
blockage by the tumour and invasion of ureteral orifices by tumour cells. In a series of 61 patients with
obstructive uraemia RC was not an option in 23 patients and obstruction was relieved using permanent
nephrostomy tubes [422]. Another ten patients underwent palliative cystectomy, but local pelvic recurrence
occurred in all ten patients within the first year of follow-up. Another small study (n = 20) showed that primary
cystectomy for T4 BC was technically feasible and associated with a very tolerable therapy-related morbidity
and mortality [423].

7.4.1.1 Guidelines for unresectable tumours

Recommendations Strength rating


Offer radical cystectomy as a palliative treatment to patients with inoperable locally Weak
advanced tumours (T4b).
Offer palliative cystectomy to patients with symptoms. Weak

7.4.1.2 EAU-ESMO consensus statements on the management of advanced- and variant bladder cancer [8, 9]*

Consensus statement
In patients with clinical T4 or clinical N+ disease (regional), radical chemoradiation can be offered accepting
that this may be palliative rather than curative in outcome.
Chemoradiation should be given to improve local control in cases of inoperable locally advanced tumours.
*Only statements which met the a priori consensus threshold across all three stakeholder groups are listed
(defined as ≥ 70% agreement and ≤ 15% disagreement, or vice versa).

7.4.2 Supportive care


7.4.2.1 Obstruction of the upper urinary tract
Unilateral (best kidney) or bilateral nephrostomy tubes provide the easiest solution for UUT obstruction, but
patients find the tubes inconvenient and prefer ureteral stenting. However, stenting can be difficult to achieve.
Stents must be regularly replaced and there is the risk of stent obstruction or displacement. Another possible
solution is a urinary diversion with, or without, a palliative cystectomy.

7.4.2.2 Bleeding and pain


In the case of bleeding, the patient must be screened first for coagulation disorders or the patient’s use of
anticoagulant drugs must be reviewed. Transurethral (laser) coagulation may be difficult in a bladder full of
tumour or with a bleeding tumour. Intravesical rinsing of the bladder with 1% silver nitrate or 1–2% alum can
be effective [424]. This can usually be done without any anaesthesia. The instillation of formalin (2.5–4% for 30
minutes) is a more aggressive and painful procedure, requiring anaesthesia. Formalin instillation has a higher
risk of side-effects, e.g., bladder fibrosis, but is more likely to control the bleeding [424]. Vesicoureteral reflux
should be excluded to prevent renal complications.

Radiation therapy is another common strategy to control bleeding and is also used to control pain. An older
study reported control of haematuria in 59% of patients and pain control in 73% [425]. Irritative bladder
and bowel complaints due to irradiation are possible, but are usually mild. Non-conservative options are
embolisation of specific arteries in the small pelvis, with success rates as high as 90% [424]. Radical surgery is
a last resort and includes cystectomy and diversion (see above, Section 7.4.1).

7.5 Bladder-sparing treatments for localised disease


7.5.1 Transurethral resection of bladder tumour
Transurethral resection of bladder tumour alone in MIBC patients is only possible as a therapeutic option if
tumour growth is limited to the superficial muscle layer and if re-staging biopsies are negative for residual
(invasive) tumour [426]. In general, approximately 50% of patients will still have to undergo RC for recurrent

42 LIMITED UPDATE MARCH 2021


MIBC with a disease-specific mortality rate of up to 47% within this group [427]. A disease-free status at
re-staging TURB appears to be crucial in making the decision not to perform RC [428, 429]. A prospective
study by Solsona et al. including 133 patients with radical TURB and re-staging negative biopsies, reported a
15-year follow-up [429]. Thirty per cent of patients had recurrent NMIBC and went on to intravesical therapy,
and 30% (n = 40) progressed, of which 27 died of BC. After five, ten, and fifteen years, the results showed CSS
rates of 81.9%, 79.5%, and 76.7%, respectively and PFS rates with an intact bladder of 75.5%, 64.9%, and
57.8%, respectively.
In conclusion, TURB alone should only be considered as a therapeutic option for muscle-invasive
disease after radical TURB, when the patient is unfit for cystectomy, or refuses open surgery, or as part of a
multimodality bladder-preserving approach.

7.5.1.1 Guideline for transurethral resection of bladder tumour

Recommendation Strength rating


Do not offer transurethral resection of bladder tumour alone as a curative treatment option Strong
as most patients will not benefit.

7.5.1.2 EAU-ESMO consensus statements on the management of advanced- and variant bladder cancer [8, 9]*

Consensus statement
Candidates for curative treatment, such as cystectomy or bladder preservation, should be clinically assessed
by at least an oncologist, a urologist, a radiation oncologist (in case adjuvant radiotherapy or bladder
preservation is considered) and a neutral HCP such as a specialist nurse.
An important determinant for patient eligibility in case of bladder-preserving treatment is absence of
carcinoma in situ.
An important determinant for patient eligibility in case of bladder-preserving treatment is absence or presence
of hydronephrosis.
When assessing patient eligibility for bladder preservation, the likelihood of successful debulking surgery
should be taken into consideration (optimal debulking).
*Only statements which met the a priori consensus threshold across all three stakeholder groups are listed
(defined as ≥ 70% agreement and ≤ 15% disagreement, or vice versa).
HCP = healthcare professional.

7.5.2 External beam radiotherapy


Current RT techniques with soft-tissue matching and image guidance result in superior bladder coverage and
a reduced integral dose to the surrounding tissues. The target total dose (to bladder and/or bladder tumour) for
curative EBRT in BC is 64–66 Gy [430, 431]. A reasonable alternative is moderately hypofractionated EBRT to
55 Gy in 20 fractions which has been suggested to be non-inferior to 64 Gy in 32 fractions in terms of invasive
locoregional control, OS, and late toxicity. In a phase II study, 55 patients (median age 86) with BC, unfit for
cystectomy or even daily RT, were treated with 6-weekly doses of 6 Gy [432]. Forty-eight patients completed
EBRT with acceptable toxicity and 17% showed local progression after two years demonstrating good local
control with this more ultra-hypofractionated schedule.
Elective treatment to the LNs is optional and should take into account patient comorbidities and the
risks of toxicity to adjacent critical structures. For node-positive disease, consider boosting grossly involved
nodes to the highest achievable dose that does not violate normal tissue constraints based on the clinical
scenario.
The use of modern standard EBRT techniques results in major related late morbidity of the urinary
bladder or bowel in less than 5% of patients [433]. Acute diarrhoea is reduced even more with intensity-
modulated RT [434]. Important prognostic factors for outcome include response to EBRT, tumour size,
hydronephrosis, presence of CIS, and completeness of the initial TURB. Additional prognostic factors reported
are age and stage [435].

With the use of modern EBRT techniques, efficacy and safely results seem to have improved over time. A 2002
Cochrane analysis demonstrated that RC has an OS benefit compared to RT [436], although this was not the
case in a 2014 retrospective review using a propensity score analysis [437]. In a 2017 retrospective cohort
study of U.S. National Cancer Data Base data, patients over 80 were identified with cT2–4, N0–3, M0 BC, who
were treated with curative EBRT (60–70 Gy, n = 739) or concurrent chemoradiotherapy (n = 630) between 2004
and 2013 [438]. The 2-year OS was 42% for EBRT vs. 56% for chemoradiotherapy (p < 0.001). For EBRT a
higher RT dose and a low stage were associated with improved OS.

LIMITED UPDATE MARCH 2021 43


In conclusion, although EBRT results seem to improve over time, EBRT alone does not seem to be as
effective as surgery or trimodality therapy (see Section 7.5.4). Factors that influence outcome should be
considered. However, EBRT can be an alternative treatment in patients unfit for radical surgery or concurrent
chemotherapy, and it can also be quite effective in helping control bleeding.

7.5.2.1 Summary of evidence and guideline for external beam radiotherapy

Summary of evidence LE
External beam radiotherapy alone should only be considered as a therapeutic option when the patient 3
is unfit for cystectomy.
Radiotherapy can also be used to stop bleeding from the tumour when local control cannot be 3
achieved by transurethral manipulation because of extensive local tumour growth.

Recommendation Strength rating


Do not offer radiotherapy alone as primary therapy for localised bladder cancer. Strong

7.5.2.2 EAU-ESMO consensus statements on the management of advanced- and variant bladder cancer [8, 9]*

Consensus statement
Radiotherapy alone (single block) is not the preferred radiotherapeutic schedule.
Radiotherapy for bladder preservation should be performed with IMRT and IGRT to reduce side effects.
Dose escalation above standard radical doses to the primary site in case of bladder preservation, either by
IMRT or brachytherapy, is not recommended.
*Only statements which met the a priori consensus threshold across all three stakeholder groups are listed
(defined as ≥ 70% agreement and ≤ 15% disagreement, or vice versa).
IGRT = image-guided radiotherapy; IMRT = intensity-modulated radiotherapy.

7.5.3 Chemotherapy
Chemotherapy alone rarely produces durable complete remissions. In general, a clinical complete response
rate of up to 56% is reported in some series, which must be weighed against a staging error of > 60% [439,
440]. Response to chemotherapy is a prognostic factor for treatment outcome and eventual survival although it
may be confounded by patient selection [441].

Several groups have reported the effect of chemotherapy on resectable tumours (neoadjuvant approach), as
well as unresectable primary tumours [223, 240, 442, 443]. Neoadjuvant chemotherapy with two to three cycles
of MVAC or CMV has led to a down-staging of the primary tumour in various prospective series [223, 240, 442].
A bladder-conserving strategy with TURB and systemic cisplatin-based chemotherapy has been
reported several years ago and could lead to long-term survival with intact bladder in a highly selected patient
population [441].
A recent large retrospective analysis of a National Cancer Database cohort reported on 1,538
patients treated with TURB and multi-agent chemotherapy [444]. The two and 5-year OS for all patients was
49% and 32.9% and for cT2 patients it was 52.6% and 36.2%, respectively. While these data show that long-
term survival with intact bladder can be achieved in a subset of patients it is not recommended for routine use.

7.5.3.1 Summary of evidence and guideline for chemotherapy

Summary of evidence LE
Complete and partial local responses have been reported with cisplatin-based chemotherapy as 2b
primary therapy for locally advanced tumours in highly selected patients.

Recommendation Strength rating


Do not offer chemotherapy alone as primary therapy for localised bladder cancer. Strong

44 LIMITED UPDATE MARCH 2021


7.5.4 Trimodality bladder-preserving treatment
Trimodality therapy (TMT) combines TURB, chemotherapy and RT. The rationale to combine TURB with RT is
to maximally achieve local tumour control in the bladder and adjacent nodes. The addition of radiosensitising
chemotherapy or other radiosensitisers (mentioned below) is aimed at the potentiation of RT. Micrometastases
are targeted by platinum-based combination chemotherapy (for details see Section 7.1). The aim of TMT is to
preserve the bladder and QoL without compromising oncological outcome.

There are no successfully completed RCTs comparing the outcome of TMT with RC, but TMT using
chemoradiation has been shown to be superior to RT alone [445, 446]. Many of the reported series have
differing characteristics as compared to the larger surgical series, which typically have median ages in the mid-
to-late 60s compared to mid-70s for some large RT series (reviewed by James, et al. [445]). In the case of TMT,
two distinct patterns of care emerge; treatment aimed at patients fit for cystectomy who elect TMT or refuse
cystectomy, and treatment aimed at older, less fit, patients. For the former category, TMT presents selective
bladder preservation and in this case the initial step is a radical TURB where as much tumour as possible
should be resected. In this case appropriate patient selection (e.g., T2 tumours, no CIS) is critical [447, 448].
Even in case of an initial presumed complete resection, a second TUR has been suggested to reveal tumour
in > 50% of patients and subsequently improves 5-year OS in case of TMT [449]. For patients who are not
candidates for cystectomy, less stringent criteria can be applied, but extensive CIS and poor bladder function
should both be regarded as relative contraindications.

A collaborative review has described the principles of TMT [450]. For radiation, two schedules are most
commonly used; historically within the RTOG a split-course format with interval cystoscopy [446] and single-
phase treatment which is now more commonly used [445]. A conventional radiation schedule includes EBRT to
the bladder and limited pelvic LNs with an initial dose of 40-45 Gy, with a boost to the whole bladder of 50–54
Gy and a further tumour boost to a total dose of 60–66 Gy. If not boosting the tumour, it is also reasonable for
the whole bladder to be treated to 59.4–66 Gy. For node-positive disease, consider boosting grossly involved
nodes to the highest achievable dose that does not violate normal tissue constraints. Therefore, elective
treatment to the LNs (when node negative) is optional and should take into account patient comorbidities and
the risks of toxicity to adjacent critical structures.

In summary, reasonable radiation fields include pelvis (with bladder and/or bladder tumour boost), bladder only
or partial bladder (tumour) only [445]. A reasonable radiation dosing alternative to conventional fractionation
when treating the bladder-only fields is moderately hypofractionated EBRT to 55 Gy in 20 fractions which has
been suggested to be non-inferior to 64 Gy in 32 fractions (fx) in terms of invasive loco-regional control, OS
and late toxicity [430, 451].

Different chemotherapy regimens have been used, but most evidence exists for cisplatin [452] and mitomycin
C plus 5-FU [445]. In addition to these agents, other regimens have also been used such as gemcitabine
and hypoxic cell sensitisation with nicotinamide and carbogen, without clear preference for a specific
radiosensitiser [8, 9]. In a recently published phase II RCT, twice-a-day radiation plus fluorouracil/cisplatin was
compared to once-daily radiation plus gemcitabine [453]. Both arms were found to result in a > 75% freedom
of distant metastases at 3 years (78% and 84%, respectively). Therefore, there are options for non-cisplatin
candidates such as 5-FU/mitomycin C or low-dose gemcitabine.

To detect non-responders who should be offered salvage cystectomy, bladder biopsies should be performed
after TMT and life-long cystoscopic surveillance is recommended.

Five-year CSS and OS rates vary between 50%–84% and 36%–74%, respectively, with salvage cystectomy
rates of 10–30% [445, 447, 450, 452, 454, 455]. The Boston group reported on their experience in 66 patients
with mixed variant histologies treated with TMT and found similar complete response, OS, DSS and salvage
cystectomy rates as in UC [456]. The majority of recurrences post-TMT are non-invasive and can be managed
conservatively [445]. In contemporary experiences, salvage cystectomy is required in about 10–15% of patients
treated with TMT and can be curative [445, 447, 455]. Current data suggest that major late complication rates
are slightly higher but remain acceptable for salvage- vs. primary cystectomy [457, 458].

A sub-analysis of two RTOG trials looked at complete response (T0) and near-complete response (Ta or
Tis) after TMT [459]. After a median follow-up of 5.9 years 41/119 (35%) of patients experienced a bladder
recurrence, and fourteen required salvage cystectomy. There was no difference between complete and near-
complete responders. Non-muscle-invasive BC recurrences after complete response to TMT were reported
in 25% of patients by the Boston group, sometimes over a decade after initial treatment [460]. A NMIBC

LIMITED UPDATE MARCH 2021 45


recurrence was associated with a lower DSS, although in properly selected patients, intravesical BCG could
avoid immediate salvage cystectomy.

The differential impact of RC vs. TMT on long-term OS is lacking a randomised comparison and rigorous
prospective data. A propensity score matched institutional analysis has suggested similar DSS and OS
between TMT and RC [455]. Two retrospective analyses of the National Cancer Database from 2004–2013
with propensity score matching compared RC to TMT. Ritch et al. identified 6,606 RC and 1,773 TMT patients
[461]. Worse survival was linked to higher age, comorbidity and tumour stage. After modelling, TMT resulted
in a lower mortality at one year (HR: 0.84, 95% CI: 0.74–0.96, p = 0.01). However, in years 2 and onwards,
there was a significant and persistent higher mortality after TMT (year 2: HR: 1.4, 95% CI: 1.2–1.6, p < 0.001;
and year 3 onwards: HR: 1.5, 95% CI: 1.2–1.8, p < 0.001). The second analysis was based on a larger cohort,
with 22,680 patients undergoing RC; 2,540 patients received definitive EBRT and 1,489 TMT [462]. Survival
after modelling was significantly better for RC compared to any EBRT, definitive EBRT and TMT (HR: 1.4, 95%
CI: 1.2–1.6) at any time point. In older patients which are potentially less ideal candidates for radical surgery,
Williams et al. found a significantly lower OS (HR :1.49, 1.31–1.69) and CSS (1.55, 1.32–1.83) for TMT as
compared to surgery as well as increased costs [463]. This was a retrospective SEER database study which
included 687 propensity-matched patients in each arm, however, the median number of radiation fractions
was well below what is considered adequate for definitive therapy and as such the radiation patients may have
been treated inadequately or palliatively. In general, such population-based studies are limited by confounding,
misclassification, and selection bias. A systematic review including 57 studies and over 30,000 patients
comparing RC and TMT found improved 10-year OS and DSS for TMT, but for the entire cohort OS and DSS
did not significantly differ between RC and TMT [464]. Complete response after TMT resulted in significantly
better survival, as did down-staging after TURB or NAC in case of RC.

Overall significant late pelvic (GI/genitourinary [GU]) toxicity rates after TMT are low and QoL is good [445, 465,
466]. A combined analysis of survivors from four RTOG trials with a median follow-up of 5.4 years showed that
combined-modality therapy was associated with low rates of late grade 3 toxicity (5.7% GU and 1.9% GI). No
late grade 4 toxicities or treatment-related deaths were recorded [465]. A retrospective study showed QoL to
be good after TMT and in most domains better than after cystectomy, although prospective validations are
needed [467]. One option to reduce side effects after TMT is the use of IMRT and image-guided radiotherapy
(IGRT) [8, 9, 468].

A collaborative review came to the conclusion that data are accumulating, suggesting that bladder preservation
with TMT leads to acceptable outcomes and therefore TMT may be considered a reasonable treatment option
in well-selected patients as compared to RC [450]. Bladder preservation as an alternative to RC is generally
reserved for patients with smaller solitary tumours, negative nodes, no extensive or multifocal CIS, no tumour-
related hydronephrosis, and good pre-treatment bladder function. Trimodality bladder-preserving treatment
should also be considered in all patients with a contraindication for surgery, either a relative or absolute
contraindication since the factors that determine fitness for surgery and chemoradiotherapy differ. There are no
definitive contemporary data supporting the benefit of using neoadjuvant or adjuvant chemotherapy combined
with chemoradiation. Patient selection is critical in achieving good outcomes [450]. Whether a node dissection
should be performed before TMT as in RC remains unclear [8, 9].

A bladder-preserving trimodality strategy requires very close multidisciplinary cooperation [8, 9]. This was also
highlighted by a Canadian group [469]. In Ontario between 1994 and 2008 only 10% (370/3,759) of patients
with cystectomy had a pre-operative radiation oncology consultation, with high geographical variations.
Independent factors associated with this consultation included advanced age (p < 0.001), greater comorbidity
(p < 0.001) and earlier year of diagnosis (p < 0.001). A bladder-preserving trimodality strategy also requires
a high level of patient compliance. Even if a patient has shown a clinical response to a trimodality bladder-
preserving strategy, the bladder remains a potential source of recurrence, hence long-term life-long bladder
monitoring is essential and patients should be counselled that this will be required.

7.5.4.1 Summary of evidence and guidelines for trimodality bladder-preserving treatment

Summary of evidence LE
In a selected patient population, long-term survival rates of trimodality bladder-preserving treatment 2b
are comparable to those of early cystectomy.

46 LIMITED UPDATE MARCH 2021


Recommendations Strength rating
Offer surgical intervention or trimodality bladder-preserving treatments (TMT) to appropriate Strong
candidates as primary curative therapeutic approaches since they are more effective than
radiotherapy alone.
Offer TMT as an alternative to selected, well-informed and compliant patients, especially for Strong
whom radical cystectomy is not an option or not acceptable.

7.5.4.2 EAU-ESMO consensus statements on the management of advanced- and variant bladder cancer [8, 9]*

Consensus statement
Candidates for curative treatment, such as cystectomy or bladder preservation, should be clinically assessed
by at least an oncologist, a urologist, a radiation oncologist (in case adjuvant radiotherapy or bladder
preservation is considered) and a neutral HCP such as a specialist nurse.
An important determinant for patient eligibility in case of bladder-preserving treatment is absence of
carcinoma in situ.
An important determinant for patient eligibility in case of bladder-preserving treatment is absence or presence
of hydronephrosis.
When assessing patient eligibility for bladder preservation, the likelihood of successful debulking surgery
should be taken into consideration (optimal debulking).
Bladder urothelial carcinoma with small cell neuroendocrine variant should be treated with neoadjuvant
chemotherapy followed by consolidating local therapy.
In case of bladder preservation with radiotherapy, combination with a radiosensitiser is always recommended
to improve clinical outcomes, such as cisplatin, 5FU/TMC, carbogen/nicotinamide or gemcitabine.
Radiotherapy for bladder preservation should be performed with IMRT and IGRT to reduce side effects.
Dose escalation above standard radical doses to the primary site in case of bladder preservation, either by
IMRT or by brachytherapy, is not recommended.
*Only statements which met the a priori consensus threshold across all three stakeholder groups are listed
(defined as ≥ 70% agreement and ≤ 15% disagreement, or vice versa).
HCP = healthcare professional; IGRT = image-guided radiotherapy; IMRT = intensity-modulated radiotherapy;
5FU = 5-fluorouracil; MMC = mitomycin-C.

7.6 Adjuvant therapy


7.6.1 Role of adjuvant platinum-based chemotherapy
Adjuvant chemotherapy after RC for patients with pT3/4 and/or LN positive (N+) disease without clinically
detectable metastases (M0) is still under debate [457, 470].

The general benefits of adjuvant chemotherapy include:


• chemotherapy is administered after accurate pathological staging, therefore, treatment in patients at low
risk for micrometastases is avoided;
• no delay in definitive surgical treatment.
The drawbacks of adjuvant chemotherapy are:
• assessment of in vivo chemosensitivity of the tumour is not possible and overtreatment is an unavoidable
problem;
• delay or intolerability of chemotherapy, due to post-operative morbidity [471].

There is limited evidence from adequately conducted and accrued phase III RCTs in favour of the routine use
of adjuvant chemotherapy [470, 472-477]. An individual patient data meta-analysis [472] of survival data from
six RCTs of adjuvant chemotherapy [454, 478-481] included 491 patients (unpublished data from Otto et al.,
were included in the analysis). All included trials suffered from significant methodological flaws including small
sample size (underpowered), incomplete accrual, use of inadequate statistical methods and design flaws
(irrelevant endpoints and failing to address salvage chemotherapy in case of relapse or metastases) [470]. In
these trials, three or four cycles of CMV, cisplatin, cyclophosphamide, and adriamycin (CISCA), methotrexate,
vinblastine, adriamycin or epirubicin, and cisplatin (MVA(E)C) and cisplatin and methotrexate (CM) were used
[482], and one trial used cisplatin monotherapy [480]. The data were not convincing to support an unequivocal
recommendation for the use of adjuvant chemotherapy. In 2014, this meta-analysis was updated with an
additional three studies [474-476] resulting in the inclusion of 945 patients from nine trials [473]. None of the
trials had fully accrued and individual patient data were not used in the analysis [473]. For one trial only an
abstract was available at the time of the meta-analysis [475] and none of the included individual trials were
significantly positive for OS in favour of adjuvant chemotherapy. In two of the trials more modern chemotherapy

LIMITED UPDATE MARCH 2021 47


regimens were used (gemcitabine/cisplatin and paclitaxel/gemcitabine/cisplatin) [474, 475]. The HR for OS was
0.77 (95% CI: 0.59–0.99, p = 0.049) and for DFS 0.66 (95% CI: 0.45–0.91, p = 0.014) with a stronger impact on
DFS in case of nodal positivity.

A retrospective cohort analysis including 3,974 patients after cystectomy and LND showed an OS benefit in
high-risk subgroups (extravesical extension and nodal involvement) (HR: 0.75, CI: 0.62–0.90) [483]. A recent
publication of the largest RCT (EORTC 30994), although not fully accrued, showed a significant improvement
of PFS for immediate, compared with deferred, cisplatin-based chemotherapy (HR: 0.54, 95% CI: 0.4–0.73,
p < 0.0001), but there was no significant OS benefit [484].
Furthermore, a large observational study including 5,653 patients with pathological T3–4 and/
or pathological node-positive BC, treated between 2003 and 2006 compared the effectiveness of adjuvant
chemotherapy vs. observation. Twenty-three percent of patients received adjuvant chemotherapy with a 5-year
OS of 37% for the adjuvant arm vs. 29.1% (HR: 0.70, 95% CI: 0.64–0.76) in the observation group [485].
Another large retrospective analysis based on National Cancer Data Base including 15,397 patients
with locally advanced (pT3/4) or LN-positive disease also demonstrated an OS benefit in patients with UC
histology [486]. In patients with concomitant variant or pure variant histology, however, no benefit was found.

From the currently available evidence it is still unclear whether immediate adjuvant chemotherapy or
chemotherapy at the time of relapse is superior, or if the two approaches are equivalent with respect to
the endpoint of OS. The most recent meta-analysis from 2014 showed a therapeutic benefit of adjuvant
chemotherapy, but the level of evidence of this review is still very low, with significant heterogeneity and
methodological flaws in the only nine included trials [473]. Patients should be informed about potential
chemotherapy options before RC, including neoadjuvant and adjuvant chemotherapy, and the limited evidence
for adjuvant chemotherapy.

7.6.2 Role of adjuvant immunotherapy


To evaluate the benefit of PD-1/PD-L1 checkpoint inhibitors, a number of randomised phase III trials
comparing checkpoint inhibitor monotherapy with atezolizumab, nivolumab or pembrolizumab are under way.
Preliminary results of two phase III trials have been presented (atezolizumab at ASCO 2020; nivolumab at
ASCO GU 2021): the primary endpoint of improved DFS was not achieved with atezolizumab but was achieved
with adjuvant nivolumab. Further results and longer follow-up have to be awaited.
So far, adjuvant immunotherapy is not standard of care and should only be given within a clinical
trial [487].

7.6.3 Guidelines for adjuvant therapy

Recommendations Strength rating


Offer adjuvant cisplatin-based combination chemotherapy to patients with pT3/4 and/or Strong
pN+ disease if no neoadjuvant chemotherapy has been given.
Only offer immunotherapy with a checkpoint inhibitor in a clinical trial setting. Strong

7.7 Metastatic disease


7.7.1 Introduction
Approximately 50% of patients with muscle-invasive UC relapse after RC, depending on the pathological stage
of the primary tumour and the nodal status. Local recurrence accounts for 30% of relapses, whereas distant
metastases are more common. Ten to fifteen percent of patients are already metastatic at diagnosis [488].
Before the development of effective chemotherapy patients with metastatic UC had a median survival rarely
exceeding three to six months [489].

7.7.1.1 Prognostic factors and treatment decisions


Prognostic factors are crucial for assessing phase II study results and stratifying phase III trials [490, 491]. In a
multivariate analysis, Karnofsky PS of ≤ 80% and presence of visceral metastases were independent prognostic
factors of poor survival after treatment with MVAC [491]. These prognostic factors have also been validated for
newer combination chemotherapy regimens [492-496]. Two additional prognostic models have been developed
in the past years including the variables leukocyte count, number of sites of visceral metastases, site of primary
tumour, PS, LN metastasis [496] and visceral metastasis, albumin and haemoglobin [495], respectively.
For patients refractory to, or progressing shortly after, platinum-based combination chemotherapy,
four prognostic groups have been established based on three adverse factors that have developed in patients
treated with vinflunine and that have been validated in an independent data set; Hb < 10 g/dL presence of liver

48 LIMITED UPDATE MARCH 2021


metastases and ECOG PS ≥ 1 [497]. It is important to acknowledge that these prognostic models have not
been validated in the context of newer agents including immunotherapy.

7.7.1.2 Comorbidity in metastatic disease


Comorbidity is defined as “the presence of one or more disease(s) in addition to an index disease” (see
Section 5.3). Comorbidity increases with age. However, chronological age does not necessarily correlate with
functional impairment. Different evaluation systems are being used to screen patients as potentially fit or unfit
for chemotherapy, but age alone should not be used as the basis for treatment selection [498].

7.7.2 First-line systemic therapy for metastatic disease


In general, patients with untreated metastatic BC can be divided into three broad categories: fit for cisplatin-
based chemotherapy, fit for carboplatin-based chemotherapy (but unfit for cisplatin) and unfit for any platinum-
based chemotherapy.

7.7.2.1 Definitions: ‘Fit for cisplatin, fit for carboplatin, unfit for any platinum-based chemotherapy’
An international survey among BC experts [499] was the basis for a consensus statement on how to classify
patients unfit for cisplatin-based chemotherapy. At least one of the following criteria has to be present: PS > 1;
GFR ≤ 60 mL/min; grade ≥ 2 audiometric loss; grade ≥ 2 peripheral neuropathy or New York Heart Association
(NYHA) class III heart failure [500]. More than 50% of patients with UC are not eligible for cisplatin-based
chemotherapy [501-504]. Renal function assessment in UC is of utmost importance for treatment selection
[501, 505]. In case of doubt, measuring GFR with radioisotopes (99mTc DTPA or 51Cr-EDTA) is recommended.
Cisplatin has also been administered in patients with a lower GFR (40–60 mL/min) using different split-dose
schedules. The respective studies were mostly small phase I and II trials in different settings (neoadjuvant
and advanced disease) demonstrating that the use of split-dose cisplatin is feasible and appears to result in
encouraging efficacy [506-509]. However, no prospective randomised trial has compared split-dose cisplatin
with conventional dosing.

Most patients that are deemed unfit for cisplatin are able to receive carboplatin-based chemotherapy. However,
some patients are deemed unfit for any platinum-based chemotherapy (both cisplatin and carboplatin) in case
of PS > 2, impaired renal function with a GFR < 30 mL/min or the combination of PS 2 and GFR < 60 mL/min
since the outcome in this patient population is poor regardless of platinum-based treatment or not [510].
Patients with multiple comorbidities may also be poor candidates for platinum-based chemotherapy.

7.7.2.2 Chemotherapy in patients fit for cisplatin


Cisplatin-containing combination chemotherapy has been the standard of care since the late 1980s
demonstrating an OS of twelve to fourteen months in different series (for a review see [511]). Methotrexate,
vinblastine, adriamycin plus cisplatin (MVAC) and GC prolonged survival to up to 14.8 and 13.8 months,
respectively, compared to monotherapy and older chemotherapy combinations. Neither of the two
combinations is superior to the other but equivalence has not been tested. Response rates were 46% and 49%
for MVAC and GC, respectively. The long-term survival results have confirmed the efficacy of the two regimens
[512]. The major difference between the above-mentioned combinations is toxicity. The lower toxicity of GC
[174] compared to standard MVAC has resulted in it becoming a new standard regimen [513]. Methotrexate,
vinblastine, adriamycin plus cisplatin is better tolerated when combined with granulocyte colony-stimulating
factor (G-CSF) [513, 514].

High-dose intensity MVAC (HD-MVAC) combined with G-CSF is less toxic and more efficacious than standard
MVAC in terms of dose density, complete response (CR), and 2-year survival rate. However, there is no
significant difference in median survival between the two regimens [515, 516]. In general, all disease sites have
been shown to respond to cisplatin-based combination chemotherapy. A response rate of 66% and 77% with
MVAC and HD-MVAC, respectively, has been reported in retroperitoneal LNs vs. 29% and 33% at extranodal
sites [515]. The disease sites also have an impact on long-term survival. In LN-only disease, 20.9% of patients
were alive at five years compared to only 6.8% of patients with visceral metastases [512].
Further intensification of treatment using paclitaxel, cisplatin and gemcitabine (PCG) triple regimen
did not result in a significant improvement in OS in the intention-to-treat (ITT) population of a large phase III
RCT, comparing PCG triple regimen to GC [517]. However, the overall response rate (ORR) was higher with the
triple regimen (56% vs. 44%, p = 0.0031), and the trend for OS improvement in the ITT population (15.8 vs.
12.7 months; HR = 0.85, p = 0.075) became significant in the eligible population.

Carboplatin-containing chemotherapy has not been proven to be equivalent to cisplatin combinations, and
should not be considered interchangeable or standard in patients fit for cisplatin. Several phase II RCTs of

LIMITED UPDATE MARCH 2021 49


carboplatin vs. cisplatin combination chemotherapy have produced lower CR rates and shorter OS for the
carboplatin arms [518]. Recently, a retrospective International Study of Advanced/Metastatic Cancer of the
Urothelium (RISC) group highlighted the importance of applying the cisplatin-eligibility criteria in order to
maintain benefit [519].

7.7.2.3 Chemotherapy in patients fit for carboplatin (but unfit for cisplatin)
Up to 50% of patients are not fit for cisplatin-containing chemotherapy but may be candidates for carboplatin
[500]. The first randomised phase II/III trial in this setting was conducted by the EORTC and compared two
carboplatin-containing regimens (methotrexate/carboplatin/vinblastine [M-CAVI] and carboplatin/gemcitabine
[GemCarbo]) in patients unfit for cisplatin. The EORTC definitions for eligibility were GFR < 60 mL/min and/or
PS 2. Both regimens were active. Severe acute toxicity was 13.6% in patients treated with GemCarbo vs. 23%
with M-CAVI, while the ORR was 42% for GemCarbo and 30% for M-CAVI. Further analysis showed that in
patients with PS 2 and impaired renal function, combination chemotherapy provided very limited benefit [510].
The ORR and severe acute toxicity were both 26% for the former group, and 20% and 24%, respectively, for
the latter group [510]. Phase III data have confirmed these results [494]. The combination of carboplatin and
gemcitabine can be considered a standard of care in this patient group.
A randomised, international phase II trial (JASINT1; JAVLOR Association Study in CDDP-unfit
Patients With Advanced Transitional Cell Carcinoma: Gemcitabine Versus Carboplatin) assessed the efficacy
and tolerability profile of two vinflunine-based regimens (vinflunine/gemcitabine vs. vinflunine/carboplatin).
Both regimens showed equal ORR and OS with less haematologic toxicity for the combination vinflunine/
gemcitabine [520].

7.7.2.4 Integration of immunotherapy in the first-line treatment of patients fit for platinum-based
chemotherapy
7.7.2.4.1 Immunotherapy combination approaches
In 2020 the results of three phase III trials have been presented and published investigating the use of
immunotherapy in the first-line setting for platinum-eligible patients.
The first trial to report was IMvigor130 investigating the combination of the PD-L1 inhibitor
atezolizumab plus platinum-gemcitabine chemotherapy vs. chemotherapy plus placebo vs. atezolizumab
alone [521]. The primary endpoint of PFS benefit for the combination vs. chemotherapy alone in the ITT group
was reached (8.2 months vs. 6.3 months [HR: 0.82, 95% CI: 0.70–0.96; one-sided, p = 0.007]) while OS was
not significant at the interim analysis after a median follow-up of 11.8 months. The small PFS benefit in the
absence of an OS benefit has raised questions of its clinical significance. Due to the sequential testing design,
the comparison of chemotherapy vs. atezolizumab alone has not yet been formally performed.
The KEYNOTE 361 study had a very similar design using the PD-1 inhibitor pembrolizumab plus
platinum-gemcitabine vs. chemotherapy alone vs. pembrolizumab alone. The results of the primary endpoints
of PFS and OS for the comparison of pembrolizumab plus chemotherapy vs. chemotherapy in the ITT
population have been presented but are not yet fully published and show no benefit for the combination [522].
DANUBE compared the immunotherapy combination (IO-IO) of CTLA-4 inhibitor tremelimumab and
PD-L1 inhibitor durvalumab with chemotherapy alone or durvalumab alone [523]. The co-primary endpoint of
improved OS for the IO-IO combination vs. chemotherapy was not reached in the ITT group nor was the OS
improved for durvalumab monotherapy vs. chemotherapy in the PD-L1-positive population.

In conclusion; at this time these three trials do not support the use of combination of PD-1/L1 checkpoint
inhibitors plus chemotherapy or the IO-IO combination.

7.7.2.4.2 Use of single-agent immunotherapy


Based on the results of two single-arm phase II trials [524, 525] the checkpoint inhibitors pembrolizumab and
atezolizumab have been approved by the FDA and the EMA for first-line treatment in cisplatin-unfit patients
in case of positive PD-L1 status. Programmed death-ligand-1 positivity for use of pembrolizumab is defined
by immunohistochemistry as a combined positive score (CPS) of ≥ 10 using the Dako 22C33 platform and for
atezolizumab as positivity of ≥ 5% tumour-infiltrating immune cells using Ventana SP142.
The PD-1 inhibitor pembrolizumab was tested in 370 patients with advanced or metastatic UC
ineligible for cisplatin, showing an ORR of 29% and complete remission in 7% of patients [524]. The PD-L1
inhibitor atezolizumab was also evaluated in the same patient population in a phase II trial (n = 119) showing
an ORR of 23% with 9% of patients achieving a complete remission; median OS was 15.9 months [525]. The
results are difficult to interpret due to the missing control arm and the heterogeneity of the study population
with regards to PD-L1 status.
The trials IMvigor 130, Keynote 361 and DANUBE all included an experimental arm with
immunotherapy alone using atezolizumab, pembrolizumab and durvalumab, respectively [521-523]. The

50 LIMITED UPDATE MARCH 2021


results presented so far have not demonstrated a benefit in terms of PFS or OS compared to platinum-based
chemotherapy but no subgroup comparisons for cisplatin-unfit patients were analysed.

7.7.2.4.3 Switch maintenance with immunotherapy


The JAVELIN Bladder 100 study investigated the impact of switch maintenance with the PD-L1 inhibitor
avelumab after initial treatment with platinum-gemcitabine combination [526]. Patients achieving at least stable
disease, or better, after 4–6 cycles of platinum-gemcitabine were randomised to avelumab or best supportive
care (BSC). Overall survival was the primary endpoint which improved to 21.4 months with avelumab compared
to 14.3 months with BSC (HR: 0.69, 95% CI: 0.56–0.86; p < 0.001). Of patients who discontinued BSC and
received subsequent treatment 53% received immunotherapy. Grade 3 or higher side effects occurred in 47%
of avelumab patients compared to 25% of BSC patients. Immune-related adverse events occurred in 29% of
all patients and 7% experienced grade 3 complications which included colitis, pneumonitis, rash, increased
liver enzymes, hyperglycaemia, myositis and hypothyroidism.
A randomised phase II trial evaluated switch maintenance treatment with pembrolizumab in patients
achieving at least stable disease on platinum-based first-line chemotherapy [527]. One hundred and eight
patients were randomised to pembrolizumab or placebo. At a median follow-up of 12.9 months (range 0.9–34.5
months), the primary endpoint of PFS was met (5.4 months vs. 3.0 months, HR: 0.65, p = 0.04) but not the
secondary endpoint of OS (22 months vs. 18.7 months, HR: 0.91, 95% CI: 0.52–1.59).

7.7.2.5 Treatment of patients unfit for any platinum-based chemotherapy


Limited data exists regarding the optimal treatment for this patient population which is characterised by
impaired PS (PS > 2) and/or impaired renal function (GFR < 30 mL/min). Historically, the outcome in this
patient group has been poor. Often BSC has been chosen instead of systemic therapy. Most trials evaluating
alternative treatment options to cisplatinum-based chemotherapy did not focus specifically on this patient
population thereby making interpretation of data difficult. The FDA (but not EMA) has approved pembrolizumab
and atezolizumab as first-line treatment for patients not fit to receive any platinum-based chemotherapy
regardless of PD-L1 status based on the results of two single-arm phase II trials [524, 525]. It has not been
reported how many patients in these two studies were unfit for any platinum-based chemotherapy.

7.7.2.6 Non-platinum combination chemotherapy


The use of single-agent chemotherapy has been associated with varying response rates. Responses with
single agents are usually short-lived, complete responses are rare, and no long-term DFS/OS has been
reported.

Different combinations of gemcitabine and paclitaxel have been studied as first- and second-line treatments.
Apart from severe pulmonary toxicity with a weekly schedule of both drugs, this combination is well tolerated
and produces response rates between 38% and 60% in both lines. Non-platinum combination chemotherapy
has not been compared to standard platinum-based chemotherapy in RCTs; therefore, it is not recommended
for first-line use in platinum-eligible patients [528-535].

7.7.3 Second-line systemic therapy for metastatic disease


7.7.3.1 Second-line chemotherapy
Second-line chemotherapy data are highly variable and mainly derive from small single-arm phase II trials
apart from a single randomised phase III study [497]. A reasonable strategy has been to re-challenge former
cisplatin-sensitive patients if progression occurred at least six to twelve months after first-line cisplatin-based
combination chemotherapy. Second-line response rates of single-agent treatment with paclitaxel (weekly),
docetaxel, nab-paclitaxel (nanoparticle albumin-bound) [536] oxaliplatin, ifosfamide, topotecan, pemetrexed,
lapatinib, gefitinib and bortezomib have ranged between 0% and 28% in small phase II trials [537-539].
Gemcitabine has also shown response in second-line use but most patients receive this drug as part of their
first-line treatment [535].
The paclitaxel/gemcitabine combination has shown response rates of 38–60% in small single-arm
studies. No phase III RCT with an adequate comparator arm has been conducted to assess the true value and
OS benefit of this second-line combination [489, 533, 540].
Vinflunine, a third-generation vinca alkaloid, was tested in a phase III RCT and compared against
BSC in patients progressing after first-line treatment with platinum-containing combination chemotherapy for
metastatic disease [541]. The results showed a modest ORR (8.6%), a clinical benefit with a favourable safety
profile and a survival benefit in favour of vinflunine, which was, however, only statistically significant in the
eligible patient population (not in the ITT population).
Vinflunine was approved as second-line treatment in Europe (not in the U.S.). More recently,
second-line therapy with PD-1/PD-L1 checkpoint inhibitors has been established as standard second-line

LIMITED UPDATE MARCH 2021 51


therapy and vinflunine is reserved for patients with contraindications to immunotherapy and may be considered
as third- or later-line treatment option although no randomised data for these indications exist.
A randomised phase III trial evaluated the addition of the angiogenesis inhibitor ramucirumab to
docetaxel chemotherapy vs. docetaxel alone, which resulted in improved PFS (4.07 vs. 2.76 months) and
higher response rates (24.5% vs. 14%), respectively [542]. While the primary endpoint of PFS prolongation was
reached, the clinical benefit appears small and no OS benefit has been shown [543].

7.7.3.2 Second-line immunotherapy for platinum-pre-treated patients


The immune checkpoint inhibitors pembrolizumab, nivolumab, atezolizumab, avelumab, and durvalumab have
demonstrated similar efficacy and safety in patients progressing during, or after, standard platinum-based
chemotherapy in phase I, II and III trials.

Pembrolizumab, a PD-1 inhibitor, has been tested in patients progressing during or after platinum-based first-
line chemotherapy in a phase III RCT and demonstrated a significant OS benefit leading to approval. In the trial,
patients (n = 542) were randomised to receive either pembrolizumab monotherapy or chemotherapy (paclitaxel,
docetaxel or vinflunine). The median OS in the pembrolizumab arm was 10.3 months (95% CI: 8.0–11.8) vs.
7.4 months (95% CI: 6.1–8.3) for the chemotherapy arm (HR for death, 0.73, 95% CI: 0.59–0.91, p = 0.002)
independent of PD-L1 expression levels [544]. This trial was recently updated with a longer follow-up of 27.7
months showing consistent improvement of OS [545]. In addition, HRQoL analysis showed that patients on
pembrolizumab experience stable, or improved, HRQoL, whereas it deteriorated on chemotherapy [546].

Atezolizumab, a PD-L1 inhibitor, tested in patients progressing during, or after, previous platinum-based
chemotherapy in phase I, phase II and phase III trials, was the first checkpoint inhibitor approved for BC [204,
547, 548]. The phase III RCT (IMvigor211) included 931 patients comparing atezolizumab with second-line
chemotherapy (either paclitaxel, docetaxel or vinflunine) did not meet its primary endpoint of improved OS for
patients with high PD-L1 expression (immune cells [IC] score 2/3) with 11.1 months (atezolizumab) vs. 10. 6
(chemotherapy) months (stratified HR: 0.87, 95% CI: 0.63–1.21, p = 0.41) but OS was numerically improved
in the ITT population in an exploratory analysis (8.6 months vs. 8.0 months, HR: 0.85, 95% CI: 0.73–0.99). A
phase IV single-arm safety study was conducted with atezolizumab including 1,004 patients confirming the
efficacy and tolerability profile [549].

The PD-1 inhibitor nivolumab was approved based on the results of a single-arm phase II trial (CheckMate
275), enrolling 270 platinum pre-treated patients. The first endpoint was ORR. Objective response rate was
19.6%, and OS was 8.74 months for the entire group [550]. Based on results of phase I/II and phase IB trials,
two additional PD-L1 inhibitors, durvalumab and avelumab, are currently only approved for this indication in the
U.S. [551-553].

7.7.3.2.1 Side-effect profile of immunotherapy


Checkpoint inhibitors including PD-1 or PD-L1 antibodies and CTLA-4 antibodies have a distinct side effect
profile associated with their mechanism of action leading to enhanced immune system activity. These adverse
events can affect any organ in the body leading to mild, moderate or severe side effects. The most common
organs affected are the skin, gastrointestinal tract, liver, lung, thyroid, adrenal and pituitary gland. Other
systems that may be affected include musculoskeletal, renal, nervous, haematologic, ocular and cardiovascular
system. Any change during immunotherapy treatment should raise suspicion about a possible relation to the
treatment. The nature of immune-related adverse events has been very well characterised and published [554].
The timely and appropriate treatment of immune-related side effects is crucial to achieve optimal
benefit from the treatment while maintaining safety. Clear guidelines for side effect management have been
published [555]. Immunotherapy treatment should be applied and supervised by trained clinicians only to
ensure early side effect recognition and treatment.
In case of interruption of immunotherapy, re-challenge will require close monitoring for adverse
events [556].

7.7.4 Novel agents for second- or later-line therapy


Genomic profiling of urothelial carcinoma has revealed common potentially actionable genomic alterations
including alterations in FGFR [557]. Erdafitinib is a pan-FGFR tyrosine kinase inhibitor and the first FDA-
approved targeted therapy for metastatic urothelial carcinoma with susceptible FGFR2/3 alterations following
platinum-containing chemotherapy. The phase II trial of erdafitinib included 99 patients whose tumour
harboured an FGFR3 mutation or FGFR2/3 fusion and who had disease progression following chemotherapy
[202]. The confirmed ORR was 40% and an additional 39% of patients had stable disease. A total of 22
patients had previously received immunotherapy with only one patient achieving a response, yet the response

52 LIMITED UPDATE MARCH 2021


rate for erdafitinib for this subgroup was 59%. At a median follow-up of 24 months, the median PFS was
5.5 months (95% CI: 4.0–6.0) and the median OS was 11.3 months (95% CI: 9.7–15.2) [202]. Treatment-
related adverse events of ≥ grade 3 occurred in 46% of patients. Common adverse events of ≥ grade 3 were
hyponatraemia (11%), stomatitis (10%), and asthenia (7%) and 13 patients discontinued erdafitinib due to
adverse events, including retinal pigment epithelial detachment, hand-foot syndrome, dry mouth, and skin/nail
events. In addition to erdafitinib, several other FGFR inhibitors are being evaluated including infigratinib which
has demonstrated promising activity [203]. The increased identification of FGFR3 mutations/fusion in UTUCs
and in NMIBC has led to several ongoing trials.

Another promising drug is enfortumab vedotin, an antibody-drug conjugate (ADC) targeting Nectin-4, a
cell adhesion molecule which is highly expressed in UC conjugated to monomethyl auristatin E (MMAE). A
published phase-II single-arm study (n = 125) in patients previously treated with platinum chemotherapy and
checkpoint inhibition showed a confirmed objective response rate of 44%, including 12% complete responses
[558]. Responses were seen across patient subgroups including 41% in checkpoint inhibitor non-responders
and 38% in patients with liver metastases. The most common treatment-related AEs included fatigue (50%),
alopecia (48%), and decreased appetite (41%). Treatment-related AEs of interest included any rash (48% all
grade, 11% ≥ G3) and any peripheral neuropathy (50% all grade, 3% ≥ G3). This data led to the accelerated
FDA approval for enfortumab vedotin in locally advanced or metastatic UC patients who have previously
received a PD-1 or PD-L1 inhibitor, and platinum-containing chemotherapy in the neoadjuvant/adjuvant,
locally advanced or metastatic setting [559]. A phase III RCT comparing enfortumab vedotin with single-agent
chemotherapy has reported preliminary results (ASCO GU 2021), reporting a significant survival benefit [558].
Another ongoing trial has reported promising activity for the combination of enfortumab vedotin in combination
with pembrolizumab as first-line treatment in cisplatin-ineligible patients with locally advanced/metastatic UC
(ORR: 73.3% with 15.6% complete responses) [560]. Another promising ADC is sacituzumab govitecan, targeting
trophoblast cell surface antigen 2 (Trop-2) conjugated to SN-38, the active metabolite of irinotecan [561].

7.7.5 Post-chemotherapy surgery and oligometastatic disease


With cisplatin-containing combination chemotherapy, excellent response rates may be obtained in patients
with LN metastases only, good PS and adequate renal function, including a high number of complete
responses with up to 20% of patients achieving long-term DFS [512, 516, 562, 563]. The role of surgery of
residual LNs after chemotherapy is still unclear. Although some studies suggest a survival benefit and QoL
improvement, the level of evidence supporting this practice is mainly anecdotal [564-578]. Retrospective
studies of post-chemotherapy surgery after partial or complete remission have indicated that surgery may
contribute to long-term DFS in selected patients [579-582]. These findings have been confirmed in a recent
systematic review including 28 studies [582].
In the absence of data from RCTs, patients should be evaluated on an individual basis and
discussed by an interdisciplinary tumour board [582].

7.7.6 Treatment of patients with bone metastases


The prevalence of metastatic bone disease (MBD) in patients with advanced/metastatic UC is 30–40% [583].
Skeletal complications due to MBD have a detrimental effect on pain and QoL and are also associated
with increased mortality [584]. Bisphosphonates such as zoledronic acid reduce and delay skeletal-related
events (SREs) due to bone metastases by inhibiting bone resorption, as shown in a small pilot study [585].
Denosumab, a fully human monoclonal antibody that binds to and neutralises RANKL (receptor activator
of nuclear factor κB ligand), was shown to be non-inferior to zoledronic acid in preventing or delaying SREs
in patients with solid tumours and advanced MBD, including patients with UC [586]. Patients with MBD,
irrespective of the cancer type, should be considered for bone-targeted treatment [584].

Patients treated with zoledronic acid or denosumab should be informed about possible side effects including
osteonecrosis of the jaw and hypocalcaemia. Supplementation with calcium and vitamin D is mandatory.
Dosing regimens of zoledronic acid should follow regulatory recommendations and have to be adjusted
according to pre-existing medical conditions, especially renal function [587]. For denosumab, no dose
adjustments are required for variations in renal function.

LIMITED UPDATE MARCH 2021 53


7.7.7 Summary of evidence and guidelines for metastatic disease

Summary of evidence LE
In a first-line setting, performance status (PS) and the presence or absence of visceral metastases are 1b
independent prognostic factors for survival.
In a second-line setting, negative prognostic factors are: liver metastasis, PS ≥ 1 and low haemoglobin 1b
(< 10 g/dL).
Cisplatin-containing combination chemotherapy can achieve median survival of up to 14 months, with 1b
long-term disease-free survival (DFS) reported in ~15% of patients with nodal disease and good PS.
Single-agent chemotherapy provides low response rates of usually short duration. 2a
Carboplatin combination chemotherapy is less effective than cisplatin-based chemotherapy in terms 2a
of complete response and survival.
Non-platinum combination chemotherapy has not been tested against standard chemotherapy in 4
patients who are fit or unfit for cisplatin-combination chemotherapy.
There is no defined standard therapy for platinum chemotherapy-unfit patients with advanced or 2b
metastatic urothelial cancer (UC).
Post-chemotherapy surgery after partial or complete response may contribute to long-term DFS in 3
selected patients.
Zoledronic acid and denosumab have been approved for supportive treatment in case of bone 1b
metastases of all cancer types including UC, as they reduce and delay skeletal related events.
PD-1 inhibitor pembrolizumab has been approved for patients that have progressed during or after 1b
previous platinum-based chemotherapy based on the results of a phase III trial.
PD-L1 inhibitors atezolizumab, nivolumab, durvalumab and avelumab have been FDA approved for 2a
patients that have progressed during or after previous platinum-based chemotherapy based on the
results of a phase II trial.
PD-1 inhibitor pembrolizumab and PD-L1 inhibitor atezolizumab have been approved for patients 2a
with advanced or metastatic UC unfit for cisplatinum-based first-line chemotherapy and with
overexpression of PD-L1 based on the results of single-arm phase II trials.
The combination of chemotherapy plus pembrolizumab or atezolizumab and the combination of 1b
durvalumab and tremelimumab have not demonstrated OS survival benefit compared to platinum-
based chemotherapy alone.
Switch maintenance with the PD-L1 inhibitor avelumab has demonstrated significant OS benefit in 1b
patients achieving at least stable disease on first-line platinum-based chemotherapy.

Recommendations Strength rating


First-line treatment for platinum-fit patients
Use cisplatin-containing combination chemotherapy with GC or HD-MVAC. Strong
In patients unfit for cisplatin but fit for carboplatin use the combination of carboplatin and Strong
gemcitabine.
In patients achieving stable disease, or better, after first-line platinum-based chemotherapy Strong
use maintenance treatment with PD-L1 inhibitor avelumab.
First-line treatment in patients unfit for platinum-based chemotherapy
Consider checkpoint inhibitors pembrolizumab or atezolizumab. Weak
Second-line treatment
Offer checkpoint inhibitor pembrolizumab to patients progressing during, or after, platinum- Strong
based combination chemotherapy for metastatic disease. If this is not possible, offer
atezolizumab, nivolumab (EMA, FDA approved); avelumab or durvalumab (FDA approved).
Further treatment after platinum- and immunotherapy
Offer treatment in clinical trials testing novel antibody drug conjugates (enfortumab vedotin, Strong
sacituzumab govitecan); or in case of patients with FGFR3 alterations, FGFR tyrosine kinase
inhibitors.
GC = gemcitabine plus cisplatin; FGFR = fibroblast growth factor receptor; HD-MVAC = high-dose intensity
methotrexate, vinblastine, adriamycin plus cisplatin.

54 LIMITED UPDATE MARCH 2021


Figure 7.2: Flow chart for the management of metastatic urothelial cancer*

PLATINUM-ELIGIBLE PLATINUM-INELIGIBLE
PS 2 and GFR < 60 mL/min
PS > 2; GFR < 30 mL/min

cisplatin carboplatin
PS 0-1 and PS 2 or GFR 30-60
GFR > 50-60 mL/min mL/min

cisplatin/gemcitabine carboplatin/gemcitabine
or DD-MVAC 4-6 cycles 4-6 cycles
PD-L1 + PD-L1 -

PD CR/PR/SD

PD Watchful
2nd line therapy Maintenance
waiting

• pembrolizumab
Switch Immunotherapy Best
(atezolizumab, avelumab,
maintenance: • atezolizumab supportive
durvalumab, nivolumab)
avelumab • pembrolizumab care
• Trials

LATER-LINE THERAPY

FGFR3 mutation UC progressing on UC patients refractory to platinum-based


platinum-based chemotherapy ± prior IO chemotherapy and IO

• erdafitinib (FDA) – in trial • enfortumab vedotin (FDA) – in trial


• chemotherapy: • chemotherapy:
o paclitaxel o paclitaxel
o Docetaxel o docetaxel
o vinflunine o vinflunine
• Trials • Trials

*Treatment within clinical trials is highly encouraged.


BSC = best supportive care; CR = complete response; DD-MVAC = dose dense methotrexate vinblastine
doxorubicin cisplatin; EV = enfortumab vedotin; FDA = US Food and Drug Administration;
FGFR = pan-fibroblast growth factor receptor tyrosine kinase inhibitor; GFR = glomerular filtration rate;
IO = immunotherapy; PR = partial response; PS = performance status; SD = stable disease.

7.8 Quality of life


7.8.1 Introduction
The evaluation of HRQoL considers physical, psychological, emotional and social functioning. In patients with
MIBC, HRQoL declines in particular in the physical and social functioning domains [588]. Several questionnaires
have been validated for assessing HRQoL in patients with BC, including FACT (Functional Assessment of
Cancer Therapy)-G [589], EORTC QLQ-C30 [590], EORTC QLQ-BLM (MIBC module) [591], and SF (Short Form)-
36 [592, 593] and recently the BCI questionnaire specifically designed and validated for BC patients [594].

7.8.2 Neoadjuvant chemotherapy


The impact of NAC on patient-reported outcomes (using EORTC QLQ questionnaires) was investigated by
Feuerstein et al. [595]. A propensity-matched analysis of 101 patients who completed NAC and 54 patients
who did not undergo NAC, showed no negative effect of NAC on patient-reported outcomes. Similar results
were reported by Huddart et al. based on data from the BC2001 trial [466].

LIMITED UPDATE MARCH 2021 55


7.8.3 Radical cystectomy and urinary diversion
Two systematic reviews and meta-analyses focused on HRQoL after RC and urinary diversion [351, 596].
Yang et al. compared HRQoL of incontinent and continent urinary diversions (all types) including
29 studies (n = 3,754) of which 9 had a prospective design (one of which was randomised) [351]. Only three
studies reported HRQoL data both pre- and post-operatively. In these three studies, an initial deterioration in
overall HRQoL was reported but general health, functional and emotional domains at 12 months post-surgery
were equal or better than baseline. After 12 months, the HRQoL benefits diminished in all domains. Overall,
no difference in HRQoL between continent and incontinent urinary diversion was reported although an ileal
conduit may confer a small physical health benefit [596].
Cerruto et al. reported HRQoL comparing ileal conduit with orthotopic neobladder reconstruction
[596]. A pooled analysis was performed including 18 studies (n = 1,553) of which the vast majority
were retrospective studies. The analysis showed no statistical significant difference in overall HRQoL, but
methodological limitations need to be considered.

Clifford et al. prospectively evaluated continence outcomes in male patients undergoing orthotopic neobladder
diversion [597]. Day-time continence increased from 59% at less than three months post-operatively to 92%
after 12 to 18 months. Night-time continence increased from 28% at less than three months post-operatively
to 51% after 18 to 36 months. Also of interest is the urinary bother in females with an orthotopic neobladder.
Bartsch and co-workers reported day-time and night-time continence rates of 70.4% and 64.8%, respectively,
in 56 female neobladder patients. Thirty-five patients (62.5%) performed clean intermittent catheterisation,
which is much worse when compared to male neobladder patients. Moreover, patients with non-organ-
confined disease (p = 0.04) and patients with a college degree (p = 0.001) showed worse outcomes on HRQoL
scores [598].

Altogether, there is no superior type of urinary diversion in terms of overall HRQoL in unselected patients.
Health-related QoL outcomes are most likely a result of good patient selection. An older, more isolated, patient
is probably better served with an ileal conduit, whereas a younger patient with a likely higher level of interest
in body image and sexuality is better off with an orthotopic diversion. The patient’s choice is the key to the
selection of reconstruction method [351].

7.8.4 Bladder sparing trimodality therapy


The only HRQoL data in bladder sparing treatment collected in an RCT setting was published by Huddart et al.
[466]. The primary endpoint was the change in the Bladder Cancer Subscale (BLCS), as part of the FACT-BL
questionnaire, at one year post-treatment. Questionnaire return rate at one and five years was 70% and 60%,
respectively. The remaining patients did mostly not respond as a result of recurrence or RC. The data show a
reduction in HRQoL in the majority of the domains immediately following radiotherapy but in most patients the
HRQoL scores returned to baseline 6 months after RT and maintained at this level for five years. Approximately
33% of patients reported persistent lower Bladder Cancer Subscale scores after five years. Addition of
chemotherapy did not affect the HRQoL outcomes.

7.8.5 Non-curative or metastatic bladder cancer


In non-curative or metastatic BC, HRQoL is reduced because of associated micturition problems, bleeding,
pain and therefore disturbance of social and sexual life [599]. There is limited literature describing HRQoL in BC
patients receiving palliative care [600] but there are reports of bladder-related symptoms relieved by palliative
surgery [423], RT [601], and/or chemotherapy [602].

A HRQoL analysis was performed in platinum-refractory patients who were randomised to pembrolizumab
vs. another line of chemotherapy (KEYNOTE-45 trial) [546]. It was reported that patients treated with
pembrolizumab had stable or improved global health status/QoL, whereas those treated with investigators’
choice of chemotherapy experienced declines in global health [546].

7.8.6 Summary of evidence and recommendations for health-related quality of life

Summary of evidence LE
Compared to non-cancer controls, the diagnosis and treatment of bladder cancer has a negative 2a
impact on HRQoL.
There is no significant difference in overall QoL between patients with continent or incontinent 2a
diversion.
In patients with MIBC treated with RC, overall HRQoL declines immediately after treatment and 1a
recovers to baseline at 12 months post-operatively.

56 LIMITED UPDATE MARCH 2021


In patients with MIBC treated with radiotherapy, overall HRQoL declines immediately after treatment. 1b
In most patients, overall HRQoL then recovers to baseline at 6 months and maintains at this level to 5
years.
In patients with MIBC treated with radiotherapy, concomitant chemotherapy or neo-adjuvant 1b
chemotherapy has no significant impact on HRQoL.
In patients with platinum-refractory advanced urothelial carcinoma, pembrolizumab may be superior in 1b
terms of HRQoL compared to another line of chemotherapy.

Recommendations Strength rating


Use validated questionnaires to assess health-related quality of life in patients with muscle- Strong
invasive bladder cancer.
Discuss the type of urinary diversion taking into account a patient preference, existing Strong
comorbidities, tumour variables and coping abilities.

8. FOLLOW-UP
8.1 Follow-up in muscle invasive bladder cancer
An appropriate schedule for disease monitoring should be based on natural timing of recurrence; probability
and site of recurrence; functional monitoring after urinary diversion and the potential available management
options [603].
Nomograms on CSS following RC have been developed and externally validated, but their wider
use cannot be recommended until further data become available [604, 605].
Current surveillance protocols are based on patterns of recurrence drawn from retrospective series
only. Combining this data is not possible since most retrospective studies use different follow-up regimens and
imaging techniques. Additionally, reports of asymptomatic recurrences diagnosed during routine oncological
follow-up and results from retrospective studies are contradictory [606-608]. From the Volkmer B, et al.
series of 1,270 RC patients, no differences in OS were observed between asymptomatic and symptomatic
recurrences [607]. Conversely, in the Giannarini, et al. series of 479 patients; those with recurrences detected
during routine follow-up (especially in the lungs) and with secondary urothelial tumours as the site of
recurrence, had a slightly higher survival [606]. Boorjian, et al. included 1,599 RC patients in their series, with
77% symptomatic recurrences. On multivariate analysis, patients who were symptomatic at recurrence had a
60% increased risk of death as compared to asymptomatic patients [608].
However, at this time, no data from prospective trials demonstrating the potential benefit of early
detection of recurrent disease and its impact on OS are available [609]. For details see Section 7.5.4.

8.2 Site of recurrence


8.2.1 Local recurrence
Local recurrence takes place in the soft tissues of the original surgical site or in LNs. Contemporary cystectomy
has a 5–15% probability of pelvic recurrence which usually occurs during the first 24 months, most often
within 6 to 18 months after surgery. However, late recurrences can occur up to five years after RC. Risk factors
described are pathological stage, LNs, positive margins, extent of LND and peri-operative chemotherapy [610].
Patients generally have a poor prognosis after pelvic recurrence. Even with treatment, median
survival ranges from four to eight months following diagnosis. Definitive therapy can prolong survival, but
mostly provides significant palliation of symptoms. Trimodality management generally involves a combination
of chemotherapy, radiation and surgery [609].

8.2.2 Distant recurrence


Distant recurrence is seen in up to 50% of patients treated with RC for MIBC. As with local recurrence,
pathological stage and nodal involvement are risk factors [611]. Systemic recurrence is more common in locally
advanced disease (pT3/4), ranging from 32 to 62%, and in patients with LN involvement (range 52–70%) [612].
The most likely sites for distant recurrence are LNs, lungs, liver and bone. Nearly 90% of distant
recurrences appear within the first three years after RC, mainly in the first two years, although late recurrence
has been described after more than 10 years. Median survival of patients with progressive disease treated
with platinum-based chemotherapy is 9–26 months [613-615]. However, longer survival (28–33% at 5 years)
has been reported in patients with minimal metastatic disease undergoing trimodality management, including
metastasectomy [565, 573].

LIMITED UPDATE MARCH 2021 57


8.2.3 Urothelial recurrences
After RC, the incidence of new urethral tumours was 4.4% (1.3–13.7%). Risk factors for secondary urethral
tumours are urethral malignancy in the prostatic urethra/prostate (in men) and bladder neck (in women).
Orthotopic neobladder was associated with a significant lower risk of urethral tumours after RC (OR: 0.44)
[616].

There is limited data, and agreement, about urethral follow-up, with some authors recommending routine
surveillance with urethral wash and urine cytology and others doubting the need for routine urethral
surveillance. However, there is a significant survival advantage in men with urethral recurrence diagnosed
asymptomatically vs. symptomatically, so follow-up of the male urethra is indicated in patients at risk of urethral
recurrence [609]. Treatment is influenced by local stage and grade of urethral occurrence. In urethral CIS, BCG
instillations have success rates of 83% [617]. In invasive disease, urethrectomy should be performed if the
urethra is the only site of disease; in case of distant disease, systemic chemotherapy is indicated [3].

Upper urinary tract UCs occur in 4–10% of cases and represent the most common sites of late recurrence
(3-year DFS following RC) [618]. Median OS is 10–55 months, and 60–67% of patients die of metastatic
disease [609]. A meta-analysis found that 38% of UTUC recurrence was diagnosed by follow-up investigations,
whereas in the remaining 62%, diagnosis was based on symptoms. When urine cytology was used during
surveillance, the rate of primary detection was 7% vs. 29.6% with UUT imaging. The meta-analysis concluded
that patients with non-invasive cancer are twice as likely to have UTUC as patients with invasive disease [619].
Multifocality increases the risk of recurrence by three-fold, while positive ureteral or urethral margins increase
the risk by seven-fold. Radical nephroureterectomy can prolong survival [620].

8.3 Time schedule for surveillance


Although, based on low level evidence only, some follow-up schedules have been suggested, guided by the
principle that recurrences tend to occur within the first years following initial treatment. A schedule suggested
by the EAU Guidelines Panel includes a CT scan (every 6 months) until the third year, followed by annual
imaging thereafter. Patients with multifocal disease, NMIBC with CIS or positive ureteral margins are at
higher risk of developing UTUC, which can develop late (> 3 years). In those cases, monitoring of the UUT is
mandatory during follow-up. Computed tomography is to be used for imaging of the UUT [619].

The exact time to stop follow-up is not well known and recently a risk-adapted schedule has been proposed,
based on the interaction between recurrence risk and competing health factors that could lead to individualised
recommendations and may increase recurrence detection. Elderly and very low-risk patients (those with NMIBC
or pT0 disease at final cystectomy report) showed a higher competing risk of non-BC mortality when compared
with their level of BC recurrence risk. On the other hand, patients with locally advanced disease or LN
involvement are at a higher risk of recurrence for more than 20 years [621]. However, this model has not been
validated and does not incorporate several risk factors related to non-BC mortality. Furthermore, the prognostic
implications of the different sites of recurrence should be considered. Local and systemic recurrences have a
poor prognosis and early detection of the disease will not influence survival [622]. Despite this, the rationale for
a risk-adapted schedule for BC surveillance appears to be promising and deserves further investigation.
Since data for follow-up strategies are sparse, a number of key questions were included in a
recently held consensus project [8, 9]. Outcomes for all statements for which consensus was achieved are
listed in Section 8.6.

8.4 Follow-up of functional outcomes and complications


Apart from oncological surveillance, patients with a urinary diversion need functional follow-up. Complications
related to urinary diversion are detected in 45% of patients during the first five years of follow-up. This rate
increases over time, and exceeds 54% after 15 years of follow-up. In a single-centre series of 259 male
patients, long-term follow-up after orthotopic bladder substitution (median 121 months [range 60–267]),
showed that excellent long-term functional outcomes can be achieved in high-volume centres with dedicated
teams [623]. A smaller multi-centre series including women only (n = 102) showed complication rates between
5–12% after orthotopic neobladder (median follow up of 24 months [range 1.5–100 months]). Both early (5%)
and late (12%) complications related to the urinary diversion [624].

The functional complications are diverse and include: vitamin B12 deficiency, metabolic acidosis, worsening of
renal function, urinary infections, urolithiasis, stenosis of uretero-intestinal anastomosis, stoma complications
in patients with ileal conduit, neobladder continence problems, and emptying dysfunction [609]. Functional
complications are especially common in women: approximately two-thirds need to catheterise their
neobladder, while almost 45% do not void spontaneously at all [598]. There seems to be a correlation between

58 LIMITED UPDATE MARCH 2021


voiding patterns and nerve preservation; in 66 women bilateral preservation of autonomic nerves decreased the
need for catheterisation to between 3.4–18.7% (CI: 95%) [624].

Recently a 21% increased risk of fractures was also described as compared to no RC due to chronic metabolic
acidosis and subsequent long-term bone loss [622].
Since low vitamin B12 levels have been reported in 17% of patients with bowel diversion, in case of
cystectomy and bowel diversion, vitamin B12 levels should be measured annually [8, 9, 378].

8.5 Summary of evidence and recommendations for specific recurrence sites

Site of recurrence Summary of evidence Recommendation Strength rating


Local recurrence Poor prognosis. Offer radiotherapy, chemotherapy Strong
Treatment should be and possibly surgery as options
individualised depending on for treatment, either alone or in
the local extent of tumour. combination.
Distant recurrence Poor prognosis. Offer chemotherapy as the first Strong
option, and consider metastasectomy
or radiotherapy in case of unique
metastasis site.
Upper urinary tract Risk factors are multifocal See EAU Guidelines on Upper Urinary Strong
recurrence disease (NMIBC/CIS or Tract Urothelial Carcinomas.
positive ureteral margins).
Secondary urethral Staging and treatment should See EAU Guidelines on Primary Strong
tumour be done as for primary Urethral Carcinoma.
urethral tumour.

8.6 EAU-ESMO consensus statements on the management of advanced- and variant


bladder cancer [8, 9]*

Consensus statement
After radical cystectomy with curative intent, regular follow-up is needed.
After radical cystectomy with curative intent, follow-up for the detection of second cancers in the urothelium
is recommended.
After radical cystectomy with curative intent, follow-up of the urethra with cytology and/or cystoscopy is
recommended in selected patients (e.g., multifocality, carcinoma in situ and tumour in the prostatic urethra).
After trimodality treatment with curative intent, follow-up for the detection of relapse is recommended every
3–4 mos initially; then after 3 yrs, every 6 mos in the majority of patients.
After trimodality treatment with curative intent, regular follow-up for the detection of relapse is needed in the
majority of patients.
After trimodality treatment with curative intent, follow-up imaging to assess distant recurrence or recurrence
outside the bladder is needed.
After trimodality treatment with curative intent, assessment of the urothelium to detect recurrence is
recommended every 6 mos in the majority of patients.
After trimodality treatment with curative intent, in addition to a CT scan, other investigations of the bladder are
recommended.
In patients with a partial or complete response after chemotherapy for metastatic urothelial cancer, regular
follow-up is needed. Imaging studies may be done according to signs/symptoms.
To detect relapse (outside the bladder) after trimodality treatment with curative intent, CT of the thorax and
abdomen is recommended as the imaging method for follow-up in the majority of patients.
To detect relapse (outside the bladder) after trimodality treatment with curative intent, routine imaging with CT
of the thorax and abdomen should be stopped after 5 yrs in the majority of patients.
In patients treated with radical cystectomy with curative intent and who have a neobladder, management of
acid bases household includes regular measurements of pH and sodium bicarbonate substitution according
to the measured value.
To detect relapse after radical cystectomy with curative intent, routine imaging with CT of the thorax and
abdomen should be stopped after 5 yrs in the majority of patients.

LIMITED UPDATE MARCH 2021 59


To detect relapse after radical cystectomy with curative intent, a CT of the thorax and abdomen is
recommended as the imaging method for follow-up in the majority of patients.
Levels of LDH and CEA are not essential in the follow-up of patients with urothelial cancer to detect
recurrence.
Vitamin B12 levels have to be measured annually in the follow-up of patients treated with radical cystectomy
and bowel diversion with curative intent.
*Only statements which met the a priori consensus threshold across all three stakeholder groups are listed
(defined as ≥ 70% agreement and ≤ 15% disagreement, or vice versa).
CEA = carcinoembryonic antigen; CT = computed tomography; LDH = lactate dehydrogenase; mos = months;
yrs = years.

9. REFERENCES
1. Rouprêt, M., et al., Guidelines on Upper Urinary Tract Urothelial Cell Carcinoma. Edn. presented at
the 36th EAU Congress Milan 2021. EAU Guidelines Office, Arnhem, The Netherlands.
https://uroweb.org/guideline/upper-urinary-tract-urothelial-cell-carcinoma
2. Babjuk, M., et al., Guidelines on Non-muscle-invasive bladder cancer (Ta, T1 and CIS). Edn.
presented at the 36th EAU Congress Milan 2021. EAU Guidelines Office, Arnhem, The Netherlands.
https://uroweb.org/guideline/non-muscle-invasive-bladder-cancer
3. Gakis, G., et al., Guidelines on Primary Urethral Carcinoma. Edn. presented at the 36th EAU
Congress Milan 2021. EAU Guidelines Office, Arnhem, The Netherlands.
https://uroweb.org/guideline/primary-urethral-carcinoma
4. Witjes, J.A., et al. European Association of Urology Guidelines on Muscle-invasive and Metastatic
Bladder Cancer: Summary of the 2020 Guidelines. Eur Urol, 2021. 79: 82.
https://uroweb.org/guideline/bladder-cancer-muscle-invasive-and-metastatic
5. Guyatt, G.H., et al. What is “quality of evidence” and why is it important to clinicians? BMJ, 2008.
336: 995.
https://pubmed.ncbi.nlm.nih.gov/18456631
6. Guyatt, G.H., et al. GRADE: an emerging consensus on rating quality of evidence and strength of
recommendations. BMJ, 2008. 336: 924.
https://pubmed.ncbi.nlm.nih.gov/18436948
7. Phillips B, et al. Oxford Centre for Evidence-based Medicine Levels of Evidence 1998. Updated by
Jeremy Howick March 2009.
https://www.cebm.net/2009/06/oxford-centre-evidence-based-medicine-levels-evidence-march-2009/
8. Horwich, A., et al. EAU-ESMO consensus statements on the management of advanced and variant
bladder cancer-an international collaborative multi-stakeholder effort: under the auspices of the EAU
and ESMO Guidelines Committees. Ann Oncol, 2019. 30: 1697.
https://pubmed.ncbi.nlm.nih.gov/31740927
9. Witjes, J.A., et al. EAU-ESMO Consensus Statements on the Management of Advanced and Variant
Bladder Cancer-An International Collaborative Multistakeholder Effort: Under the Auspices of the
EAU-ESMO Guidelines Committees. Eur Urol, 2020. 77: 223.
https://pubmed.ncbi.nlm.nih.gov/31753752
10. IARC, Cancer Today. Estimated number of new cases in 2020, worldwide, both sexes, all ages.
2021 [access date March 2021].
https://gco.iarc.fr/today/online-analysis-table
11. Burger, M., et al. Epidemiology and risk factors of urothelial bladder cancer. Eur Urol, 2013. 63: 234.
https://pubmed.ncbi.nlm.nih.gov/22877502
12. Bosetti, C., et al. Trends in mortality from urologic cancers in Europe, 1970-2008. Eur Urol, 2011. 60: 1.
https://pubmed.ncbi.nlm.nih.gov/21497988
13. Chavan, S., et al. International variations in bladder cancer incidence and mortality. Eur Urol, 2014.
66: 59.
https://pubmed.ncbi.nlm.nih.gov/24451595
14. Comperat, E., et al. Clinicopathological characteristics of urothelial bladder cancer in patients less
than 40 years old. Virchows Arch, 2015. 466: 589.
https://pubmed.ncbi.nlm.nih.gov/25697540

60 LIMITED UPDATE MARCH 2021


15. Freedman, N.D., et al. Association between smoking and risk of bladder cancer among men and
women. JAMA, 2011. 306: 737.
https://pubmed.ncbi.nlm.nih.gov/21846855
16. Tobacco smoke and involuntary smoking. IARC Monogr Eval Carcinog Risks Hum, 2004. 83: 1.
https://pubmed.ncbi.nlm.nih.gov/15285078
17. Brennan, P., et al. Cigarette smoking and bladder cancer in men: a pooled analysis of 11 case-
control studies. Int J Cancer, 2000. 86: 289.
https://pubmed.ncbi.nlm.nih.gov/10738259
18. Gandini, S., et al. Tobacco smoking and cancer: a meta-analysis. Int J Cancer, 2008. 122: 155.
https://pubmed.ncbi.nlm.nih.gov/17893872
19. Al Hussein Al Awamlh, B., et al. Association of Smoking and Death from Genitourinary Malignancies:
Analysis of the National Longitudinal Mortality Study. J Urol, 2019. 202: 1248.
https://pubmed.ncbi.nlm.nih.gov/31290707
20. Pashos, C.L., et al. Bladder cancer: epidemiology, diagnosis, and management. Cancer Pract, 2002.
10: 311.
https://pubmed.ncbi.nlm.nih.gov/12406054
21. Harling, M., et al. Bladder cancer among hairdressers: a meta-analysis. Occup Environ Med, 2010.
67: 351.
https://pubmed.ncbi.nlm.nih.gov/20447989
22. Weistenhofer, W., et al. N-acetyltransferase-2 and medical history in bladder cancer cases with a
suspected occupational disease (BK 1301) in Germany. J Toxicol Environ Health A, 2008. 71: 906.
https://pubmed.ncbi.nlm.nih.gov/18569594
23. Rushton, L., et al. Occupation and cancer in Britain. Br J Cancer, 2010. 102: 1428.
https://pubmed.ncbi.nlm.nih.gov/20424618
24. Chrouser, K., et al. Bladder cancer risk following primary and adjuvant external beam radiation for
prostate cancer. J Urol, 2005. 174: 107.
https://pubmed.ncbi.nlm.nih.gov/18405759
25. Nieder, A.M., et al. Radiation therapy for prostate cancer increases subsequent risk of bladder and
rectal cancer: a population based cohort study. J Urol, 2008. 180: 2005.
https://pubmed.ncbi.nlm.nih.gov/18801517
26. Zelefsky, M.J., et al. Incidence of secondary cancer development after high-dose intensity-
modulated radiotherapy and image-guided brachytherapy for the treatment of localized prostate
cancer. Int J Radiat Oncol Biol Phys, 2012. 83: 953.
https://pubmed.ncbi.nlm.nih.gov/22172904
27. Zamora-Ros, R., et al. Flavonoid and lignan intake in relation to bladder cancer risk in the European
Prospective Investigation into Cancer and Nutrition (EPIC) study. Br J Cancer, 2014. 111: 1870.
https://pubmed.ncbi.nlm.nih.gov/25121955
28. Teleka, S., et al. Risk of bladder cancer by disease severity in relation to metabolic factors and smoking:
A prospective pooled cohort study of 800,000 men and women. Int J Cancer, 2018. 143: 3071.
https://pubmed.ncbi.nlm.nih.gov/29756343
29. Xu, Y. Diabetes mellitus and the risk of bladder cancer: A PRISMA-compliant meta-analysis of
cohort studies. Medicine (Baltimore). 2017, 96: e8588.
https://pubmed.ncbi.nlm.nih.gov/29145273
30. Adil, M., et al. Pioglitazone and risk of bladder cancer in type 2 diabetes mellitus patients: A
systematic literature review and meta-analysis of observational studies using real-world data.
Clinical Epidemiology and Global Health, 2018. 6: 61.
https://www.sciencedirect.com/science/article/abs/pii/S2213398417300544
31. Schistosomes, liver flukes and Helicobacter pylori. IARC Working Group on the Evaluation of
Carcinogenic Risks to Humans. Lyon, 7-14 June 1994. IARC Monogr Eval Carcinog Risks Hum,
1994. 61: 1.
https://pubmed.ncbi.nlm.nih.gov/7715068
32. Gouda, I., et al. Bilharziasis and bladder cancer: a time trend analysis of 9843 patients. J Egypt Natl
Canc Inst, 2007. 19: 158.
https://pubmed.ncbi.nlm.nih.gov/19034337
33. Salem, H.K., et al. Changing patterns (age, incidence, and pathologic types) of schistosoma-
associated bladder cancer in Egypt in the past decade. Urology, 2012. 79: 379.
https://pubmed.ncbi.nlm.nih.gov/22112287
34. Pelucchi, C., et al. Mechanisms of disease: The epidemiology of bladder cancer. Nat Clin Pract Urol,
2006. 3: 327.
https://pubmed.ncbi.nlm.nih.gov/16763645

LIMITED UPDATE MARCH 2021 61


35. Bayne, C.E., et al. Role of urinary tract infection in bladder cancer: a systematic review and meta-
analysis. World J Urol, 2018. 36: 1181.
https://pubmed.ncbi.nlm.nih.gov/29520590
36. Yu, Z., et al. The risk of bladder cancer in patients with urinary calculi: a meta-analysis. Urolithiasis,
2018. 46: 573.
https://pubmed.ncbi.nlm.nih.gov/29305631
37. Liu, S., et al. The impact of female gender on bladder cancer-specific death risk after radical
cystectomy: a meta-analysis of 27,912 patients. Int Urol Nephrol, 2015. 47: 951.
https://pubmed.ncbi.nlm.nih.gov/25894962
38. Waldhoer, T., et al. Sex Differences of >/= pT1 Bladder Cancer Survival in Austria: A Descriptive,
Long-Term, Nation-Wide Analysis Based on 27,773 Patients. Urol Int, 2015. 94: 383.
https://pubmed.ncbi.nlm.nih.gov/25833466
39. Krimphove, M.J., et al. Sex-specific Differences in the Quality of Treatment of Muscle-invasive
Bladder Cancer Do Not Explain the Overall Survival Discrepancy. Eur Urol Focus, 2019. 7: 124.
https://pubmed.ncbi.nlm.nih.gov/31227463
40. Patafio, F.M., et al. Is there a gender effect in bladder cancer? A population-based study of practice
and outcomes. Can Urol Assoc J, 2015. 9: 269.
https://pubmed.ncbi.nlm.nih.gov/26316913
41. Andreassen, B.K., et al. Bladder cancer survival: Women better off in the long run. Eur J Cancer,
2018. 95: 52.
https://pubmed.ncbi.nlm.nih.gov/29635144
42. Cohn, J.A., et al. Sex disparities in diagnosis of bladder cancer after initial presentation with
hematuria: a nationwide claims-based investigation. Cancer, 2014. 120: 555.
https://pubmed.ncbi.nlm.nih.gov/24496869
43. Dietrich, K., et al. Parity, early menopause and the incidence of bladder cancer in women: a case-
control study and meta-analysis. Eur J Cancer, 2011. 47: 592.
https://pubmed.ncbi.nlm.nih.gov/21067913
44. Scosyrev, E., et al. Sex and racial differences in bladder cancer presentation and mortality in the US.
Cancer, 2009. 115: 68.
https://pubmed.ncbi.nlm.nih.gov/19072984
45. Stenzl, A. Words of wisdom. Re: sex and racial differences in bladder cancer presentation and
mortality in the US. Eur Urol, 2010. 57: 729.
https://pubmed.ncbi.nlm.nih.gov/20965044
46. Abufaraj, M., et al. The impact of hormones and reproductive factors on the risk of bladder cancer
in women: results from the Nurses’ Health Study and Nurses’ Health Study II. Int J Epidemiol, 2020.
49: 599.
https://pubmed.ncbi.nlm.nih.gov/31965144
47. Martin, C., et al. Familial Cancer Clustering in Urothelial Cancer: A Population-Based Case-Control
Study. J Natl Cancer Inst, 2018. 110: 527.
https://pubmed.ncbi.nlm.nih.gov/29228305
48. Murta-Nascimento, C., et al. Risk of bladder cancer associated with family history of cancer: do
low-penetrance polymorphisms account for the increase in risk? Cancer Epidemiol Biomarkers Prev,
2007. 16: 1595.
https://pubmed.ncbi.nlm.nih.gov/17684133
49. Figueroa, J.D., et al. Genome-wide association study identifies multiple loci associated with bladder
cancer risk. Hum Mol Genet, 2014. 23: 1387.
https://pubmed.ncbi.nlm.nih.gov/24163127
50. Rothman, N., et al. A multi-stage genome-wide association study of bladder cancer identifies
multiple susceptibility loci. Nat Genet, 2010. 42: 978.
https://pubmed.ncbi.nlm.nih.gov/20972438
51. Kiemeney, L.A., et al. Sequence variant on 8q24 confers susceptibility to urinary bladder cancer. Nat
Genet, 2008. 40: 1307.
https://pubmed.ncbi.nlm.nih.gov/18794855
52. Varma, M., et al. Dataset for the reporting of urinary tract carcinoma-biopsy and transurethral
resection specimen: recommendations from the International Collaboration on Cancer Reporting
(ICCR). Mod Pathol, 2020. 33: 700.
https://pubmed.ncbi.nlm.nih.gov/31685965
53. Paner, G.P., et al. Further characterization of the muscle layers and lamina propria of the urinary
bladder by systematic histologic mapping: implications for pathologic staging of invasive urothelial
carcinoma. Am J Surg Pathol, 2007. 31: 1420.
https://pubmed.ncbi.nlm.nih.gov/17721199

62 LIMITED UPDATE MARCH 2021


54. Weiner, A.B., et al. Tumor Location May Predict Adverse Pathology and Survival Following Definitive
Treatment for Bladder Cancer: A National Cohort Study. Eur Urol Oncol, 2019. 2: 304.
https://pubmed.ncbi.nlm.nih.gov/31200845
55. Stenzl, A. Current concepts for urinary diversion in women. Eur Urol (EAU Update series 1), 2003: 91.
https://link.springer.com/chapter/10.1007/978-1-59259-097-1_1
56. Varinot, J., et al. Full analysis of the prostatic urethra at the time of radical cystoprostatectomy for
bladder cancer: impact on final disease stage. Virchows Arch, 2009. 455: 449.
https://pubmed.ncbi.nlm.nih.gov/19841937
57. Hansel, D.E., et al. A contemporary update on pathology standards for bladder cancer: transurethral
resection and radical cystectomy specimens. Eur Urol, 2013. 63: 321.
https://pubmed.ncbi.nlm.nih.gov/23088996
58. Herr, H.W. Pathologic evaluation of radical cystectomy specimens. Cancer, 2002. 95: 668.
https://pubmed.ncbi.nlm.nih.gov/12209761
59. Fajkovic, H., et al. Extranodal extension is a powerful prognostic factor in bladder cancer patients
with lymph node metastasis. Eur Urol, 2013. 64: 837.
https://pubmed.ncbi.nlm.nih.gov/22877503
60. Fritsche, H.M., et al. Prognostic value of perinodal lymphovascular invasion following radical
cystectomy for lymph node-positive urothelial carcinoma. Eur Urol, 2013. 63: 739.
https://pubmed.ncbi.nlm.nih.gov/23079053
61. Neuzillet, Y., et al. Positive surgical margins and their locations in specimens are adverse prognosis
features after radical cystectomy in non-metastatic carcinoma invading bladder muscle: results from
a nationwide case-control study. BJU Int, 2013. 111: 1253.
https://pubmed.ncbi.nlm.nih.gov/23331375
62. Baltaci, S., et al. Reliability of frozen section examination of obturator lymph nodes and impact on
lymph node dissection borders during radical cystectomy: results of a prospective multicentre study
by the Turkish Society of Urooncology. BJU Int, 2011. 107: 547.
https://pubmed.ncbi.nlm.nih.gov/20633004
63. Jimenez, R.E., et al. Grading the invasive component of urothelial carcinoma of the bladder and its
relationship with progression-free survival. Am J Surg Pathol, 2000. 24: 980.
https://pubmed.ncbi.nlm.nih.gov/10895820
64. Veskimae, E., et al. What Is the Prognostic and Clinical Importance of Urothelial and Nonurothelial
Histological Variants of Bladder Cancer in Predicting Oncological Outcomes in Patients with
Muscle-invasive and Metastatic Bladder Cancer? A European Association of Urology Muscle
Invasive and Metastatic Bladder Cancer Guidelines Panel Systematic Review. Eur Urol Oncol, 2019.
2: 625.
https://pubmed.ncbi.nlm.nih.gov/31601522
65. Sjodahl, G., et al. A molecular taxonomy for urothelial carcinoma. Clin Cancer Res, 2012. 18: 3377.
https://pubmed.ncbi.nlm.nih.gov/22553347
66. Choi, W., et al. Identification of distinct basal and luminal subtypes of muscle-invasive bladder
cancer with different sensitivities to frontline chemotherapy. Cancer Cell, 2014. 25: 152.
https://pubmed.ncbi.nlm.nih.gov/24525232
67. Sauter G, et al., Tumours of the urinary system: non-invasive urothelial neoplasias., In: WHO
classification of classification of tumors of the urinary system and male genital organs. 2004, IARCC
Press: Lyon.
https://publications.iarc.fr/Book-And-Report-Series/Who-Classification-Of-Tumours/WHO-
Classification-Of-Tumours-Of-The-Urinary-System-And-Male-Genital-Organs-2016
68. Richard, P.O., et al. Active Surveillance for Renal Neoplasms with Oncocytic Features is Safe. J Urol,
2016. 195: 581.
https://pubmed.ncbi.nlm.nih.gov/26388501
69. Comperat, E.M., et al. Grading of Urothelial Carcinoma and The New “World Health Organisation
Classification of Tumours of the Urinary System and Male Genital Organs 2016”. Eur Urol Focus,
2019. 5: 457.
https://pubmed.ncbi.nlm.nih.gov/29366854
70. Compérat, E., et al. Dataset for the reporting of carcinoma of the bladder-cystectomy,
cystoprostatectomy and diverticulectomy specimens: recommendations from the International
Collaboration on Cancer Reporting (ICCR). Virchows Arch, 2020. 476: 521.
https://pubmed.ncbi.nlm.nih.gov/31915958
71. Brierley JD, G.M., Wittekind C, TNM classification of malignant tumors. UICC International Union
Against Cancer. 8th edn. 2016, Oxford.

LIMITED UPDATE MARCH 2021 63


72. Jensen, J.B., et al. Incidence of occult lymph-node metastasis missed by standard pathological
examination in patients with bladder cancer undergoing radical cystectomy. Scand J Urol Nephrol,
2011. 45: 419.
https://pubmed.ncbi.nlm.nih.gov/21767245
73. Mariappan, P., et al. Good quality white-light transurethral resection of bladder tumours
(GQ-WLTURBT) with experienced surgeons performing complete resections and obtaining detrusor
muscle reduces early recurrence in new non-muscle-invasive bladder cancer: validation across time
and place and recommendation for benchmarking. BJU Int, 2012. 109: 1666.
https://pubmed.ncbi.nlm.nih.gov/22044434
74. Magers, M.J., et al. Clinicopathological characteristics of ypT0N0 urothelial carcinoma following
neoadjuvant chemotherapy and cystectomy. J Clin Pathol, 2019. 72: 550.
https://pubmed.ncbi.nlm.nih.gov/31164444
75. Martini, A., et al. Tumor downstaging as an intermediate endpoint to assess the activity of
neoadjuvant systemic therapy in patients with muscle-invasive bladder cancer. Cancer, 2019.
125: 3155.
https://pubmed.ncbi.nlm.nih.gov/31150110
76. Fossa, S.D., et al. Clinical significance of the “palpable mass” in patients with muscle-infiltrating
bladder cancer undergoing cystectomy after pre-operative radiotherapy. Br J Urol, 1991. 67: 54.
https://pubmed.ncbi.nlm.nih.gov/1993277
77. Wijkstrom, H., et al. Evaluation of clinical staging before cystectomy in transitional cell bladder
carcinoma: a long-term follow-up of 276 consecutive patients. Br J Urol, 1998. 81: 686.
https://pubmed.ncbi.nlm.nih.gov/9634042
78. Ploeg, M., et al. Discrepancy between clinical staging through bimanual palpation and pathological
staging after cystectomy. Urol Oncol, 2012. 30: 247.
https://pubmed.ncbi.nlm.nih.gov/20451418
79. Blick, C.G., et al. Evaluation of diagnostic strategies for bladder cancer using computed
tomography (CT) urography, flexible cystoscopy and voided urine cytology: results for 778 patients
from a hospital haematuria clinic. BJU Int, 2012. 110: 84.
https://pubmed.ncbi.nlm.nih.gov/22122739
80. Knox, M.K., et al. Evaluation of multidetector computed tomography urography and ultrasonography
for diagnosing bladder cancer. Clin Radiol, 2008. 63: 1317.
https://pubmed.ncbi.nlm.nih.gov/18996261
81. Lokeshwar, V.B., et al. Bladder tumor markers beyond cytology: International Consensus Panel on
bladder tumor markers. Urology, 2005. 66: 35.
https://pubmed.ncbi.nlm.nih.gov/16399415
82. Raitanen, M.P., et al. Differences between local and review urinary cytology in diagnosis of bladder
cancer. An interobserver multicenter analysis. Eur Urol, 2002. 41: 284.
https://pubmed.ncbi.nlm.nih.gov/12180229
83. van Rhijn, B.W., et al. Urine markers for bladder cancer surveillance: a systematic review. Eur Urol,
2005. 47: 736.
https://pubmed.ncbi.nlm.nih.gov/15925067
84. Barkan, G.A., et al. The Paris System for Reporting Urinary Cytology: The Quest to Develop a
Standardized Terminology. Adv Anat Pathol, 2016. 23: 193.
https://pubmed.ncbi.nlm.nih.gov/27233050
85. Mariappan, P., et al. Detrusor muscle in the first, apparently complete transurethral resection
of bladder tumour specimen is a surrogate marker of resection quality, predicts risk of early
recurrence, and is dependent on operator experience. Eur Urol, 2010. 57: 843.
https://pubmed.ncbi.nlm.nih.gov/19524354
86. Stenzl, A., et al. Hexaminolevulinate guided fluorescence cystoscopy reduces recurrence in patients
with nonmuscle invasive bladder cancer. J Urol, 2010. 184: 1907.
https://pubmed.ncbi.nlm.nih.gov/20850152
87. Burger, M., et al. Photodynamic diagnosis of non-muscle-invasive bladder cancer with
hexaminolevulinate cystoscopy: a meta-analysis of detection and recurrence based on raw data.
Eur Urol, 2013. 64: 846.
https://pubmed.ncbi.nlm.nih.gov/23602406
88. Matzkin, H., et al. Transitional cell carcinoma of the prostate. J Urol, 1991. 146: 1207.
https://www.auajournals.org/doi/abs/10.1016/S0022-5347%2817%2938047-3
89. Mungan, M.U., et al. Risk factors for mucosal prostatic urethral involvement in superficial transitional
cell carcinoma of the bladder. Eur Urol, 2005. 48: 760.
https://pubmed.ncbi.nlm.nih.gov/16005563

64 LIMITED UPDATE MARCH 2021


90. Kassouf, W., et al. Prostatic urethral biopsy has limited usefulness in counseling patients regarding
final urethral margin status during orthotopic neobladder reconstruction. J Urol, 2008. 180: 164.
https://pubmed.ncbi.nlm.nih.gov/18485384
91. Walsh, D.L., et al. Dilemmas in the treatment of urothelial cancers of the prostate. Urol Oncol, 2009.
27: 352.
https://pubmed.ncbi.nlm.nih.gov/18439852
92. Lebret, T., et al. Urethral recurrence of transitional cell carcinoma of the bladder. Predictive value of
preoperative latero-montanal biopsies and urethral frozen sections during prostatocystectomy. Eur
Urol, 1998. 33: 170.
https://pubmed.ncbi.nlm.nih.gov/9519359
93. Donat, S.M., et al. The efficacy of transurethral biopsy for predicting the long-term clinical impact of
prostatic invasive bladder cancer. J Urol, 2001. 165: 1580.
https://pubmed.ncbi.nlm.nih.gov/11342921
94. von Rundstedt, F.C., et al. Transurethral biopsy of the prostatic urethra is associated with final apical
margin status at radical cystoprostatectomy. J Clin Urol, 2016. 9: 404.
https://pubmed.ncbi.nlm.nih.gov/27818773
95. Kates, M., et al. Accuracy of urethral frozen section during radical cystectomy for bladder cancer.
Urol Oncol, 2016. 34: 532.e1.
https://pubmed.ncbi.nlm.nih.gov/27432433
96. Damiano, R., et al. Clinicopathologic features of prostate adenocarcinoma incidentally discovered at
the time of radical cystectomy: an evidence-based analysis. Eur Urol, 2007. 52: 648.
https://pubmed.ncbi.nlm.nih.gov/17600614
97. Gakis, G., et al. Incidental prostate cancer at radical cystoprostatectomy: implications for apex-
sparing surgery. BJU Int, 2010. 105: 468.
https://pubmed.ncbi.nlm.nih.gov/20102366
98. Bruins, H.M., et al. Incidental prostate cancer in patients with bladder urothelial carcinoma:
comprehensive analysis of 1,476 radical cystoprostatectomy specimens. J Urol, 2013. 190: 1704.
https://pubmed.ncbi.nlm.nih.gov/23707451
99. Kaelberer, J.B., et al. Incidental prostate cancer diagnosed at radical cystoprostatectomy for
bladder cancer: disease-specific outcomes and survival. Prostate Int, 2016. 4: 107.
https://pubmed.ncbi.nlm.nih.gov/27689068
100. Mottet, N., et al, EAU-EANM-ESTRO-ESUR-SIOG Prostate Cancer Guidelines, 2021. Edn. presented
at the 36th Annual Congress Milan. EAU Guidelines Office Arnhem, The Netherlands.
https://pubmed.ncbi.nlm.nih.gov/33172724
101. AJCC Cancer Staging Manual. 8th Edn. 2017, Cham, Switzerland.
https://www.springer.com/gp/book/9783319406176
102. Amin, M.B., et al. The Eighth Edition AJCC Cancer Staging Manual: Continuing to build a bridge
from a population-based to a more “personalized” approach to cancer staging. CA Cancer J Clin,
2017. 67: 93.
https://pubmed.ncbi.nlm.nih.gov/28094848
103. Woo, S., et al. Diagnostic Performance of Vesical Imaging Reporting and Data System for the
Prediction of Muscle-invasive Bladder Cancer: A Systematic Review and Meta-analysis. Eur Urol
Oncol, 2020. 3: 306.
https://pubmed.ncbi.nlm.nih.gov/32199915
104. Gandhi, N., et al. Diagnostic accuracy of magnetic resonance imaging for tumour staging of bladder
cancer: systematic review and meta-analysis. BJU Int, 2018. 122: 744.
https://pubmed.ncbi.nlm.nih.gov/29727910
105. Paik, M.L., et al. Limitations of computerized tomography in staging invasive bladder cancer before
radical cystectomy. J Urol, 2000. 163: 1693.
https://pubmed.ncbi.nlm.nih.gov/10799162
106. Mallampati, G.K., et al. MR imaging of the bladder. Magn Reson Imaging Clin N Am, 2004. 12: 545.
https://pubmed.ncbi.nlm.nih.gov/15271370
107. Rajesh, A., et al. Bladder cancer: evaluation of staging accuracy using dynamic MRI. Clin Radiol,
2011. 66: 1140.
https://pubmed.ncbi.nlm.nih.gov/21924408
108. Thomsen, H.S. Nephrogenic systemic fibrosis: history and epidemiology. Radiol Clin North Am,
2009. 47: 827.
https://pubmed.ncbi.nlm.nih.gov/19744597

LIMITED UPDATE MARCH 2021 65


109. Kundra, V., et al. Imaging in oncology from the University of Texas M. D. Anderson Cancer
Center. Imaging in the diagnosis, staging, and follow-up of cancer of the urinary bladder. AJR Am
J Roentgenol, 2003. 180: 1045.
https://pubmed.ncbi.nlm.nih.gov/17885053
110. Kim, B., et al. Bladder tumor staging: comparison of contrast-enhanced CT, T1- and T2-weighted
MR imaging, dynamic gadolinium-enhanced imaging, and late gadolinium-enhanced imaging.
Radiology, 1994. 193: 239.
https://pubmed.ncbi.nlm.nih.gov/8090898
111. Kim, J.K., et al. Bladder cancer: analysis of multi-detector row helical CT enhancement pattern and
accuracy in tumor detection and perivesical staging. Radiology, 2004. 231: 725.
https://pubmed.ncbi.nlm.nih.gov/15118111
112. Yang, W.T., et al. Comparison of dynamic helical CT and dynamic MR imaging in the evaluation of
pelvic lymph nodes in cervical carcinoma. AJR Am J Roentgenol, 2000. 175: 759.
https://pubmed.ncbi.nlm.nih.gov/10954463
113. Kim, S.H., et al. Uterine cervical carcinoma: comparison of CT and MR findings. Radiology, 1990.
175: 45.
https://pubmed.ncbi.nlm.nih.gov/2315503
114. Kim, S.H., et al. Uterine cervical carcinoma: evaluation of pelvic lymph node metastasis with MR
imaging. Radiology, 1994. 190: 807.
https://pubmed.ncbi.nlm.nih.gov/8115631
115. Oyen, R.H., et al. Lymph node staging of localized prostatic carcinoma with CT and CT-guided fine-
needle aspiration biopsy: prospective study of 285 patients. Radiology, 1994. 190: 315.
https://pubmed.ncbi.nlm.nih.gov/8284375
116. Barentsz, J.O., et al. MR imaging of the male pelvis. Eur Radiol, 1999. 9: 1722.
https://pubmed.ncbi.nlm.nih.gov/10602944
117. Dorfman, R.E., et al. Upper abdominal lymph nodes: criteria for normal size determined with CT.
Radiology, 1991. 180: 319.
https://pubmed.ncbi.nlm.nih.gov/2068292
118. Vind-Kezunovic, S., et al. Detection of Lymph Node Metastasis in Patients with Bladder Cancer
using Maximum Standardised Uptake Value and (18)F-fluorodeoxyglucose Positron Emission
Tomography/Computed Tomography: Results from a High-volume Centre Including Long-term
Follow-up. Eur Urol Focus, 2019. 5: 90.
https://pubmed.ncbi.nlm.nih.gov/28753817
119. Ito, Y., et al. Preoperative hydronephrosis grade independently predicts worse pathological
outcomes in patients undergoing nephroureterectomy for upper tract urothelial carcinoma. J Urol,
2011. 185: 1621.
https://pubmed.ncbi.nlm.nih.gov/21419429
120. Cowan, N.C., et al. Multidetector computed tomography urography for diagnosing upper urinary
tract urothelial tumour. BJU Int, 2007. 99: 1363.
https://pubmed.ncbi.nlm.nih.gov/17428251
121. Messer, J.C., et al. Multi-institutional validation of the ability of preoperative hydronephrosis to predict
advanced pathologic tumor stage in upper-tract urothelial carcinoma. Urol Oncol, 2013. 31: 904.
https://pubmed.ncbi.nlm.nih.gov/21906967
122. Hurel, S., et al. Influence of preoperative factors on the oncologic outcome for upper urinary tract
urothelial carcinoma after radical nephroureterectomy. World J Urol, 2015. 33: 335.
https://pubmed.ncbi.nlm.nih.gov/24810657
123. Verhoest, G., et al. Predictive factors of recurrence and survival of upper tract urothelial carcinomas.
World J Urol, 2011. 29: 495.
https://pubmed.ncbi.nlm.nih.gov/21681525
124. Takahashi, N., et al. Gadolinium enhanced magnetic resonance urography for upper urinary tract
malignancy. J Urol, 2010. 183: 1330.
https://pubmed.ncbi.nlm.nih.gov/20171676
125. Girvin, F., et al. Pulmonary nodules: detection, assessment, and CAD. AJR Am J Roentgenol, 2008.
191: 1057.
https://pubmed.ncbi.nlm.nih.gov/18806142
126. Heidenreich, A., et al. Imaging studies in metastatic urogenital cancer patients undergoing systemic
therapy: recommendations of a multidisciplinary consensus meeting of the Association of Urological
Oncology of the German Cancer Society. Urol Int, 2010. 85: 1.
https://pubmed.ncbi.nlm.nih.gov/20693823

66 LIMITED UPDATE MARCH 2021


127. Braendengen, M., et al. Clinical significance of routine pre-cystectomy bone scans in patients with
muscle-invasive bladder cancer. Br J Urol, 1996. 77: 36.
https://pubmed.ncbi.nlm.nih.gov/8653315
128. Brismar, J., et al. Bone scintigraphy in staging of bladder carcinoma. Acta Radiol, 1988. 29: 251.
https://pubmed.ncbi.nlm.nih.gov/2965914
129. Lauenstein, T.C., et al. Whole-body MR imaging: evaluation of patients for metastases. Radiology,
2004. 233: 139.
https://pubmed.ncbi.nlm.nih.gov/15317952
130. Schmidt, G.P., et al. Whole-body MR imaging of bone marrow. Eur J Radiol, 2005. 55: 33.
https://pubmed.ncbi.nlm.nih.gov/15950099
131. Yang, Z., et al. Is whole-body fluorine-18 fluorodeoxyglucose PET/CT plus additional pelvic images
(oral hydration-voiding-refilling) useful for detecting recurrent bladder cancer? Ann Nucl Med, 2012.
26: 571.
https://pubmed.ncbi.nlm.nih.gov/22763630
132. Maurer, T., et al. Diagnostic efficacy of [11C]choline positron emission tomography/computed
tomography compared with conventional computed tomography in lymph node staging of patients
with bladder cancer prior to radical cystectomy. Eur Urol, 2012. 61: 1031.
https://pubmed.ncbi.nlm.nih.gov/22196847
133. Yoshida, S., et al. Role of diffusion-weighted magnetic resonance imaging in predicting sensitivity to
chemoradiotherapy in muscle-invasive bladder cancer. Int J Radiat Oncol Biol Phys, 2012. 83: e21.
https://pubmed.ncbi.nlm.nih.gov/24976935
134. Game, X., et al. Radical cystectomy in patients older than 75 years: assessment of morbidity and
mortality. Eur Urol, 2001. 39: 525.
https://pubmed.ncbi.nlm.nih.gov/11464032
135. Clark, P.E., et al. Radical cystectomy in the elderly: comparison of clincal outcomes between
younger and older patients. Cancer, 2005. 104: 36.
https://pubmed.ncbi.nlm.nih.gov/15912515
136. May, M., et al. Results from three municipal hospitals regarding radical cystectomy on elderly
patients. Int Braz J Urol, 2007. 33: 764.
https://pubmed.ncbi.nlm.nih.gov/18199344
137. Ethun, C.G., et al. Frailty and cancer: Implications for oncology surgery, medical oncology, and
radiation oncology. CA Cancer J Clin, 2017. 67: 362.
https://pubmed.ncbi.nlm.nih.gov/28731537
138. Miller, D.C., et al. The impact of co-morbid disease on cancer control and survival following radical
cystectomy. J Urol, 2003. 169: 105.
https://pubmed.ncbi.nlm.nih.gov/12478114
139. Haden, T.D., et al. Comparative Perioperative Outcomes in Septuagenarians and Octogenarians
Undergoing Radical Cystectomy for Bladder Cancer-Do Outcomes Differ? Eur Urol Focus, 2018.
4: 895.
https://pubmed.ncbi.nlm.nih.gov/28865996
140. Brown, A.S., et al. National Institutes of Health Consensus Development Conference Statement:
Geriatric Assessment Methods for Clinical Decision-making. J Am Geriatr Soc, 1988. 36: 342.
https://agsjournals.onlinelibrary.wiley.com/doi/abs/10.1111/j.1532-5415.1988.tb02362.x
141. Mayr, R., et al. Sarcopenia as a comorbidity-independent predictor of survival following radical
cystectomy for bladder cancer. J Cachexia Sarcopenia Muscle, 2018. 9: 505.
https://pubmed.ncbi.nlm.nih.gov/29479839
142. Lyon, T.D., et al. Sarcopenia and Response to Neoadjuvant Chemotherapy for Muscle-Invasive
Bladder Cancer. Clin Genitourin Cancer, 2019. 17: 216.
https://pubmed.ncbi.nlm.nih.gov/31060857
143. Lawrentschuk, N., et al. Prevention and management of complications following radical cystectomy
for bladder cancer. Eur Urol, 2010. 57: 983.
https://pubmed.ncbi.nlm.nih.gov/20227172
144. Donahue, T.F., et al. Risk factors for the development of parastomal hernia after radical cystectomy.
J Urol, 2014. 191: 1708.
https://pubmed.ncbi.nlm.nih.gov/24384155
145. Djaladat, H., et al. The association of preoperative serum albumin level and American Society of
Anesthesiologists (ASA) score on early complications and survival of patients undergoing radical
cystectomy for urothelial bladder cancer. BJU Int, 2014. 113: 887.
https://pubmed.ncbi.nlm.nih.gov/23906037

LIMITED UPDATE MARCH 2021 67


146. Garg, T., et al. Preoperative serum albumin is associated with mortality and complications after
radical cystectomy. BJU Int, 2014. 113: 918.
https://pubmed.ncbi.nlm.nih.gov/24053616
147. Rochon, P.A., et al. Comorbid illness is associated with survival and length of hospital stay in
patients with chronic disability. A prospective comparison of three comorbidity indices. Med Care,
1996. 34: 1093.
https://pubmed.ncbi.nlm.nih.gov/8911426
148. Williams, S.B., et al. Systematic Review of Comorbidity and Competing-risks Assessments for
Bladder Cancer Patients. Eur Urol Oncol, 2018. 1: 91.
https://pubmed.ncbi.nlm.nih.gov/30345422
149. Zietman, A.L., et al. Organ-conserving approaches to muscle-invasive bladder cancer: future
alternatives to radical cystectomy. Ann Med, 2000. 32: 34.
https://pubmed.ncbi.nlm.nih.gov/10711576
150. Lughezzani, G., et al. A population-based competing-risks analysis of the survival of patients treated
with radical cystectomy for bladder cancer. Cancer, 2011. 117: 103.
https://pubmed.ncbi.nlm.nih.gov/20803606
151. Froehner, M., et al. Complications following radical cystectomy for bladder cancer in the elderly. Eur
Urol, 2009. 56: 443.
https://pubmed.ncbi.nlm.nih.gov/19481861
152. Korc-Grodzicki, B., et al. Prevention of post-operative delirium in older patients with cancer
undergoing surgery. J Geriatr Oncol, 2015. 6: 60.
https://pubmed.ncbi.nlm.nih.gov/25454768
153. Soubeyran, P., et al. Screening for vulnerability in older cancer patients: the ONCODAGE
Prospective Multicenter Cohort Study. PLoS One, 2014. 9: e115060.
https://pubmed.ncbi.nlm.nih.gov/25503576
154. Rockwood, K., et al. A global clinical measure of fitness and frailty in elderly people. Cmaj, 2005.
173: 489.
https://pubmed.ncbi.nlm.nih.gov/16129869
155. de Groot, V., et al. How to measure comorbidity. a critical review of available methods. J Clin
Epidemiol, 2003. 56: 221.
https://pubmed.ncbi.nlm.nih.gov/12725876
156. Linn, B.S., et al. Cumulative illness rating scale. J Am Geriatr Soc, 1968. 16: 622.
https://pubmed.ncbi.nlm.nih.gov/5646906
157. Charlson, M.E., et al. A new method of classifying prognostic comorbidity in longitudinal studies:
development and validation. J Chronic Dis, 1987. 40: 373.
https://pubmed.ncbi.nlm.nih.gov/3558716
158. Litwin, M.S., et al. Assessment of prognosis with the total illness burden index for prostate cancer:
aiding clinicians in treatment choice. Cancer, 2007. 109: 1777.
https://pubmed.ncbi.nlm.nih.gov/17354226
159. Paleri, V., et al. Applicability of the adult comorbidity evaluation - 27 and the Charlson indexes to
assess comorbidity by notes extraction in a cohort of United Kingdom patients with head and neck
cancer: a retrospective study. J Laryngol Otol, 2002. 116: 200.
https://pubmed.ncbi.nlm.nih.gov/11893262
160. Greenfield, S., et al. The importance of co-existent disease in the occurrence of postoperative
complications and one-year recovery in patients undergoing total hip replacement. Comorbidity and
outcomes after hip replacement. Med Care, 1993. 31: 141.
https://pubmed.ncbi.nlm.nih.gov/8433577
161. Kaplan, M.H., et al. The importance of classifying initial co-morbidity in evaluating the outcome of
diabetes mellitus. J Chronic Dis, 1974. 27: 387.
https://pubmed.ncbi.nlm.nih.gov/4436428
162. Farhat, J.S., et al. Are the frail destined to fail? Frailty index as predictor of surgical morbidity and
mortality in the elderly. J Trauma Acute Care Surg, 2012. 72: 1526.
https://pubmed.ncbi.nlm.nih.gov/22695416
163. Mayr, R., et al. Predictive capacity of four comorbidity indices estimating perioperative mortality
after radical cystectomy for urothelial carcinoma of the bladder. BJU Int, 2012. 110: E222.
https://pubmed.ncbi.nlm.nih.gov/22314129
164. Morgan, T.M., et al. Predicting the probability of 90-day survival of elderly patients with bladder
cancer treated with radical cystectomy. J Urol, 2011. 186: 829.
https://pubmed.ncbi.nlm.nih.gov/21788035

68 LIMITED UPDATE MARCH 2021


165. Abdollah, F., et al. Development and validation of a reference table for prediction of postoperative
mortality rate in patients treated with radical cystectomy: a population-based study. Ann Surg
Oncol, 2012. 19: 309.
https://pubmed.ncbi.nlm.nih.gov/21701925
166. Koppie, T.M., et al. Age-adjusted Charlson comorbidity score is associated with treatment decisions
and clinical outcomes for patients undergoing radical cystectomy for bladder cancer. Cancer, 2008.
112: 2384.
https://pubmed.ncbi.nlm.nih.gov/18404699
167. Bolenz, C., et al. Management of elderly patients with urothelial carcinoma of the bladder: guideline
concordance and predictors of overall survival. BJU Int, 2010. 106: 1324.
https://pubmed.ncbi.nlm.nih.gov/20500510
168. Yoo, S., et al. Does radical cystectomy improve overall survival in octogenarians with muscle-
invasive bladder cancer? Korean J Urol, 2011. 52: 446.
https://pubmed.ncbi.nlm.nih.gov/21860763
169. Mayr, R., et al. Comorbidity and performance indices as predictors of cancer-independent mortality
but not of cancer-specific mortality after radical cystectomy for urothelial carcinoma of the bladder.
Eur Urol, 2012. 62: 662.
https://pubmed.ncbi.nlm.nih.gov/22534059
170. Hall, W.H., et al. An electronic application for rapidly calculating Charlson comorbidity score. BMC
Cancer, 2004. 4: 94.
https://pubmed.ncbi.nlm.nih.gov/15610554
171. Extermann, M., et al. Comorbidity and functional status are independent in older cancer patients.
J Clin Oncol, 1998. 16: 1582.
https://pubmed.ncbi.nlm.nih.gov/9552069
172. Blagden, S.P., et al. Performance status score: do patients and their oncologists agree? Br J Cancer,
2003. 89: 1022.
https://pubmed.ncbi.nlm.nih.gov/12966419
173. Logothetis, C.J., et al. Escalated MVAC with or without recombinant human granulocyte-
macrophage colony-stimulating factor for the initial treatment of advanced malignant urothelial
tumors: results of a randomized trial. J Clin Oncol, 1995. 13: 2272.
https://pubmed.ncbi.nlm.nih.gov/7666085
174. von der Maase, H., et al. Gemcitabine and cisplatin versus methotrexate, vinblastine, doxorubicin,
and cisplatin in advanced or metastatic bladder cancer: results of a large, randomized,
multinational, multicenter, phase III study. J Clin Oncol, 2000. 18: 3068.
https://pubmed.ncbi.nlm.nih.gov/11001674
175. Niegisch, G., et al. Prognostic factors in second-line treatment of urothelial cancers with
gemcitabine and paclitaxel (German Association of Urological Oncology trial AB20/99). Eur Urol,
2011. 60: 1087.
https://pubmed.ncbi.nlm.nih.gov/21839579
176. Cohen, H.J., et al. A controlled trial of inpatient and outpatient geriatric evaluation and management.
N Engl J Med, 2002. 346: 905.
https://pubmed.ncbi.nlm.nih.gov/11907291
177. Balducci, L., et al. General guidelines for the management of older patients with cancer. Oncology
(Williston Park), 2000. 14: 221.
https://pubmed.ncbi.nlm.nih.gov/11195414
178. Castagneto, B., et al. Single-agent gemcitabine in previously untreated elderly patients with
advanced bladder carcinoma: response to treatment and correlation with the comprehensive
geriatric assessment. Oncology, 2004. 67: 27.
https://pubmed.ncbi.nlm.nih.gov/15459492
179. Dutta, R., et al. Effect of tumor location on survival in urinary bladder adenocarcinoma: A
population-based analysis. Urol Oncol, 2016. 34: 531.e1.
https://pubmed.ncbi.nlm.nih.gov/27427223
180. Mathieu, R., et al. The prognostic role of lymphovascular invasion in urothelial carcinoma of the
bladder. Nat Rev Urol, 2016. 13: 471.
https://pubmed.ncbi.nlm.nih.gov/27431340
181. Kimura, S., et al. Prognostic value of concomitant carcinoma in situ in the radical cystectomy
specimen: A systematic review and meta-analysis. J Urol, 2019. 201: 46.
https://pubmed.ncbi.nlm.nih.gov/30077559
182. Svatek, R.S., et al. Intravesical tumor involvement of the trigone is associated with nodal metastasis
in patients undergoing radical cystectomy. Urology, 2014. 84: 1147.
https://pubmed.ncbi.nlm.nih.gov/25174656

LIMITED UPDATE MARCH 2021 69


183. Donat, S.M., et al. Mechanisms of prostatic stromal invasion in patients with bladder cancer: clinical
significance. J Urol, 2001. 165: 1117.
https://pubmed.ncbi.nlm.nih.gov/11257650
184. Paner, G.P., et al. Challenges in Pathologic Staging of Bladder Cancer: Proposals for Fresh
Approaches of Assessing Pathologic Stage in Light of Recent Studies and Observations Pertaining
to Bladder Histoanatomic Variances. Adv Anat Pathol, 2017. 24: 113.
https://pubmed.ncbi.nlm.nih.gov/28398951
185. Moschini, M., et al. Impact of the Level of Urothelial Carcinoma Involvement of the Prostate on
Survival after Radical Cystectomy. Bladder Cancer, 2017. 3: 161.
https://pubmed.ncbi.nlm.nih.gov/28824943
186. Wu, S., et al. Pretreatment Neutrophil-Lymphocyte Ratio as a Predictor in Bladder Cancer and
Metastatic or Unresectable Urothelial Carcinoma Patients: a Pooled Analysis of Comparative
Studies. Cell Physiol Biochem, 2018. 46: 1352.
https://pubmed.ncbi.nlm.nih.gov/29689562
187. Ojerholm, E., et al. Neutrophil-to-lymphocyte ratio as a bladder cancer biomarker: Assessing
prognostic and predictive value in SWOG 8710. Cancer, 2017. 123: 794.
https://pubmed.ncbi.nlm.nih.gov/27787873
188. Ku, J.H., et al. Lymph node density as a prognostic variable in node-positive bladder cancer: a
meta-analysis. BMC Cancer, 2015. 15: 447.
https://pubmed.ncbi.nlm.nih.gov/26027955
189. Lee, D., et al. Lymph node density vs. the American Joint Committee on Cancer TNM nodal staging
system in node-positive bladder cancer in patients undergoing extended or super-extended pelvic
lymphadenectomy. Urol Oncol, 2017. 35: 151.e1.
https://pubmed.ncbi.nlm.nih.gov/28139370
190. Jensen, J.B., et al. Evaluation of different lymph node (LN) variables as prognostic markers in
patients undergoing radical cystectomy and extended LN dissection to the level of the inferior
mesenteric artery. BJU Int, 2012. 109: 388.
https://pubmed.ncbi.nlm.nih.gov/21851538
191. Bruins, H.M., et al. Critical evaluation of the American Joint Committee on Cancer TNM nodal
staging system in patients with lymph node-positive disease after radical cystectomy. Eur Urol,
2012. 62: 671.
https://pubmed.ncbi.nlm.nih.gov/22575915
192. Robertson, A.G., et al. Comprehensive Molecular Characterization of Muscle-Invasive Bladder
Cancer. Cell, 2017. 171: 540.
https://pubmed.ncbi.nlm.nih.gov/28988769
193. Choi, W., et al. Intrinsic basal and luminal subtypes of muscle-invasive bladder cancer. Nat Rev Urol,
2014. 11: 400.
https://pubmed.ncbi.nlm.nih.gov/24960601
194. Kamoun, A., et al. A Consensus Molecular Classification of Muscle-invasive Bladder Cancer. Eur
Urol, 2020. 77: 420.
https://pubmed.ncbi.nlm.nih.gov/31563503
195. Abudurexiti, M., et al. Development and External Validation of a Novel 12-Gene Signature for
Prediction of Overall Survival in Muscle-Invasive Bladder Cancer. Front Oncol, 2019. 9: 856.
https://pubmed.ncbi.nlm.nih.gov/31552180
196. Morera, D.S., et al. Clinical Parameters Outperform Molecular Subtypes for Predicting Outcome in
Bladder Cancer: Results from Multiple Cohorts, Including TCGA. J Urol, 2020. 203: 62.
https://pubmed.ncbi.nlm.nih.gov/31112107
197. Pietzak, E.J., et al. Genomic Differences Between “Primary” and “Secondary” Muscle-invasive
Bladder Cancer as a Basis for Disparate Outcomes to Cisplatin-based Neoadjuvant Chemotherapy.
Eur Urol, 2019. 75: 231.
https://pubmed.ncbi.nlm.nih.gov/30290956
198. Motterle, G., et al. Predicting Response to Neoadjuvant Chemotherapy in Bladder Cancer. Eur Urol
Focus, 2019.
https://pubmed.ncbi.nlm.nih.gov/31708469
199. Shariat, S.F., et al. Association of angiogenesis related markers with bladder cancer outcomes and
other molecular markers. J Urol, 2010. 183: 1744.
https://pubmed.ncbi.nlm.nih.gov/20299037
200. Plimack, E.R., et al. Defects in DNA Repair Genes Predict Response to Neoadjuvant Cisplatin-based
Chemotherapy in Muscle-invasive Bladder Cancer. Eur Urol, 2015. 68: 959.
https://pubmed.ncbi.nlm.nih.gov/26238431

70 LIMITED UPDATE MARCH 2021


201. Van Allen, E.M., et al. Somatic ERCC2 mutations correlate with cisplatin sensitivity in muscle-
invasive urothelial carcinoma. Cancer Discov, 2014. 4: 1140.
https://pubmed.ncbi.nlm.nih.gov/25096233
202. Loriot, Y., et al. Erdafitinib in Locally Advanced or Metastatic Urothelial Carcinoma. N Engl J Med,
2019. 381: 338.
https://pubmed.ncbi.nlm.nih.gov/31340094
203. Pal, S.K., et al. Efficacy of BGJ398, a Fibroblast Growth Factor Receptor 1-3 Inhibitor, in Patients
with Previously Treated Advanced Urothelial Carcinoma with FGFR3 Alterations. Cancer Discov,
2018. 8: 812.
https://pubmed.ncbi.nlm.nih.gov/29848605
204. Rosenberg, J.E., et al. Atezolizumab in patients with locally advanced and metastatic urothelial
carcinoma who have progressed following treatment with platinum-based chemotherapy: a single-
arm, multicentre, phase 2 trial. Lancet, 2016. 387: 1909.
https://pubmed.ncbi.nlm.nih.gov/26952546
205. European Medicine Agency. EMA restricts use of Keytruda and Tecentriq in bladder cancer. 2018
[access date March 2021].
https://www.ema.europa.eu/en/news/ema-restricts-use-keytruda-tecentriq-bladder-cancer
206. Kandoth, C., et al. Mutational landscape and significance across 12 major cancer types. Nature,
2013. 502: 333.
https://pubmed.ncbi.nlm.nih.gov/24132290
207. Sharma, P., et al. Nivolumab monotherapy in recurrent metastatic urothelial carcinoma (CheckMate
032): a multicentre, open-label, two-stage, multi-arm, phase 1/2 trial. Lancet Oncol, 2016. 17: 1590.
https://pubmed.ncbi.nlm.nih.gov/27733243
208. Powles, T., et al. Clinical efficacy and biomarker analysis of neoadjuvant atezolizumab in operable
urothelial carcinoma in the ABACUS trial. Nat Med, 2019. 25: 1706.
https://pubmed.ncbi.nlm.nih.gov/31686036
209. Necchi, A., et al. Pembrolizumab as Neoadjuvant Therapy Before Radical Cystectomy in Patients
With Muscle-Invasive Urothelial Bladder Carcinoma (PURE-01): An Open-Label, Single-Arm, Phase
II Study. J Clin Oncol, 2018. 36: 3353.
https://pubmed.ncbi.nlm.nih.gov/30343614
210. Mariathasan, S., et al. TGFβ attenuates tumour response to PD-L1 blockade by contributing to
exclusion of T cells. Nature, 2018. 554: 544.
https://pubmed.ncbi.nlm.nih.gov/29443960
211. Wang, L., et al. EMT- and stroma-related gene expression and resistance to PD-1 blockade in
urothelial cancer. Nature Com, 2018. 9: 3503.
https://pubmed.ncbi.nlm.nih.gov/30158554
212. Stein, J.P., et al. Radical cystectomy in the treatment of invasive bladder cancer: long-term results in
1,054 patients. J Clin Oncol, 2001. 19: 666.
https://pubmed.ncbi.nlm.nih.gov/11157016
213. Stein, J.P., et al. Radical cystectomy for invasive bladder cancer: long-term results of a standard
procedure. World J Urol, 2006. 24: 296.
https://pubmed.ncbi.nlm.nih.gov/16518661
214. Dalbagni, G., et al. Cystectomy for bladder cancer: a contemporary series. J Urol, 2001. 165: 1111.
https://pubmed.ncbi.nlm.nih.gov/11257649
215. Bassi, P., et al. Prognostic factors of outcome after radical cystectomy for bladder cancer: a
retrospective study of a homogeneous patient cohort. J Urol, 1999. 161: 1494.
https://pubmed.ncbi.nlm.nih.gov/10210380
216. Ghoneim, M.A., et al. Radical cystectomy for carcinoma of the bladder: critical evaluation of the
results in 1,026 cases. J Urol, 1997. 158: 393.
https://pubmed.ncbi.nlm.nih.gov/9224310
217. David, K.A., et al. Low incidence of perioperative chemotherapy for stage III bladder cancer 1998 to
2003: a report from the National Cancer Data Base. J Urol, 2007. 178: 451.
https://pubmed.ncbi.nlm.nih.gov/17561135
218. Porter, M.P., et al. Patterns of use of systemic chemotherapy for Medicare beneficiaries with
urothelial bladder cancer. Urol Oncol, 2011. 29: 252.
https://pubmed.ncbi.nlm.nih.gov/19450992
219. Sanchez-Ortiz, R.F., et al. An interval longer than 12 weeks between the diagnosis of muscle invasion
and cystectomy is associated with worse outcome in bladder carcinoma. J Urol, 2003. 169: 110.
https://pubmed.ncbi.nlm.nih.gov/12478115

LIMITED UPDATE MARCH 2021 71


220. Stein, J.P. Contemporary concepts of radical cystectomy and the treatment of bladder cancer.
J Urol, 2003. 169: 116.
https://pubmed.ncbi.nlm.nih.gov/12478116
221. Boeri, L., et al. Delaying Radical Cystectomy After Neoadjuvant Chemotherapy for Muscle-invasive
Bladder Cancer is Associated with Adverse Survival Outcomes. Eur Urol Oncol, 2019. 2: 390.
https://pubmed.ncbi.nlm.nih.gov/31277775
222. Pfister, C., et al. Randomized Phase III Trial of Dose-dense Methotrexate, Vinblastine, Doxorubicin,
and Cisplatin, or Gemcitabine and Cisplatin as Perioperative Chemotherapy for Patients with
Muscle-invasive Bladder Cancer. Analysis of the GETUG/AFU V05 VESPER Trial Secondary
Endpoints: Chemotherapy Toxicity and Pathological Responses. Eur Urol, 2021. 79: 214.
https://pubmed.ncbi.nlm.nih.gov/32868138
223. Grossman, H.B., et al. Neoadjuvant chemotherapy plus cystectomy compared with cystectomy
alone for locally advanced bladder cancer. N Engl J Med, 2003. 349: 859.
https://pubmed.ncbi.nlm.nih.gov/12944571
224. Sherif, A., et al. Neoadjuvant cisplatinum based combination chemotherapy in patients with invasive
bladder cancer: a combined analysis of two Nordic studies. Eur Urol, 2004. 45: 297.
https://pubmed.ncbi.nlm.nih.gov/15036674
225. Kimura, S., et al. Impact of Gender on Chemotherapeutic Response and Oncologic Outcomes in
Patients Treated With Radical Cystectomy and Perioperative Chemotherapy for Bladder Cancer: A
Systematic Review and Meta-Analysis. Clin Genitourin Cancer, 2020. 18: 78.
https://pubmed.ncbi.nlm.nih.gov/31889669
226. D’Andrea, D., et al. Impact of sex on response to neoadjuvant chemotherapy in patients with
bladder cancer. Urol Oncol, 2020. 38: 639.e1.
https://pubmed.ncbi.nlm.nih.gov/32057595
227. Griffiths, G., et al. International phase III trial assessing neoadjuvant cisplatin, methotrexate, and
vinblastine chemotherapy for muscle-invasive bladder cancer: long-term results of the BA06 30894
trial. J Clin Oncol, 2011. 29: 2171.
https://pubmed.ncbi.nlm.nih.gov/21502557
228. Sherif, A., et al. Neoadjuvant cisplatin-methotrexate chemotherapy for invasive bladder cancer --
Nordic cystectomy trial 2. Scand J Urol Nephrol, 2002. 36: 419.
https://pubmed.ncbi.nlm.nih.gov/12623505
229. Sengelov, L., et al. Neoadjuvant chemotherapy with cisplatin and methotrexate in patients with
muscle-invasive bladder tumours. Acta Oncol, 2002. 41: 447.
https://pubmed.ncbi.nlm.nih.gov/12442921
230. Orsatti, M., et al. Alternating chemo-radiotherapy in bladder cancer: a conservative approach. Int
J Radiat Oncol Biol Phys, 1995. 33: 173.
https://pubmed.ncbi.nlm.nih.gov/7642415
231. Shipley, W.U., et al. Phase III trial of neoadjuvant chemotherapy in patients with invasive bladder
cancer treated with selective bladder preservation by combined radiation therapy and chemotherapy:
initial results of Radiation Therapy Oncology Group 89-03. J Clin Oncol, 1998. 16: 3576.
https://pubmed.ncbi.nlm.nih.gov/9817278
232. Abol-Enein H, et al. Neo-adjuvant chemotherapy in the treatment of invasive transitional bladder
cancer. A controlled prospective randomized study. Br J Urol 1997. 79: 174. [No abstract available].
233. Collaboration, A.B.C.M.-a. Neoadjuvant chemotherapy in invasive bladder cancer: a systematic
review and meta-analysis. Lancet, 2003. 361: 1927.
https://pubmed.ncbi.nlm.nih.gov/12801735
234. Winquist, E., et al. Neoadjuvant chemotherapy for transitional cell carcinoma of the bladder: a
systematic review and meta-analysis. J Urol, 2004. 171: 561.
https://pubmed.ncbi.nlm.nih.gov/14713760
235. Neoadjuvant chemotherapy in invasive bladder cancer: update of a systematic review and meta-
analysis of individual patient data advanced bladder cancer (ABC) meta-analysis collaboration. Eur
Urol, 2005. 48: 202.
https://pubmed.ncbi.nlm.nih.gov/15939524
236. Wallace, D.M., et al. Neo-adjuvant (pre-emptive) cisplatin therapy in invasive transitional cell
carcinoma of the bladder. Br J Urol, 1991. 67: 608.
https://pubmed.ncbi.nlm.nih.gov/2070206
237. Martinez-Pineiro, J.A., et al. Neoadjuvant cisplatin chemotherapy before radical cystectomy in
invasive transitional cell carcinoma of the bladder: a prospective randomized phase III study. J Urol,
1995. 153: 964.
https://pubmed.ncbi.nlm.nih.gov/7853584

72 LIMITED UPDATE MARCH 2021


238. Rintala, E., et al. Neoadjuvant chemotherapy in bladder cancer: a randomized study. Nordic
Cystectomy Trial I. Scand J Urol Nephrol, 1993. 27: 355.
https://pubmed.ncbi.nlm.nih.gov/8290916
239. Malmstrom, P.U., et al. Five-year followup of a prospective trial of radical cystectomy and
neoadjuvant chemotherapy: Nordic Cystectomy Trial I. The Nordic Cooperative Bladder Cancer
Study Group. J Urol, 1996. 155: 1903.
https://pubmed.ncbi.nlm.nih.gov/8618283
240. Neoadjuvant cisplatin, methotrexate, and vinblastine chemotherapy for muscle-invasive bladder
cancer: a randomised controlled trial. International collaboration of trialists. Lancet, 1999. 354: 533.
https://pubmed.ncbi.nlm.nih.gov/10470696
241. Yin, M., et al. Neoadjuvant Chemotherapy for Muscle-Invasive Bladder Cancer: A Systematic
Review and Two-Step Meta-Analysis. Oncologist, 2016. 21: 708.
https://pubmed.ncbi.nlm.nih.gov/27053504
242. Galsky, M.D., et al. Comparative effectiveness of gemcitabine plus cisplatin versus methotrexate,
vinblastine, doxorubicin, plus cisplatin as neoadjuvant therapy for muscle-invasive bladder cancer.
Cancer, 2015. 121: 2586.
https://pubmed.ncbi.nlm.nih.gov/25872978
243. Yuh, B.E., et al. Pooled analysis of clinical outcomes with neoadjuvant cisplatin and gemcitabine
chemotherapy for muscle invasive bladder cancer. J Urol, 2013. 189: 1682.
https://pubmed.ncbi.nlm.nih.gov/23123547
244. Lee, F.C., et al. Pathologic Response Rates of Gemcitabine/Cisplatin versus Methotrexate/
Vinblastine/Adriamycin/Cisplatin Neoadjuvant Chemotherapy for Muscle Invasive Urothelial Bladder
Cancer. Adv Urol, 2013. 2013: 317190.
https://pubmed.ncbi.nlm.nih.gov/24382958
245. Dash, A., et al. A role for neoadjuvant gemcitabine plus cisplatin in muscle-invasive urothelial
carcinoma of the bladder: a retrospective experience. Cancer, 2008. 113: 2471.
https://pubmed.ncbi.nlm.nih.gov/18823036
246. Choueiri, T.K., et al. Neoadjuvant dose-dense methotrexate, vinblastine, doxorubicin, and cisplatin
with pegfilgrastim support in muscle-invasive urothelial cancer: pathologic, radiologic, and
biomarker correlates. J Clin Oncol, 2014. 32: 1889.
https://pubmed.ncbi.nlm.nih.gov/24821883
247. Plimack, E.R., et al. Accelerated methotrexate, vinblastine, doxorubicin, and cisplatin is safe, effective,
and efficient neoadjuvant treatment for muscle-invasive bladder cancer: results of a multicenter phase
II study with molecular correlates of response and toxicity. J Clin Oncol, 2014. 32: 1895.
https://pubmed.ncbi.nlm.nih.gov/24821881
248. Peyton, C.C., et al. Downstaging and Survival Outcomes Associated With Neoadjuvant
Chemotherapy Regimens Among Patients Treated With Cystectomy for Muscle-Invasive Bladder
Cancer. JAMA Oncol, 2018. 4: 1535.
https://pubmed.ncbi.nlm.nih.gov/30178038
249. Anari, F., et al. Neoadjuvant Dose-dense Gemcitabine and Cisplatin in Muscle-invasive Bladder
Cancer: Results of a Phase 2 Trial. Eur Urol Oncol, 2018. 1: 54.
https://pubmed.ncbi.nlm.nih.gov/30420974
250. Iyer, G., et al. Multicenter Prospective Phase II Trial of Neoadjuvant Dose-Dense Gemcitabine Plus
Cisplatin in Patients With Muscle-Invasive Bladder Cancer. J Clin Oncol, 2018. 36: 1949.
https://pubmed.ncbi.nlm.nih.gov/29742009
251. Osterman, C.K., et al. Efficacy of Split Schedule Versus Conventional Schedule Neoadjuvant
Cisplatin-Based Chemotherapy for Muscle-Invasive Bladder Cancer. Oncologist, 2019. 24: 688.
https://pubmed.ncbi.nlm.nih.gov/30728277
252. Vetterlein, M.W., et al. Neoadjuvant chemotherapy prior to radical cystectomy for muscle-invasive
bladder cancer with variant histology. Cancer, 2017. 123: 4346.
https://pubmed.ncbi.nlm.nih.gov/28743155
253. Hanna, N., et al. Effectiveness of Neoadjuvant Chemotherapy for Muscle-invasive Bladder Cancer in
the Current Real World Setting in the USA. Eur Urol Oncol, 2018. 1: 83.
https://pubmed.ncbi.nlm.nih.gov/31100232
254. Panebianco, V., et al. Multiparametric Magnetic Resonance Imaging for Bladder Cancer:
Development of VI-RADS (Vesical Imaging-Reporting And Data System). Eur Urol, 2018. 74: 294.
https://pubmed.ncbi.nlm.nih.gov/29755006
255. Letocha, H., et al. Positron emission tomography with L-methyl-11C-methionine in the monitoring
of therapy response in muscle-invasive transitional cell carcinoma of the urinary bladder. Br J Urol,
1994. 74: 767.
https://pubmed.ncbi.nlm.nih.gov/7827849

LIMITED UPDATE MARCH 2021 73


256. Nishimura, K., et al. The effects of neoadjuvant chemotherapy and chemo-radiation therapy on MRI
staging in invasive bladder cancer: comparative study based on the pathological examination of
whole layer bladder wall. Int Urol Nephrol, 2009. 41: 869.
https://pubmed.ncbi.nlm.nih.gov/19396568
257. Barentsz, J.O., et al. Evaluation of chemotherapy in advanced urinary bladder cancer with fast
dynamic contrast-enhanced MR imaging. Radiology, 1998. 207: 791.
https://pubmed.ncbi.nlm.nih.gov/9609906
258. Krajewski, K.M., et al. Optimisation of the size variation threshold for imaging evaluation of response
in patients with platinum-refractory advanced transitional cell carcinoma of the urothelium treated
with vinflunine. Eur J Cancer, 2012. 48: 1495.
https://pubmed.ncbi.nlm.nih.gov/22176867
259. Rosenblatt, R., et al. Pathologic downstaging is a surrogate marker for efficacy and increased
survival following neoadjuvant chemotherapy and radical cystectomy for muscle-invasive urothelial
bladder cancer. Eur Urol, 2012. 61: 1229.
https://pubmed.ncbi.nlm.nih.gov/22189383
260. Voskuilen, C.S., et al. Multicenter Validation of Histopathologic Tumor Regression Grade After
Neoadjuvant Chemotherapy in Muscle-invasive Bladder Carcinoma. Am J Surg Pathol, 2019.
43: 1600.
https://pubmed.ncbi.nlm.nih.gov/31524642
261. Takata, R., et al. Predicting response to methotrexate, vinblastine, doxorubicin, and cisplatin
neoadjuvant chemotherapy for bladder cancers through genome-wide gene expression profiling.
Clin Cancer Res, 2005. 11: 2625.
https://pubmed.ncbi.nlm.nih.gov/15814643
262. Takata, R., et al. Validation study of the prediction system for clinical response of M-VAC
neoadjuvant chemotherapy. Cancer Sci, 2007. 98: 113.
https://pubmed.ncbi.nlm.nih.gov/17116130
263. Miron, B., et al. Defects in DNA Repair Genes Confer Improved Long-term Survival after Cisplatin-
based Neoadjuvant Chemotherapy for Muscle-invasive Bladder Cancer. Eur Urol Oncol, 2020. 3: 544.
https://pubmed.ncbi.nlm.nih.gov/32165095
264. Zaghloul, M.S., et al. Adjuvant Sandwich Chemotherapy Plus Radiotherapy vs Adjuvant
Chemotherapy Alone for Locally Advanced Bladder Cancer After Radical Cystectomy: A
Randomized Phase 2 Trial. JAMA Surg, 2018. 153: e174591.
https://pubmed.ncbi.nlm.nih.gov/29188298
265. Iwata, T., et al. The role of adjuvant radiotherapy after surgery for upper and lower urinary tract
urothelial carcinoma: A systematic review. Urol Oncol, 2019. 37: 659.
https://pubmed.ncbi.nlm.nih.gov/31255542
266. Slack, N.H., et al. Five-year follow-up results of a collaborative study of therapies for carcinoma of
the bladder. J Surg Oncol, 1977. 9: 393.
https://pubmed.ncbi.nlm.nih.gov/330958
267. Smith, J.A., Jr., et al. Treatment of advanced bladder cancer with combined preoperative irradiation
and radical cystectomy versus radical cystectomy alone: a phase III intergroup study. J Urol, 1997.
157: 805.
https://pubmed.ncbi.nlm.nih.gov/9072571
268. Ghoneim, M.A., et al. Randomized trial of cystectomy with or without preoperative radiotherapy for
carcinoma of the bilharzial bladder. J Urol, 1985. 134: 266.
https://pubmed.ncbi.nlm.nih.gov/3894693
269. Anderstrom, C., et al. A prospective randomized study of preoperative irradiation with cystectomy or
cystectomy alone for invasive bladder carcinoma. Eur Urol, 1983. 9: 142.
https://pubmed.ncbi.nlm.nih.gov/6861819
270. Blackard, C.E., et al. Results of a clinical trial of surgery and radiation in stages II and 3 carcinoma of
the bladder. J Urol, 1972. 108: 875.
https://pubmed.ncbi.nlm.nih.gov/5082739
271. Huncharek, M., et al. Planned preoperative radiation therapy in muscle invasive bladder cancer;
results of a meta-analysis. Anticancer Res, 1998. 18: 1931.
https://pubmed.ncbi.nlm.nih.gov/9677446
272. El-Monim, H.A., et al. A prospective randomized trial for postoperative vs. preoperative adjuvant
radiotherapy for muscle-invasive bladder cancer. Urol Oncol, 2013. 31: 359.
https://pubmed.ncbi.nlm.nih.gov/21353794
273. Hautmann, R.E., et al. Urinary diversion. Urology, 2007. 69: 17.
https://pubmed.ncbi.nlm.nih.gov/17280907

74 LIMITED UPDATE MARCH 2021


274. Gore, J.L., et al. Mortality increases when radical cystectomy is delayed more than 12 weeks: results
from a Surveillance, Epidemiology, and End Results-Medicare analysis. Cancer, 2009. 115: 988.
https://pubmed.ncbi.nlm.nih.gov/19142878
275. Russell, B., et al. A Systematic Review and Meta-analysis of Delay in Radical Cystectomy and the
Effect on Survival in Bladder Cancer Patients. Eur Urol Oncol, 2020. 3: 239.
https://pubmed.ncbi.nlm.nih.gov/31668714
276. Hernandez, V., et al. Oncological and functional outcomes of sexual function-preserving cystectomy
compared with standard radical cystectomy in men: A systematic review. Urol Oncol, 2017. 35: 539.
e17.
https://pubmed.ncbi.nlm.nih.gov/28495555
277. Veskimae, E., et al. Systematic review of the oncological and functional outcomes of pelvic organ-
preserving radical cystectomy (RC) compared with standard RC in women who undergo curative
surgery and orthotopic neobladder substitution for bladder cancer. BJU Int, 2017. 120: 12.
https://pubmed.ncbi.nlm.nih.gov/28220653
278. Stenzl, A., et al. Cystectomy – Technical Considerations in Male and Female Patients. EAU Update
Series, 2005. 3: 138.
https://www.sciencedirect.com/science/article/abs/pii/S1570912405000310
279. Mertens, L.S., et al. Prostate sparing cystectomy for bladder cancer: 20-year single center
experience. J Urol, 2014. 191: 1250.
https://pubmed.ncbi.nlm.nih.gov/24286830
280. Kessler, T.M., et al. Attempted nerve sparing surgery and age have a significant effect on urinary
continence and erectile function after radical cystoprostatectomy and ileal orthotopic bladder
substitution. J Urol, 2004. 172: 1323.
https://pubmed.ncbi.nlm.nih.gov/15371833
281. de Vries, R.R., et al. Prostate-sparing cystectomy: long-term oncological results. BJU Int, 2009.
104: 1239.
https://pubmed.ncbi.nlm.nih.gov/19549261
282. Basiri, A., et al. Overall survival and functional results of prostate-sparing cystectomy: a matched
case-control study. Urol J, 2012. 9: 678.
https://pubmed.ncbi.nlm.nih.gov/23235973
283. Wang, X.H., et al. [Impact of preservation of distal prostatic capsula and seminal vesicle on
functions of orthotopic ideal neobladder and erectile function of bladder cancer patients]. Ai Zheng,
2008. 27: 62.
https://pubmed.ncbi.nlm.nih.gov/18184466
284. Moon, H., et al. Nerve and Seminal Sparing Cystectomy for Bladder Cancer. Korean J Urol 2005: 555.
https://www.researchgate.net/publication/291150065
285. Vilaseca, A., et al. Erectile function after cystectomy with neurovascular preservation. Actas Urol
Esp, 2013. 37: 554.
https://pubmed.ncbi.nlm.nih.gov/23790714
286. el-Bahnasawy, M.S., et al. Urethral pressure profile following orthotopic neobladder: differences
between nerve sparing and standard radical cystectomy techniques. J Urol, 2006. 175: 1759.
https://pubmed.ncbi.nlm.nih.gov/16600753
287. Hekal, I.A., et al. Recoverability of erectile function in post-radical cystectomy patients: subjective
and objective evaluations. Eur Urol, 2009. 55: 275.
https://pubmed.ncbi.nlm.nih.gov/18603350
288. Jacobs, B.L., et al. Prostate capsule sparing versus nerve sparing radical cystectomy for bladder
cancer: results of a randomized, controlled trial. J Urol, 2015. 193: 64.
https://pubmed.ncbi.nlm.nih.gov/25066875
289. Colombo, R., et al. Fifteen-year single-centre experience with three different surgical procedures of
nerve-sparing cystectomy in selected organ-confined bladder cancer patients. World J Urol, 2015.
33: 1389.
https://pubmed.ncbi.nlm.nih.gov/25577131
290. Gotsadze, D.T., et al. [Why and how to modify standard cystectomy]. Urologiia, 2008: 22.
https://pubmed.ncbi.nlm.nih.gov/18572764
291. Rozet F, et al. Oncological evaluation of prostate sparing cystectomy: the Montsouris long-term
results. J Urol 2008. 179: 2170.
https://pubmed.ncbi.nlm.nih.gov/18423740
292. Muto, G., et al. Seminal-sparing cystectomy: technical evolution and results over a 20-year period.
Urology, 2014. 83: 856.
https://pubmed.ncbi.nlm.nih.gov/24485363

LIMITED UPDATE MARCH 2021 75


293. Voigt, M., et al. Influence of Simple and Radical Cystectomy on Sexual Function and Pelvic Organ
Prolapse in Female Patients: A Scoping Review of the Literature. Sex Med Rev, 2019. 7: 408.
https://pubmed.ncbi.nlm.nih.gov/31029621
294. Ali-El-Dein, B., et al. Preservation of the internal genital organs during radical cystectomy in selected
women with bladder cancer: a report on 15 cases with long term follow-up. Eur J Surg Oncol, 2013.
39: 358.
https://pubmed.ncbi.nlm.nih.gov/23422323
295. Ali-El-Dein, B., et al. Secondary malignant involvement of gynecologic organs in radical cystectomy
specimens in women: is it mandatory to remove these organs routinely? J Urol, 2004. 172: 885.
https://pubmed.ncbi.nlm.nih.gov/15310990
296. Temkin, S.M., et al. Ovarian Cancer Prevention in High-risk Women. Clin Obstet Gynecol, 2017.
60: 738.
https://pubmed.ncbi.nlm.nih.gov/28957949
297. Bai, S., et al. The Feasibility and Safety of Reproductive Organ Preserving Radical Cystectomy for
Elderly Female Patients With Muscle-Invasive Bladder Cancer: A Retrospective Propensity Score-
matched Study. Urology, 2019. 125: 138.
https://pubmed.ncbi.nlm.nih.gov/30445122
298. Wallmeroth, A., et al. Patterns of metastasis in muscle-invasive bladder cancer (pT2-4): An autopsy
study on 367 patients. Urol Int, 1999. 62: 69.
https://pubmed.ncbi.nlm.nih.gov/10461106
299. Davies, J.D., et al. Anatomic basis for lymph node counts as measure of lymph node dissection
extent: a cadaveric study. Urology, 2013. 81: 358.
https://pubmed.ncbi.nlm.nih.gov/23374802
300. Jensen, J.B., et al. Lymph node mapping in patients with bladder cancer undergoing radical
cystectomy and lymph node dissection to the level of the inferior mesenteric artery. BJU Int, 2010.
106: 199.
https://pubmed.ncbi.nlm.nih.gov/20002670
301. Vazina, A., et al. Stage specific lymph node metastasis mapping in radical cystectomy specimens.
J Urol, 2004. 171: 1830.
https://pubmed.ncbi.nlm.nih.gov/15076287
302. Leissner, J., et al. Extended radical lymphadenectomy in patients with urothelial bladder cancer:
results of a prospective multicenter study. J Urol, 2004. 171: 139.
https://pubmed.ncbi.nlm.nih.gov/14665862
303. Roth, B., et al. A new multimodality technique accurately maps the primary lymphatic landing sites
of the bladder. Eur Urol, 2010. 57: 205.
https://pubmed.ncbi.nlm.nih.gov/19879039
304. Dorin, R.P., et al. Lymph node dissection technique is more important than lymph node count in
identifying nodal metastases in radical cystectomy patients: a comparative mapping study. Eur Urol,
2011. 60: 946.
https://pubmed.ncbi.nlm.nih.gov/21802833
305. Wiesner, C., et al. Cancer-specific survival after radical cystectomy and standardized extended
lymphadenectomy for node-positive bladder cancer: prediction by lymph node positivity and
density. BJU Int, 2009. 104: 331.
https://pubmed.ncbi.nlm.nih.gov/19220265
306. Simone, G., et al. Stage-specific impact of extended versus standard pelvic lymph node dissection
in radical cystectomy. Int J Urol, 2013. 20: 390.
https://pubmed.ncbi.nlm.nih.gov/22970939
307. Holmer, M., et al. Extended lymph node dissection in patients with urothelial cell carcinoma of the
bladder: can it make a difference? World J Urol, 2009. 27: 521.
https://pubmed.ncbi.nlm.nih.gov/19145436
308. Poulsen, A.L., et al. Radical cystectomy: extending the limits of pelvic lymph node dissection
improves survival for patients with bladder cancer confined to the bladder wall. J Urol, 1998.
160: 2015.
https://pubmed.ncbi.nlm.nih.gov/9817313
309. Jensen, J.B., et al. Extended versus limited lymph node dissection in radical cystectomy: impact on
recurrence pattern and survival. Int J Urol, 2012. 19: 39.
https://pubmed.ncbi.nlm.nih.gov/22050425
310. Dhar, N.B., et al. Outcome after radical cystectomy with limited or extended pelvic lymph node
dissection. J Urol, 2008. 179: 873.
https://pubmed.ncbi.nlm.nih.gov/18221953

76 LIMITED UPDATE MARCH 2021


311. Zlotta, A.R. Limited, extended, superextended, megaextended pelvic lymph node dissection at the
time of radical cystectomy: what should we perform? Eur Urol, 2012. 61: 243.
https://pubmed.ncbi.nlm.nih.gov/22119158
312. Zehnder, P., et al. Super extended versus extended pelvic lymph node dissection in patients
undergoing radical cystectomy for bladder cancer: a comparative study. J Urol, 2011. 186: 1261.
https://pubmed.ncbi.nlm.nih.gov/21849183
313. Bruins, H.M., et al. The impact of the extent of lymphadenectomy on oncologic outcomes in patients
undergoing radical cystectomy for bladder cancer: a systematic review. Eur Urol, 2014. 66: 1065.
https://pubmed.ncbi.nlm.nih.gov/25074764
314. Brossner, C., et al. Does extended lymphadenectomy increase the morbidity of radical cystectomy?
BJU Int, 2004. 93: 64.
https://pubmed.ncbi.nlm.nih.gov/14678370
315. Finelli, A., et al. Laparoscopic extended pelvic lymphadenectomy for bladder cancer: technique and
initial outcomes. J Urol, 2004. 172: 1809.
https://pubmed.ncbi.nlm.nih.gov/15540725
316. Abd El Latif, A., et al. Impact of extended versus standard lymph node dissection on overall survival
among patients with urothelial cancer of bladder. J Urol. 187: e707.
https://www.auajournals.org/doi/pdf/10.1016/j.juro.2012.02.1768
317. El-Latif, A.A., et al. 1896 Impact of extended (E) versus standard lymph node dissection (SLND) on
post-cystectomy survival (PCS) among patients withln-negative urothelial bladder cancer (UBC).
J Urol. 185: e759.
https://www.researchgate.net/publication/251483166
318. Abol-Enein, H., et al. Does the extent of lymphadenectomy in radical cystectomy for bladder cancer
influence disease-free survival? A prospective single-center study. Eur Urol, 2011. 60: 572.
https://pubmed.ncbi.nlm.nih.gov/21684070
319. Dharaskar, A., et al. Does extended lymph node dissection affect the lymph node density and
survival after radical cystectomy? Indian J Cancer, 2011. 48: 230.
https://pubmed.ncbi.nlm.nih.gov/21768672
320. Abdollah, F., et al. Stage-specific impact of pelvic lymph node dissection on survival in patients with
non-metastatic bladder cancer treated with radical cystectomy. BJU Int, 2012. 109: 1147.
https://pubmed.ncbi.nlm.nih.gov/21883849
321. Liu, J.-J., et al. 1404 Practice patterns of pelvic lymph node dissection for radical cystectomy from
the veterans affairs central cancer (VACCR). J Urol. 185: e562.
https://www.auajournals.org/doi/pdf/10.1016/j.juro.2011.02.1295
322. Isaka, S., et al. [Pelvic lymph node dissection for invasive bladder cancer]. Nihon Hinyokika Gakkai
Zasshi, 1989. 80: 402.
https://pubmed.ncbi.nlm.nih.gov/2733302
323. Miyakawa, M., et al. [Results of the multidisciplinary treatment of invasive bladder cancer].
Hinyokika Kiyo, 1986. 32: 1931.
https://pubmed.ncbi.nlm.nih.gov/3825830
324. Simone, G., et al. 1755 Extended versus super-extended PLND during radical cystectomy:
comparison of two prospective series. J Urol. 187: e708.
https://www.auajournals.org/doi/full/10.1016/j.juro.2012.02.1771
325. Bostrom, P.J., et al. 1595 Extended lymphadenectomy and chemotherapie offer survival advantage
in muscle-invasive bladder cancer. J Urol. 185: e640.
https://www.auajournals.org/doi/full/10.1016/j.juro.2011.02.1645
326. Yuasa, M., et al. [Clinical evaluation of total cystectomy for bladder carcinoma: a ten-year
experience]. Hinyokika Kiyo, 1988. 34: 975.
https://pubmed.ncbi.nlm.nih.gov/3223462
327. Mandel, P., et al. Extent of lymph node dissection and recurrence-free survival after radical
cystectomy: a meta-analysis. Urol Oncol, 2014. 32: 1184.
https://pubmed.ncbi.nlm.nih.gov/25027683
328. Bi, L., et al. Extended vs non-extended pelvic lymph node dissection and their influence on
recurrence-free survival in patients undergoing radical cystectomy for bladder cancer: a systematic
review and meta-analysis of comparative studies. BJU Int, 2014. 113: E39.
https://pubmed.ncbi.nlm.nih.gov/24053715
329. Gschwend, J.E., et al. Extended Versus Limited Lymph Node Dissection in Bladder Cancer Patients
Undergoing Radical Cystectomy: Survival Results from a Prospective, Randomized Trial. Eur Urol,
2019. 75: 604.
https://pubmed.ncbi.nlm.nih.gov/30337060

LIMITED UPDATE MARCH 2021 77


330. Koppie, T.M., et al. Standardization of pelvic lymphadenectomy performed at radical cystectomy: can
we establish a minimum number of lymph nodes that should be removed? Cancer, 2006. 107: 2368.
https://pubmed.ncbi.nlm.nih.gov/17041887
331. Wright, J.L., et al. The association between extent of lymphadenectomy and survival among
patients with lymph node metastases undergoing radical cystectomy. Cancer, 2008. 112: 2401.
https://pubmed.ncbi.nlm.nih.gov/18383515
332. Zehnder, P., et al. Radical cystectomy with super-extended lymphadenectomy: impact of separate
vs en bloc lymph node submission on analysis and outcomes. BJU Int, 2016. 117: 253.
https://pubmed.ncbi.nlm.nih.gov/25307941
333. Wang, Y.C., et al. Extended versus non-extended lymphadenectomy during radical cystectomy for
patients with bladder cancer: a meta-analysis of the effect on long-term and short-term outcomes.
World J Surg Oncol, 2019. 17: 225.
https://pubmed.ncbi.nlm.nih.gov/31864368
334. Faraj, K., et al. Robotic vs. open cystectomy: How length-of-stay differences relate conditionally to
age. Urol Oncol, 2019. 37: 354.e1.
https://pubmed.ncbi.nlm.nih.gov/30770298
335. Rai, B.P., et al. Robotic versus open radical cystectomy for bladder cancer in adults. Cochrane
Database Syst Rev, 2019. 4: Cd011903.
https://pubmed.ncbi.nlm.nih.gov/31016718
336. Wilson, T.G., et al. Best practices in robot-assisted radical cystectomy and urinary reconstruction:
recommendations of the Pasadena Consensus Panel. Eur Urol, 2015. 67: 363.
337. Al Hussein Al Awamlh, B., et al. The safety of robot-assisted cystectomy in patients with previous
history of pelvic irradiation. BJU Int, 2016. 118: 437.
https://pubmed.ncbi.nlm.nih.gov/25582930
338. Goh, A.C., et al. A Population-based Study of Ureteroenteric Strictures After Open and Robot-
assisted Radical Cystectomy. Urology, 2020. 135: 57.
https://pubmed.ncbi.nlm.nih.gov/31618656
339. Venkatramani, V., et al. Predictors of Recurrence, and Progression-Free and Overall Survival
following Open versus Robotic Radical Cystectomy: Analysis from the RAZOR Trial with a 3-Year
Followup. J Urol, 2020. 203: 522.
https://pubmed.ncbi.nlm.nih.gov/31549935
340. Faraj, K.S., et al. Robot Assisted Radical Cystectomy vs Open Radical Cystectomy: Over 10 years
of the Mayo Clinic Experience. Urol Oncol, 2019. 37: 862.
https://pubmed.ncbi.nlm.nih.gov/31526651
341. Mantica, G., et al. Port-site metastasis and atypical recurrences after robotic-assisted radical
cystectomy (RARC): an updated comprehensive and systematic review of current evidences.
J Robot Surg, 2020.
https://pubmed.ncbi.nlm.nih.gov/32152900
342. Wei, L., et al. Accurate Quantification of Residual Cancer Cells in Pelvic Washing Reveals
Association with Cancer Recurrence Following Robot-Assisted Radical Cystectomy. J Urol, 2019.
201: 1105.
https://pubmed.ncbi.nlm.nih.gov/30730413
343. Parekh, D.J., et al. Robot-assisted radical cystectomy versus open radical cystectomy in patients
with bladder cancer (RAZOR): an open-label, randomised, phase 3, non-inferiority trial. Lancet,
2018. 391: 2525.
https://pubmed.ncbi.nlm.nih.gov/29976469
344. Hussein, A.A., et al. Outcomes of Intracorporeal Urinary Diversion after Robot-Assisted Radical
Cystectomy: Results from the International Robotic Cystectomy Consortium. J Urol, 2018. 199: 1302.
https://pubmed.ncbi.nlm.nih.gov/29275112
345. Zhang, J.H., et al. Large Single Institution Comparison of Perioperative Outcomes and
Complications of Open Radical Cystectomy, Intracorporeal Robot-Assisted Radical Cystectomy and
Robotic Extracorporeal Approach. J Urol, 2020. 203: 512.
https://pubmed.ncbi.nlm.nih.gov/31580189
346. Tang, K., et al. Laparoscopic versus open radical cystectomy in bladder cancer: a systematic review
and meta-analysis of comparative studies. PLoS One, 2014. 9: e95667.
https://pubmed.ncbi.nlm.nih.gov/24835573
347. Albisinni, S., et al. Long-term analysis of oncological outcomes after laparoscopic radical
cystectomy in Europe: results from a multicentre study by the European Association of Urology
(EAU) section of Uro-technology. BJU Int, 2015. 115: 937.
https://pubmed.ncbi.nlm.nih.gov/25294421

78 LIMITED UPDATE MARCH 2021


348. Khan, M.S., et al. A Single-centre Early Phase Randomised Controlled Three-arm Trial of Open,
Robotic, and Laparoscopic Radical Cystectomy (CORAL). Eur Urol, 2016. 69: 613.
https://pubmed.ncbi.nlm.nih.gov/26272237
349. Khan, M.S., et al. Long-term Oncological Outcomes from an Early Phase Randomised Controlled
Three-arm Trial of Open, Robotic, and Laparoscopic Radical Cystectomy (CORAL). Eur Urol, 2020.
77: 110.
https://pubmed.ncbi.nlm.nih.gov/31740072
350. Stenzl, A. Bladder substitution. Curr Opin Urol, 1999. 9: 241.
https://pubmed.ncbi.nlm.nih.gov/10726098
351. Yang, L.S., et al. A systematic review and meta-analysis of quality of life outcomes after radical
cystectomy for bladder cancer. Surg Oncol, 2016. 25: 281.
https://pubmed.ncbi.nlm.nih.gov/27566035
352. Check, D.K., et al. Decision Regret Related to Urinary Diversion Choice among Patients Treated with
Cystectomy. J Urol, 2020. 203: 159.
https://pubmed.ncbi.nlm.nih.gov/31441673
353. Roth, B., et al. Positive Pre-cystectomy Biopsies of the Prostatic Urethra or Bladder Neck Do Not
Necessarily Preclude Orthotopic Bladder Substitution. J Urol, 2019. 201: 909.
https://pubmed.ncbi.nlm.nih.gov/30694935
354. Lebret, T., et al. After cystectomy, is it justified to perform a bladder replacement for patients with
lymph node positive bladder cancer? Eur Urol, 2002. 42: 344.
https://pubmed.ncbi.nlm.nih.gov/12361899
355. Gerharz, E.W., et al. Metabolic and functional consequences of urinary reconstruction with bowel.
BJU Int, 2003. 91: 143.
https://pubmed.ncbi.nlm.nih.gov/12519116
356. Madersbacher, S., et al. Contemporary cystectomy and urinary diversion. World J Urol, 2002. 20: 151.
https://pubmed.ncbi.nlm.nih.gov/12196898
357. Pruthi, R.S., et al. Fast track program in patients undergoing radical cystectomy: results in 362
consecutive patients. J Am Coll Surg, 2010. 210: 93.
https://pubmed.ncbi.nlm.nih.gov/20123338
358. Kouba, E.J., et al. Gum chewing stimulates bowel motility in patients undergoing radical cystectomy
with urinary diversion. Urology, 2007. 70: 1053.
https://pubmed.ncbi.nlm.nih.gov/18158012
359. Karl, A., et al. A new concept for early recovery after surgery for patients undergoing radical
cystectomy for bladder cancer: results of a prospective randomized study. J Urol, 2014. 191: 335.
https://pubmed.ncbi.nlm.nih.gov/23968966
360. Xu, W., et al. Postoperative Pain Management after Radical Cystectomy: Comparing Traditional
versus Enhanced Recovery Protocol Pathway. J Urol, 2015. 194: 1209.
https://pubmed.ncbi.nlm.nih.gov/26021824
361. Lee, C.T., et al. Alvimopan accelerates gastrointestinal recovery after radical cystectomy: a
multicenter randomized placebo-controlled trial. Eur Urol, 2014. 66: 265.
https://pubmed.ncbi.nlm.nih.gov/24630419
362. Chiang, H.A., et al. Implementation of a Perioperative Venous Thromboembolism Prophylaxis
Program for Patients Undergoing Radical Cystectomy on an Enhanced Recovery After Surgery
Protocol. Eur Urol Focus, 2018. 6: 74.
https://pubmed.ncbi.nlm.nih.gov/30228076
363. Brennan, K., et al. Venous Thromboembolism and Peri-Operative Chemotherapy for Muscle-Invasive
Bladder Cancer: A Population-based Study. Bladder Cancer, 2018. 4: 419.
https://pubmed.ncbi.nlm.nih.gov/30417053
364. Tikkinen , K.A.O., et al. EAU Guidelines Thromboprophylaxis in Urological Surgery. Edn presented at
the 32th EAU Annual Congress London, 2017. EAU Guidelines Office, Arnhem, The Netherlands.
https://uroweb.org/guideline/thromboprophylaxis/
365. Hautmann, R.E., et al. Long-term results of standard procedures in urology: the ileal neobladder.
World J Urol, 2006. 24: 305.
https://pubmed.ncbi.nlm.nih.gov/16830152
366. Hautmann, R.E., et al. Lessons learned from 1,000 neobladders: the 90-day complication rate.
J Urol, 2010. 184: 990.
https://pubmed.ncbi.nlm.nih.gov/20643429
367. Stein, J.P., et al. Indications and technique of the orthotopic neobladder in women. Urol Clin North
Am, 2002. 29: 725.
https://pubmed.ncbi.nlm.nih.gov/12476536

LIMITED UPDATE MARCH 2021 79


368. Hautmann, R.E., et al. Radical cystectomy for urothelial carcinoma of the bladder without
neoadjuvant or adjuvant therapy: long-term results in 1100 patients. Eur Urol, 2012. 61: 1039.
https://pubmed.ncbi.nlm.nih.gov/22381169
369. Jentzmik, F., et al. The ileal neobladder in female patients with bladder cancer: long-term clinical,
functional, and oncological outcome. World J Urol, 2012. 30: 733.
https://pubmed.ncbi.nlm.nih.gov/22322390
370. Ahmadi, H., et al. Urinary functional outcome following radical cystoprostatectomy and ileal
neobladder reconstruction in male patients. J Urol, 2013. 189: 1782.
https://pubmed.ncbi.nlm.nih.gov/23159582
371. Neuzillet, Y., et al. The Z-shaped ileal neobladder after radical cystectomy: an 18 years experience
with 329 patients. BJU Int, 2011. 108: 596.
https://pubmed.ncbi.nlm.nih.gov/21223470
372. Gershman, B., et al. Comparative impact of continent and incontinent urinary diversion on long-term
renal function after radical cystectomy in patients with preoperative chronic kidney disease 2 and
chronic kidney disease 3a. Int J Urol, 2015. 22: 651.
https://pubmed.ncbi.nlm.nih.gov/25881721
373. Longo, N., et al. Complications and quality of life in elderly patients with several comorbidities
undergoing cutaneous ureterostomy with single stoma or ileal conduit after radical cystectomy. BJU
Int, 2016. 118: 521.
https://pubmed.ncbi.nlm.nih.gov/26935245
374. Deliveliotis, C., et al. Urinary diversion in high-risk elderly patients: modified cutaneous ureterostomy
or ileal conduit? Urology, 2005. 66: 299.
https://pubmed.ncbi.nlm.nih.gov/16040096
375. Kilciler, M., et al. Comparison of ileal conduit and transureteroureterostomy with
ureterocutaneostomy urinary diversion. Urol Int, 2006. 77: 245.
https://pubmed.ncbi.nlm.nih.gov/17033213
376. Figueroa, A.J., et al. Radical cystectomy for elderly patients with bladder carcinoma: an updated
experience with 404 patients. Cancer, 1998. 83: 141.
https://pubmed.ncbi.nlm.nih.gov/9655304
377. Berger, I., et al. Impact of the use of bowel for urinary diversion on perioperative complications and
90-day mortality in patients aged 75 years or older. Urol Int, 2015. 94: 394.
https://pubmed.ncbi.nlm.nih.gov/25612612
378. Nieuwenhuijzen, J.A., et al. Urinary diversions after cystectomy: the association of clinical factors,
complications and functional results of four different diversions. Eur Urol, 2008. 53: 834.
https://pubmed.ncbi.nlm.nih.gov/17904276
379. Wood, D.N., et al. Stomal complications of ileal conduits are significantly higher when formed in
women with intractable urinary incontinence. J Urol, 2004. 172: 2300.
https://pubmed.ncbi.nlm.nih.gov/15538253
380. Neal, D.E. Complications of ileal conduit diversion in adults with cancer followed up for at least five
years. Br Med J (Clin Res Ed), 1985. 290: 1695.
https://pubmed.ncbi.nlm.nih.gov/3924218
381. Mues, A.C., et al. Contemporary experience in the management of angiomyolipoma. J Endourol,
2010. 24: 1883.
https://pubmed.ncbi.nlm.nih.gov/20919915
382. Donat, S.M., et al. Radical cystectomy in octogenarians--does morbidity outweigh the potential
survival benefits? J Urol, 2010. 183: 2171.
https://pubmed.ncbi.nlm.nih.gov/20399461
383. Hautmann, R.E., et al. 25 years of experience with 1,000 neobladders: long-term complications.
J Urol, 2011. 185: 2207.
https://pubmed.ncbi.nlm.nih.gov/21497841
384. Stein, J.P., et al. The orthotopic T pouch ileal neobladder: experience with 209 patients. J Urol,
2004. 172: 584.
https://pubmed.ncbi.nlm.nih.gov/15247737
385. Abol-Enein, H., et al. Functional results of orthotopic ileal neobladder with serous-lined extramural
ureteral reimplantation: experience with 450 patients. J Urol, 2001. 165: 1427.
https://pubmed.ncbi.nlm.nih.gov/11342891
386. Stein, J.P., et al. Results with radical cystectomy for treating bladder cancer: a ‘reference standard’
for high-grade, invasive bladder cancer. BJU Int, 2003. 92: 12.
https://pubmed.ncbi.nlm.nih.gov/12823375

80 LIMITED UPDATE MARCH 2021


387. Yossepowitch, O., et al. Orthotopic urinary diversion after cystectomy for bladder cancer:
implications for cancer control and patterns of disease recurrence. J Urol, 2003. 169: 177.
https://pubmed.ncbi.nlm.nih.gov/12478130
388. Stein, J.P., et al. Urethral tumor recurrence following cystectomy and urinary diversion: clinical and
pathological characteristics in 768 male patients. J Urol, 2005. 173: 1163.
https://pubmed.ncbi.nlm.nih.gov/15758728
389. Gerharz, E.W., et al. Quality of life after cystectomy and urinary diversion: an evidence based
analysis. J Urol, 2005. 174: 1729.
https://pubmed.ncbi.nlm.nih.gov/16217273
390. Porter, M.P., et al. Health related quality of life after radical cystectomy and urinary diversion for
bladder cancer: a systematic review and critical analysis of the literature. J Urol, 2005. 173: 1318.
https://pubmed.ncbi.nlm.nih.gov/15758789
391. Wiesner, C., et al. Continent cutaneous urinary diversion: long-term follow-up of more than 800
patients with ileocecal reservoirs. World J Urol, 2006. 24: 315.
https://pubmed.ncbi.nlm.nih.gov/16676186
392. Thoeny, H.C., et al. Is ileal orthotopic bladder substitution with an afferent tubular segment
detrimental to the upper urinary tract in the long term? J Urol, 2002. 168: 2030.
https://pubmed.ncbi.nlm.nih.gov/12394702
393. Gakis, G., et al. [Benefits and risks of orthotopic neobladder reconstruction in female patients].
Aktuelle Urol, 2011. 42: 109.
https://pubmed.ncbi.nlm.nih.gov/21437834
394. Stein, J.P., et al. Pathological guidelines for orthotopic urinary diversion in women with bladder
cancer: a review of the literature. J Urol, 2007. 178: 756.
https://pubmed.ncbi.nlm.nih.gov/17631333
395. Stein, J.P., et al. Indications for lower urinary tract reconstruction in women after cystectomy for
bladder cancer: a pathological review of female cystectomy specimens. J Urol, 1995. 154: 1329.
https://pubmed.ncbi.nlm.nih.gov/7658531
396. Vallancien, G., et al. Cystectomy with prostate sparing for bladder cancer in 100 patients: 10-year
experience. J Urol, 2002. 168: 2413.
https://pubmed.ncbi.nlm.nih.gov/12441929
397. Stenzl, A., et al. Radical cystectomy with orthotopic neobladder for invasive bladder cancer: a critical
analysis of long term oncological, functional and quality of life results. Int Braz J Urol, 2010. 36: 537.
https://pubmed.ncbi.nlm.nih.gov/21044370
398. Nielsen, M.E., et al. Association of hospital volume with conditional 90-day mortality after
cystectomy: an analysis of the National Cancer Data Base. BJU Int, 2014. 114: 46.
https://pubmed.ncbi.nlm.nih.gov/24219110
399. Porter, M.P., et al. Hospital volume and 90-day mortality risk after radical cystectomy: a population-
based cohort study. World J Urol, 2011. 29: 73.
https://pubmed.ncbi.nlm.nih.gov/21132553
400. Hautmann, R.E., et al. ICUD-EAU International Consultation on Bladder Cancer 2012: Urinary
diversion. Eur Urol, 2013. 63: 67.
https://pubmed.ncbi.nlm.nih.gov/22995974
401. Cookson, M.S., et al. Complications of radical cystectomy for nonmuscle invasive disease:
comparison with muscle invasive disease. J Urol, 2003. 169: 101.
https://pubmed.ncbi.nlm.nih.gov/12478113
402. Sabir, E.F., et al. Impact of hospital volume on local recurrence and distant metastasis in bladder
cancer patients treated with radical cystectomy in Sweden. Scand J Urol, 2013. 47: 483.
https://pubmed.ncbi.nlm.nih.gov/23590830
403. Morgan, T.M., et al. Volume outcomes of cystectomy--is it the surgeon or the setting? J Urol, 2012.
188: 2139.
https://pubmed.ncbi.nlm.nih.gov/23083864
404. Finks, J.F., et al. Trends in hospital volume and operative mortality for high-risk surgery. N Engl
J Med, 2011. 364: 2128.
https://pubmed.ncbi.nlm.nih.gov/21631325
405. Corcoran, A.T., et al. Variation in performance of candidate surgical quality measures for muscle-
invasive bladder cancer by hospital type. BJU Int, 2015. 115: 230.
https://pubmed.ncbi.nlm.nih.gov/24447637
406. Ravi, P., et al. Benefit in regionalisation of care for patients treated with radical cystectomy: a
nationwide inpatient sample analysis. BJU Int, 2014. 113: 733.
https://pubmed.ncbi.nlm.nih.gov/24007240

LIMITED UPDATE MARCH 2021 81


407. Shabsigh, A., et al. Defining early morbidity of radical cystectomy for patients with bladder cancer
using a standardized reporting methodology. Eur Urol, 2009. 55: 164.
https://pubmed.ncbi.nlm.nih.gov/18675501
408. Wang, Y.L., et al. Perioperative Blood Transfusion Promotes Worse Outcomes of Bladder Cancer
after Radical Cystectomy: A Systematic Review and Meta-Analysis. PLoS One, 2015. 10: e0130122.
https://pubmed.ncbi.nlm.nih.gov/26080092
409. Buchner, A., et al. Dramatic impact of blood transfusion on cancer-specific survival after radical
cystectomy irrespective of tumor stage. Scand J Urol, 2017. 51: 130.
https://pubmed.ncbi.nlm.nih.gov/28332428
410. Zaid, H.B., et al. Efficacy and Safety of Intraoperative Tranexamic Acid Infusion for Reducing Blood
Transfusion During Open Radical Cystectomy. Urology, 2016. 92: 57.
https://pubmed.ncbi.nlm.nih.gov/26968489
411. Hammond, J., et al. Rates of venous thromboembolism among patients with major surgery for
cancer. Ann Surg Oncol, 2011. 18: 3240.
https://pubmed.ncbi.nlm.nih.gov/21584837
412. Potretzke, A.M., et al. Highest risk of symptomatic venous thromboembolic events after radical
cystectomy occurs in patients with obesity or nonurothelial cancers. Urol Ann, 2015. 7: 355.
https://pubmed.ncbi.nlm.nih.gov/26229325
413. Shariat, S.F., et al. Outcomes of radical cystectomy for transitional cell carcinoma of the bladder: a
contemporary series from the Bladder Cancer Research Consortium. J Urol, 2006. 176: 2414.
https://pubmed.ncbi.nlm.nih.gov/17085118
414. Nuhn, P., et al. External validation of postoperative nomograms for prediction of all-cause mortality,
cancer-specific mortality, and recurrence in patients with urothelial carcinoma of the bladder. Eur
Urol, 2012. 61: 58.
https://pubmed.ncbi.nlm.nih.gov/21840642
415. Madersbacher, S., et al. Radical cystectomy for bladder cancer today--a homogeneous series
without neoadjuvant therapy. J Clin Oncol, 2003. 21: 690.
https://pubmed.ncbi.nlm.nih.gov/12586807
416. Bruins, H.M., et al. Clinical outcomes and recurrence predictors of lymph node positive urothelial
cancer after cystectomy. J Urol, 2009. 182: 2182.
https://pubmed.ncbi.nlm.nih.gov/19758623
417. Abdollah, F., et al. Incidence, survival and mortality rates of stage-specific bladder cancer in United
States: a trend analysis. Cancer Epidemiol, 2013. 37: 219.
https://pubmed.ncbi.nlm.nih.gov/23485480
418. Bruins, H.M., et al. The Importance of Hospital and Surgeon Volume as Major Determinants of
Morbidity and Mortality After Radical Cystectomy for Bladder Cancer: A Systematic Review and
Recommendations by the European Association of Urology Muscle-invasive and Metastatic Bladder
Cancer Guideline Panel. Eur Urol Oncol, 2020. 3: 131.
https://pubmed.ncbi.nlm.nih.gov/31866215
419. Ok, J.H., et al. Medical and surgical palliative care of patients with urological malignancies. J Urol,
2005. 174: 1177.
https://pubmed.ncbi.nlm.nih.gov/16145365
420. Ubrig, B., et al. Extraperitoneal bilateral cutaneous ureterostomy with midline stoma for palliation of
pelvic cancer. Urology, 2004. 63: 973.
https://pubmed.ncbi.nlm.nih.gov/15134993
421. Zebic, N., et al. Radical cystectomy in patients aged > or = 75 years: an updated review of patients
treated with curative and palliative intent. BJU Int, 2005. 95: 1211.
https://pubmed.ncbi.nlm.nih.gov/15892803
422. El-Tabey, N.A., et al. Bladder cancer with obstructive uremia: oncologic outcome after definitive
surgical management. Urology, 2005. 66: 531.
https://pubmed.ncbi.nlm.nih.gov/16140072
423. Nagele, U., et al. The rationale for radical cystectomy as primary therapy for T4 bladder cancer.
World J Urol, 2007. 25: 401.
https://pubmed.ncbi.nlm.nih.gov/17525849
424. Ghahestani, S.M., et al. Palliative treatment of intractable hematuria in context of advanced bladder
cancer: a systematic review. Urol J, 2009. 6: 149.
https://pubmed.ncbi.nlm.nih.gov/19711266
425. Srinivasan, V., et al. A comparison of two radiotherapy regimens for the treatment of symptoms from
advanced bladder cancer. Clin Oncol (R Coll Radiol), 1994. 6: 11.
https://pubmed.ncbi.nlm.nih.gov/7513538

82 LIMITED UPDATE MARCH 2021


426. Herr, H.W. Conservative management of muscle-infiltrating bladder cancer: prospective experience.
J Urol, 1987. 138: 1162.
https://pubmed.ncbi.nlm.nih.gov/3669160
427. Herr, H.W. Transurethral resection of muscle-invasive bladder cancer: 10-year outcome. J Clin
Oncol, 2001. 19: 89.
https://pubmed.ncbi.nlm.nih.gov/11134199
428. Holmäng, S., et al. Long-term followup of all patients with muscle invasive (stages T2, T3 and T4)
bladder carcinoma in a geographical region. J Urol, 1997. 158: 389.
https://pubmed.ncbi.nlm.nih.gov/9224309
429. Solsona, E., et al. Feasibility of radical transurethral resection as monotherapy for selected patients
with muscle invasive bladder cancer. J Urol, 2010. 184: 475.
https://pubmed.ncbi.nlm.nih.gov/20620402
430. Choudhury, A., et al. Hypofractionated radiotherapy in locally advanced bladder cancer: an
individual patient data meta-analysis of the BC2001 and BCON trials. Lancet Oncol, 2021. 22: 246.
https://pubmed.ncbi.nlm.nih.gov/33539743
431. Korpics, M., et al. Maximizing survival in patients with muscle-invasive bladder cancer undergoing
curative bladder-preserving radiotherapy: the impact of radiotherapy dose escalation. J Radiat
Oncol, 2017. 6: 387.
https://www.researchgate.net/publication/318459921
432. Hafeez, S., et al. Clinical Outcomes of Image Guided Adaptive Hypofractionated Weekly Radiation
Therapy for Bladder Cancer in Patients Unsuitable for Radical Treatment. Int J Radiat Oncol Biol
Phys, 2017. 98: 115.
https://pubmed.ncbi.nlm.nih.gov/28586948
433. Milosevic, M., et al. Radiotherapy for bladder cancer. Urology, 2007. 69: 80.
https://pubmed.ncbi.nlm.nih.gov/17280910
434. Sondergaard, J., et al. A comparison of morbidity following conformal versus intensity-modulated
radiotherapy for urinary bladder cancer. Acta Oncol, 2014. 53: 1321.
https://pubmed.ncbi.nlm.nih.gov/24980045
435. Tonoli, S., et al. Radical radiotherapy for bladder cancer: retrospective analysis of a series of 459
patients treated in an Italian institution. Clin Oncol (R Coll Radiol), 2006. 18: 52.
https://pubmed.ncbi.nlm.nih.gov/16477920
436. Shelley, M.D., et al. Surgery versus radiotherapy for muscle invasive bladder cancer. Cochrane
Database Syst Rev, 2002: Cd002079.
https://pubmed.ncbi.nlm.nih.gov/11869621
437. Booth, C.M., et al. Curative therapy for bladder cancer in routine clinical practice: a population-
based outcomes study. Clin Oncol (R Coll Radiol), 2014. 26: 506.
https://pubmed.ncbi.nlm.nih.gov/24954284
438. Korpics, M.C., et al. Concurrent chemotherapy is associated with improved survival in elderly
patients with bladder cancer undergoing radiotherapy. Cancer, 2017. 123: 3524.
https://pubmed.ncbi.nlm.nih.gov/28581675
439. Scher, H.I., et al. Neoadjuvant M-VAC (methotrexate, vinblastine, doxorubicin and cisplatin) effect on
the primary bladder lesion. J Urol, 1988. 139: 470.
https://pubmed.ncbi.nlm.nih.gov/3343728
440. Herr, H.W., et al. Neoadjuvant chemotherapy and bladder-sparing surgery for invasive bladder
cancer: ten-year outcome. J Clin Oncol, 1998. 16: 1298.
https://pubmed.ncbi.nlm.nih.gov/9552029
441. Sternberg, C.N., et al. Can patient selection for bladder preservation be based on response to
chemotherapy? Cancer, 2003. 97: 1644.
https://pubmed.ncbi.nlm.nih.gov/12655521
442. Kachnic, L.A., et al. Bladder preservation by combined modality therapy for invasive bladder cancer.
J Clin Oncol, 1997. 15: 1022.
https://pubmed.ncbi.nlm.nih.gov/9060542
443. Als, A.B., et al. Long-term survival after gemcitabine and cisplatin in patients with locally advanced
transitional cell carcinoma of the bladder: focus on supplementary treatment strategies. Eur Urol,
2007. 52: 478.
https://pubmed.ncbi.nlm.nih.gov/17383078
444. Audenet, F., et al. Effectiveness of Transurethral Resection plus Systemic Chemotherapy as
Definitive Treatment for Muscle Invasive Bladder Cancer in Population Level Data. J Urol, 2018.
200: 996.
https://pubmed.ncbi.nlm.nih.gov/29879397

LIMITED UPDATE MARCH 2021 83


445. James, N.D., et al. Radiotherapy with or without chemotherapy in muscle-invasive bladder cancer.
New Engl J Med, 2012. 366: 1477.
https://pubmed.ncbi.nlm.nih.gov/22512481
446. Efstathiou, J.A., et al. Long-term outcomes of selective bladder preservation by combined-modality
therapy for invasive bladder cancer: the MGH experience. Eur Urol, 2012. 61: 705.
https://pubmed.ncbi.nlm.nih.gov/22101114
447. Giacalone, N.J., et al. Long-term Outcomes After Bladder-preserving Tri-modality Therapy for
Patients with Muscle-invasive Bladder Cancer: An Updated Analysis of the Massachusetts General
Hospital Experience. Eur Urol, 2017. 71: 952.
https://pubmed.ncbi.nlm.nih.gov/28081860
448. Mak, R.H., et al. Long-term outcomes in patients with muscle-invasive bladder cancer after
selective bladder-preserving combined-modality therapy: a pooled analysis of Radiation Therapy
Oncology Group protocols 8802, 8903, 9506, 9706, 9906, and 0233. J Clin Oncol, 2014. 32: 3801.
https://pubmed.ncbi.nlm.nih.gov/25366678
449. Suer, E., et al. Significance of second transurethral resection on patient outcomes in muscle-
invasive bladder cancer patients treated with bladder-preserving multimodal therapy. World J Urol,
2016. 34: 847.
https://pubmed.ncbi.nlm.nih.gov/26462931
450. Ploussard, G., et al. Critical analysis of bladder sparing with trimodal therapy in muscle-invasive
bladder cancer: a systematic review. Eur Urol, 2014. 66: 120.
https://pubmed.ncbi.nlm.nih.gov/24613684
451. Amestoy, F., et al. Review of hypo-fractionated radiotherapy for localized muscle invasive bladder
cancer. Crit Rev Oncol Hematol, 2019. 142: 76.
https://pubmed.ncbi.nlm.nih.gov/31377435
452. Hoskin, P.J., et al. Radiotherapy with concurrent carbogen and nicotinamide in bladder carcinoma.
J Clin Oncol, 2010. 28: 4912.
https://pubmed.ncbi.nlm.nih.gov/20956620
453. Coen, J.J., et al. Bladder Preservation With Twice-a-Day Radiation Plus Fluorouracil/Cisplatin or
Once Daily Radiation Plus Gemcitabine for Muscle-Invasive Bladder Cancer: NRG/RTOG 0712-A
Randomized Phase II Trial. J Clin Oncol, 2019. 37: 44.
https://pubmed.ncbi.nlm.nih.gov/30433852
454. Ramani, V.A., et al. Differential complication rates following radical cystectomy in the irradiated and
nonirradiated pelvis. Eur Urol, 2010. 57: 1058.
https://pubmed.ncbi.nlm.nih.gov/20022162
455. Kulkarni, G.S., et al. Propensity Score Analysis of Radical Cystectomy Versus Bladder-Sparing
Trimodal Therapy in the Setting of a Multidisciplinary Bladder Cancer Clinic. J Clin Oncol, 2017.
35: 2299.
https://pubmed.ncbi.nlm.nih.gov/28410011
456. Krasnow, R.E., et al. Clinical Outcomes of Patients with Histologic Variants of Urothelial Cancer
Treated with Trimodality Bladder-sparing Therapy. Eur Urol, 2017. 72: 54.
https://pubmed.ncbi.nlm.nih.gov/28040351
457. Cohen, S.M., et al. The role of perioperative chemotherapy in the treatment of urothelial cancer.
Oncologist, 2006. 11: 630.
https://pubmed.ncbi.nlm.nih.gov/16794242
458. Eswara, J.R., et al. Complications and long-term results of salvage cystectomy after failed bladder
sparing therapy for muscle invasive bladder cancer. J Urol, 2012. 187: 463.
https://pubmed.ncbi.nlm.nih.gov/22177159
459. Mitin, T., et al. Long-Term Outcomes Among Patients Who Achieve Complete or Near-Complete
Responses After the Induction Phase of Bladder-Preserving Combined-Modality Therapy for
Muscle-Invasive Bladder Cancer: A Pooled Analysis of NRG Oncology/RTOG 9906 and 0233. Int
J Radiat Oncol Biol Phys, 2016. 94: 67.
https://pubmed.ncbi.nlm.nih.gov/26700703
460. Sanchez, A., et al. Incidence, Clinicopathological Risk Factors, Management and Outcomes of
Nonmuscle Invasive Recurrence after Complete Response to Trimodality Therapy for Muscle
Invasive Bladder Cancer. J Urol, 2018. 199: 407.
https://pubmed.ncbi.nlm.nih.gov/28870862
461. Ritch, C.R., et al. Propensity matched comparative analysis of survival following chemoradiation or
radical cystectomy for muscle-invasive bladder cancer. BJU Int, 2018. 121: 745.
https://pubmed.ncbi.nlm.nih.gov/29281848

84 LIMITED UPDATE MARCH 2021


462. Cahn, D.B., et al. Contemporary use trends and survival outcomes in patients undergoing radical
cystectomy or bladder-preservation therapy for muscle-invasive bladder cancer. Cancer, 2017.
123: 4337.
https://pubmed.ncbi.nlm.nih.gov/28743162
463. Williams, S.B., et al. Comparing Survival Outcomes and Costs Associated With Radical Cystectomy
and Trimodal Therapy for Older Adults With Muscle-Invasive Bladder Cancer. JAMA Surg, 2018.
153: 881.
https://pubmed.ncbi.nlm.nih.gov/29955780
464. Fahmy, O., et al. A systematic review and meta-analysis on the oncological long-term outcomes
after trimodality therapy and radical cystectomy with or without neoadjuvant chemotherapy for
muscle-invasive bladder cancer. Urol Oncol, 2018. 36: 43.
https://pubmed.ncbi.nlm.nih.gov/29102254
465. Efstathiou, J.A., et al. Late pelvic toxicity after bladder-sparing therapy in patients with invasive
bladder cancer: RTOG 89-03, 95-06, 97-06, 99-06. J Clin Oncol, 2009. 27: 4055.
https://pubmed.ncbi.nlm.nih.gov/19636019
466. Huddart, R.A., et al. Patient-reported Quality of Life Outcomes in Patients Treated for Muscle-
invasive Bladder Cancer with Radiotherapy ± Chemotherapy in the BC2001 Phase III Randomised
Controlled Trial. Eur Urol, 2020. 77: 260.
https://pubmed.ncbi.nlm.nih.gov/31843338
467. Mak, K.S., et al. Quality of Life in Long-term Survivors of Muscle-Invasive Bladder Cancer. Int
J Radiat Oncol Biol Phys, 2016. 96: 1028.
https://pubmed.ncbi.nlm.nih.gov/27727064
468. Sherry, A.D., et al. Intensity-modulated radiotherapy is superior to three-dimensional conformal
radiotherapy in the trimodality management of muscle-invasive bladder cancer with daily cone
beam computed tomography optimization. J Radiat Oncol, 2019. 8: 395.
https://pubmed.ncbi.nlm.nih.gov/33343830
469. Quirt, J.S., et al. Patterns of Referral to Radiation Oncology among Patients with Bladder Cancer: a
Population-based Study. Clin Oncol (R Coll Radiol), 2017. 29: 171.
https://pubmed.ncbi.nlm.nih.gov/27829531
470. Sylvester, R., et al. The role of adjuvant combination chemotherapy after cystectomy in locally
advanced bladder cancer: what we do not know and why. Ann Oncol, 2000. 11: 851.
https://pubmed.ncbi.nlm.nih.gov/10997813
471. Donat, S.M., et al. Potential impact of postoperative early complications on the timing of adjuvant
chemotherapy in patients undergoing radical cystectomy: a high-volume tertiary cancer center
experience. Eur Urol, 2009. 55: 177.
https://pubmed.ncbi.nlm.nih.gov/18640770
472. Advanced Bladder Cancer (ABC) Meta-analysis Collaboration. Adjuvant chemotherapy in invasive
bladder cancer: a systematic review and meta-analysis of individual patient data Advanced Bladder
Cancer (ABC) Meta-analysis Collaboration. Eur Urol, 2005. 48: 189.
https://pubmed.ncbi.nlm.nih.gov/15939530
473. Leow, J.J., et al. Adjuvant chemotherapy for invasive bladder cancer: a 2013 updated systematic
review and meta-analysis of randomized trials. Eur Urol, 2014. 66: 42.
https://pubmed.ncbi.nlm.nih.gov/24018020
474. Cognetti, F., et al. Adjuvant chemotherapy with cisplatin and gemcitabine versus chemotherapy at
relapse in patients with muscle-invasive bladder cancer submitted to radical cystectomy: an Italian,
multicenter, randomized phase III trial. Ann Oncol, 2012. 23: 695.
https://pubmed.ncbi.nlm.nih.gov/21859900
475. Paz-Ares L.G., et al. Randomized phase III trial comparing adjuvant paclitaxel/gemcitabine/cisplatin
(PGC) to observation in patients with resected invasive bladder cancer: Results of the Spanish
Oncology Genitourinary Group (SOGUG) 99/01 study. J Clin Oncol (Meeting Abstracts), 2010. 28: 18
suppl LBA4518.
https://ascopubs.org/doi/abs/10.1200/jco.2010.28.18_suppl.lba4518
476. Stadler, W.M., et al. Phase III study of molecularly targeted adjuvant therapy in locally advanced
urothelial cancer of the bladder based on p53 status. J Clin Oncol, 2011. 29: 3443.
https://pubmed.ncbi.nlm.nih.gov/21810677
477. Lehmann, J., et al. Complete long-term survival data from a trial of adjuvant chemotherapy vs
control after radical cystectomy for locally advanced bladder cancer. BJU Int, 2006. 97: 42.
https://pubmed.ncbi.nlm.nih.gov/16336326

LIMITED UPDATE MARCH 2021 85


478. Freiha, F., et al. A randomized trial of radical cystectomy versus radical cystectomy plus cisplatin,
vinblastine and methotrexate chemotherapy for muscle invasive bladder cancer. J Urol, 1996.
155: 495.
https://pubmed.ncbi.nlm.nih.gov/8558644
479. Stockle, M., et al. Adjuvant polychemotherapy of nonorgan-confined bladder cancer after radical
cystectomy revisited: long-term results of a controlled prospective study and further clinical
experience. J Urol, 1995. 153: 47.
https://pubmed.ncbi.nlm.nih.gov/7966789
480. Studer, U.E., et al. Adjuvant cisplatin chemotherapy following cystectomy for bladder cancer: results
of a prospective randomized trial. J Urol, 1994. 152: 81.
https://pubmed.ncbi.nlm.nih.gov/8201695
481. Skinner, D.G., et al. Adjuvant chemotherapy following cystectomy benefits patients with deeply
invasive bladder cancer. Semin Urol, 1990. 8: 279.
https://pubmed.ncbi.nlm.nih.gov/2284533
482. Lehmann, J., et al. Adjuvant cisplatin plus methotrexate versus methotrexate, vinblastine, epirubicin,
and cisplatin in locally advanced bladder cancer: results of a randomized, multicenter, phase III trial
(AUO-AB 05/95). J Clin Oncol, 2005. 23: 4963.
https://pubmed.ncbi.nlm.nih.gov/15939920
483. Svatek, R.S., et al. The effectiveness of off-protocol adjuvant chemotherapy for patients with
urothelial carcinoma of the urinary bladder. Clin Cancer Res, 2010. 16: 4461.
https://pubmed.ncbi.nlm.nih.gov/20651056
484. Sternberg, C.N., et al. Immediate versus deferred chemotherapy after radical cystectomy in patients
with pT3-pT4 or N+ M0 urothelial carcinoma of the bladder (EORTC 30994): an intergroup, open-
label, randomised phase 3 trial. Lancet Oncol, 2015. 16: 76.
https://pubmed.ncbi.nlm.nih.gov/25498218
485. Galsky, M.D., et al. Effectiveness of Adjuvant Chemotherapy for Locally Advanced Bladder Cancer.
J Clin Oncol, 2016. 34: 825.
https://pubmed.ncbi.nlm.nih.gov/26786930
486. Berg, S., et al. Impact of adjuvant chemotherapy in patients with adverse features and variant
histology at radical cystectomy for muscle-invasive carcinoma of the bladder: Does histologic
subtype matter? Cancer, 2019. 125: 1449.
https://pubmed.ncbi.nlm.nih.gov/30620387
487. Hussain, M.A., et al. IMvigor010: Primary analysis from a phase III randomized study of adjuvant
atezolizumab (atezo) versus observation (obs) in high-risk muscle-invasive urothelial carcinoma
(MIUC). J Clin Oncol 2020. 38: Abstr 5000.
https://ascopubs.org/doi/10.1200/JCO.2020.38.15_suppl.5000
488. Rosenberg, J.E., et al. Update on chemotherapy for advanced bladder cancer. J Urol, 2005. 174: 14.
https://pubmed.ncbi.nlm.nih.gov/15947569
489. Sternberg, C.N., et al. Gemcitabine, paclitaxel, pemetrexed and other newer agents in urothelial and
kidney cancers. Crit Rev Oncol Hematol, 2003. 46 Suppl: S105.
https://pubmed.ncbi.nlm.nih.gov/12850531
490. Loehrer, P.J., Sr., et al. A randomized comparison of cisplatin alone or in combination with
methotrexate, vinblastine, and doxorubicin in patients with metastatic urothelial carcinoma: a
cooperative group study. J Clin Oncol, 1992. 10: 1066.
https://pubmed.ncbi.nlm.nih.gov/1607913
491. Bajorin, D.F., et al. Long-term survival in metastatic transitional-cell carcinoma and prognostic
factors predicting outcome of therapy. J Clin Oncol, 1999. 17: 3173.
https://pubmed.ncbi.nlm.nih.gov/10506615
492. Bellmunt, J., et al. Pretreatment prognostic factors for survival in patients with advanced urothelial
tumors treated in a phase I/II trial with paclitaxel, cisplatin, and gemcitabine. Cancer, 2002. 95: 751.
https://pubmed.ncbi.nlm.nih.gov/12209718
493. Sengelov, L., et al. Metastatic urothelial cancer: evaluation of prognostic factors and change in
prognosis during the last twenty years. Eur Urol, 2001. 39: 634.
https://pubmed.ncbi.nlm.nih.gov/11464051
494. De Santis, M., et al. Randomized phase II/III trial assessing gemcitabine/carboplatin and
methotrexate/carboplatin/vinblastine in patients with advanced urothelial cancer who are unfit for
cisplatin-based chemotherapy: EORTC study 30986. J Clin Oncol, 2012. 30: 191.
https://pubmed.ncbi.nlm.nih.gov/22162575
495. Apolo, A.B., et al. Prognostic model for predicting survival of patients with metastatic urothelial
cancer treated with cisplatin-based chemotherapy. J Natl Cancer Inst, 2013. 105: 499.
https://pubmed.ncbi.nlm.nih.gov/23411591

86 LIMITED UPDATE MARCH 2021


496. Galsky, M.D., et al. Nomogram for predicting survival in patients with unresectable and/or metastatic
urothelial cancer who are treated with cisplatin-based chemotherapy. Cancer, 2013. 119: 3012.
https://pubmed.ncbi.nlm.nih.gov/23720216
497. Bellmunt, J., et al. Prognostic factors in patients with advanced transitional cell carcinoma of the
urothelial tract experiencing treatment failure with platinum-containing regimens. J Clin Oncol, 2010.
28: 1850.
https://pubmed.ncbi.nlm.nih.gov/20231682
498. Galsky, M.D., et al. Cisplatin-based combination chemotherapy in septuagenarians with metastatic
urothelial cancer. Urol Oncol, 2014. 32: 30.e15.
https://pubmed.ncbi.nlm.nih.gov/23428534
499. Galsky, M.D., et al. A consensus definition of patients with metastatic urothelial carcinoma who are
unfit for cisplatin-based chemotherapy. Lancet Oncol, 2011. 12: 211.
https://pubmed.ncbi.nlm.nih.gov/21376284
500. Galsky, M.D., et al. Treatment of patients with metastatic urothelial cancer “unfit” for Cisplatin-based
chemotherapy. J Clin Oncol, 2011. 29: 2432.
https://pubmed.ncbi.nlm.nih.gov/21555688
501. Dash, A., et al. Impact of renal impairment on eligibility for adjuvant cisplatin-based chemotherapy in
patients with urothelial carcinoma of the bladder. Cancer, 2006. 107: 506.
https://pubmed.ncbi.nlm.nih.gov/16773629
502. Nogue-Aliguer, M., et al. Gemcitabine and carboplatin in advanced transitional cell carcinoma of the
urinary tract: an alternative therapy. Cancer, 2003. 97: 2180.
https://pubmed.ncbi.nlm.nih.gov/12712469
503. Balducci, L., et al. Management of cancer in the older person: a practical approach. Oncologist,
2000. 5: 224.
https://pubmed.ncbi.nlm.nih.gov/10884501
504. De Santis, M., et al. New developments in first- and second-line chemotherapy for transitional cell,
squamous cell and adenocarcinoma of the bladder. Curr Opin Urol, 2007. 17: 363.
https://pubmed.ncbi.nlm.nih.gov/17762632
505. Raj, G.V., et al. Formulas calculating creatinine clearance are inadequate for determining eligibility
for Cisplatin-based chemotherapy in bladder cancer. J Clin Oncol, 2006. 24: 3095.
https://pubmed.ncbi.nlm.nih.gov/16809735
506. Carles, J., et al. Feasiblity study of gemcitabine and cisplatin administered every two weeks in
patients with advanced urothelial tumors and impaired renal function. Clin Transl Oncol, 2006. 8: 755.
https://pubmed.ncbi.nlm.nih.gov/17074675
507. Hussain, S.A., et al. A study of split-dose cisplatin-based neo-adjuvant chemotherapy in muscle-
invasive bladder cancer. Oncol Lett, 2012. 3: 855.
https://pubmed.ncbi.nlm.nih.gov/22741006
508. Hussain, S.A., et al. A phase I/II study of gemcitabine and fractionated cisplatin in an outpatient
setting using a 21-day schedule in patients with advanced and metastatic bladder cancer. Br
J Cancer, 2004. 91: 844.
https://pubmed.ncbi.nlm.nih.gov/15292922
509. Morales-Barrera, R., et al. Cisplatin and gemcitabine administered every two weeks in patients with
locally advanced or metastatic urothelial carcinoma and impaired renal function. Eur J Cancer, 2012.
48: 1816.
https://pubmed.ncbi.nlm.nih.gov/22595043
510. De Santis, M., et al. Randomized phase II/III trial assessing gemcitabine/ carboplatin and
methotrexate/carboplatin/vinblastine in patients with advanced urothelial cancer “unfit” for cisplatin-
based chemotherapy: phase II--results of EORTC study 30986. J Clin Oncol, 2009. 27: 5634.
https://pubmed.ncbi.nlm.nih.gov/19786668
511. Bellmunt, J., et al. New therapeutic challenges in advanced bladder cancer. Semin Oncol, 2012.
39: 598.
https://pubmed.ncbi.nlm.nih.gov/23040256
512. von der Maase, H., et al. Long-term survival results of a randomized trial comparing gemcitabine
plus cisplatin, with methotrexate, vinblastine, doxorubicin, plus cisplatin in patients with bladder
cancer. J Clin Oncol, 2005. 23: 4602.
https://pubmed.ncbi.nlm.nih.gov/16034041
513. Gabrilove, J.L., et al. Effect of granulocyte colony-stimulating factor on neutropenia and associated
morbidity due to chemotherapy for transitional-cell carcinoma of the urothelium. N Engl J Med,
1988. 318: 1414.
https://pubmed.ncbi.nlm.nih.gov/2452983

LIMITED UPDATE MARCH 2021 87


514. Bamias, A., et al. Docetaxel and cisplatin with granulocyte colony-stimulating factor (G-CSF) versus
MVAC with G-CSF in advanced urothelial carcinoma: a multicenter, randomized, phase III study
from the Hellenic Cooperative Oncology Group. J Clin Oncol, 2004. 22: 220.
https://pubmed.ncbi.nlm.nih.gov/14665607
515. Sternberg, C.N., et al. Randomized phase III trial of high-dose-intensity methotrexate, vinblastine,
doxorubicin, and cisplatin (MVAC) chemotherapy and recombinant human granulocyte colony-
stimulating factor versus classic MVAC in advanced urothelial tract tumors: European Organization
for Research and Treatment of Cancer Protocol no. 30924. J Clin Oncol, 2001. 19: 2638.
https://pubmed.ncbi.nlm.nih.gov/11352955
516. Sternberg, C.N., et al. Seven year update of an EORTC phase III trial of high-dose intensity M-VAC
chemotherapy and G-CSF versus classic M-VAC in advanced urothelial tract tumours. Eur J Cancer,
2006. 42: 50.
https://pubmed.ncbi.nlm.nih.gov/16330205
517. Bellmunt, J., et al. Randomized phase III study comparing paclitaxel/cisplatin/gemcitabine and
gemcitabine/cisplatin in patients with locally advanced or metastatic urothelial cancer without prior
systemic therapy: EORTC Intergroup Study 30987. J Clin Oncol, 2012. 30: 1107.
https://pubmed.ncbi.nlm.nih.gov/22370319
518. Galsky, M.D., et al. Comparative effectiveness of cisplatin-based and carboplatin-based
chemotherapy for treatment of advanced urothelial carcinoma. Ann Oncol, 2012. 23: 406.
https://pubmed.ncbi.nlm.nih.gov/21543626
519. Bamias, A., et al. Impact of contemporary patterns of chemotherapy utilization on survival in
patients with advanced cancer of the urinary tract: a Retrospective International Study of Invasive/
Advanced Cancer of the Urothelium (RISC). Ann Oncol, 2018. 29: 361.
https://pubmed.ncbi.nlm.nih.gov/29077785
520. De Santis, M., et al. Vinflunine-gemcitabine versus vinflunine-carboplatin as first-line chemotherapy
in cisplatin-unfit patients with advanced urothelial carcinoma: results of an international randomized
phase II trial (JASINT1). Ann Oncol, 2016. 27: 449.
https://pubmed.ncbi.nlm.nih.gov/26673352
521. Galsky, M.D., et al. Atezolizumab with or without chemotherapy in metastatic urothelial cancer
(IMvigor130): a multicentre, randomised, placebo-controlled phase 3 trial. Lancet, 2020. 395: 1547.
https://pubmed.ncbi.nlm.nih.gov/32416780
522. Alva, A., et al. LBA23 - Pembrolizumab (P) combined with chemotherapy (C) vs C alone as first-line
(1L) therapy for advanced urothelial carcinoma (UC): KEYNOTE-361. Ann Oncol, 2020. 31: S1142.
https://www.annalsofoncology.org/article/S0923-7534(20)42334-8/abstract
523. Powles, T., et al. Durvalumab alone and durvalumab plus tremelimumab versus chemotherapy in
previously untreated patients with unresectable, locally advanced or metastatic urothelial carcinoma
(DANUBE): a randomised, open-label, multicentre, phase 3 trial. Lancet Oncol, 2020. 21: 1574.
https://pubmed.ncbi.nlm.nih.gov/32971005
524. O’Donnell, P.H., et al. Pembrolizumab (Pembro; MK-3475) for advanced urothelial cancer: Results of
a phase IB study. J Clin Oncol, 2015. 33: 296.
https://ascopubs.org/doi/abs/10.1200/jco.2015.33.7_suppl.296
525. Balar, A.V., et al. Atezolizumab as first-line treatment in cisplatin-ineligible patients with locally
advanced and metastatic urothelial carcinoma: a single-arm, multicentre, phase 2 trial. Lancet,
2017. 389: 67.
https://pubmed.ncbi.nlm.nih.gov/27939400
526. Powles, T., et al. Avelumab Maintenance Therapy for Advanced or Metastatic Urothelial Carcinoma.
N Engl J Med, 2020. 383: 1218.
https://pubmed.ncbi.nlm.nih.gov/32945632
527. Galsky, M.D., et al. Randomized Double-Blind Phase II Study of Maintenance Pembrolizumab
Versus Placebo After First-Line Chemotherapy in Patients With Metastatic Urothelial Cancer. J Clin
Oncol, 2020. 38: 1797.
https://pubmed.ncbi.nlm.nih.gov/32271672
528. Albers, P., et al. Gemcitabine monotherapy as second-line treatment in cisplatin-refractory
transitional cell carcinoma - prognostic factors for response and improvement of quality of life.
Onkologie, 2002. 25: 47.
https://pubmed.ncbi.nlm.nih.gov/11893883
529. Sternberg, C.N., et al. Chemotherapy with an every-2-week regimen of gemcitabine and paclitaxel
in patients with transitional cell carcinoma who have received prior cisplatin-based therapy. Cancer,
2001. 92: 2993.
https://pubmed.ncbi.nlm.nih.gov/11753976

88 LIMITED UPDATE MARCH 2021


530. Meluch, A.A., et al. Paclitaxel and gemcitabine chemotherapy for advanced transitional-cell
carcinoma of the urothelial tract: a phase II trial of the Minnie pearl cancer research network. J Clin
Oncol, 2001. 19: 3018.
https://pubmed.ncbi.nlm.nih.gov/11408496
531. Parameswaran R, et al. A Hoosier Oncology Group phase II study of weekly paclitaxel and
gemcitabine in advanced transitional cell (TCC) carcinoma of the bladder. Proc Am Soc Clin Oncol,
2001. 200. [No abstract available].
532. Guardino AE, et al. Gemcitabine and paclitaxel as second line chemotherapy for advanced urothelial
malignancies. Proc Am Soc Clin Oncol 2002. 21. [No abstract available].
533. Fechner, G., et al. Randomised phase II trial of gemcitabine and paclitaxel second-line
chemotherapy in patients with transitional cell carcinoma (AUO Trial AB 20/99). Int J Clin Pract,
2006. 60: 27.
https://pubmed.ncbi.nlm.nih.gov/16409425
534. Kaufman DS, et al. Gemcitabine (G) and paclitaxel (P) every two weeks (GP2w):a completed
multicenter phase II trial in locally advanced or metastatic urothelial cancer (UC). Proc Am Soc Clin
Oncol 2002. 21. [No abstract available].
535. Calabro, F., et al. Gemcitabine and paclitaxel every 2 weeks in patients with previously untreated
urothelial carcinoma. Cancer, 2009. 115: 2652.
https://pubmed.ncbi.nlm.nih.gov/19396817
536. Ko, Y.J., et al. Nanoparticle albumin-bound paclitaxel for second-line treatment of metastatic
urothelial carcinoma: a single group, multicentre, phase 2 study. Lancet Oncol, 2013. 14: 769.
https://pubmed.ncbi.nlm.nih.gov/23706985
537. von der Maase, H. Gemcitabine in transitional cell carcinoma of the urothelium. Expert Rev
Anticancer Ther, 2003. 3: 11.
https://pubmed.ncbi.nlm.nih.gov/12597345
538. Oing, C., et al. Second Line Chemotherapy for Advanced and Metastatic Urothelial Carcinoma:
Vinflunine and Beyond-A Comprehensive Review of the Current Literature. J Urol, 2016. 195: 254.
https://pubmed.ncbi.nlm.nih.gov/26410730
539. Raggi, D., et al. Second-line single-agent versus doublet chemotherapy as salvage therapy for
metastatic urothelial cancer: a systematic review and meta-analysis. Ann Oncol, 2016. 27: 49.
https://pubmed.ncbi.nlm.nih.gov/26487582
540. Albers, P., et al. Randomized phase III trial of 2nd line gemcitabine and paclitaxel chemotherapy
in patients with advanced bladder cancer: short-term versus prolonged treatment [German
Association of Urological Oncology (AUO) trial AB 20/99]. Ann Oncol, 2011. 22: 288.
https://pubmed.ncbi.nlm.nih.gov/20682548
541. Bellmunt, J., et al. Phase III trial of vinflunine plus best supportive care compared with best
supportive care alone after a platinum-containing regimen in patients with advanced transitional cell
carcinoma of the urothelial tract. J Clin Oncol, 2009. 27: 4454.
https://pubmed.ncbi.nlm.nih.gov/19687335
542. Petrylak, D.P., et al. Ramucirumab plus docetaxel versus placebo plus docetaxel in patients with
locally advanced or metastatic urothelial carcinoma after platinum-based therapy (RANGE): a
randomised, double-blind, phase 3 trial. Lancet, 2017. 390: 2266.
https://pubmed.ncbi.nlm.nih.gov/28916371
543. Petrylak, D.P., et al. Ramucirumab plus docetaxel versus placebo plus docetaxel in patients with
locally advanced or metastatic urothelial carcinoma after platinum-based therapy (RANGE): overall
survival and updated results of a randomised, double-blind, phase 3 trial. Lancet Oncol, 2020.
21: 105.
https://pubmed.ncbi.nlm.nih.gov/31753727
544. Bellmunt, J., et al. Pembrolizumab as Second-Line Therapy for Advanced Urothelial Carcinoma. N
Engl J Med, 2017. 376: 1015.
https://pubmed.ncbi.nlm.nih.gov/28212060
545. Fradet, Y., et al. Randomized phase III KEYNOTE-045 trial of pembrolizumab versus paclitaxel,
docetaxel, or vinflunine in recurrent advanced urothelial cancer: results of > 2 years of follow-up.
Ann Oncol, 2019.
https://pubmed.ncbi.nlm.nih.gov/31050707
546. Vaughn, D.J., et al. Health-Related Quality-of-Life Analysis From KEYNOTE-045: A Phase III Study
of Pembrolizumab Versus Chemotherapy for Previously Treated Advanced Urothelial Cancer. J Clin
Oncol, 2018. 36: 1579.
https://pubmed.ncbi.nlm.nih.gov/29590008

LIMITED UPDATE MARCH 2021 89


547. Powles, T., et al. MPDL3280A (anti-PD-L1) treatment leads to clinical activity in metastatic bladder
cancer. Nature, 2014. 515: 558.
https://pubmed.ncbi.nlm.nih.gov/25428503
548. Powles, T., et al. Atezolizumab versus chemotherapy in patients with platinum-treated locally
advanced or metastatic urothelial carcinoma (IMvigor211): a multicentre, open-label, phase 3
randomised controlled trial. Lancet, 2018. 391: 748.
https://pubmed.ncbi.nlm.nih.gov/29268948
549. Sternberg, C.N., et al. Primary Results from SAUL, a Multinational Single-arm Safety Study of
Atezolizumab Therapy for Locally Advanced or Metastatic Urothelial or Nonurothelial Carcinoma of
the Urinary Tract. Eur Urol, 2019. 76: 73.
https://pubmed.ncbi.nlm.nih.gov/30910346
550. Sharma, P., et al. Nivolumab in metastatic urothelial carcinoma after platinum therapy (CheckMate
275): a multicentre, single-arm, phase 2 trial. Lancet Oncol, 2017. 18: 312.
https://pubmed.ncbi.nlm.nih.gov/28131785
551. Farina, M.S., et al. Immunotherapy in Urothelial Cancer: Recent Results and Future Perspectives.
Drugs, 2017. 77: 1077.
https://pubmed.ncbi.nlm.nih.gov/28493171
552. Apolo, A.B., et al. Avelumab, an Anti-Programmed Death-Ligand 1 Antibody, In Patients With
Refractory Metastatic Urothelial Carcinoma: Results From a Multicenter, Phase Ib Study. J Clin
Oncol, 2017. 35: 2117.
https://pubmed.ncbi.nlm.nih.gov/28375787
553. Powles, T., et al. Efficacy and Safety of Durvalumab in Locally Advanced or Metastatic Urothelial
Carcinoma: Updated Results From a Phase 1/2 Open-label Study. JAMA Oncol, 2017. 3: e172411.
https://pubmed.ncbi.nlm.nih.gov/28817753
554. Postow, M.A., et al. Immune-Related Adverse Events Associated with Immune Checkpoint
Blockade. N Engl J Med, 2018. 378: 158.
https://pubmed.ncbi.nlm.nih.gov/29320654
555. Brahmer, J.R., et al. Management of Immune-Related Adverse Events in Patients Treated With
Immune Checkpoint Inhibitor Therapy: American Society of Clinical Oncology Clinical Practice
Guideline. J Clin Oncol, 2018. 36: 1714.
https://pubmed.ncbi.nlm.nih.gov/29442540
556. Maher, V.E., et al. Analysis of the Association Between Adverse Events and Outcome in Patients
Receiving a Programmed Death Protein 1 or Programmed Death Ligand 1 Antibody. J Clin Oncol,
2019. 37: 2730.
https://pubmed.ncbi.nlm.nih.gov/31116675
557. Robertson, A.G., et al. Comprehensive Molecular Characterization of Muscle-Invasive Bladder
Cancer. Cell, 2018. 174: 1033.
https://pubmed.ncbi.nlm.nih.gov/30096301
558. Rosenberg, J.E., et al. Pivotal Trial of Enfortumab Vedotin in Urothelial Carcinoma After Platinum and
Anti-Programmed Death 1/Programmed Death Ligand 1 Therapy. J Clin Oncol, 2019. 37: 2592.
https://pubmed.ncbi.nlm.nih.gov/31356140
559. Chang, E., et al. FDA Approval Summary: Enfortumab Vedotin for Locally Advanced or Metastatic
Urothelial Carcinoma. Clin Cancer Res, 2021. 27: 922.
https://pubmed.ncbi.nlm.nih.gov/32962979
560. Rosenberg, J.E., et al. Study EV-103: Preliminary durability results of enfortumab vedotin plus
pembrolizumab for locally advanced or metastatic urothelial carcinoma. J Clin Oncol, 2020. 38: 441.
https://ascopubs.org/doi/abs/10.1200/JCO.2020.38.6_suppl.441
561. Tagawa, S.T., et al. Sacituzumab govitecan (IMMU-132) in patients with previously treated
metastatic urothelial cancer (mUC): Results from a phase I/II study. J Clin Oncol, 2019. 37: 354.
https://ascopubs.org/doi/abs/10.1200/JCO.2019.37.7_suppl.354
562. Stadler, W.M. Gemcitabine doublets in advanced urothelial cancer. Semin Oncol, 2002. 29: 15.
https://pubmed.ncbi.nlm.nih.gov/11894003
563. Hussain, M., et al. Combination paclitaxel, carboplatin, and gemcitabine is an active treatment for
advanced urothelial cancer. J Clin Oncol, 2001. 19: 2527.
https://pubmed.ncbi.nlm.nih.gov/11331332
564. Abe, T., et al. Impact of multimodal treatment on survival in patients with metastatic urothelial
cancer. Eur Urol, 2007. 52: 1106.
https://pubmed.ncbi.nlm.nih.gov/17367917
565. Bekku, K., et al. Could salvage surgery after chemotherapy have clinical impact on cancer survival
of patients with metastatic urothelial carcinoma? Int J Clin Oncol, 2013. 18: 110.
https://pubmed.ncbi.nlm.nih.gov/22095246

90 LIMITED UPDATE MARCH 2021


566. Cowles, R.S., et al. Long-term results following thoracotomy for metastatic bladder cancer. Urology,
1982. 20: 390.
https://pubmed.ncbi.nlm.nih.gov/7147508
567. de Vries, R.R., et al. Long-term survival after combined modality treatment in metastatic bladder
cancer patients presenting with supra-regional tumor positive lymph nodes only. Eur J Surg Oncol,
2009. 35: 352.
https://pubmed.ncbi.nlm.nih.gov/18722076
568. Dodd, P.M., et al. Outcome of postchemotherapy surgery after treatment with methotrexate,
vinblastine, doxorubicin, and cisplatin in patients with unresectable or metastatic transitional cell
carcinoma. J Clin Oncol, 1999. 17: 2546.
https://pubmed.ncbi.nlm.nih.gov/10561321
569. Donat, S.M., et al. Methotrexate, vinblastine, doxorubicin and cisplatin chemotherapy and
cystectomy for unresectable bladder cancer. J Urol, 1996. 156: 368.
https://pubmed.ncbi.nlm.nih.gov/8683681
570. Gowardhan, B., et al. Twenty-three years of disease-free survival following cutaneous metastasis
from a primary bladder transitional cell carcinoma. Int J Urol, 2004. 11: 1031.
https://pubmed.ncbi.nlm.nih.gov/15509212
571. Kanzaki, R., et al. Outcome of surgical resection of pulmonary metastasis from urinary tract
transitional cell carcinoma. Interact Cardiovasc Thorac Surg, 2010. 11: 60.
https://pubmed.ncbi.nlm.nih.gov/20395251
572. Ku, J.H., et al. Metastasis of transitional cell carcinoma to the lower abdominal wall 20 years after
cystectomy. Yonsei Med J, 2005. 46: 181.
https://pubmed.ncbi.nlm.nih.gov/15744826
573. Lehmann, J., et al. Surgery for metastatic urothelial carcinoma with curative intent: the German
experience (AUO AB 30/05). Eur Urol, 2009. 55: 1293.
https://pubmed.ncbi.nlm.nih.gov/19058907
574. Matsuguma, H., et al. Is there a role for pulmonary metastasectomy with a curative intent in patients
with metastatic urinary transitional cell carcinoma? Ann Thorac Surg, 2011. 92: 449.
https://pubmed.ncbi.nlm.nih.gov/21801905
575. Miller, R.S., et al. Cisplatin, methotrexate and vinblastine plus surgical restaging for patients with
advanced transitional cell carcinoma of the urothelium. J Urol, 1993. 150: 65.
https://pubmed.ncbi.nlm.nih.gov/8510277
576. Otto, T., et al. Impact of surgical resection of bladder cancer metastases refractory to systemic
therapy on performance score: a phase II trial. Urology, 2001. 57: 55.
https://pubmed.ncbi.nlm.nih.gov/11164143
577. Sarmiento, J.M., et al. Solitary cerebral metastasis from transitional cell carcinoma after a 14-year
remission of urinary bladder cancer treated with gemcitabine: Case report and literature review. Surg
Neurol Int, 2012. 3: 82.
https://pubmed.ncbi.nlm.nih.gov/22937482
578. Tanis, P.J., et al. Surgery for isolated lung metastasis in two patients with bladder cancer. Urology,
2005. 66: 881.
https://pubmed.ncbi.nlm.nih.gov/16230169
579. Herr, H.W., et al. Post-chemotherapy surgery in patients with unresectable or regionally metastatic
bladder cancer. J Urol, 2001. 165: 811.
https://pubmed.ncbi.nlm.nih.gov/11176475
580. Sweeney, P., et al. Is there a therapeutic role for post-chemotherapy retroperitoneal lymph node
dissection in metastatic transitional cell carcinoma of the bladder? J Urol, 2003. 169: 2113.
https://pubmed.ncbi.nlm.nih.gov/12771730
581. Siefker-Radtke, A.O., et al. Is there a role for surgery in the management of metastatic urothelial
cancer? The M. D. Anderson experience. J Urol, 2004. 171: 145.
https://pubmed.ncbi.nlm.nih.gov/14665863
582. Abufaraj, M., et al. The Role of Surgery in Metastatic Bladder Cancer: A Systematic Review. Eur
Urol, 2018. 73: 543.
https://pubmed.ncbi.nlm.nih.gov/29122377
583. Coleman, R.E. Metastatic bone disease: clinical features, pathophysiology and treatment strategies.
Cancer Treat Rev, 2001. 27: 165.
https://pubmed.ncbi.nlm.nih.gov/11417967
584. Aapro, M., et al. Guidance on the use of bisphosphonates in solid tumours: recommendations of an
international expert panel. Ann Oncol, 2008. 19: 420.
https://pubmed.ncbi.nlm.nih.gov/17906299

LIMITED UPDATE MARCH 2021 91


585. Zaghloul, M.S., et al. A prospective, randomized, placebo-controlled trial of zoledronic acid in bony
metastatic bladder cancer. Int J Clin Oncol, 2010. 15: 382.
https://pubmed.ncbi.nlm.nih.gov/20354750
586. Henry, D.H., et al. Randomized, double-blind study of denosumab versus zoledronic acid in the
treatment of bone metastases in patients with advanced cancer (excluding breast and prostate
cancer) or multiple myeloma. J Clin Oncol, 2011. 29: 1125.
https://pubmed.ncbi.nlm.nih.gov/21343556
587. Rosen, L.S., et al. Long-term efficacy and safety of zoledronic acid in the treatment of skeletal
metastases in patients with nonsmall cell lung carcinoma and other solid tumors: a randomized,
Phase III, double-blind, placebo-controlled trial. Cancer, 2004. 100: 2613.
https://pubmed.ncbi.nlm.nih.gov/15197804
588. Smith, A.B., et al. Impact of bladder cancer on health-related quality of life. BJU Int, 2018. 121: 549.
https://pubmed.ncbi.nlm.nih.gov/28990272
589. Cella, D.F., et al. The Functional Assessment of Cancer Therapy scale: development and validation
of the general measure. J Clin Oncol, 1993. 11: 570.
https://pubmed.ncbi.nlm.nih.gov/8445433
590. Aaronson, N.K., et al. The European Organization for Research and Treatment of Cancer QLQ-C30:
a quality-of-life instrument for use in international clinical trials in oncology. J Natl Cancer Inst, 1993.
85: 365.
https://pubmed.ncbi.nlm.nih.gov/8433390
591. Sogni, F., et al. Morbidity and quality of life in elderly patients receiving ileal conduit or orthotopic
neobladder after radical cystectomy for invasive bladder cancer. Urology, 2008. 71: 919.
https://pubmed.ncbi.nlm.nih.gov/18355900
592. Ware, J.E., Jr., et al. The MOS 36-item short-form health survey (SF-36). I. Conceptual framework
and item selection. Med Care, 1992. 30: 473.
https://pubmed.ncbi.nlm.nih.gov/1593914
593. Ware, J.E., Jr., et al. Evaluating translations of health status questionnaires. Methods from the
IQOLA project. International Quality of Life Assessment. Int J Technol Assess Health Care, 1995.
11: 525.
https://pubmed.ncbi.nlm.nih.gov/7591551
594. Gilbert, S.M., et al. Development and validation of the Bladder Cancer Index: a comprehensive,
disease specific measure of health related quality of life in patients with localized bladder cancer.
J Urol, 2010. 183: 1764.
https://pubmed.ncbi.nlm.nih.gov/20299056
595. Feuerstein, M.A., et al. Propensity-matched analysis of patient-reported outcomes for neoadjuvant
chemotherapy prior to radical cystectomy. World J Urol, 2019. 37: 2401.
https://pubmed.ncbi.nlm.nih.gov/30798382
596. Cerruto, M.A., et al. Systematic review and meta-analysis of non RCT’s on health related quality
of life after radical cystectomy using validated questionnaires: Better results with orthotopic
neobladder versus ileal conduit. Eur J Surg Oncol, 2016. 42: 343.
https://pubmed.ncbi.nlm.nih.gov/26620844
597. Clifford, T.G., et al. Prospective Evaluation of Continence Following Radical Cystectomy and
Orthotopic Urinary Diversion Using a Validated Questionnaire. J Urol, 2016. 196: 1685.
https://pubmed.ncbi.nlm.nih.gov/27256205
598. Bartsch, G., et al. Urinary functional outcomes in female neobladder patients. World J Urol, 2014.
32: 221.
https://pubmed.ncbi.nlm.nih.gov/24317553
599. Fossa, S.D., et al. Quality of life in patients with muscle-infiltrating bladder cancer and hormone-
resistant prostatic cancer. Eur Urol, 1989. 16: 335.
https://pubmed.ncbi.nlm.nih.gov/2476317
600. Mommsen, S., et al. Quality of life in patients with advanced bladder cancer. A randomized study
comparing cystectomy and irradiation--the Danish Bladder Cancer Study Group (DAVECA protocol
8201). Scand J Urol Nephrol Suppl, 1989. 125: 115.
https://pubmed.ncbi.nlm.nih.gov/2699072
601. Fokdal, L., et al. Radical radiotherapy for urinary bladder cancer: treatment outcomes. Expert Rev
Anticancer Ther, 2006. 6: 269.
https://pubmed.ncbi.nlm.nih.gov/16445379
602. Rodel, C., et al. Combined-modality treatment and selective organ preservation in invasive bladder
cancer: long-term results. J Clin Oncol, 2002. 20: 3061.
https://pubmed.ncbi.nlm.nih.gov/12118019

92 LIMITED UPDATE MARCH 2021


603. Malkowicz, S.B., et al. Muscle-invasive urothelial carcinoma of the bladder. Urology, 2007. 69: 3.
https://pubmed.ncbi.nlm.nih.gov/17280906
604. Karakiewicz, P.I., et al. Nomogram for predicting disease recurrence after radical cystectomy for
transitional cell carcinoma of the bladder. J Urol, 2006. 176: 1354.
https://pubmed.ncbi.nlm.nih.gov/16952631
605. Zaak, D., et al. Predicting individual outcomes after radical cystectomy: an external validation of
current nomograms. BJU Int, 2010. 106: 342.
https://pubmed.ncbi.nlm.nih.gov/20002664
606. Giannarini, G., et al. Do patients benefit from routine follow-up to detect recurrences after radical
cystectomy and ileal orthotopic bladder substitution? Eur Urol, 2010. 58: 486.
https://pubmed.ncbi.nlm.nih.gov/20541311
607. Volkmer, B.G., et al. Oncological followup after radical cystectomy for bladder cancer-is there any
benefit? J Urol, 2009. 181: 1587.
https://pubmed.ncbi.nlm.nih.gov/19233433
608. Boorjian, S.A., et al. Detection of asymptomatic recurrence during routine oncological followup after
radical cystectomy is associated with improved patient survival. J Urol, 2011. 186: 1796.
https://pubmed.ncbi.nlm.nih.gov/21944088
609. Soukup, V., et al. Follow-up after surgical treatment of bladder cancer: a critical analysis of the
literature. Eur Urol, 2012. 62: 290.
https://pubmed.ncbi.nlm.nih.gov/22609313
610. Huguet, J. Follow-up after radical cystectomy based on patterns of tumour recurrence and its risk
factors. Actas Urol Esp, 2013. 37: 376.
https://pubmed.ncbi.nlm.nih.gov/23611464
611. Ghoneim, M.A., et al. Radical cystectomy for carcinoma of the bladder: 2,720 consecutive cases
5 years later. J Urol, 2008. 180: 121.
https://pubmed.ncbi.nlm.nih.gov/18485392
612. Donat, S.M. Staged based directed surveillance of invasive bladder cancer following radical
cystectomy: valuable and effective? World J Urol, 2006. 24: 557.
https://pubmed.ncbi.nlm.nih.gov/17009050
613. Mathers, M.J., et al. Is there evidence for a multidisciplinary follow-up after urological cancer? An
evaluation of subsequent cancers. World J Urol, 2008. 26: 251.
https://pubmed.ncbi.nlm.nih.gov/18421461
614. Vrooman, O.P., et al. Follow-up of patients after curative bladder cancer treatment: guidelines vs.
practice. Curr Opin Urol, 2010. 20: 437.
https://pubmed.ncbi.nlm.nih.gov/20657286
615. Cagiannos, I., et al. Surveillance strategies after definitive therapy of invasive bladder cancer. Can
Urol Assoc J, 2009. 3: S237.
https://pubmed.ncbi.nlm.nih.gov/20019993
616. Fahmy, O., et al. Urethral recurrence after radical cystectomy for urothelial carcinoma: A systematic
review and meta-analysis. Urol Oncol, 2018. 36: 54.
https://pubmed.ncbi.nlm.nih.gov/29196179
617. Varol, C., et al. Treatment of urethral recurrence following radical cystectomy and ileal bladder
substitution. J Urol, 2004. 172: 937.
https://pubmed.ncbi.nlm.nih.gov/15311003
618. Gakis, G., et al. Systematic Review on the Fate of the Remnant Urothelium after Radical
Cystectomy. Eur Urol, 2017. 71: 545.
https://pubmed.ncbi.nlm.nih.gov/27720534
619. Picozzi, S., et al. Upper urinary tract recurrence following radical cystectomy for bladder cancer: a
meta-analysis on 13,185 patients. J Urol, 2012. 188: 2046.
https://pubmed.ncbi.nlm.nih.gov/23083867
620. Sanderson, K.M., et al. Upper tract urothelial recurrence following radical cystectomy for transitional
cell carcinoma of the bladder: an analysis of 1,069 patients with 10-year followup. J Urol, 2007.
177: 2088.
https://pubmed.ncbi.nlm.nih.gov/17509294
621. Stewart-Merrill, S.B., et al. Evaluation of current surveillance guidelines following radical cystectomy
and proposal of a novel risk-based approach. Urol Oncol, 2015. 33: 339 e1.
https://pubmed.ncbi.nlm.nih.gov/26031371
622. Gupta, A., et al. Risk of fracture after radical cystectomy and urinary diversion for bladder cancer.
J Clin Oncol, 2014. 32: 3291.
https://pubmed.ncbi.nlm.nih.gov/25185104

LIMITED UPDATE MARCH 2021 93


623. Hautmann, R.E., et al. Functional Outcome and Complications following Ileal Neobladder
Reconstruction in Male Patients without Tumor Recurrence. More than 35 Years of Experience from
a Single Center. J Urol, 2021. 205: 174.
https://pubmed.ncbi.nlm.nih.gov/32856988
624. Stenzl, A., et al. Urethra-sparing cystectomy and orthotopic urinary diversion in women with
malignant pelvic tumors. Cancer, 2001. 92: 1864.
https://pubmed.ncbi.nlm.nih.gov/11745259

10. CONFLICT OF INTEREST


All members of the Muscle-invasive and Metastatic Bladder Cancer Guidelines Working Group have provided
disclosure statements of all relationships that they have that might be perceived as a potential source of
a conflict of interest. This information is publicly accessible through the European Association of Urology
website: https://uroweb.org/guideline/bladder-cancer-muscle-invasive-and-metastatic/?type=panel.
This guidelines document was developed with the financial support of the European Association of
Urology. No external sources of funding and support have been involved. The EAU is a non-profit organization
and funding is limited to administrative assistance and travel and meeting expenses. No honoraria or other
reimbursements have been provided.

11. CITATION INFORMATION


The format in which to cite the EAU Guidelines will vary depending on the style guide of the journal in which the
citation appears. Accordingly, the number of authors or whether, for instance, to include the publisher, location,
or an ISBN number may vary.

The compilation of the complete Guidelines should be referenced as:


EAU Guidelines. Edn. presented at the EAU Annual Congress Milan 2021. ISBN 978-94-92671-13-4.

If a publisher and/or location is required, include:


EAU Guidelines Office, Arnhem, The Netherlands. http://uroweb.org/guidelines/compilations-of-all-guidelines/

References to individual guidelines should be structured in the following way:


Contributors’ names. Title of resource. Publication type. ISBN. Publisher and publisher location, year.

94 LIMITED UPDATE MARCH 2021


EAU Guidelines on
Upper Urinary
Tract Urothelial
Carcinoma
M. Rouprêt, M. Babjuk (Chair), M. Burger (Vice-chair),
E. Compérat, N.C. Cowan, P. Gontero, F. Liedberg,
A. Masson-Lecomte, A.H. Mostafid, J. Palou,
B.W.G. van Rhijn, S.F. Shariat, R. Sylvester
Guidelines Associates: O. Capoun, D. Cohen,
J.L. Dominguez-Escrig, T. Seisen, V. Soukup

© European Association of Urology 2021


TABLE OF CONTENTS PAGE
1. INTRODUCTION 4
1.1 Aim and scope 4
1.2 Panel composition 4
1.3 Available publications 4
1.4 Publication history & summary of changes 4
1.4.1 Summary of changes 4

2. METHODS 5
2.1 Data identification 5
2.2 Review 6

3. EPIDEMIOLOGY, AETIOLOGY AND PATHOLOGY 6


3.1 Epidemiology 6
3.2 Risk factors 7
3.3 Histology and classification 8
3.3.1 Histological types 8
3.4 Summary of evidence and recommendations for epidemiology, aetiology and pathology 8

4. STAGING AND CLASSIFICATION SYSTEMS 8


4.1 Classification 8
4.2 Tumour Node Metastasis staging 9
4.3 Tumour grade 9
4.4 Future developments 9

5. DIAGNOSIS 9
5.1 Symptoms 9
5.2 Imaging 9
5.2.1 Computed tomography urography 9
5.2.2 Magnetic resonance urography 10
5.3 Cystoscopy 10
5.4 Cytology 10
5.5 Diagnostic ureteroscopy 10
5.6 Distant metastases 10
5.6.1 18F-Fluorodeoxglucose positron emission tomography/computed tomography 10
5.7 Summary of evidence and guidelines for the diagnosis of UTUC 11

6. PROGNOSIS 11
6.1 Prognostic factors 11
6.1.1 Patient-related factors 11
6.1.1.1 Age and gender 11
6.1.1.2 Ethnicity 11
6.1.1.3 Tobacco consumption 11
6.1.1.4 Surgical delay 11
6.1.1.5 Other factors 11
6.1.2 Tumour-related factors 12
6.1.2.1 Tumour stage and grade 12
6.1.2.2 Tumour location, multifocality, size and hydronephrosis 12
6.1.2.3 Variant histology 12
6.1.2.4 Lymph node involvement 12
6.1.2.5 Lymphovascular invasion 12
6.1.2.6 Surgical margins 12
6.1.2.7 Other pathological factors 12
6.1.3 Molecular markers 12
6.2 Risk stratification for clinical decision making 12
6.2.1 Low- versus high-risk UTUC 12
6.2.2 Peri-operative predictive tools for high-risk disease 13
6.3 Bladder recurrence 14
6.4 Summary of evidence and guidelines for the prognosis of UTUC 14

2 UPPER URINARY TRACT UROTHELIAL CARCINOMA - LIMITED UPDATE 2021


7. DISEASE MANAGEMENT 14
7.1 Localised non-metastatic disease 14
7.1.1 Kidney-sparing surgery 14
7.1.1.1 Ureteroscopy 14
7.1.1.2 Percutaneous access 14
7.1.1.3 Ureteral resection 15
7.1.1.4 Upper urinary tract instillation of topical agents 15
7.1.1.5 Guidelines for kidney-sparing management of UTUC 15
7.1.2 Management of high-risk non-metastatic UTUC 15
7.1.2.1 Surgical approach 15
7.1.2.1.1 Open radical nephroureterectomy 15
7.1.2.1.2 Minimal invasive radical nephroureterectomy 15
7.1.2.1.3 Management of bladder cuff 16
7.1.2.1.4 Lymph node dissection 16
7.1.3 Peri-operative chemotherapy 16
7.1.3.1 Neoadjuvant chemotherapy 16
7.1.3.2 Adjuvant chemotherapy 16
7.1.4 Adjuvant radiotherapy after radical nephroureterectomy 16
7.1.5 Post-operative bladder instillation 17
7.1.6 Summary of evidence and guidelines for the management of high-risk non-
metastatic UTUC 17
7.2 Metastatic disease 20
7.2.1 Radical nephroureterectomy 20
7.2.2 Metastasectomy 20
7.2.3 Systemic treatments 20
7.2.3.1 First-line setting 20
7.2.3.2 Second-line setting 20
7.2.4 Summary of evidence and guidelines for the treatment of metastatic UTUC 21

8. FOLLOW-UP 22
8.1 Summary of evidence and guidelines for the follow-up of UTUC 22

9. REFERENCES 23

10. CONFLICT OF INTEREST 38

11. CITATION INFORMATION 38

UPPER URINARY TRACT UROTHELIAL CARCINOMA - LIMITED UPDATE 2021 3


1. INTRODUCTION
1.1 Aim and scope
The European Association of Urology (EAU) Non-muscle-invasive Bladder Cancer (NMIBC) Guidelines
Panel has compiled these clinical guidelines to provide urologists with evidence-based information and
recommendations for the management of upper urinary tract urothelial carcinoma (UTUC). Separate EAU
guidelines documents are available addressing non-muscle-invasive bladder cancer [1], muscle-invasive and
metastatic bladder cancer (MIBC) [2], and primary urethral carcinoma [3].
It must be emphasised that clinical guidelines present the best evidence available to the experts
but following guideline recommendations will not necessarily result in the best outcome. Guidelines can never
replace clinical expertise when making treatment decisions for individual patients, but rather help to focus
decisions - also taking personal values and preferences/individual circumstances of patients into account.
Guidelines are not mandates and do not purport to be a legal standard of care.

1.2 Panel composition


The European Association of Urology (EAU) Guidelines Panel on NMIBC consists of an international
multidisciplinary group of clinicians, including urologists, uro-oncologists, a radiologist, a pathologist and
a statistician. Members of this panel have been selected based on their expertise and to represent the
professionals treating patients suspected of harbouring urothelial carcinoma (UC). All experts involved in the
production of this document have submitted potential conflict of interest statements, which can be viewed on
the EAU website Uroweb: https://uroweb.org/guideline/upper-urinary-tract-urothelial-cell-carcinoma/.

1.3 Available publications


A quick reference document (Pocket guidelines) is available in print and as an app for iOS and Android
devices, presenting the main findings of the UTUC Guidelines. These are abridged versions which may require
consultation together with the full text version. Several scientific publications are available as are a number of
translations of all versions of the EAU UTUC Guidelines, the most recent scientific summary was published in
2020 [4]. All documents are accessible through the EAU website Uroweb: https://uroweb.org/guideline/upper-
urinary-tract-urothelial-cell-carcinoma/.

1.4 Publication history & summary of changes


The first EAU Guidelines on UTUC were published in 2011. This 2021 publication presents a substantial update
of the 2020 version.

1.4.1 Summary of changes


The literature for the complete document has been assessed and updated, whenever relevant. Conclusions
and recommendations have been rephrased and added to throughout the current document.

Key changes for the 2021 print:


• Chapter 5 – Diagnosis, new section 5.3 – 18F-Fluorodeoxglucose positron emission tomography/
computed tomography (FDT-PET/CT) was added resulting in a change of a recommendation.

5.7 Summary of evidence and guidelines for the diagnosis of UTUC

Recommendation Strength rating


Magnetic resonance urography or FDG-PET/CT may be used when CT is contra- Weak
indicated.

• Chapter 6 – Prognosis, additional information has been addeded and this section was restructured,
resulting in changes to Figures 6.1 and 6.2 and the Summary of evidence. New sections 6.2.2 – Peri-
operative predictive tools for high risk disease and 6.3 – Bladder recurrence were added.

6.4 Summary of evidence and guidelines for the prognosis of UTUC

Summary of evidence LE
Models are available to predict non-organ confined disease and altered prognosis after RNU. 3
Patient, tumour and treatment-related factors impact risk of bladder recurrence. 3

4 UPPER URINARY TRACT UROTHELIAL CARCINOMA - LIMITED UPDATE 2021


• Chapter 7 – Treatment, Sections 7.1.2 – Management of high-risk non-metastatic UTUC and 7.1.3.3 –
Induction chemotherapy were added. Section 7.1.5 – Post-operative bladder instillation was expanded,
resulting in a changed recommendation.

7.1.6 Summary of evidence and guidelines for the management of high-risk non-metastatic UTUC

Summary of evidence LE
Post-operative chemotherapy improves disease-free survival. 1b

Recommendation Strength rating


Offer post-operative systemic platinum-based chemotherapy to patients with Strong
muscle-invasive UTUC.

• Section 7.2 – Metastatic disease, systemic treatments, considerable new data has been added in both
the first-line and second-line setting, not resulting in a change to the recommendations. A change was
made to Figure 7.2 – Surgical treatment according to location and status, to include post-operative
chemotherapy as an option for high-risk tumours.

2. METHODS
2.1 Data identification
Standard procedure for EAU Guidelines includes an annual assessment of newly published literature in the
field to guide future updates. For the 2021 UTUC Guidelines, new and relevant evidence has been identified,
collated and appraised through a structured assessment of the literature. The search was restricted to articles
published between May 30th 2019 and May 29th 2020. Databases searched included Pubmed, Ovid, EMBASE
and both the Cochrane Central Register of Controlled Trials and the Cochrane Database of Systematic Reviews.
After deduplication, a total of 614 unique records were identified, retrieved and screened for relevance.
Excluded from the search were basic research studies, case series, reports and editorial comments.
Only articles published in the English language, addressing adults were included. The publications identified
were mainly retrospective, including some large multicentre studies. Owing to the scarcity of randomised data,
articles were selected based on the following criteria: evolution of concepts, intermediate- and long-term clinical
outcomes, study quality, and relevance. Older studies were only included if they were historically relevant.
A total of 35 new publications were added to the 2021 UTUC Guidelines print. A detailed search strategy is
available online: https://uroweb.org/guideline/upper-urinary-tract-urothelial-cell-carcinoma/?type=appendices-
publications.

For Chapters 3-6 (Epidemiology, Aetiology and Pathology, Staging and Classification systems, Diagnosis and
Prognosis) references used in this text are assessed according to their level of evidence (LE) based on the 2009
Oxford Centre for Evidence-Based Medicine (CEBM) Levels of Evidence [5]. For the Disease Management and
Follow-up chapters (Chapters 7 and 8) a system modified from the 2009 CEBM LEs has been used [5]. For
each recommendation within the guidelines there is an accompanying online strength rating form, based on a
modified GRADE methodology [6, 7]. These forms address a number of key elements, namely:

1.  he overall quality of the evidence which exists for the commendation references used in
T
this text are graded according to the CEBM Levels of Evidence (see above) [5];
2. the magnitude of the effect (individual or combined effects);
3. the certainty of the results (precision, consistency, heterogeneity and other statistical or
study related factors);
4. the balance between desirable and undesirable outcomes;
5. the impact of patient values and preferences on the intervention;
6. the certainty of those patient values and preferences.

These key elements are the basis which panels use to define the strength rating of each recommendation.
The strength of each recommendation is represented by the words ‘strong’ or ‘weak’. The strength of each
recommendation is determined by the balance between desirable and undesirable consequences of alternative
management strategies, the quality of the evidence (including certainty of estimates), and nature and variability
of patient values and preferences [8].

UPPER URINARY TRACT UROTHELIAL CARCINOMA - LIMITED UPDATE 2021 5


Additional information can be found in the general Methodology section of this print, and online at
the EAU website; https://uroweb.org/guidelines/policies-and-methodological-documents/.
A list of Associations endorsing the EAU Guidelines can also be viewed online at the above address.

2.2 Review
The 2021 UTUC Guidelines have been peer-reviewed prior to publication.

3. EPIDEMIOLOGY, AETIOLOGY AND


PATHOLOGY
3.1 Epidemiology
Urothelial carcinomas (UCs) are the sixth most common tumours in developed countries [9]. They can be
located in the lower (bladder and urethra) and/or the upper (pyelocaliceal cavities and ureter) urinary tract.
Bladder tumours account for 90–95% of UCs and are the most common urinary tract malignancy [1]. Upper
urinary tract UCs are uncommon and account for only 5–10% of UCs [9] with an estimated annual incidence
in Western countries of almost two cases per 100,000 inhabitants. This rate has risen in the past few decades
as a result of improved detection and improved bladder cancer survival [10]. Pyelocaliceal tumours are
approximately twice as common as ureteral tumours and multifocal tumours are found in approximately
10–20% of cases [11]. The presence of concomitant carcinoma in situ of the upper tract is between 11 and
36% [10]. In 17% of cases, concurrent bladder cancer is present [12] whilst a prior history of bladder cancer
is found in 41% of American men but in only 4% of Chinese men [13]. This, along with genetic and epigenetic
factors, may explain why Asian patients present with more advanced and higher grade disease compared to
other ethnic groups [10]. Following treatment, recurrence in the bladder occurs in 22–47% of UTUC patients,
depending on initial tumour grade [14] compared with 2–5% in the contralateral upper tract [15].
With regards to UTUC occurring following an initial diagnosis of bladder cancer, a series of 82
patients treated with bacillus Calmette-Guérin (BCG) who had regular upper tract imaging between years 1
and 3 showed a 13% incidence of UTUC, all of which were asymptomatic [16], whilst in another series of
307 patients without routine upper tract imaging the incidence was 25% [17]. A multicentre cohort study
(n = 402) with a 50 month follow-up has demonstrated a UTUC incidence of 7.5% in NMIBC receiving BCG
with predictors being intravesical recurrence and non-papillary tumour at transurethral resection of the bladder
[18]. Following radical cystectomy for MIBC, 3–5% of patients develop a metachronous UTUC [19, 20].
Approximately two-thirds of patients who present with UTUCs have invasive disease at diagnosis
compared to 15–25% of patients presenting with muscle-invasive bladder tumours [21]. This is probably due
to the absence of muscularis propria layer in the upper tract, so tumours are more likely to upstage at an earlier
time-point. Approximately 9% of patients present with metastasis [10, 22]. Upper urinary tract UCs have a peak
incidence in individuals aged 70–90 years and are twice as common in men [23].
Upper tract UC and bladder cancer exhibit significant differences in the prevalence of common
genomic alterations. In individual patients with a history of both tumours, bladder cancer and UTUC were
always clonally related. Genomic characterisation of UTUC provides information regarding the risk of bladder
recurrence and can identify tumours associated with Lynch syndrome [24].
The Amsterdam criteria are a set of diagnostic criteria used by doctors to help identify families
which are likely to have Lynch syndrome [25]. In Lynch-related UTUC, immunohistochemistry analysis showed
loss of protein expression corresponding to the disease-predisposing MMR (mismatch repair) gene mutation
in 98% of the samples (46% were microsatellite instable and 54% microsatellite stable) [26]. The majority of
tumours develop in MSH2 mutation carriers [26]. Patients identified at high risk for Lynch syndrome should
undergo DNA sequencing for patient and family counselling [27, 28]. Germline mutations in DNA MMR genes
defining Lynch syndrome, are found in 9% of patients with UTUC compared to 1% of patients with bladder
cancer, linking UTUC to Lynch syndrome [29]. A study of 115 consecutive UTUC patients, reported that 13.9%
screened positive for potential Lynch syndrome and 5.2% had confirmed Lynch syndrome [30]. This is one of
the highest rates of undiagnosed genetic disease in urological cancers, which justifies screening of all patients
under 65 presenting with UTUC and those with a family history of UTUC (see Figure 3.1) [31, 32].

6 UPPER URINARY TRACT UROTHELIAL CARCINOMA - LIMITED UPDATE 2021


Figure 3.1: S
 election of patients with UTUC for Lynch syndrome screening during the first medical
Interview

UTUC

Systemac screening during medical interview

Suspicion of hereditary UTUC Sporadic UTUC


- Age < 65 yr
- Personal history of Lynch-spectrum cancer
or
- First-degree relave < 50 yr with Lynch-spectrum cancer
or
- Two first-degree relaves with Lynch-spectrum cancer

Germ-line DNA sequencing: mutaon (5-9%)

- Clinical evaluaon for other Lynch-related cancer:


colorectal, gastrointesnal, endometrial ovarian and skin
- Close monitoring and follow-up
- Familial genec counselling

UTUC = upper urinary tract urothelial carcinoma.

3.2 Risk factors


A number of environmental factors have been implicated in the development of UTUC [11, 33]. Published
evidence in support of a causative role for these factors is not strong, with the exception of smoking and
aristolochic acid. Tobacco exposure increases the relative risk of developing UTUC from 2.5 to 7.0 [34-36].
A large population-based study assessing familial clustering in relatives of UC patients, including 229,251
relatives of case subjects and 1197,552 relatives of matched control subjects, has demonstrated genetic
or environmental roots independent of smoking-related behaviours. With more than 9% of the cohort being
UTUC patients, clustering was not seen in upper tract disease. This may suggest that the familial clustering of
urothelial cancer is specific to lower tract cancers [37].
In Taiwan, the presence of arsenic in drinking water has been tentatively linked to UTUC [38].
Aristolochic acid, a nitrophenanthrene carboxylic acid produced by Aristolochia plants, exerts multiple
effects on the urinary system. Aristolochic acid irreversibly injures renal proximal tubules resulting in chronic
tubulointerstitial disease, while the mutagenic properties of this chemical carcinogen lead predominantly
to UTUC [39-41]. Aristolochic acid has been linked to bladder cancer, renal cell carcinoma, hepatocellular
carcinoma and intrahepatic cholangiocarcinoma [42]. Two routes of exposure to aristolochic acid are known:
(i) environmental contamination of agricultural products by Aristolochia plants, as reported for Balkan endemic
nephropathy [43]; and (ii) ingestion of Aristolochia-based herbal remedies [44, 45]. Aristolochia herbs are
used worldwide, especially in China and Taiwan [41]. Following bioactivation, aristolochic acid reacts with
genomic DNA to form aristolactam-deoxyadenosine adducts [46]; these lesions persist for decades in target
tissues, serving as robust biomarkers of exposure [9]. These adducts generate a unique mutational spectrum,

UPPER URINARY TRACT UROTHELIAL CARCINOMA - LIMITED UPDATE 2021 7


characterised by A>T transversions located predominately on the non-transcribed strand of DNA [42, 47].
However, fewer than 10% of individuals exposed to aristolochic acid develop UTUC [40].
Two retrospective series found that aristolochic acid-associated UTUC is more common in females
[48, 49]. However, females with aristolochic acid UTUC have a better prognosis than their male counterparts.

Alcohol consumption is associated with development of UTUC. A large case-control study (1,569 cases
and 506,797 controls) has evidenced a significantly higher risk of UTUC in ever-drinkers compared to never-
drinkers (OR: 1.23; 95% CI: 1.08–1.40, p = 0.001). Compared to never-drinkers, the risk threshold for UTUC
was > 15 g of alcohol/day. A dose-response was observed [50].

Differences in the ability to counteract carcinogens may contribute to host susceptibility to UTUC. Some
genetic polymorphisms are associated with an increased risk of cancer or faster disease progression that
introduces variability in the inter-individual susceptibility to the risk factors previously mentioned. Upper urinary
tract UCs may share some risk factors and described molecular pathways with bladder UC [24]. So far, two
UTUC-specific polymorphisms have been reported [51].
A history of bladder cancer is associated with higher risk of UTUCs (see Section 3.1). Patients who
undergo ureteral stenting prior to radical cystectomy are at higher risk for upper tract recurrence [52].

3.3 Histology and classification


3.3.1 Histological types
Upper urinary tract tumours are almost always urothelial carcinomas and pure non-urothelial histology is rare
[53, 54]. However, variants are present in approximately 25% of UTUCs [55, 56]. Pure squamous cell carcinoma
of the urinary tract is often assumed to be associated with chronic inflammatory diseases and infections
arising from urolithiasis [57, 58]. Urothelial carcinoma with divergent squamous differentiation is present in
approximately 15% of cases [57]. Keratinising squamous metaplasia of urothelium is a risk factor for squamous
cell cancers and therefore mandates surveillance. Upper urinary tract UCs with variant histology are high-
grade and have a worse prognosis compared with pure UC [56, 59, 60]. Other variants, although rare, include
sarcomatoid and UCs with inverted growth [60].
However, collecting duct carcinomas, which may seem to share similar characteristics with UCs,
display a unique transcriptomic signature similar to renal cancer, with a putative cell of origin in the distal
convoluted tubules. Therefore, collecting duct carcinomas are considered as renal tumours [61].

3.4 Summary of evidence and recommendations for epidemiology, aetiology and


pathology

Summary of evidence LE
Aristolochic acid and smoking exposure increases the risk for UTUC. 2a
Patients with Lynch syndrome are at risk for UTUC. 3

Recommendations Strength rating


Evaluate patient and family history based on the Amsterdam criteria to identify patients with Weak
upper tract urothelial carcinoma.
Evaluate patient exposure to smoking and aristolochic acid. Weak

4. STAGING AND CLASSIFICATION SYSTEMS


4.1 Classification
The classification and morphology of UTUC and bladder carcinoma are similar [1]. However, it is not possible
to distinguish between non-invasive papillary tumours (papillary urothelial tumours of low malignant potential
and low- and high-grade papillary UC) [62], flat lesions (carcinoma in situ [CIS]), and invasive carcinoma. This
is in most cases due to the size of biopsy specimens that do not include deep tissue required for pathological
staging. With UTUC, histological grade is a surrogate for pathological stage, as it strongly correlates [63].

8 UPPER URINARY TRACT UROTHELIAL CARCINOMA - LIMITED UPDATE 2021


4.2 Tumour Node Metastasis staging
The tumour, node, metastasis (TNM) classification is shown in Table 1 [64]. The regional lymph nodes (LNs)
are the hilar and retroperitoneal nodes and, for the mid- and distal ureter, the pelvic nodes. Laterality does not
affect N classification.

4.3 Tumour grade


In 2004, the WHO and the International Society of Urological Pathology published a new histological
classification of UCs which provides a different patient stratification between individual categories compared
to the older 1973 WHO classification [65, 66]. In 2016, an update of the 2004 WHO grading classification was
published without major changes [65]. These guidelines are still based on both the 1973 and 2004/2016 WHO
classifications since most published data use the 1973 classification [62].

4.4 Future developments


A number of studies focussing on molecular classification have been able to demonstrate genetically different
groups of UTUC by evaluating DNA, RNA and protein expression. Four molecular subtypes with distinct
clinical behaviours were identified, but, as yet, it is unclear whether these subtypes will respond differently to
treatment [67].

Table 1: TNM classification 2017 for upper tract urothelial cell carcinoma [64]

T - Primary tumour
TX Primary tumour cannot be assessed
T0 No evidence of primary tumour
Ta Non-invasive papillary carcinoma
Tis Carcinoma in situ
T1 Tumour invades subepithelial connective tissue
T2 Tumour invades muscularis
T3 (Renal pelvis) Tumour invades beyond muscularis into peripelvic fat or renal parenchyma
(Ureter) Tumour invades beyond muscularis into periureteric fat
T4 Tumour invades adjacent organs or through the kidney into perinephric fat
N - Regional lymph nodes
NX Regional lymph nodes cannot be assessed
N0 No regional lymph node metastasis
N1 Metastasis in a single lymph node 2 cm or less in the greatest dimension
N2 Metastasis in a single lymph node more than 2 cm, or multiple lymph nodes
M - Distant metastasis
M0 No distant metastasis
M1 Distant metastasis
TNM = Tumour, Node, Metastasis (classification).

5. DIAGNOSIS
5.1 Symptoms
The diagnosis of UTUC may be incidental or symptom related. The most common symptom is visible or
nonvisible haematuria (70–80%) [68, 69]. Flank pain, due to clot or tumour tissue obstruction or less often due
to local growth, occurs in approximately 20–32% of cases [69, 70]. Systemic symptoms (including anorexia,
weight loss, malaise, fatigue, fever, night sweats, or cough) associated with UTUC should prompt evaluation for
metastases associated with a worse prognosis [70].

5.2 Imaging
5.2.1 Computed tomography urography
Computed tomography (CT) urography has the highest diagnostic accuracy of the available imaging techniques
[71]. A meta-analysis of 13 studies comprising 1,233 patients revealed a pooled sensitivity of CT urography for
UTUC of 92% (CI: 0.85–0.96) and a pooled specificity of 95% (CI: 0.88–0.98) [72].

UPPER URINARY TRACT UROTHELIAL CARCINOMA - LIMITED UPDATE 2021 9


Rapid acquisition of thin sections allows high-resolution isotropic images that can be viewed in
multiple planes to assist with diagnosis without loss of resolution. Epithelial “flat lesions” without mass effect or
urothelial thickening are generally not visible with CT.
The presence of enlarged LNs is highly predictive of metastases in UTUC [73, 74].

5.2.2 Magnetic resonance urography


Magnetic resonance (MR) urography is indicated in patients who cannot undergo CT urography, usually when
radiation or iodinated contrast media are contraindicated [75]. The sensitivity of MR urography is 75% after
contrast injection for tumours < 2 cm [75]. The use of MR urography with gadolinium-based contrast media
should be limited in patients with severe renal impairment (< 30 mL/min creatinine clearance), due to the risk
of nephrogenic systemic fibrosis. Computed tomography urography is more sensitive and specific for the
diagnosis and staging of UTUC as compared with MR urography [76].

5.3 Cystoscopy
Urethrocystocopy is an integral part of UTUC diagnosis to rule out concomitant bladder cancer [10, 12].

5.4 Cytology
Abnormal cytology may indicate high-grade UTUC when bladder cystoscopy is normal, and in the absence of
CIS in the bladder and prostatic urethra [1, 77, 78]. Cytology is less sensitive for UTUC than bladder tumours
and should be performed selectively for the affected upper tract [79]. In a recent study, barbotage cytology
detected up to 91% of cancers [80]. Barbotage cytology taken from the renal cavities and ureteral lumina is
preferred before application of a contrast agent for retrograde ureteropyelography as it may cause deterioration
of cytological specimens [81]. Retrograde ureteropyelography remains an option to detect UTUCs [71, 82, 83].
The sensitivity of fluorescence in situ hybridisation (FISH) for molecular abnormalities characteristic
of UTUCs is approximately 50% and therefore its use in clinical practice remains unproven [84, 85].

5.5 Diagnostic ureteroscopy


Flexible ureteroscopy (URS) is used to visualise the ureter, renal pelvis and collecting system and for biopsy
of suspicious lesions. Presence, appearance and size of tumour can be determined using URS. In addition,
ureteroscopic biopsies can determine tumour grade in more than 90% of cases with a low false-negative
rate, regardless of sample size [86]. Undergrading may occur following diagnostic biopsy, making intensive
follow-up necessary if kidney-sparing treatment is chosen [87]. Ureteroscopy also facilitates selective ureteral
sampling for cytology in situ [83, 88, 89]. Stage assessment using ureteroscopic biopsy is inaccurate.
Combining ureteroscopic biopsy grade, imaging findings such as hydronephrosis, and urinary
cytology may help in the decision-making process between radical nephroureterectomy (RNU) and kidney-
sparing therapy [89, 90]. While some studies suggest a higher rate of intravesical recurrence after RNU in
patients who underwent diagnostic URS pre-operatively [91, 92], one study did not [93].

Technical developments in flexible ureteroscopes and the use of novel imaging techniques improve
visualisation and diagnosis of flat lesions [94]. Narrow-band imaging is a promising technique, but results are
preliminary [95]. Optical coherence tomography and confocal laser endomicroscopy (Cellvizio®) have been
used in vivo to evaluate tumour grade and/or for staging purposes, with a promising correlation with definitive
histology in high-grade UTUC [96, 97]. Recommendations for the diagnosis of UTUC are listed in Section 5.6.

5.6 Distant metastases


Prior to any treatment with curative intent, it is essential to rule out distant metastases. Computed tomography
is the diagnostic technique of choice for lung- and abdominal staging for metastases [72].

5.6.1 18F-Fluorodeoxglucose positron emission tomography/computed tomography

A retrospective multi-centre publication on the use of 18F-Fluorodeoxglucose positron emission tomography/


computed tomography (FDG-PET/CT) for the detection of nodal metastasis in 117 surgically-treated UTUC
patients reported a promising sensitivity and specificity of 82% and 84%, respectively. Suspicious LNs on
FDG-PET/CT were associated with worse recurrence-free survival [98]. These results warrant further validation
and comparison to MR and CT.

10 UPPER URINARY TRACT UROTHELIAL CARCINOMA - LIMITED UPDATE 2021


5.7 Summary of evidence and guidelines for the diagnosis of UTUC

Summary of evidence LE
The diagnosis and staging of UTUC is best done with computed tomography urography and URS. 2a
Selective urinary cytology has high sensitivity in high-grade tumours, including carcinoma in situ. 3
Urethrocystoscopy can detect concomitant bladder cancer. 2a

Recommendations Strength rating


Perform a urethrocystoscopy to rule out bladder tumour. Strong
Perform a computed tomography (CT) urography for diagnosis and staging. Strong
Use diagnostic ureteroscopy and biopsy if imaging and cytology are not sufficient for the Strong
diagnosis and/or risk-stratification of the tumour.
Magnetic resonance urography or 18F-Fluorodeoxglucose positron emission tomography/ Weak
computed tomography may be used when CT is contra-indicated.

6. PROGNOSIS
6.1 Prognostic factors
Upper urinary tract UCs that invade the muscle wall usually have a very poor prognosis. The 5-year specific
survival is < 50% for pT2/pT3 and < 10% for pT4 UTUC [99-102]. Many prognostic factors have been identifed
and can be used to risk-stratify patients in order to decide on the most appropriate local treatment (radical vs.
conservative) and discuss peri-operative systemic therapy. Factors can be divided into patient-related factors
and tumour-related factors.

6.1.1 Patient-related factors


6.1.1.1 Age and gender
Older age at the time of RNU is independently associated with decreased cancer-specific survival (CSS) [100,
103, 104] (LE: 3). However, even elderly patients can be cured with RNU [105]. Therefore, chronological age alone
should not be a contraindication to RNU [104, 105]. Gender has no impact on prognosis of UTUC [23, 100, 106].

6.1.1.2 Ethnicity
One multicentre study of academic centres did not show any difference in outcomes between races [107],
but U.S. population-based studies have indicated that African-American patients have worse outcomes
than other ethnicities (LE: 3). Whether this is related to access to care or biological and/or patterns of care
remains unknown. Another study has demonstrated differences between Chinese and American patients at
presentation (risk factor, disease characteristics and predictors of adverse oncologic outcomes) [13].

6.1.1.3 Tobacco consumption


Being a smoker at diagnosis increases the risk for disease recurrence and mortality after RNU [108, 109] and
recurrence within the bladder [110] (LE: 3). There is a close relationship between tobacco consumption and
prognosis [111] (LE: 3); smoking cessation improves cancer control [109].

6.1.1.4 Surgical delay


A delay between diagnosis of an invasive tumour and its removal may increase the risk of disease progression.
Once a decision regarding RNU has been made, the procedure should be carried out within twelve weeks,
when possible [112-116] (LE: 3).

6.1.1.5 Other factors


A higher American Society of Anesthesiologists score confers worse CSS after RNU [117] (LE: 3), as does poor
performance status [118]. Obesity and higher body mass index adversely affect cancer-specific outcomes in
patients treated with RNU [119] (LE: 3), with potential differences between races [120]. Several blood-based
biomarkers have been associated with locally advanced disease and cancer-specific mortality such as high
pre-treatment-derived neutrophil-lymphocyte ratio [121-124], low albumin [123, 125], high C-reactive protein
[123] or modified Glasgow score [126], high De Ritis (AST/ALT) ratio [127], altered renal function [123, 128] and
high fibrinogen [123, 129] (LE: 3).

UPPER URINARY TRACT UROTHELIAL CARCINOMA - LIMITED UPDATE 2021 11


6.1.2 Tumour-related factors
6.1.2.1 Tumour stage and grade
The main prognostic factors are tumour stage and grade [21, 89, 100, 130, 131].

6.1.2.2 Tumour location, multifocality, size and hydronephrosis


Initial location of the UTUC is a prognostic factor in some studies [132, 133] (LE: 3). After adjustment for the
effect of tumour stage, patients with ureteral and/or multifocal tumours seem to have a worse prognosis than
patients diagnosed with renal pelvic tumours [100, 132-138]. Hydronephrosis is associated with advanced
disease and poor oncological outcome [73, 81, 139]. A large multi-institutional retrospective study including
932 RNUs for non-metastatic UTUC demonstrated that a 2 cm cut-off appears to be most useful in identifying
patients at risk of harbouring ≥ pT2 UTUC [140].

6.1.2.3 Variant histology


Pathological variants are associated with worse cancer-specific and overall survival (OS) [56] (LE: 3). Most
studied variants are micropapillary [59], squamous [141] and carcomatoid [59] which are consistently
associated with locally advanced disease and worse outcome.

6.1.2.4 Lymph node involvement


Patients with nodal metastasis experience very poor survival after surgery [142]. Lymph node density (cut-off
30%) and extranodal extension are powerful predictors of survival outcomes in N+ UTUC [143-145]. Lymph
node dissection (LND) performed at the time of RNU allows for optimal tumour staging, although its curative
role remains controversial [102, 144, 146, 147] (LE: 3).

6.1.2.5 Lymphovascular invasion


Lymphovascular invasion (LVI) is present in approximately 20% of UTUCs and is an independent predictor of
survival [148-150]. Lymphovascular invasion status should be specifically reported in the pathological reports
of all UTUC specimens [148, 151, 152] (LE: 3).

6.1.2.6 Surgical margins


Positive soft tissue surgical margin is associated with a higher disease recurrence after RNU. Pathologists
should look for and report positive margins at the level of ureteral transection, bladder cuff, and around the
tumour [153] (LE: 3).

6.1.2.7 Other pathological factors


Extensive tumour necrosis (> 10% of the tumour area) is an independent prognostic predictor in patients who
undergo RNU [154, 155] (LE: 3). In case neoadjuvant treatment was administered, pathological downstaging is
associated with better OS [156] (LE: 3). The architecture of UTUC is also a strong prognosticator with sessile
growth pattern being associated with worse outcome [157, 158] (LE: 3). Concomitant CIS in organ-confined
UTUC and a history of bladder CIS are associated with a higher risk of recurrence and cancer-specific mortality
[159, 160] (LE: 3). Macroscopic infiltration or invasion of peri-pelvic adipose tissue confers a higher risk of
disease recurrence after RNU compared to microscopic infiltration of renal parenchyma [55, 161].

6.1.3 Molecular markers


Because of the rarity of UTUC, the main limitations of molecular studies are their retrospective design and, for
most studies, small sample size. None of the investigated markers have been validated yet to support their
introduction in daily clinical decision making [100, 162].

6.2 Risk stratification for clinical decision making


6.2.1 Low- versus high-risk UTUC
As tumour stage is difficult to assess clinically in UTUC, it is useful to “risk stratify” UTUC between low-
and high risk of progression to identify those patients who are more likely to benefit from kidney-sparing
treatment and those who should be treated radically [163, 164] (see Figure 6.2). The factors to consider for risk
stratification are presented in Figure 6.1.

12 UPPER URINARY TRACT UROTHELIAL CARCINOMA - LIMITED UPDATE 2021


Figure 6.1: Risk stratification of non-metastatic UTUC

UTUC

Low-risk UTUC* High-risk UTUC**

• Unifocal disease • Mulfocal disease


• Tumour size < 2 cm • Tumour size ≥ 2 cm
• Negave for high-grade cytology • High-grade cytology
• Low-grade URS biopsy • High-grade URS biopsy
• No invasive aspect on CT • Local invasion on CT
• Hydronephrosis
• Previous radical cystectomy for high-
grade bladder cancer
• Variant histology

CT = computed tomography; URS = ureteroscopy; UTUC = upper urinary tract urothelial carcinoma.
* All these factors need to be present.
**Any of these factors need to be present.

6.2.2 Peri-operative predictive tools for high-risk disease


There are several pre-RNU models aiming at predicting which patient has muscle-invasive/non-organ-
confined disease [165-168] (LE: 3). Prognostic nomograms based on pre-operative factors and pathological
characteristics are available [102, 168-174]. The main factors included in these models, which may be used
when counselling patients regarding follow-up and administration of peri-operative chemotherapy, are detailed
in Figure 6.2.

Figure 6.2: U
 pper urinary tract urothelial cell carcinoma - prognostic factors included in prognostic
models

High-risk UTUC

Prognosc factors

Pre-operave Post-operave

• Advanced age • Advanced stage


• Poor social status • High grade
• Chronic kidney disease • Carcinoma in situ
• Ureteral tumour locaon • Lymphovascular invasion
• Sessile tumor architecture • Lymph node involvement
• Advanced clinical stage • Sessile tumour architecture
• High grade (biopsy, cytology)
• Blood based biomarkers

UTUC = upper tract urothelial carcinoma.

UPPER URINARY TRACT UROTHELIAL CARCINOMA - LIMITED UPDATE 2021 13


6.3 Bladder recurrence
A meta-analysis of available data has identified significant predictors of bladder recurrence after RNU [175]
(LE: 3). Three categories of predictors for increased risk of bladder recurrence were identified:
1. Patient-specific factors such as male gender, previous bladder cancer, smoking and pre-
operative chronic kidney disease.
2. Tumour-specific factors such as positive pre-operative urinary cytology, tumour grade,
ureteral location, multifocality, tumour diameter, invasive pT stage, and necrosis [176].
3. Treatment-specific factors such as laparoscopic approach, extravesical bladder cuff
removal, and positive surgical margins [175].

In addition, the use of diagnostic URS has been associated with a higher risk of developing bladder recurrence
after RNU [91, 92] (LE: 3). Based on low-level evidence only, a single dose of intravesical chemotherapy
after diagnostic/therapeutic ureteroscopy of non-metastatic UTUC has been suggested to lower the rate of
intravesical recurrence, similarly to that after RNU [175].

6.4 Summary of evidence and guidelines for the prognosis of UTUC

Summary of evidence LE
Important prognostic factors for risk stratification include tumour multifocality, size, stage, grade, 3
hydronephrosis and variant histology.
Models are available to predict non-organ confined disease and altered prognosis after RNU. 3
Patient, tumour and treatment-related factors impact risk of bladder recurrence. 3
Currently, no prognostic biomarkers are validated for clinical use. 3

Recommendation Strength rating


Use prognostic factors to risk-stratify patients for therapeutic guidance. Weak

7. DISEASE MANAGEMENT
7.1 Localised non-metastatic disease
7.1.1 Kidney-sparing surgery
Kidney-sparing surgery for low-risk UTUC reduces the morbidity associated with radical surgery (e.g., loss
of kidney function), without compromising oncological outcomes [177]. In low-risk cancers, it is the preferred
approach as survival is similar to that after RNU [177]. This option should therefore be discussed in all low-
risk cases, irrespective of the status of the contralateral kidney. In addition, it can also be considered in select
patients with a serious renal insufficiency or having a solitary kidney (LE: 3). Recommendations for kidney-
sparing management of UTUC are listed in Section 7.1.1.5.

7.1.1.1 Ureteroscopy
Endoscopic ablation should be considered in patients with clinically low-risk cancer [178, 179]. A flexible
ureteroscope is useful in the management of pelvicalyceal tumours [180]. The patient should be informed of
the need and be willing to comply with an early second-look URS [181] and stringent surveillance; complete
tumour resection or destruction is necessary [181]. Nevertheless, a risk of disease progression remains with
endoscopic management due to the suboptimal performance of imaging and biopsy for risk stratification and
tumour biology [182].

7.1.1.2 Percutaneous access


Percutaneous management can be considered for low-risk UTUC in the renal pelvis [179, 183] (LE: 3). This may
also be offered for low-risk tumours in the lower caliceal system that are inaccessible or difficult to manage by
flexible URS. However, this approach is being used less due to the availability of improved endoscopic tools
such as distal-tip deflection of recent ureteroscopes [179, 183]. Moreover, a risk of tumour seeding remains
with a percutaneous access [183].

14 UPPER URINARY TRACT UROTHELIAL CARCINOMA - LIMITED UPDATE 2021


7.1.1.3 Ureteral resection
Segmental ureteral resection with wide margins provides adequate pathological specimens for staging and
grading while preserving the ipsilateral kidney. Lymphadenectomy can also be performed during segmental
ureteral resection [177]. Segmental resection of the proximal two-thirds of ureter is associated with higher
failure rates than for the distal ureter [184, 185] (LE: 3).
Distal ureterectomy with ureteroneocystostomy are indicated for low-risk tumours in the distal ureter
that cannot be removed completely endoscopically and for high-risk tumours when kidney-sparing surgery
for renal function preservation is desired (in case of an imperative indication) [101, 184, 185] (LE: 3). A total
ureterectomy with an ileal-ureteral substitution is technically feasible, but only in selected cases when a renal-
sparing procedure is mandatory and the tumour is low risk [186].

7.1.1.4 Upper urinary tract instillation of topical agents


The antegrade instillation of BCG or mitomycin C in the upper urinary tract via percutaneous nephrostomy
after complete tumour eradication has been studied for CIS after kidney-sparing management [160, 187]
(LE: 3). Retrograde instillation through a single J open-ended ureteric stent is also used. Both the antegrade
and retrograde approach can be dangerous due to possible ureteric obstruction and consecutive pyelovenous
influx during instillation/perfusion. The reflux obtained from a double-J stent has been used but this approach
is suboptimal because the drug often does not reach the renal pelvis [188-191].
A systematic review and meta-analysis assessing the oncologic outcomes of patients with papillary
UTUC or CIS of the upper tract treated with kidney-sparing surgery and adjuvant endocavitary treatment
analysed the effect of adjuvant therapies (i.e., chemotherapeutic agents and/or immunotherapy with BCG)
after kidney-sparing surgery for papillary non-invasive (Ta-T1) UTUCs and of adjuvant BCG for the treatment
of upper tract CIS, finding no difference between the method of drug administration (antegrade vs. retrograde
vs. combined approach) in terms of recurrence, progression, CSS, and OS. Furthermore, the recurrence
rates following adjuvant instillations are comparable to those reported in the literature in untreated patients,
questioning their efficacy [192]. The analyses were based on retrospective small studies suffering from
publication and reporting bias.
A single-arm phase III RCT showed that the use of mitomycin-containing reverse thermal gel
(UGN-101) instillations via retrograde catheter to the renal pelvis and calyces was associated with a complete
response rate in a total of 42 patients (59%) in biopsy-proven low-grade UTUC measuring less than 15 mm.
The median follow-up of patients with a complete response was 11 months. The most frequently reported
all-cause adverse events were ureteric stenosis in 31 (44%) of 71 patients, urinary tract infection in 23 (32%),
haematuria in 22 (31%), flank pain in 21 (30%), and nausea in 17 (24%). A total of 19 (27%) of 71 patients had
drug-related or procedure-related serious adverse events. No deaths were regarded as related to treatment
[193].

7.1.1.5 Guidelines for kidney-sparing management of UTUC

Recommendations Strength rating


Offer kidney-sparing management as primary treatment option to patients with low-risk Strong
tumours.
Offer kidney-sparing management (distal ureterectomy) to patients with high-risk tumours Weak
limited to the distal ureter.
Offer kidney-sparing management to patients with solitary kidney and/or impaired renal Strong
function, providing that it will not compromise survival. This decision will have to be made
on a case-by-case basis in consultation with the patient.

7.1.2 Management of high-risk non-metastatic UTUC


7.1.2.1 Surgical approach
7.1.2.1.1 Open radical nephroureterectomy
Open RNU with bladder cuff excision is the standard treatment of high-risk UTUC, regardless of tumour
location [21] (LE: 3). Radical nephroureterectomy must be performed according to oncological principles
preventing tumour seeding [21]. Section 7.1.6 lists the recommendations for RNU.

7.1.2.1.2 Minimal invasive radical nephroureterectomy


Retroperitoneal metastatic dissemination and metastasis along the trocar pathway following manipulation
of large tumours in a pneumoperitoneal environment have been reported in few cases [194, 195]. Several
precautions may lower the risk of tumour spillage:

UPPER URINARY TRACT UROTHELIAL CARCINOMA - LIMITED UPDATE 2021 15


1. avoid entering the urinary tract;
2. avoid direct contact between instruments and the tumour;
3. perform the procedure in a closed system. Avoid morcellation of the tumour and use an
endobag for tumour extraction;
4. the kidney and ureter must be removed en bloc with the bladder cuff;
5. Invasive or large (T3/T4 and/or N+/M+) tumours are contraindications for minimal-invasive
RNU as the outcome is worse compared to an open approach [196, 197].

Laparoscopic RNU is safe in experienced hands when adhering to strict oncological principles. There is a
tendency towards equivalent oncological outcomes after laparoscopic or open RNU [195, 198-201] (LE:
3). One prospective randomised study has shown that laparoscopic RNU is inferior to open RNU for non-
organ confined UTUC. However, this was a small trial (n = 80), which was likely underpowered [197] (LE: 2).
Oncological outcomes after RNU have not changed significantly over the past three decades despite staging
and surgical refinements [202] (LE: 3). A robot-assisted laparoscopic approach can be considered with recent
data suggesting oncologic equivalence with the other approaches [203-205].

7.1.2.1.3 Management of bladder cuff


Resection of the distal ureter and its orifice is performed because there is a considerable risk of tumour
recurrence in this area and in the bladder [175, 184, 206-208]. Several techniques have been considered to
simplify distal ureter resection, including the pluck technique, stripping, transurethral resection of the intramural
ureter, and intussusception. None of these techniques has convincingly been shown to be equal to complete
bladder cuff excision [15, 206, 207] (LE: 3).

7.1.2.1.4 Lymph node dissection


The use of a LND template is likely to have a greater impact on patient survival than the number of removed
LNs [209]. Template-based and completeness of LND improves CSS in patients with muscle-invasive disease
and reduces the risk of local recurrence [210]. Even in clinically [211] and pathologically [212] node-negative
patients, LND improves survival. The risk of LN metastasis increases with advancing tumour stage [146].
Lymph node dissection appears to be unnecessary in cases of TaT1 UTUC because of the low risk of LN
metastasis [213-216], however, tumour staging is inaccurate pre-operatively; therefore a template-based LND
should be offered to all patients who are scheduled for RNU. The templates for LND have been described [210,
217, 218].

7.1.3 Peri-operative chemotherapy


7.1.3.1 Neoadjuvant chemotherapy
In patients treated prior to losing their renal reserve several retrospective studies evaluating the role of
neoadjuvant chemotherapy have shown promising pathological downstaging and complete response rates
[156, 219-222]. In addition, neoadjuvant chemotherapy has been shown to result in lower disease recurrence
and mortality rates compared to RNU alone without compromising the use of definitive surgical treatment
[221, 223-225]. No RCTs have been published yet but prospective data from a phase II trial showed that the
use of neoadjuvant chemotherapy was associated with a 14% pathological complete response rate for high-
grade UTUC [226]. In addition, final pathological stage was ≤ ypT1 in more than 60% of included patients with
acceptable toxicity profile.

7.1.3.2 Adjuvant chemotherapy


A phase III prospective randomised trial (n = 261) evaluating the benefit of adjuvant gemcitabine-platinum
combination chemotherapy initiated within 90 days after RNU vs. surveillance has reported a significant
improvement in disease-free survival in patients with pT2–pT4, N (any) or LN-positive (pT any, N1–3) M0 UTUC
[227] (LE: 1).
The main limitation of using adjuvant chemotherapy for advanced UTUC remains the limited ability
to deliver full dose cisplatin-based regimen after RNU, given that this surgical procedure is likely to impact renal
function [228, 229]. In a retrospective study histological variants of UTUC exhibit different survival rates and
adjuvant chemotherapy was only associated with an OS benefit in patients with pure urothelial carcinoma [230]
(LE: 3).

7.1.4 Adjuvant radiotherapy after radical nephroureterectomy


Adjuvant radiation therapy has been suggested to control loco-regional disease after surgical removal.
The data remains controversial and insufficient for conclusions [232-235]. Moreover, its added value to
chemotherapy remains questionable [234].

16 UPPER URINARY TRACT UROTHELIAL CARCINOMA - LIMITED UPDATE 2021


7.1.5 Post-operative bladder instillation
The rate of bladder recurrence after RNU for UTUC is 22–47% [164, 207]. Two prospective randomised
trials [236, 237] and two meta-analyses [238, 239] have demonstrated that a single post-operative dose of
intravesical chemotherapy (mitomycin C, pirarubicin) 2–10 days after surgery reduces the risk of bladder
tumour recurrence within the first years post-RNU (LE: 2). Prior to instillation, a cystogram might be considered
in case of any concerns about drug extravasation.

Based on current evidence it is unlikely that additional instillations beyond one peri-operative instillation of
chemotherapy further substantially reduces the risk of intravesical recurrence [240]. Whilst there is no direct
evidence supporting the use of intravesical instillation of chemotherapy after kidney-sparing surgery, single-
dose chemotherapy might be effective in that setting as well (LE: 4). Management is outlined in Figures 7.1 and
7.2. One low-level evidence study suggested that bladder irrigation might reduce the risk of bladder recurrence
after RNU [241].

7.1.6 Summary of evidence and guidelines for the management of high-risk non-metastatic UTUC

Summary of evidence LE
Radical nephroureterectomy is the standard treatment for high-risk UTUC, regardless of tumour 2a
location.
Open, laparoscopic and robotic approaches have similar oncological outcomes for organ-confined 2a
UTUC.
Failure to completely remove the bladder cuff increases the risk of bladder cancer recurrence. 3
Lymphadenectomy improves survival in muscle-invasive UTUC. 3
Post-operative chemotherapy improves disease-free survival. 1b
Single post-operative intravesical instillation of chemotherapy lowers the bladder cancer recurrence 1
rate.

Recommendations Strength rating


Perform radical nephroureterectomy (RNU) in patients with high-risk non-metastatic upper Strong
tract urothelial carcinoma (UTUC).
Perform open RNU in non-organ confined UTUC. Weak
Remove the bladder cuff in its entirety. Strong
Perform a template-based lymphadenectomy in patients with muscle-invasive UTUC. Strong
Offer post-operative systemic platinum-based chemotherapy to patients with muscle- Strong
invasive UTUC.
Deliver a post-operative bladder instillation of chemotherapy to lower the intravesical Strong
recurrence rate.

UPPER URINARY TRACT UROTHELIAL CARCINOMA - LIMITED UPDATE 2021 17


Figure 7.1: Proposed flowchart for the management of UTUC

UTUC

Diagnosc evaluaon:
CTU, urinary cytology, cystoscopy

+/- Flexible ureteroscopy with biopsies

Low-risk UTUC High-risk UTUC*

RNU +/- template lymphadenectomy


+/- peri-operave planum-based
combinaon chemotherapy
Kidney-sparing surgery:
flexible ureteroscopy or segmental Open Laparoscopic
resecon (prefer open in cT3, cN+)
or percutaneous approach

Recurrence

Single post-operave dose of intravesical


Close and stringent follow-up chemotherapy

*In patients with solitary kidney, consider a more conservative approach.


CTU = computed tomography urography; RNU = radical nephroureterectomy;
UTUC = upper urinary tract urothelial carcinoma.

18 UPPER URINARY TRACT UROTHELIAL CARCINOMA - LIMITED UPDATE 2021


UTUC

Ureter Kidney

UTUC = upper urinary tract urothelial carcinoma.


*In case not amendable to endoscopic management.
Mid & Proximal Distal Calyx Renal pelvis

1 = first treatment option; 2 = secondary treatment option.

UPPER URINARY TRACT UROTHELIAL CARCINOMA - LIMITED UPDATE 2021


Figure 7.2: Surgical treatment according to location and risk status

Low-risk High-risk Low-risk High-risk Low-risk High-risk Low-risk High-risk

LND = lymph node dissection; RNU = radical nephroureterectomy; URS = ureteroscopy;


1. URS • RNU 1. URS 1. RNU 1. URS • RNU 1. URS • RNU
2. Uretero- +/- LND or or 2. RNU* +/- LND 2. Percutaneous +/- LND
ureterostomy* +/- peri- 2. Distal 2. Distal +/- post- +/- post-
operave ureterec- ureterectomy operave operave
chemo tomy +/- LND chemo chemo
+/- peri-
operave
chemo

19
7.2 Metastatic disease
7.2.1 Radical nephroureterectomy
The role of RNU in the treatment of patients with metastatic UTUC has recently been explored in several
observational studies [242-245]. Although evidence remains very limited, RNU may be associated with cancer-
specific [242, 244, 245] and OS benefit in selected patients, especially those fit enough to receive cisplatin-
based chemotherapy [243, 244]. It is noteworthy that these benefits may be limited to those patients with only
one metastatic site [244]. Nonetheless, given the high risk of bias of the observational studies addressing RNU
for metastatic UTUC, indications for RNU in this setting should mainly be reserved for palliative patients, aimed
at controlling symptomatic disease [17, 108] (LE: 3).

7.2.2 Metastasectomy
There is no UTUC-specific study supporting the role of metastasectomy in patients with advanced disease.
However, several reports including both UTUC and bladder cancer patients suggested that resection of
metastatic lesions could be safe and oncologically beneficial in selected patients with a life expectancy of more
than six months [246-248]. This was confirmed in the most recent and largest study to date [249]. Nonetheless,
in the absence of data from randomised controlled trials, patients should be evaluated on an individual basis
and the decision to perform a metastasectomy (surgically or otherwise) should be done in a shared decision-
making process with the patient.

7.2.3 Systemic treatments


7.2.3.1 First-line setting
Extrapolating from the bladder cancer literature and small, single-centre, UTUC studies, platinum-based
combination chemotherapy, especially using cisplatin, is likely to be efficacious as first-line treatment of
metastatic UTUC. A retrospective analysis of three RCTs showed that primary tumour location in the lower- or
upper urinary tract had no impact on progression-free or OS in patients with locally advanced or metastatic UC
treated with platinum-based combination chemotherapy [250].

The efficacy of immunotherapy using programmed death-1 (PD1) or programmed death-ligand 1 (PD-L1)
inhibitors has been evaluated in the first-line setting for the treatment of patients with metastactic urothelial
carcinoma, including those with UTUC.
Data from a phase III RCT showed that the use of avelumab maintenance therapy after 4 to 6
cycles of gemcitabin plus cisplatin or carboplatin significantly prolonged OS as compared to best supportive
care alone in those patients with advanced or metastatic urothelial carcinoma who did not progress during,
or responded to, first-line chemotherapy [251]. Although no subgroup analysis based on tumour location was
available in this study, almost 30% of the included patients had UTUC.
Atezolizumab was associated with an objective response rate of 39% in 33 (28%) cisplatin-ineligible
patients with metastatic UTUC included in a single-arm phase II trial (n = 119) [252]. Median OS in the overall
cohort was 15.9 months and toxicity was acceptable.
However, data from a phase III RCT including 1,213 patients with metastatic urothelial carcinoma,
of which 312 (26%) were diagnosed with UTUC patients showed that the combination of atezolizumab with
platinum-based chemotherapy was associated with limited PFS benefit as compared to platinum-based
chemotherapy alone and no significant impact was observed on OS.
In a single-arm phase II trial (n = 370) pembrolizumab was associated with an objective response
rate of 22% in 69 (19%) cisplatin-ineligible patients with metastatic UTUC [253]. In the overall cohort, a PD-L1
expression of 10% was associated with a greater response rate to pembrolizumab, which had acceptable
toxicity. However, addition of pembrolizumab to platinum-based chemotherapy did not reach statistical
significance for PFS and OS benefits as compared to platinum-based chemotherapy alone in a phase III RCT
in 351 patients with metastatic urothelial carcinoma including 64 (18%) UTUC patients [254]. Another negative
phase III RCT showed that durvalumab alone, or in combination with tremelimumab, did not prolong OS as
compared to platinum-based chemotherapy for metastatic urothelial carcinoma [255]. This study included 221
(21%) UTUC patients and subgroup analyses suggested that durvalumab alone may have better efficacy for
these individuals as compared to those with bladder cancer.

No other data are currently available in the first-line setting but several trials are currently testing
immunotherapy combinations with nivolumab (NCT03036098 [256]), pembrolizumab (NCT02178722 [257]) or
durvalumab (NCT03682068 [258]) in patients with metastatic UC, including those with UTUC.

7.2.3.2 Second-line setting


Similar to the bladder cancer setting, second-line treatment of metastatic UTUC remains challenging. In a
post-hoc subgroup analysis of locally advanced or metastatic UC, vinflunine was reported to be as effective in
UTUC as for bladder cancer progressing after cisplatin-based chemotherapy [259].

20 UPPER URINARY TRACT UROTHELIAL CARCINOMA - LIMITED UPDATE 2021


More importantly, a phase III RCT including 542 patients who received prior platinum-based
chemotherapy for advanced UC showed that pembrolizumab could decrease the risk of death by almost 50%
in those with UTUC (n = 75, 13.8%), although these results were borderline significant [260]. The objective
response rate was 21.1% in the overall cohort and median OS was 10.3 months. Interestingly, although no
subgroup analysis was available for UTUC patients (n = 65/21%), a single-arm phase II trial demonstrated
that atezolizumab has durable activity associated with PD-L1 expression on immune cells in patients with
metastatic UC [261]. The objective response rate was 26% in the group of those overexpressing PD-L1 and
15% in the overall population. However, a phase III RCT, including 51 (21.8%) patients with UTUC, showed
that it was not associated with prolonged OS as compared to chemotherapy in patients overexpressing PD-L1,
despite a more favourable safety profile [262].
Other immunotherapies such as nivolumab [263], avelumab [264, 265] and durvalumab [266] have
shown objective response rates ranging from 17.8% [266] to 19.6% [263] and median OS ranging from 7.7
months to 18.2 months in patients with platinum-resistant metastatic UC, overall. These results were obtained
from single-arm phase I or II trials only and the number of UTUC patients included in these studies was only
specified for avelumab (n = 7/15.9%) [265] without any subgroup analysis based on primary tumour location.
The immunotherapy combination of nivolumab plus ipilimumab has shown significant anti-tumour
activity with objective response rate up to 38% in a phase I/II multicentre trial including 78 patients with
metastatic UC progressing after platinum-based chemotherapy [267]. Although UTUC patients were included
in this trial, no subgroup analysis was available. Other immunotherapy combinations may be effective in the
second-line setting but data are currently limited [268].

In a recent phase II study the use of erdafitinib, a tyrosine kinase inhibitor of FGFR1-4, was associated with a
40% response rate in 99 patients with metastatic urothelial carcinoma and prespecified FGFR alterations who
progressed after first-line chemotherapy [269]. This study included 23 UTUC patients with visceral metastases
showing a 43% response rate.

7.2.4 Summary of evidence and guidelines for the treatment of metastatic UTUC

Summary of evidence LE
Radical nephroureterectomy may improve quality of life and oncologic outcomes in select metastatic 3
patients.
Cisplatin-based combination chemotherapy can improve median survival. 2
Single-agent and carboplatin-based combination chemotherapy are less effective than cisplatin-based 3
combination chemotherapy in terms of complete response and survival.
Non-platinum combination chemotherapy has not been tested against standard chemotherapy in 4
patients who are fit or unfit for cisplatin combination chemotherapy.
PD-1 inhibitor pembrolizumab has been approved for patients who have progressed during or after 1b
previous platinum-based chemotherapy based on the results of a phase III trial.
PD-L1 inhibitor atezolizumab has been FDA approved for patients that have progressed during or after 2a
previous platinum-based chemotherapy based on the results of a phase II trial.
PD-1 inhibitor nivolumab has been approved for patients that have progressed during or after previous 2a
platinum-based chemotherapy based on the results of a phase II trial.
PD-1 inhibitor pembrolizumab has been approved for patients with advanced or metastatic UC 2a
ineligible for cisplatin-based first-line chemotherapy based on the results of a phase II trial but use of
pembrolizumab is restricted to PD-L1 positive patients.
PD-L1 inhibitor atezolizumab has been approved for patients with advanced or metastatic UC 2a
ineligible for cisplatin-based first-line chemotherapy based on the results of a phase II trial but use of
atezolizumab is restricted to PD-L1 positive patients.

Recommendations Strength rating


Offer radical nephroureterectomy as a palliative treatment to symptomatic patients with Weak
resectable locally advanced tumours.
First-line treatment for cisplatin-eligible patients
Use cisplatin-containing combination chemotherapy with GC or HD-MVAC. Strong
Do not offer carboplatin or non-platinum combination chemotherapy. Strong
First-line treatment in patients unfit for cisplatin
Offer checkpoint inhibitors pembrolizumab or atezolizumab depending on PD-L1 status. Weak
Offer carboplatin combination chemotherapy if PD-L1 is negative. Strong

UPPER URINARY TRACT UROTHELIAL CARCINOMA - LIMITED UPDATE 2021 21


Second-line treatment
Offer checkpoint inhibitor (pembrolizumab) to patients with disease progression during or Strong
after platinum-based combination chemotherapy for metastatic disease.
Offer checkpoint inhibitor (atezolizumab or nivolumab) to patients with disease progression Strong
during or after platinum-based combination chemotherapy for metastatic disease.
Only offer vinflunine to patients for metastatic disease as second-line treatment if Strong
immunotherapy or combination chemotherapy is not feasible. Alternatively, offer vinflunine
as third- or subsequent-line treatment.
GC = gemcitabine plus cisplatin; HD-MVAC = high-dose intensity methotrexate, vinblastine, adriamycin plus
cisplatin; PD-L1 = programmed death ligand 1; PCG = paclitaxel, cisplatin, gemcitabine.

8. FOLLOW-UP
The risk of recurrence and death evolves during the follow-up period after surgery [270]. Stringent follow-up is
mandatory to detect metachronous bladder tumours (probability increases over time [271]), local recurrence,
and distant metastases. Section 8.1 presents the summary of evidence and recommendations for follow-up of
UTUC.
Surveillance regimens are based on cystoscopy and urinary cytology for > 5 years [12, 14, 15, 164].
Bladder recurrence is not considered a distant recurrence. When kidney-sparing surgery is performed, the
ipsilateral UUT requires careful and long-term follow-up due to the high risk of disease recurrence [180, 272,
273] and progression to RNU beyond 5 years [274]. Despite endourological improvements, follow-up after
kidney-sparing management is difficult and frequent, and repeated endoscopic procedures are necessary.
Following kidney-sparing surgery, and as done in bladder cancer, an early repeated (second look) ureteroscopy
within six to eight weeks after primary endoscopic treatment has been proposed, but is not yet routine practice
[2, 181].

8.1 Summary of evidence and guidelines for the follow-up of UTUC

Summary of evidence LE
Follow-up is more frequent and more stringent in patients who have undergone kidney-sparing 3
treatment compared to radical nephroureterectomy.

Recommendations Strength rating


After radical nephroureterectomy
Low-risk tumours
Perform cystoscopy at three months. If negative, perform subsequent cystoscopy nine Weak
months later and then yearly, for five years.
High-risk tumours
Perform cystoscopy and urinary cytology at three months. If negative, repeat subsequent Weak
cystoscopy and cytology every three months for a period of two years, and every six
months thereafter until five years, and then yearly.
Perform computed tomography (CT) urography and chest CT every six months for two years, Weak
and then yearly.
After kidney-sparing management
Low-risk tumours
Perform cystoscopy and CT urography at three and six months, and then yearly for five Weak
years.
Perform ureteroscopy (URS) at three months. Weak
High-risk tumours
Perform cystoscopy, urinary cytology, CT urography and chest CT at three and six months, Weak
and then yearly.
Perform URS and urinary cytology in situ at three and six months. Weak

22 UPPER URINARY TRACT UROTHELIAL CARCINOMA - LIMITED UPDATE 2021


9. REFERENCES
1. Babjuk, M., et al., EAU Guidelines on Non-muscle-invasive Bladder Cancer (T1, T1 and CIS), in EAU
Guidelines, Edn. presented at the 36th EAU Annual Congress Milan. 2021, EAU Guidelines Office.
https://uroweb.org/guideline/non-muscle-invasive-bladder-cancer
2. Witjes, J.A., et al., EAU Guidelines on Muscle-invasive and Metastatic Bladder Cancer in EAU
Guidelines, Edn. presented at the 36th EAU Annual Congress Milan. 2021, EAU Guidelines Office.
https://uroweb.org/guideline/bladder-cancer-muscle-invasive-and-metastatic
3. Gakis, G., et al., EAU Guidelines on Primary Urethral Carcinoma, in EAU Guidelines, Edn. presented
at the 36th EAU Annual Congress, Milan. 2021, EAU Guidelines Office.
https://uroweb.org/guideline/primary-urethral-carcinoma
4. Rouprêt, M., et al. European Association of Urology Guidelines on Upper Urinary Tract Urothelial
Carcinoma: 2020 Update. Eur Urol, 2021, 79: 62.
https://uroweb.org/guideline/upper-urinary-tract-urothelial-cell-carcinoma
5. Phillips B, et al. Oxford Centre for Evidence-based Medicine Levels of Evidence, 1998. Updated by
Jeremy Howick March 2009. [access date March 2021].
https://www.cebm.ox.ac.uk/resources/levels-of-evidence/oxford-centre-for-evidence-based-
medicine-levels-of-evidence-march-2009
6. Guyatt, G.H., et al. GRADE: an emerging consensus on rating quality of evidence and strength of
recommendations. BMJ, 2008. 336: 924.
https://pubmed.ncbi.nlm.nih.gov/18436948
7. Guyatt, G.H., et al. What is “quality of evidence” and why is it important to clinicians? BMJ, 2008.
336: 995.
https://pubmed.ncbi.nlm.nih.gov/18456631
8. Guyatt, G.H., et al. Going from evidence to recommendations. BMJ, 2008. 336: 1049.
https://pubmed.ncbi.nlm.nih.gov/18467413
9. Siegel, R.L., et al. Cancer Statistics, 2021. CA Cancer J Clin, 2021. 71: 7.
https://pubmed.ncbi.nlm.nih.gov/33433946
10. Soria, F., et al. Epidemiology, diagnosis, preoperative evaluation and prognostic assessment of
upper-tract urothelial carcinoma (UTUC). World J Urol, 2017. 35: 379.
https://pubmed.ncbi.nlm.nih.gov/27604375
11. Green, D.A., et al. Urothelial carcinoma of the bladder and the upper tract: disparate twins. J Urol,
2013. 189: 1214.
https://pubmed.ncbi.nlm.nih.gov/23023150
12. Cosentino, M., et al. Upper urinary tract urothelial cell carcinoma: location as a predictive factor for
concomitant bladder carcinoma. World J Urol, 2013. 31: 141.
https://pubmed.ncbi.nlm.nih.gov/22552732
13. Singla, N., et al. A Multi-Institutional Comparison of Clinicopathological Characteristics and
Oncologic Outcomes of Upper Tract Urothelial Carcinoma in China and the United States. J Urol,
2017. 197: 1208.
https://pubmed.ncbi.nlm.nih.gov/27887951
14. Xylinas, E., et al. Multifocal Carcinoma In Situ of the Upper Tract Is Associated With High Risk of
Bladder Cancer Recurrence. European Urology. 61: 1069.
https://pubmed.ncbi.nlm.nih.gov/22402109
15. Li, W.M., et al. Oncologic outcomes following three different approaches to the distal ureter and
bladder cuff in nephroureterectomy for primary upper urinary tract urothelial carcinoma. Eur Urol,
2010. 57: 963.
https://pubmed.ncbi.nlm.nih.gov/20079965
16. Miller, E.B., et al. Upper tract transitional cell carcinoma following treatment of superficial bladder
cancer with BCG. Urology, 1993. 42: 26.
https://pubmed.ncbi.nlm.nih.gov/8328123
17. Herr, H.W. Extravesical tumor relapse in patients with superficial bladder tumors. J Clin Oncol, 1998.
16: 1099.
https://pubmed.ncbi.nlm.nih.gov/9508196
18. Nishiyama, N., et al. Upper tract urothelial carcinoma following intravesical bacillus Calmette-Guérin
therapy for nonmuscle-invasive bladder cancer: Results from a multi-institutional retrospective
study. Urol Oncol, 2018. 36: 306.e9.
https://pubmed.ncbi.nlm.nih.gov/29550096

UPPER URINARY TRACT UROTHELIAL CARCINOMA - LIMITED UPDATE 2021 23


19. Sanderson, K.M., et al. Upper urinary tract tumour after radical cystectomy for transitional cell
carcinoma of the bladder: an update on the risk factors, surveillance regimens and treatments. BJU
Int, 2007. 100: 11.
https://pubmed.ncbi.nlm.nih.gov/17428248
20. Ayyathurai, R., et al. Monitoring of the upper urinary tract in patients with bladder cancer. Indian
J Urol, 2011. 27: 238.
https://pubmed.ncbi.nlm.nih.gov/21814316
21. Margulis, V., et al. Outcomes of radical nephroureterectomy: a series from the Upper Tract Urothelial
Carcinoma Collaboration. Cancer, 2009. 115: 1224.
https://pubmed.ncbi.nlm.nih.gov/19156917
22. Browne, B.M., et al. An Analysis of Staging and Treatment Trends for Upper Tract Urothelial
Carcinoma in the National Cancer Database. Clin Genitourin Cancer, 2018. 16: e743.
https://pubmed.ncbi.nlm.nih.gov/29506950
23. Shariat, S.F., et al. Gender differences in radical nephroureterectomy for upper tract urothelial
carcinoma. World J Urol, 2011. 29: 481.
https://pubmed.ncbi.nlm.nih.gov/20886219
24. Audenet, F., et al. Clonal Relatedness and Mutational Differences between Upper Tract and Bladder
Urothelial Carcinoma. Clin Cancer Res, 2019. 25: 967.
https://pubmed.ncbi.nlm.nih.gov/30352907
25. Umar, A., et al. Revised Bethesda Guidelines for hereditary nonpolyposis colorectal cancer (Lynch
syndrome) and microsatellite instability. J Natl Cancer Inst, 2004. 96: 261.
https://pubmed.ncbi.nlm.nih.gov/14970275
26. Therkildsen, C., et al. Molecular subtype classification of urothelial carcinoma in Lynch syndrome.
Mol Oncol, 2018. 12: 1286.
https://pubmed.ncbi.nlm.nih.gov/29791078
27. Roupret, M., et al. Upper urinary tract urothelial cell carcinomas and other urological malignancies
involved in the hereditary nonpolyposis colorectal cancer (lynch syndrome) tumor spectrum. Eur
Urol, 2008. 54: 1226.
https://pubmed.ncbi.nlm.nih.gov/18715695
28. Acher, P., et al. Towards a rational strategy for the surveillance of patients with Lynch syndrome
(hereditary non-polyposis colon cancer) for upper tract transitional cell carcinoma. BJU Int, 2010.
106: 300.
https://pubmed.ncbi.nlm.nih.gov/20553255
29. Ju, J.Y., et al. Universal Lynch Syndrome Screening Should be Performed in All Upper Tract
Urothelial Carcinomas. Am J Surg Pathol, 2018. 42: 1549.
https://pubmed.ncbi.nlm.nih.gov/30148743
30. Metcalfe, M.J., et al. Universal Point of Care Testing for Lynch Syndrome in Patients with Upper
Tract Urothelial Carcinoma. J Urol, 2018. 199: 60.
https://pubmed.ncbi.nlm.nih.gov/28797715
31. Pradere, B., et al. Lynch syndrome in upper tract urothelial carcinoma: significance, screening, and
surveillance. Curr Opin Urol, 2017. 27: 48.
https://pubmed.ncbi.nlm.nih.gov/27533503
32. Audenet, F., et al. A proportion of hereditary upper urinary tract urothelial carcinomas are
misclassified as sporadic according to a multi-institutional database analysis: proposal of patient-
specific risk identification tool. BJU Int, 2012. 110: E583.
https://pubmed.ncbi.nlm.nih.gov/22703159
33. Colin, P., et al. Environmental factors involved in carcinogenesis of urothelial cell carcinomas of the
upper urinary tract. BJU Int, 2009. 104: 1436.
https://pubmed.ncbi.nlm.nih.gov/19689473
34. Dickman K.G., et al. Epidemiology and Risk Factors for Upper Urinary Urothelial Cancers, In: Upper
Tract Urothelial Carcinoma. 2015, Springer: New York, NY, USA.
35. McLaughlin, J.K., et al. Cigarette smoking and cancers of the renal pelvis and ureter. Cancer Res,
1992. 52: 254.
https://pubmed.ncbi.nlm.nih.gov/1728398
36. Crivelli, J.J., et al. Effect of smoking on outcomes of urothelial carcinoma: a systematic review of the
literature. Eur Urol, 2014. 65: 742.
https://pubmed.ncbi.nlm.nih.gov/23810104
37. Martin, C., et al. Familial Cancer Clustering in Urothelial Cancer: A Population-Based Case-Control
Study. J Natl Cancer Inst, 2018. 110: 527.
https://pubmed.ncbi.nlm.nih.gov/29228305

24 UPPER URINARY TRACT UROTHELIAL CARCINOMA - LIMITED UPDATE 2021


38. Chen C-H., et al. Arsenics and urothelial carcinoma. In: Health Hazards of Environmental Arsenic
Poisoning from Epidemic to Pandemic, Chen C.J., Chiou H.Y. (Eds) 2011, World Scientific: Taipei.
39. Aristolochic acids. Rep Carcinog, 2011. 12: 45.
https://pubmed.ncbi.nlm.nih.gov/21822318
40. Cosyns, J.P. Aristolochic acid and ‘Chinese herbs nephropathy’: a review of the evidence to date.
Drug Saf, 2003. 26: 33.
https://pubmed.ncbi.nlm.nih.gov/12495362
41. Grollman, A.P. Aristolochic acid nephropathy: Harbinger of a global iatrogenic disease. Environ Mol
Mutagen, 2013. 54: 1.
https://pubmed.ncbi.nlm.nih.gov/23238808
42. Rosenquist, T.A., et al. Mutational signature of aristolochic acid: Clue to the recognition of a global
disease. DNA Repair (Amst), 2016. 44: 205.
https://pubmed.ncbi.nlm.nih.gov/27237586
43. Jelakovic, B., et al. Aristolactam-DNA adducts are a biomarker of environmental exposure to
aristolochic acid. Kidney Int, 2012. 81: 559.
https://pubmed.ncbi.nlm.nih.gov/22071594
44. Chen, C.H., et al. Aristolochic acid-associated urothelial cancer in Taiwan. Proc Natl Acad Sci U S A,
2012. 109: 8241.
https://pubmed.ncbi.nlm.nih.gov/22493262
45. Nortier, J.L., et al. Urothelial carcinoma associated with the use of a Chinese herb (Aristolochia
fangchi). N Engl J Med, 2000. 342: 1686.
https://pubmed.ncbi.nlm.nih.gov/10841870
46. Sidorenko, V.S., et al. Bioactivation of the human carcinogen aristolochic acid. Carcinogenesis,
2014. 35: 1814.
https://pubmed.ncbi.nlm.nih.gov/24743514
47. Hoang, M.L., et al. Mutational signature of aristolochic acid exposure as revealed by whole-exome
sequencing. Sci Transl Med, 2013. 5: 197ra102.
https://pubmed.ncbi.nlm.nih.gov/23926200
48. Huang, C.C., et al. Gender Is a Significant Prognostic Factor for Upper Tract Urothelial Carcinoma:
A Large Hospital-Based Cancer Registry Study in an Endemic Area. Front Oncol, 2019. 9: 157.
https://pubmed.ncbi.nlm.nih.gov/30949449
49. Xiong, G., et al. Aristolochic acid containing herbs induce gender-related oncological differences in
upper tract urothelial carcinoma patients. Cancer Manag Res, 2018. 10: 6627.
https://pubmed.ncbi.nlm.nih.gov/30584358
50. Zaitsu, M., et al. Alcohol consumption and risk of upper-tract urothelial cancer. Cancer Epidemiol,
2017. 48: 36.
https://pubmed.ncbi.nlm.nih.gov/28364670
51. Roupret, M., et al. Genetic variability in 8q24 confers susceptibility to urothelial carcinoma of the
upper urinary tract and is linked with patterns of disease aggressiveness at diagnosis. J Urol, 2012.
187: 424.
https://pubmed.ncbi.nlm.nih.gov/22177160
52. Kiss, B., et al. Stenting Prior to Cystectomy is an Independent Risk Factor for Upper Urinary Tract
Recurrence. J Urol, 2017. 198: 1263.
https://pubmed.ncbi.nlm.nih.gov/28603003
53. Sakano, S., et al. Impact of variant histology on disease aggressiveness and outcome after
nephroureterectomy in Japanese patients with upper tract urothelial carcinoma. Int J Clin Oncol,
2015. 20: 362.
https://pubmed.ncbi.nlm.nih.gov/24964974
54. Ouzzane, A., et al. Small cell carcinoma of the upper urinary tract (UUT-SCC): report of a rare entity
and systematic review of the literature. Cancer Treat Rev, 2011. 37: 366.
https://pubmed.ncbi.nlm.nih.gov/21257269
55. Rink, M., et al. Impact of histological variants on clinical outcomes of patients with upper urinary
tract urothelial carcinoma. J Urol, 2012. 188: 398.
https://pubmed.ncbi.nlm.nih.gov/22698626
56. Mori, K., et al. Prognostic Value of Variant Histology in Upper Tract Urothelial Carcinoma Treated
with Nephroureterectomy: A Systematic Review and Meta-Analysis. J Urol, 2020. 203: 1075.
https://pubmed.ncbi.nlm.nih.gov/31479406
57. Perez-Montiel, D., et al. High-grade urothelial carcinoma of the renal pelvis: clinicopathologic study
of 108 cases with emphasis on unusual morphologic variants. Mod Pathol, 2006. 19: 494.
https://pubmed.ncbi.nlm.nih.gov/16474378

UPPER URINARY TRACT UROTHELIAL CARCINOMA - LIMITED UPDATE 2021 25


58. Desai, F.S., et al. Retrospective Evaluation of Risk Factors and Immunohistochemical Findings
for Pre-Neoplastic and Neoplastic lesions of Upper Urinary Tract in Patients with Chronic
Nephrolithiasis. Asian Pac J Cancer Prev, 2015. 16: 8293.
https://pubmed.ncbi.nlm.nih.gov/26745075
59. Zamboni, S., et al. Incidence and survival outcomes in patients with upper urinary tract urothelial
carcinoma diagnosed with variant histology and treated with nephroureterectomy. BJU Int, 2019.
124: 738.
https://pubmed.ncbi.nlm.nih.gov/30908835
60. Kim, J.K., et al. Variant histology as a significant predictor of survival after radical
nephroureterectomy in patients with upper urinary tract urothelial carcinoma. Urol Oncol, 2017.
35: 458 e9.
https://pubmed.ncbi.nlm.nih.gov/28347659
61. Albadine, R., et al. PAX8 (+)/p63 (-) immunostaining pattern in renal collecting duct carcinoma
(CDC): a useful immunoprofile in the differential diagnosis of CDC versus urothelial carcinoma of
upper urinary tract. Am J Surg Pathol, 2010. 34: 965.
https://pubmed.ncbi.nlm.nih.gov/20463571
62. Soukup, V., et al. Prognostic Performance and Reproducibility of the 1973 and 2004/2016 World
Health Organization Grading Classification Systems in Non-muscle-invasive Bladder Cancer: A
European Association of Urology Non-muscle Invasive Bladder Cancer Guidelines Panel Systematic
Review. Eur Urol, 2017. 72: 801.
https://pubmed.ncbi.nlm.nih.gov/28457661
63. Subiela, J.D., et al. Diagnostic accuracy of ureteroscopic biopsy in predicting stage and grade at
final pathology in upper tract urothelial carcinoma: Systematic review and meta-analysis. Eur J Surg
Oncol, 2020. 46: 1989.
https://pubmed.ncbi.nlm.nih.gov/32674841
64. Brierley, J.D., et al., TNM Classification of Malignant Tumours. 8th ed. 2016.
65. Moch, H., Humphrey, H., Ulbright, T.M., Reuter, V.E., editors. WHO Classification of Tumours of the
Urinary System and Male Genital Organs. Fourth edition. 2016, Lyon.
https://pubmed.ncbi.nlm.nih.gov/26935559
66. Sauter, G., et al. Tumours of the urinary system: non-invasive urothelial neoplasias, in WHO
classification of classification of tumours of the urinary system and male genital organs. Eble, J.N.,
Sauter, G., Isabell, A., Sesterhenn, J.I., Epstein, M.D., Epstein, J.I. Editors. 2004, IARC Press: Lyon.
https://isbndata.org/978-92-832-2415-0/pathology-and-genetics-of-tumours-of-the-urinary-system-
and-male-genital-organs-iarc-who-classification-of-tumours
67. Moss, T.J., et al. Comprehensive Genomic Characterization of Upper Tract Urothelial Carcinoma.
Eur Urol, 2017. 72: 641.
https://pubmed.ncbi.nlm.nih.gov/28601352
68. Inman, B.A., et al. Carcinoma of the upper urinary tract: predictors of survival and competing causes
of mortality. Cancer, 2009. 115: 2853.
https://pubmed.ncbi.nlm.nih.gov/19434668
69. Cowan, N.C. CT urography for hematuria. Nat Rev Urol, 2012. 9: 218.
https://pubmed.ncbi.nlm.nih.gov/22410682
70. Raman, J.D., et al. Does preoperative symptom classification impact prognosis in patients with
clinically localized upper-tract urothelial carcinoma managed by radical nephroureterectomy? Urol
Oncol, 2011. 29: 716.
https://pubmed.ncbi.nlm.nih.gov/20056458
71. Cowan, N.C., et al. Multidetector computed tomography urography for diagnosing upper urinary
tract urothelial tumour. BJU Int, 2007. 99: 1363.
https://pubmed.ncbi.nlm.nih.gov/17428251
72. Janisch, F., et al. Diagnostic performance of multidetector computed tomographic (MDCTU) in
upper tract urothelial carcinoma (UTUC): a systematic review and meta-analysis. World J Urol, 2020.
38: 1165.
https://pubmed.ncbi.nlm.nih.gov/31321509
73. Verhoest, G., et al. Predictive factors of recurrence and survival of upper tract urothelial carcinomas.
World J Urol, 2011. 29: 495.
https://pubmed.ncbi.nlm.nih.gov/21681525/
74. Millán-Rodríguez, F., et al. Conventional CT signs in staging transitional cell tumors of the upper
urinary tract. Eur Urol, 1999. 35: 318.
https://pubmed.ncbi.nlm.nih.gov/10087395/

26 UPPER URINARY TRACT UROTHELIAL CARCINOMA - LIMITED UPDATE 2021


75. Takahashi, N., et al. Gadolinium enhanced magnetic resonance urography for upper urinary tract
malignancy. J Urol, 2010. 183: 1330.
https://pubmed.ncbi.nlm.nih.gov/20171676
76. Razavi, S.A., et al. Comparative effectiveness of imaging modalities for the diagnosis of upper and
lower urinary tract malignancy: a critically appraised topic. Acad Radiol, 2012. 19: 1134.
https://pubmed.ncbi.nlm.nih.gov/22717592
77. Witjes, J.A., et al. Hexaminolevulinate-guided fluorescence cystoscopy in the diagnosis and
follow-up of patients with non-muscle-invasive bladder cancer: review of the evidence and
recommendations. Eur Urol, 2010. 57: 607.
https://pubmed.ncbi.nlm.nih.gov/20116164
78. Rosenthal D.L., et al. The Paris System for Reporting Urinary Cytology. 2016, Switzerland.
https://www.springer.com/gp/book/9783319228631
79. Messer, J., et al. Urinary cytology has a poor performance for predicting invasive or high-grade
upper-tract urothelial carcinoma. BJU Int, 2011. 108: 701.
https://pubmed.ncbi.nlm.nih.gov/21320275
80. Malm, C., et al. Diagnostic accuracy of upper tract urothelial carcinoma: how samples are collected
matters. Scand J Urol, 2017. 51: 137.
https://pubmed.ncbi.nlm.nih.gov/28385123
81. Messer, J.C., et al. Multi-institutional validation of the ability of preoperative hydronephrosis to predict
advanced pathologic tumor stage in upper-tract urothelial carcinoma. Urol Oncol, 2013. 31: 904.
https://pubmed.ncbi.nlm.nih.gov/21906967/
82. Wang, L.J., et al. Diagnostic accuracy of transitional cell carcinoma on multidetector computerized
tomography urography in patients with gross hematuria. J Urol, 2009. 181: 524.
https://pubmed.ncbi.nlm.nih.gov/19100576
83. Lee, K.S., et al. MR urography versus retrograde pyelography/ureteroscopy for the exclusion of
upper urinary tract malignancy. Clin Radiol, 2010. 65: 185.
https://pubmed.ncbi.nlm.nih.gov/20152273
84. McHale, T., et al. Comparison of urinary cytology and fluorescence in situ hybridization in the
detection of urothelial neoplasia: An analysis of discordant results. Diagn Cytopathol, 2019. 47: 282.
https://pubmed.ncbi.nlm.nih.gov/30417563
85. Jin, H., et al. A comprehensive comparison of fluorescence in situ hybridization and cytology for
the detection of upper urinary tract urothelial carcinoma: A systematic review and meta-analysis.
Medicine (Baltimore), 2018. 97: e13859.
https://pubmed.ncbi.nlm.nih.gov/30593189
86. Rojas, C.P., et al. Low biopsy volume in ureteroscopy does not affect tumor biopsy grading in upper
tract urothelial carcinoma. Urologic oncology, 2013. 31: 1696.
https://pubmed.ncbi.nlm.nih.gov/22819696
87. Smith, A.K., et al. Inadequacy of biopsy for diagnosis of upper tract urothelial carcinoma:
implications for conservative management. Urology, 2011. 78: 82.
https://pubmed.ncbi.nlm.nih.gov/21550642
88. Ishikawa, S., et al. Impact of diagnostic ureteroscopy on intravesical recurrence and survival in
patients with urothelial carcinoma of the upper urinary tract. J Urol, 2010. 184: 883.
https://pubmed.ncbi.nlm.nih.gov/20643446
89. Clements, T., et al. High-grade ureteroscopic biopsy is associated with advanced pathology of
upper-tract urothelial carcinoma tumors at definitive surgical resection. J Endourol, 2012. 26: 398.
https://pubmed.ncbi.nlm.nih.gov/22192113
90. Brien, J.C., et al. Preoperative hydronephrosis, ureteroscopic biopsy grade and urinary cytology can
improve prediction of advanced upper tract urothelial carcinoma. J Urol, 2010. 184: 69.
https://pubmed.ncbi.nlm.nih.gov/20478585
91. Marchioni, M., et al. Impact of diagnostic ureteroscopy on intravesical recurrence in patients
undergoing radical nephroureterectomy for upper tract urothelial cancer: a systematic review and
meta-analysis. BJU Int, 2017. 120: 313.
https://pubmed.ncbi.nlm.nih.gov/28621055
92. Guo, R.Q., et al. Impact of ureteroscopy before radical nephroureterectomy for upper tract urothelial
carcinomas on oncological outcomes: a meta-analysis. BJU Int, 2018. 121: 184.
https://pubmed.ncbi.nlm.nih.gov/29032580
93. Lee, H.Y., et al. The diagnostic ureteroscopy before radical nephroureterectomy in upper urinary
tract urothelial carcinoma is not associated with higher intravesical recurrence. World J Surg Oncol,
2018. 16: 135.
https://pubmed.ncbi.nlm.nih.gov/29986730

UPPER URINARY TRACT UROTHELIAL CARCINOMA - LIMITED UPDATE 2021 27


94. Bus, M.T., et al. Optical diagnostics for upper urinary tract urothelial cancer: technology, thresholds,
and clinical applications. J Endourol, 2015. 29: 113.
https://pubmed.ncbi.nlm.nih.gov/25178057
95. Knoedler, J.J., et al. Advances in the management of upper tract urothelial carcinoma: improved
endoscopic management through better diagnostics. Ther Adv Urol, 2018. 10: 421.
https://pubmed.ncbi.nlm.nih.gov/30574202
96. Breda, A., et al. Correlation Between Confocal Laser Endomicroscopy (Cellvizio((R))) and
Histological Grading of Upper Tract Urothelial Carcinoma: A Step Forward for a Better Selection of
Patients Suitable for Conservative Management. Eur Urol Focus, 2018. 4: 954.
https://pubmed.ncbi.nlm.nih.gov/28753800
97. Bus, M.T., et al. Optical Coherence Tomography as a Tool for In Vivo Staging and Grading of Upper
Urinary Tract Urothelial Carcinoma: A Study of Diagnostic Accuracy. J Urol, 2016. 196: 1749.
https://pubmed.ncbi.nlm.nih.gov/27475968
98. Voskuilen, C.S., et al. Diagnostic Value of (18)F-fluorodeoxyglucose Positron Emission Tomography
with Computed Tomography for Lymph Node Staging in Patients with Upper Tract Urothelial
Carcinoma. Eur Urol Oncol, 2020. 3: 73.
https://pubmed.ncbi.nlm.nih.gov/31591037
99. Jeldres, C., et al. A population-based assessment of perioperative mortality after nephroureterectomy
for upper-tract urothelial carcinoma. Urology, 2010. 75: 315.
https://pubmed.ncbi.nlm.nih.gov/19963237
100. Lughezzani, G., et al. Prognostic factors in upper urinary tract urothelial carcinomas: a comprehen-
sive review of the current literature. Eur Urol, 2012. 62: 100.
https://pubmed.ncbi.nlm.nih.gov/22381168
101. Lughezzani, G., et al. Nephroureterectomy and segmental ureterectomy in the treatment of invasive
upper tract urothelial carcinoma: A population-based study of 2299 patients. European Journal of
Cancer. 45: 3291.
https://pubmed.ncbi.nlm.nih.gov/19615885
102. Roupret, M., et al. Prediction of cancer specific survival after radical nephroureterectomy for upper
tract urothelial carcinoma: development of an optimized postoperative nomogram using decision
curve analysis. J Urol, 2013. 189: 1662.
https://pubmed.ncbi.nlm.nih.gov/23103802
103. Kim, H.S., et al. Association between demographic factors and prognosis in urothelial carcinoma of
the upper urinary tract: a systematic review and meta-analysis. Oncotarget, 2017. 8: 7464.
https://pubmed.ncbi.nlm.nih.gov/27448978
104. Shariat, S.F., et al. Advanced patient age is associated with inferior cancer-specific survival after
radical nephroureterectomy. BJU Int, 2010. 105: 1672.
https://pubmed.ncbi.nlm.nih.gov/19912201
105. Chromecki, T.F., et al. Chronological age is not an independent predictor of clinical outcomes after
radical nephroureterectomy. World J Urol, 2011. 29: 473.
https://pubmed.ncbi.nlm.nih.gov/21499902
106. Fernandez, M.I., et al. Evidence-based sex-related outcomes after radical nephroureterectomy for
upper tract urothelial carcinoma: results of large multicenter study. Urology, 2009. 73: 142.
https://pubmed.ncbi.nlm.nih.gov/18845322
107. Matsumoto, K., et al. Racial differences in the outcome of patients with urothelial carcinoma of the
upper urinary tract: an international study. BJU Int, 2011. 108: E304.
https://pubmed.ncbi.nlm.nih.gov/21507184
108. Simsir, A., et al. Prognostic factors for upper urinary tract urothelial carcinomas: stage, grade, and
smoking status. Int Urol Nephrol, 2011. 43: 1039.
https://pubmed.ncbi.nlm.nih.gov/21547471
109. Rink, M., et al. Impact of smoking on oncologic outcomes of upper tract urothelial carcinoma after
radical nephroureterectomy. Eur Urol, 2013. 63: 1082.
https://pubmed.ncbi.nlm.nih.gov/22743166
110. Xylinas, E., et al. Impact of smoking status and cumulative exposure on intravesical recurrence of
upper tract urothelial carcinoma after radical nephroureterectomy. BJU Int, 2014. 114: 56.
https://pubmed.ncbi.nlm.nih.gov/24053463
111. Shigeta, K., et al. A Novel Risk-based Approach Simulating Oncological Surveillance After Radical
Nephroureterectomy in Patients with Upper Tract Urothelial Carcinoma. Eur Urol Oncol, 2019.
https://pubmed.ncbi.nlm.nih.gov/31395480
112. Sundi, D., et al. Upper tract urothelial carcinoma: impact of time to surgery. Urol Oncol, 2012. 30: 266.
https://pubmed.ncbi.nlm.nih.gov/20869888

28 UPPER URINARY TRACT UROTHELIAL CARCINOMA - LIMITED UPDATE 2021


113. Gadzinski, A.J., et al. Long-term outcomes of immediate versus delayed nephroureterectomy for
upper tract urothelial carcinoma. J Endourol, 2012. 26: 566.
https://pubmed.ncbi.nlm.nih.gov/21879886
114. Lee, J.N., et al. Impact of surgical wait time on oncologic outcomes in upper urinary tract urothelial
carcinoma. J Surg Oncol, 2014. 110: 468.
https://pubmed.ncbi.nlm.nih.gov/25059848
115. Waldert, M., et al. A delay in radical nephroureterectomy can lead to upstaging. BJU Int, 2010.
105: 812.
https://pubmed.ncbi.nlm.nih.gov/19732052
116. Xia, L., et al. Impact of surgical waiting time on survival in patients with upper tract urothelial
carcinoma: A national cancer database study. Urol Oncol, 2018. 36: 10 e15.
https://pubmed.ncbi.nlm.nih.gov/29031419
117. Berod, A.A., et al. The role of American Society of Anesthesiologists scores in predicting urothelial
carcinoma of the upper urinary tract outcome after radical nephroureterectomy: results from a
national multi-institutional collaborative study. BJU Int, 2012. 110: E1035.
https://pubmed.ncbi.nlm.nih.gov/22568669
118. Carrion, A., et al. Intraoperative prognostic factors and atypical patterns of recurrence in patients
with upper urinary tract urothelial carcinoma treated with laparoscopic radical nephroureterectomy.
Scand J Urol, 2016. 50: 305.
https://pubmed.ncbi.nlm.nih.gov/26926709
119. Ehdaie, B., et al. Obesity adversely impacts disease specific outcomes in patients with upper tract
urothelial carcinoma. J Urol, 2011. 186: 66.
https://pubmed.ncbi.nlm.nih.gov/21571333
120. Yeh, H.C., et al. Interethnic differences in the impact of body mass index on upper tract urothelial
carcinoma following radical nephroureterectomy. World J Urol, 2020.
https://pubmed.ncbi.nlm.nih.gov/32318857
121. Dalpiaz, O., et al. Validation of the pretreatment derived neutrophil-lymphocyte ratio as a prognostic
factor in a European cohort of patients with upper tract urothelial carcinoma. Br J Cancer, 2014.
110: 2531.
https://pubmed.ncbi.nlm.nih.gov/24691424
122. Vartolomei, M.D., et al. Is neutrophil-to-lymphocytes ratio a clinical relevant preoperative biomarker
in upper tract urothelial carcinoma? A meta-analysis of 4385 patients. World J Urol, 2018. 36: 1019.
https://pubmed.ncbi.nlm.nih.gov/29468284
123. Mori, K., et al. Prognostic value of preoperative blood-based biomarkers in upper tract urothelial
carcinoma treated with nephroureterectomy: A systematic review and meta-analysis. Urol Oncol,
2020. 38: 315.
https://pubmed.ncbi.nlm.nih.gov/32088103
124. Zheng, Y., et al. Combination of Systemic Inflammation Response Index and Platelet-to-
Lymphocyte Ratio as a Novel Prognostic Marker of Upper Tract Urothelial Carcinoma After Radical
Nephroureterectomy. Front Oncol, 2019. 9: 914.
https://pubmed.ncbi.nlm.nih.gov/31620369
125. Liu, J., et al. The prognostic significance of preoperative serum albumin in urothelial carcinoma: a
systematic review and meta-analysis. Biosci Rep, 2018. 38.
https://pubmed.ncbi.nlm.nih.gov/29685957
126. Soria, F., et al. Prognostic value of the systemic inflammation modified Glasgow prognostic score
in patients with upper tract urothelial carcinoma (UTUC) treated with radical nephroureterectomy:
Results from a large multicenter international collaboration. Urol Oncol, 2020. 38: 602.e11.
https://pubmed.ncbi.nlm.nih.gov/32037197
127. Mori, K., et al. Prognostic role of preoperative De Ritis ratio in upper tract urothelial carcinoma
treated with nephroureterectomy. Urol Oncol, 2020. 38: 601.e17.
https://pubmed.ncbi.nlm.nih.gov/32127252
128. Momota, M., et al. The Impact of Preoperative Severe Renal Insufficiency on Poor Postsurgical
Oncological Prognosis in Patients with Urothelial Carcinoma. Eur Urol Focus, 2019. 5: 1066.
https://pubmed.ncbi.nlm.nih.gov/29548907
129. Xu, H., et al. Pretreatment elevated fibrinogen level predicts worse oncologic outcomes in upper
tract urothelial carcinoma. Asian J Androl, 2020. 22: 177.
https://pubmed.ncbi.nlm.nih.gov/31169138
130. Mbeutcha, A., et al. Prognostic factors and predictive tools for upper tract urothelial carcinoma: a
systematic review. World J Urol, 2017. 35: 337.
https://pubmed.ncbi.nlm.nih.gov/27101100

UPPER URINARY TRACT UROTHELIAL CARCINOMA - LIMITED UPDATE 2021 29


131. Petrelli, F., et al. Prognostic Factors of Overall Survival in Upper Urinary Tract Carcinoma: A
Systematic Review and Meta-analysis. Urology, 2017. 100: 9.
https://pubmed.ncbi.nlm.nih.gov/27516121
132. Yafi, F.A., et al. Impact of tumour location versus multifocality in patients with upper tract urothelial
carcinoma treated with nephroureterectomy and bladder cuff excision: a homogeneous series
without perioperative chemotherapy. BJU Int, 2012. 110: E7.
https://pubmed.ncbi.nlm.nih.gov/22177329
133. Ouzzane, A., et al. Ureteral and multifocal tumours have worse prognosis than renal pelvic tumours
in urothelial carcinoma of the upper urinary tract treated by nephroureterectomy. Eur Urol, 2011.
60: 1258.
https://pubmed.ncbi.nlm.nih.gov/21665356
134. Chromecki, T.F., et al. The impact of tumor multifocality on outcomes in patients treated with radical
nephroureterectomy. Eur Urol, 2012. 61: 245.
https://pubmed.ncbi.nlm.nih.gov/21975249
135. Williams, A.K., et al. Multifocality rather than tumor location is a prognostic factor in upper tract
urothelial carcinoma. Urol Oncol, 2013. 31: 1161.
https://pubmed.ncbi.nlm.nih.gov/23415596
136. Hurel, S., et al. Influence of preoperative factors on the oncologic outcome for upper urinary tract
urothelial carcinoma after radical nephroureterectomy. World J Urol, 2015. 33: 335.
https://pubmed.ncbi.nlm.nih.gov/24810657
137. Isbarn, H., et al. Location of the primary tumor is not an independent predictor of cancer specific
mortality in patients with upper urinary tract urothelial carcinoma. J Urol, 2009. 182: 2177.
https://pubmed.ncbi.nlm.nih.gov/19758662
138. Lwin, A.A., et al. Urothelial Carcinoma of the Renal Pelvis and Ureter: Does Location Make a
Difference? Clin Genitourin Cancer, 2020. 18: 45.
https://pubmed.ncbi.nlm.nih.gov/31786118
139. Ito, Y., et al. Preoperative hydronephrosis grade independently predicts worse pathological
outcomes in patients undergoing nephroureterectomy for upper tract urothelial carcinoma. J Urol,
2011. 185: 1621.
https://pubmed.ncbi.nlm.nih.gov/21419429
140. Foerster, B., et al. The Performance of Tumor Size as Risk Stratification Parameter in Upper Tract
Urothelial Carcinoma (UTUC). Clin Genitourin Cancer, 2020.
https://pubmed.ncbi.nlm.nih.gov/33046411
141. Yu, J., et al. Impact of squamous differentiation on intravesical recurrence and prognosis of patients
with upper tract urothelial carcinoma. Ann Transl Med, 2019. 7: 377.
https://pubmed.ncbi.nlm.nih.gov/31555691
142. Pelcovits, A., et al. Outcomes of upper tract urothelial carcinoma with isolated lymph node
involvement following surgical resection: implications for multi-modal management. World J Urol,
2020. 38: 1243.
https://pubmed.ncbi.nlm.nih.gov/31388818
143. Fajkovic, H., et al. Prognostic value of extranodal extension and other lymph node parameters in
patients with upper tract urothelial carcinoma. J Urol, 2012. 187: 845.
https://pubmed.ncbi.nlm.nih.gov/22248522
144. Roscigno, M., et al. Lymphadenectomy at the time of nephroureterectomy for upper tract urothelial
cancer. Eur Urol, 2011. 60: 776.
https://pubmed.ncbi.nlm.nih.gov/21798659
145. Raza, S.J., et al. Lymph node density for stratification of survival outcomes with node positive upper
tract urothelial carcinoma. Can J Urol, 2019. 26: 9852.
https://pubmed.ncbi.nlm.nih.gov/31469641
146. Lughezzani, G., et al. A critical appraisal of the value of lymph node dissection at
nephroureterectomy for upper tract urothelial carcinoma. Urology, 2010. 75: 118.
https://pubmed.ncbi.nlm.nih.gov/19864000
147. Nazzani, S., et al. Rates of lymph node invasion and their impact on cancer specific mortality in
upper urinary tract urothelial carcinoma. Eur J Surg Oncol, 2019. 45: 1238.
https://pubmed.ncbi.nlm.nih.gov/30563773
148. Kikuchi, E., et al. Lymphovascular invasion predicts clinical outcomes in patients with node-negative
upper tract urothelial carcinoma. J Clin Oncol, 2009. 27: 612.
https://pubmed.ncbi.nlm.nih.gov/19075275

30 UPPER URINARY TRACT UROTHELIAL CARCINOMA - LIMITED UPDATE 2021


149. Novara, G., et al. Prognostic role of lymphovascular invasion in patients with urothelial carcinoma of
the upper urinary tract: an international validation study. Eur Urol, 2010. 57: 1064.
https://pubmed.ncbi.nlm.nih.gov/20071073
150. Liu, W., et al. Prognostic Value of Lymphovascular Invasion in Upper Urinary Tract Urothelial
Carcinoma after Radical Nephroureterectomy: A Systematic Review and Meta-Analysis. Dis
Markers, 2019. 2019: 7386140.
https://pubmed.ncbi.nlm.nih.gov/31565103
151. Godfrey, M.S., et al. Prognostic indicators for upper tract urothelial carcinoma after radical
nephroureterectomy: the impact of lymphovascular invasion. BJU Int, 2012. 110: 798.
https://pubmed.ncbi.nlm.nih.gov/22313599
152. Samaratunga, H., et al. Data Set for the Reporting of Carcinoma of the Renal Pelvis and Ureter-
Nephroureterectomy and Ureterectomy Specimens: Recommendations From the International
Collaboration on Cancer Reporting (ICCR). Am J Surg Pathol, 2019. 43: e1.
https://pubmed.ncbi.nlm.nih.gov/31192862
153. Colin, P., et al. Influence of positive surgical margin status after radical nephroureterectomy on upper
urinary tract urothelial carcinoma survival. Ann Surg Oncol, 2012. 19: 3613.
https://pubmed.ncbi.nlm.nih.gov/22843187
154. Zigeuner, R., et al. Tumour necrosis is an indicator of aggressive biology in patients with urothelial
carcinoma of the upper urinary tract. Eur Urol, 2010. 57: 575.
https://pubmed.ncbi.nlm.nih.gov/19959276
155. Seitz, C., et al. Association of tumor necrosis with pathological features and clinical outcome
in 754 patients undergoing radical nephroureterectomy for upper tract urothelial carcinoma: an
international validation study. J Urol, 2010. 184: 1895.
https://pubmed.ncbi.nlm.nih.gov/20846680
156. Martini, A., et al. Pathological downstaging as a novel endpoint for the development of neoadjuvant
chemotherapy for upper tract urothelial carcinoma. BJU Int, 2019.
https://pubmed.ncbi.nlm.nih.gov/30801918
157. Remzi, M., et al. Tumour architecture is an independent predictor of outcomes after
nephroureterectomy: a multi-institutional analysis of 1363 patients. BJU Int, 2009. 103: 307.
https://pubmed.ncbi.nlm.nih.gov/18990163
158. Fritsche, H.M., et al. Macroscopic sessile tumor architecture is a pathologic feature of biologically
aggressive upper tract urothelial carcinoma. Urol Oncol, 2012. 30: 666.
https://pubmed.ncbi.nlm.nih.gov/20933445
159. Wheat, J.C., et al. Concomitant carcinoma in situ is a feature of aggressive disease in patients
with organ confined urothelial carcinoma following radical nephroureterectomy. Urol Oncol, 2012.
30: 252.
https://pubmed.ncbi.nlm.nih.gov/20451416
160. Redrow, G.P., et al. Upper Urinary Tract Carcinoma In Situ: Current Knowledge, Future Direction.
J Urol, 2017. 197: 287.
https://pubmed.ncbi.nlm.nih.gov/27664578
161. Roscigno, M., et al. International validation of the prognostic value of subclassification for AJCC
stage pT3 upper tract urothelial carcinoma of the renal pelvis. BJU Int, 2012. 110: 674.
https://pubmed.ncbi.nlm.nih.gov/22348322
162. Scarpini, S., et al. Impact of the expression of Aurora-A, p53, and MIB-1 on the prognosis of
urothelial carcinomas of the upper urinary tract. Urol Oncol, 2012. 30: 182.
https://pubmed.ncbi.nlm.nih.gov/20189840
163. Roupret, M., et al. A new proposal to risk stratify urothelial carcinomas of the upper urinary tract
(UTUCs) in a predefinitive treatment setting: low-risk versus high-risk UTUCs. Eur Urol, 2014.
66: 181.
https://pubmed.ncbi.nlm.nih.gov/24361259
164. Seisen, T., et al. Risk-adapted strategy for the kidney-sparing management of upper tract tumours.
Nat Rev Urol, 2015. 12: 155.
https://pubmed.ncbi.nlm.nih.gov/25708579
165. Margulis, V., et al. Preoperative multivariable prognostic model for prediction of nonorgan confined
urothelial carcinoma of the upper urinary tract. J Urol, 2010. 184: 453.
https://pubmed.ncbi.nlm.nih.gov/20620397
166. Favaretto, R.L., et al. Combining imaging and ureteroscopy variables in a preoperative multivariable
model for prediction of muscle-invasive and non-organ confined disease in patients with upper tract
urothelial carcinoma. BJU Int, 2012. 109: 77.
https://pubmed.ncbi.nlm.nih.gov/21631698

UPPER URINARY TRACT UROTHELIAL CARCINOMA - LIMITED UPDATE 2021 31


167. Petros, F.G., et al. Preoperative multiplex nomogram for prediction of high-risk nonorgan-confined
upper-tract urothelial carcinoma. Urol Oncol, 2019. 37: 292 e1.
https://pubmed.ncbi.nlm.nih.gov/30584035
168. Yoshida, T., et al. Development and external validation of a preoperative nomogram for predicting
pathological locally advanced disease of clinically localized upper urinary tract carcinoma. Cancer
Med, 2020. 9: 3733.
https://pubmed.ncbi.nlm.nih.gov/32253820
169. Cha, E.K., et al. Predicting clinical outcomes after radical nephroureterectomy for upper tract
urothelial carcinoma. Eur Urol, 2012. 61: 818.
https://pubmed.ncbi.nlm.nih.gov/22284969
170. Yates, D.R., et al. Cancer-specific survival after radical nephroureterectomy for upper urinary tract
urothelial carcinoma: proposal and multi-institutional validation of a post-operative nomogram.
Br J Cancer, 2012. 106: 1083.
https://pubmed.ncbi.nlm.nih.gov/22374463
171. Seisen, T., et al. Postoperative nomogram to predict cancer-specific survival after radical
nephroureterectomy in patients with localised and/or locally advanced upper tract urothelial
carcinoma without metastasis. BJU Int, 2014. 114: 733.
https://pubmed.ncbi.nlm.nih.gov/24447471
172. Ku, J.H., et al. External validation of an online nomogram in patients undergoing radical
nephroureterectomy for upper urinary tract urothelial carcinoma. Br J Cancer, 2013. 109: 1130.
https://pubmed.ncbi.nlm.nih.gov/23949152
173. Krabbe, L.M., et al. Postoperative Nomogram for Relapse-Free Survival in Patients with High Grade
Upper Tract Urothelial Carcinoma. J Urol, 2017. 197: 580.
https://pubmed.ncbi.nlm.nih.gov/27670916
174. Zhang, G.L., et al. A Model for the Prediction of Survival in Patients With Upper Tract Urothelial
Carcinoma After Surgery. Dose Response, 2019. 17: 1559325819882872.
https://pubmed.ncbi.nlm.nih.gov/31662711
175. Seisen, T., et al. A Systematic Review and Meta-analysis of Clinicopathologic Factors Linked
to Intravesical Recurrence After Radical Nephroureterectomy to Treat Upper Tract Urothelial
Carcinoma. Eur Urol, 2015. 67: 1122.
https://pubmed.ncbi.nlm.nih.gov/25488681
176. Zhang, X., et al. Development and Validation of a Model for Predicting Intravesical Recurrence
in Organ-confined Upper Urinary Tract Urothelial Carcinoma Patients after Radical
Nephroureterectomy: a Retrospective Study in One Center with Long-term Follow-up. Pathol Oncol
Res, 2020. 26: 1741.
https://pubmed.ncbi.nlm.nih.gov/31643022
177. Seisen, T., et al. Oncologic Outcomes of Kidney-sparing Surgery Versus Radical
Nephroureterectomy for Upper Tract Urothelial Carcinoma: A Systematic Review by the EAU Non-
muscle Invasive Bladder Cancer Guidelines Panel. European Urology. 70: 1052.
https://pubmed.ncbi.nlm.nih.gov/27477528
178. Cutress, M.L., et al. Long-term endoscopic management of upper tract urothelial carcinoma:
20-year single-centre experience. BJU Int, 2012. 110: 1608.
https://pubmed.ncbi.nlm.nih.gov/22564677
179. Cutress, M.L., et al. Ureteroscopic and percutaneous management of upper tract urothelial
carcinoma (UTUC): systematic review. BJU Int, 2012. 110: 614.
https://pubmed.ncbi.nlm.nih.gov/22471401
180. Cornu, J.N., et al. Oncologic control obtained after exclusive flexible ureteroscopic management of
upper urinary tract urothelial cell carcinoma. World J Urol, 2010. 28: 151.
https://pubmed.ncbi.nlm.nih.gov/20044752
181. Villa, L., et al. Early repeated ureteroscopy within 6-8 weeks after a primary endoscopic treatment
in patients with upper tract urothelial cell carcinoma: preliminary findings. World J Urol, 2016. 34:
1201.
https://pubmed.ncbi.nlm.nih.gov/26699629
182. Vemana, G., et al. Survival Comparison Between Endoscopic and Surgical Management for Patients
With Upper Tract Urothelial Cancer: A Matched Propensity Score Analysis Using Surveillance,
Epidemiology and End Results-Medicare Data. Urology, 2016. 95: 115.
https://pubmed.ncbi.nlm.nih.gov/27233931
183. Roupret, M., et al. Upper urinary tract transitional cell carcinoma: recurrence rate after percutaneous
endoscopic resection. Eur Urol, 2007. 51: 709.
https://pubmed.ncbi.nlm.nih.gov/16911852

32 UPPER URINARY TRACT UROTHELIAL CARCINOMA - LIMITED UPDATE 2021


184. Jeldres, C., et al. Segmental ureterectomy can safely be performed in patients with transitional cell
carcinoma of the ureter. J Urol, 2010. 183: 1324.
https://pubmed.ncbi.nlm.nih.gov/20171666
185. Colin, P., et al. Comparison of oncological outcomes after segmental ureterectomy or radical
nephroureterectomy in urothelial carcinomas of the upper urinary tract: results from a large French
multicentre study. BJU Int, 2012. 110: 1134.
https://pubmed.ncbi.nlm.nih.gov/22394612
186. Ou, Y.C., et al. Long-term outcomes of total ureterectomy with ileal-ureteral substitution treatment
for ureteral cancer: a single-center experience. BMC Urol, 2018. 18: 73.
https://pubmed.ncbi.nlm.nih.gov/30170590
187. Giannarini, G., et al. Antegrade perfusion with bacillus Calmette-Guerin in patients with non-muscle-
invasive urothelial carcinoma of the upper urinary tract: who may benefit? Eur Urol, 2011. 60: 955.
https://pubmed.ncbi.nlm.nih.gov/21807456
188. Irie, A., et al. Intravesical instillation of bacille Calmette-Guerin for carcinoma in situ of the urothelium
involving the upper urinary tract using vesicoureteral reflux created by a double-pigtail catheter.
Urology, 2002. 59: 53.
https://pubmed.ncbi.nlm.nih.gov/11796281
189. Horiguchi, H., et al. Impact of bacillus Calmette-Guerin therapy of upper urinary tract carcinoma in
situ: comparison of oncological outcomes with radical nephroureterectomy. Med Oncol, 2018. 35: 41.
https://pubmed.ncbi.nlm.nih.gov/29480348
190. Tomisaki, I., et al. Efficacy and Tolerability of Bacillus Calmette-Guerin Therapy as the First-Line
Therapy for Upper Urinary Tract Carcinoma In Situ. Cancer Invest, 2018. 36: 152.
https://pubmed.ncbi.nlm.nih.gov/29393701
191. Yossepowitch, O., et al. Assessment of vesicoureteral reflux in patients with self-retaining ureteral
stents: implications for upper urinary tract instillation. J Urol, 2005. 173: 890.
https://pubmed.ncbi.nlm.nih.gov/15711312
192. Foerster, B., et al. Endocavitary treatment for upper tract urothelial carcinoma: A meta-analysis of
the current literature. Urol Oncol, 2019. 37: 430.
https://pubmed.ncbi.nlm.nih.gov/30846387
193. Kleinmann, N., et al. Primary chemoablation of low-grade upper tract urothelial carcinoma using
UGN-101, a mitomycin-containing reverse thermal gel (OLYMPUS): an open-label, single-arm,
phase 3 trial. Lancet Oncol, 2020. 21: 776.
https://pubmed.ncbi.nlm.nih.gov/32631491
194. Roupret, M., et al. Oncological risk of laparoscopic surgery in urothelial carcinomas. World J Urol,
2009. 27: 81.
https://pubmed.ncbi.nlm.nih.gov/19020880
195. Ong, A.M., et al. Trocar site recurrence after laparoscopic nephroureterectomy. J Urol, 2003.
170: 1301.
https://pubmed.ncbi.nlm.nih.gov/14501747
196. Peyronnet, B., et al. Oncological Outcomes of Laparoscopic Nephroureterectomy Versus Open
Radical Nephroureterectomy for Upper Tract Urothelial Carcinoma: An European Association of
Urology Guidelines Systematic Review. Eur Urol Focus, 2019. 5: 205.
https://pubmed.ncbi.nlm.nih.gov/29154042
197. Simone, G., et al. Laparoscopic versus open nephroureterectomy: perioperative and oncologic
outcomes from a randomised prospective study. Eur Urol, 2009. 56: 520.
https://pubmed.ncbi.nlm.nih.gov/19560259
198. Favaretto, R.L., et al. Comparison between laparoscopic and open radical nephroureterectomy in
a contemporary group of patients: are recurrence and disease-specific survival associated with
surgical technique? Eur Urol, 2010. 58: 645.
https://pubmed.ncbi.nlm.nih.gov/20724065
199. Walton, T.J., et al. Oncological outcomes after laparoscopic and open radical nephroureterectomy:
results from an international cohort. BJU Int, 2011. 108: 406.
https://pubmed.ncbi.nlm.nih.gov/21078048
200. Ni, S., et al. Laparoscopic versus open nephroureterectomy for the treatment of upper urinary tract
urothelial carcinoma: a systematic review and cumulative analysis of comparative studies. Eur Urol,
2012. 61: 1142.
https://pubmed.ncbi.nlm.nih.gov/22349569
201. Ariane, M.M., et al. Assessment of oncologic control obtained after open versus laparoscopic
nephroureterectomy for upper urinary tract urothelial carcinomas (UUT-UCs): results from a large
French multicenter collaborative study. Ann Surg Oncol, 2012. 19: 301.
https://pubmed.ncbi.nlm.nih.gov/21691878

UPPER URINARY TRACT UROTHELIAL CARCINOMA - LIMITED UPDATE 2021 33


202. Adibi, M., et al. Oncological outcomes after radical nephroureterectomy for upper tract urothelial
carcinoma: comparison over the three decades. Int J Urol, 2012. 19: 1060.
https://pubmed.ncbi.nlm.nih.gov/22882743
203. Clements, M.B., et al. Robotic-Assisted Surgery for Upper Tract Urothelial Carcinoma: A
Comparative Survival Analysis. Ann Surg Oncol, 2018. 25: 2550.
https://pubmed.ncbi.nlm.nih.gov/29948423
204. Rodriguez, J.F., et al. Utilization and Outcomes of Nephroureterectomy for Upper Tract Urothelial
Carcinoma by Surgical Approach. J Endourol, 2017. 31: 661.
https://pubmed.ncbi.nlm.nih.gov/28537436
205. Aboumohamed, A.A., et al. Oncologic Outcomes Following Robot-Assisted Laparoscopic
Nephroureterectomy with Bladder Cuff Excision for Upper Tract Urothelial Carcinoma. J Urol, 2015.
194: 1561.
https://pubmed.ncbi.nlm.nih.gov/26192256
206. Xylinas, E., et al. Impact of Distal Ureter Management on Oncologic Outcomes Following Radical
Nephroureterectomy for Upper Tract Urothelial Carcinoma. European Urology, 2014. 65: 210.
https://pubmed.ncbi.nlm.nih.gov/22579047
207. Xylinas, E., et al. Prediction of Intravesical Recurrence After Radical Nephroureterectomy:
Development of a Clinical Decision-making Tool. European Urology, 2014. 65: 650.
https://pubmed.ncbi.nlm.nih.gov/24070577
208. Phe, V., et al. Does the surgical technique for management of the distal ureter influence the outcome
after nephroureterectomy? BJU Int, 2011. 108: 130.
https://pubmed.ncbi.nlm.nih.gov/21070580
209. Kondo, T., et al. Template-based lymphadenectomy in urothelial carcinoma of the upper urinary
tract: impact on patient survival. Int J Urol, 2010. 17: 848.
https://pubmed.ncbi.nlm.nih.gov/20812922
210. Dominguez-Escrig, J.L., et al. Potential Benefit of Lymph Node Dissection During Radical
Nephroureterectomy for Upper Tract Urothelial Carcinoma: A Systematic Review by the European
Association of Urology Guidelines Panel on Non-muscle-invasive Bladder Cancer. Eur Urol Focus,
2019. 5: 224.
https://pubmed.ncbi.nlm.nih.gov/29158169
211. Dong, F., et al. Lymph node dissection could bring survival benefits to patients diagnosed with
clinically node-negative upper urinary tract urothelial cancer: a population-based, propensity score-
matched study. Int J Clin Oncol, 2019. 24: 296.
https://pubmed.ncbi.nlm.nih.gov/30334174
212. Lenis, A.T., et al. Role of surgical approach on lymph node dissection yield and survival in patients
with upper tract urothelial carcinoma. Urol Oncol, 2018. 36: 9 e1.
https://pubmed.ncbi.nlm.nih.gov/29066013
213. Moschini, M., et al. Trends of lymphadenectomy in upper tract urothelial carcinoma (UTUC) patients
treated with radical nephroureterectomy. World J Urol, 2017. 35: 1541.
https://pubmed.ncbi.nlm.nih.gov/28247066
214. Zareba, P., et al. Association between lymph node yield and survival among patients undergoing
radical nephroureterectomy for urothelial carcinoma of the upper tract. Cancer, 2017. 123: 1741.
https://pubmed.ncbi.nlm.nih.gov/28152158
215. Xylinas, E., et al. External validation of the pathological nodal staging score in upper tract urothelial
carcinoma: A population-based study. Urol Oncol, 2017. 35: 33 e21.
https://pubmed.ncbi.nlm.nih.gov/27816402
216. Xylinas, E., et al. Prediction of true nodal status in patients with pathological lymph node negative
upper tract urothelial carcinoma at radical nephroureterectomy. J Urol, 2013. 189: 468.
https://pubmed.ncbi.nlm.nih.gov/23253960
217. Matin, S.F., et al. Patterns of Lymphatic Metastases in Upper Tract Urothelial Carcinoma and
Proposed Dissection Templates. J Urol, 2015. 194: 1567.
https://pubmed.ncbi.nlm.nih.gov/26094807
218. Kondo, T., et al. Template-based lymphadenectomy in urothelial carcinoma of the renal pelvis: a
prospective study. Int J Urol, 2014. 21: 453.
https://pubmed.ncbi.nlm.nih.gov/24754341
219. Matin, S.F., et al. Incidence of downstaging and complete remission after neoadjuvant
chemotherapy for high-risk upper tract transitional cell carcinoma. Cancer, 2010. 116: 3127.
https://pubmed.ncbi.nlm.nih.gov/20564621

34 UPPER URINARY TRACT UROTHELIAL CARCINOMA - LIMITED UPDATE 2021


220. Liao, R.S., et al. Comparison of Pathological Stage in Patients Treated with and without
Neoadjuvant Chemotherapy for High Risk Upper Tract Urothelial Carcinoma. J Urol, 2018. 200: 68.
https://pubmed.ncbi.nlm.nih.gov/29307680
221. Meng, X., et al. High Response Rates to Neoadjuvant Chemotherapy in High-Grade Upper Tract
Urothelial Carcinoma. Urology, 2019. 129: 146.
https://pubmed.ncbi.nlm.nih.gov/30930207
222. Almassi, N., et al. Impact of Neoadjuvant Chemotherapy on Pathologic Response in Patients With
Upper Tract Urothelial Carcinoma Undergoing Extirpative Surgery. Clin Genitourin Cancer, 2018. 16:
e1237.
https://pubmed.ncbi.nlm.nih.gov/30217764
223. Kubota, Y., et al. Oncological outcomes of neoadjuvant chemotherapy in patients with locally
advanced upper tract urothelial carcinoma: a multicenter study. Oncotarget, 2017. 8: 101500.
https://pubmed.ncbi.nlm.nih.gov/29254181
224. Hosogoe, S., et al. Platinum-based Neoadjuvant Chemotherapy Improves Oncological Outcomes in
Patients with Locally Advanced Upper Tract Urothelial Carcinoma. Eur Urol Focus, 2018. 4: 946.
https://pubmed.ncbi.nlm.nih.gov/28753881
225. Porten, S., et al. Neoadjuvant chemotherapy improves survival of patients with upper tract urothelial
carcinoma. Cancer, 2014. 120: 1794.
https://pubmed.ncbi.nlm.nih.gov/24633966
226. Margulis, V., et al. Phase II Trial of Neoadjuvant Systemic Chemotherapy Followed by Extirpative
Surgery in Patients with High Grade Upper Tract Urothelial Carcinoma. J Urol, 2020. 203: 690.
https://pubmed.ncbi.nlm.nih.gov/31702432
227. Birtle, A., et al. Adjuvant chemotherapy in upper tract urothelial carcinoma (the POUT trial): a phase
3, open-label, randomised controlled trial. Lancet, 2020. 395: 1268.
https://pubmed.ncbi.nlm.nih.gov/32145825
228. Kaag, M.G., et al. Changes in renal function following nephroureterectomy may affect the use of
perioperative chemotherapy. Eur Urol, 2010. 58: 581.
https://pubmed.ncbi.nlm.nih.gov/20619530
229. Lane, B.R., et al. Chronic kidney disease after nephroureterectomy for upper tract urothelial
carcinoma and implications for the administration of perioperative chemotherapy. Cancer, 2010.
116: 2967.
https://pubmed.ncbi.nlm.nih.gov/20564402
230. Tully, K.H., et al. Differences in survival and impact of adjuvant chemotherapy in patients with variant
histology of tumors of the renal pelvis. World J Urol, 2020. 38: 2227.
https://pubmed.ncbi.nlm.nih.gov/31748954
231. Urakami, S., et al. Clinical response to induction chemotherapy predicts improved survival outcome
in urothelial carcinoma with clinical lymph nodal metastasis treated by consolidative surgery. Int
J Clin Oncol, 2015. 20: 1171.
https://pubmed.ncbi.nlm.nih.gov/25953680
232. Hahn, A.W., et al. Effect of Adjuvant Radiotherapy on Survival in Patients with Locoregional
Urothelial Malignancies of the Upper Urinary Tract. Anticancer Res, 2016. 36: 4051.
https://pubmed.ncbi.nlm.nih.gov/27466512
233. Huang, Y.C., et al. Adjuvant radiotherapy for locally advanced upper tract urothelial carcinoma. Sci
Rep, 2016. 6: 38175.
https://pubmed.ncbi.nlm.nih.gov/27910890
234. Czito, B., et al. Adjuvant radiotherapy with and without concurrent chemotherapy for locally
advanced transitional cell carcinoma of the renal pelvis and ureter. J Urol, 2004. 172: 1271.
https://pubmed.ncbi.nlm.nih.gov/15371822
235. Iwata, T., et al. The role of adjuvant radiotherapy after surgery for upper and lower urinary tract
urothelial carcinoma: A systematic review. Urol Oncol, 2019. 37: 659.
https://pubmed.ncbi.nlm.nih.gov/31255542
236. O’Brien, T., et al. Prevention of bladder tumours after nephroureterectomy for primary upper
urinary tract urothelial carcinoma: a prospective, multicentre, randomised clinical trial of a single
postoperative intravesical dose of mitomycin C (the ODMIT-C Trial). Eur Urol, 2011. 60: 703.
https://pubmed.ncbi.nlm.nih.gov/21684068
237. Ito, A., et al. Prospective randomized phase II trial of a single early intravesical instillation of
pirarubicin (THP) in the prevention of bladder recurrence after nephroureterectomy for upper urinary
tract urothelial carcinoma: the THP Monotherapy Study Group Trial. J Clin Oncol, 2013. 31: 1422.
https://pubmed.ncbi.nlm.nih.gov/23460707

UPPER URINARY TRACT UROTHELIAL CARCINOMA - LIMITED UPDATE 2021 35


238. Hwang, E.C., et al. Single-dose intravesical chemotherapy after nephroureterectomy for upper tract
urothelial carcinoma. Cochrane Database Syst Rev, 2019. 5: Cd013160.
https://pubmed.ncbi.nlm.nih.gov/31102534
239. Fang, D., et al. Prophylactic intravesical chemotherapy to prevent bladder tumors after
nephroureterectomy for primary upper urinary tract urothelial carcinomas: a systematic review and
meta-analysis. Urol Int, 2013. 91: 291.
https://pubmed.ncbi.nlm.nih.gov/23948770
240. Harraz, A.M., et al. Single Versus Maintenance Intravesical Chemotherapy for the Prevention of
Bladder Recurrence after Radical Nephroureterectomy for Upper Tract Urothelial Carcinoma: A
Randomized Clinical Trial. Clin Genitourin Cancer, 2019. 17: e1108.
https://pubmed.ncbi.nlm.nih.gov/31594736
241. Yamamoto, S., et al. Intravesical irrigation might prevent bladder recurrence in patients undergoing
radical nephroureterectomy for upper urinary tract urothelial carcinoma. Int J Urol, 2019. 26: 791.
https://pubmed.ncbi.nlm.nih.gov/31081198
242. Dong, F., et al. How do organ-specific metastases affect prognosis and surgical treatment for
patients with metastatic upper tract urothelial carcinoma: first evidence from population based data.
Clin Exp Metastasis, 2017. 34: 467.
https://pubmed.ncbi.nlm.nih.gov/29500709
243. Seisen, T., et al. Efficacy of Systemic Chemotherapy Plus Radical Nephroureterectomy for
Metastatic Upper Tract Urothelial Carcinoma. European Urology, 2017. 71: 714.
https://pubmed.ncbi.nlm.nih.gov/27912971
244. Moschini, M., et al. Efficacy of Surgery in the Primary Tumor Site for Metastatic Urothelial Cancer:
Analysis of an International, Multicenter, Multidisciplinary Database. Eur Urol Oncol, 2020. 3: 94.
https://pubmed.ncbi.nlm.nih.gov/31307962
245. Nazzani, S., et al. Survival Effect of Nephroureterectomy in Metastatic Upper Urinary Tract Urothelial
Carcinoma. Clin Genitourin Cancer, 2019. 17: e602.
https://pubmed.ncbi.nlm.nih.gov/31005472
246. Siefker-Radtke, A.O., et al. Is there a role for surgery in the management of metastatic urothelial
cancer? The M. D. Anderson experience. J Urol, 2004. 171: 145.
https://pubmed.ncbi.nlm.nih.gov/14665863
247. Abe, T., et al. Impact of multimodal treatment on survival in patients with metastatic urothelial
cancer. Eur Urol, 2007. 52: 1106.
https://pubmed.ncbi.nlm.nih.gov/17367917
248. Lehmann, J., et al. Surgery for metastatic urothelial carcinoma with curative intent: the German
experience (AUO AB 30/05). Eur Urol, 2009. 55: 1293.
https://pubmed.ncbi.nlm.nih.gov/19058907
249. Faltas, B.M., et al. Metastasectomy in older adults with urothelial carcinoma: Population-based
analysis of use and outcomes. Urol Oncol, 2018. 36: 9 e11.
https://pubmed.ncbi.nlm.nih.gov/28988653
250. Moschini, M., et al. Impact of Primary Tumor Location on Survival from the European Organization for
the Research and Treatment of Cancer Advanced Urothelial Cancer Studies. J Urol, 2018. 199: 1149.
https://pubmed.ncbi.nlm.nih.gov/29158104
251. Powles, T., et al. Avelumab Maintenance Therapy for Advanced or Metastatic Urothelial Carcinoma.
N Engl J Med, 2020. 383: 1218.
https://pubmed.ncbi.nlm.nih.gov/32945632
252. Balar, A.V., et al. Atezolizumab as first-line treatment in cisplatin-ineligible patients with locally
advanced and metastatic urothelial carcinoma: a single-arm, multicentre, phase 2 trial. Lancet,
2017. 389: 67.
https://pubmed.ncbi.nlm.nih.gov/27939400
253. Balar, A.V., et al. First-line pembrolizumab in cisplatin-ineligible patients with locally advanced and
unresectable or metastatic urothelial cancer (KEYNOTE-052): a multicentre, single-arm, phase 2
study. Lancet Oncol, 2017. 18: 1483.
https://pubmed.ncbi.nlm.nih.gov/28967485
254. Alva, A., et al. LBA23 Pembrolizumab (P) combined with chemotherapy (C) vs C alone as first-line
(1L) therapy for advanced urothelial carcinoma (UC): KEYNOTE-361. Ann Oncol 2020. 31: S1142.
https://www.annalsofoncology.org/article/S0923-7534(20)42334-8/abstract
255. Powles, T., et al. Durvalumab alone and durvalumab plus tremelimumab versus chemotherapy in
previously untreated patients with unresectable, locally advanced or metastatic urothelial carcinoma
(DANUBE): a randomised, open-label, multicentre, phase 3 trial. Lancet Oncol, 2020. 21: 1574.
https://pubmed.ncbi.nlm.nih.gov/32971005

36 UPPER URINARY TRACT UROTHELIAL CARCINOMA - LIMITED UPDATE 2021


256. Study of Nivolumab in Combination With Ipilimumab or Standard of Care Chemotherapy Compared
to the Standard of Care Chemotherapy Alone in Treatment of Patients With Untreated Inoperable or
Metastatic Urothelial Cancer (CheckMate901). 2019 [access date March 2021].
https://clinicaltrials.gov/ct2/show/NCT03036098
257. A Phase 1/2 Study Exploring the Safety, Tolerability, and Efficacy of Pembrolizumab (MK-3475) in
Combination With Epacadostat (INCB024360) in Subjects With Selected Cancers (INCB 24360-202
/ MK-3475-037 / KEYNOTE-037/ ECHO-202). 2019 [access date March 2021].
https://clinicaltrials.gov/ct2/show/NCT02178722
258. Study of Durvalumab Given With Chemotherapy, Durvalumab in Combination With Tremelimumab
Given With Chemotherapy, or Chemotherapy in Patients With Unresectable Urothelial Cancer (NILE).
2020. p. NCT03682068 [access date March 2021].
https://www.clinicaltrials.gov/ct2/show/NCT03682068
259. Heers, H., et al. Vinflunine in the Treatment of Upper Tract Urothelial Carcinoma - Subgroup Analysis
of an Observational Study. Anticancer Res, 2017. 37: 6437.
https://pubmed.ncbi.nlm.nih.gov/29061830
260. Bellmunt, J., et al. Pembrolizumab as Second-Line Therapy for Advanced Urothelial Carcinoma. N
Engl J Med, 2017. 376: 1015.
https://pubmed.ncbi.nlm.nih.gov/28212060
261. Rosenberg, J.E., et al. Atezolizumab in patients with locally advanced and metastatic urothelial
carcinoma who have progressed following treatment with platinum-based chemotherapy: a single-
arm, multicentre, phase 2 trial. Lancet, 2016. 387: 1909.
https://pubmed.ncbi.nlm.nih.gov/26952546
262. Powles, T., et al. Atezolizumab versus chemotherapy in patients with platinum-treated locally
advanced or metastatic urothelial carcinoma (IMvigor211): a multicentre, open-label, phase 3
randomised controlled trial. Lancet, 2018. 391: 748.
https://pubmed.ncbi.nlm.nih.gov/29268948
263. Sharma, P., et al. Nivolumab in metastatic urothelial carcinoma after platinum therapy (CheckMate
275): a multicentre, single-arm, phase 2 trial. Lancet Oncol, 2017. 18: 312.
https://pubmed.ncbi.nlm.nih.gov/28131785
264. Patel, M.R., et al. Avelumab in metastatic urothelial carcinoma after platinum failure (JAVELIN Solid
Tumor): pooled results from two expansion cohorts of an open-label, phase 1 trial. Lancet Oncol,
2018. 19: 51.
https://pubmed.ncbi.nlm.nih.gov/29217288
265. Apolo, A.B., et al. Avelumab, an Anti-Programmed Death-Ligand 1 Antibody, In Patients With
Refractory Metastatic Urothelial Carcinoma: Results From a Multicenter, Phase Ib Study. J Clin
Oncol, 2017. 35: 2117.
https://pubmed.ncbi.nlm.nih.gov/28375787
266. Powles, T., et al. Efficacy and Safety of Durvalumab in Locally Advanced or Metastatic Urothelial
Carcinoma: Updated Results From a Phase 1/2 Open-label Study. JAMA Oncol, 2017. 3: e172411.
https://pubmed.ncbi.nlm.nih.gov/28817753
267. Sharma, P., et al. Nivolumab Alone and With Ipilimumab in Previously Treated Metastatic Urothelial
Carcinoma: CheckMate 032 Nivolumab 1 mg/kg Plus Ipilimumab 3 mg/kg Expansion Cohort
Results. J Clin Oncol, 2019. 37: 1608.
https://pubmed.ncbi.nlm.nih.gov/31100038
268. Siefker-Radtke, A., et al. Immunotherapy in metastatic urothelial carcinoma: focus on immune
checkpoint inhibition. Nat Rev Urol, 2018. 15: 112.
https://pubmed.ncbi.nlm.nih.gov/29205200
269. Loriot, Y., et al. Erdafitinib in Locally Advanced or Metastatic Urothelial Carcinoma. New England
Journal of Medicine, 2019. 381: 338.
https://pubmed.ncbi.nlm.nih.gov/31340094
270. Ploussard, G., et al. Conditional survival after radical nephroureterectomy for upper tract carcinoma.
Eur Urol, 2015. 67: 803.
https://pubmed.ncbi.nlm.nih.gov/25145551
271. Shigeta, K., et al. The Conditional Survival with Time of Intravesical Recurrence of Upper Tract
Urothelial Carcinoma. J Urol, 2017. 198: 1278.
https://pubmed.ncbi.nlm.nih.gov/28634017
272. Mandalapu, R.S., et al. Update of the ICUD-SIU consultation on upper tract urothelial carcinoma
2016: treatment of low-risk upper tract urothelial carcinoma. World J Urol, 2017. 35: 355.
https://pubmed.ncbi.nlm.nih.gov/27233780

UPPER URINARY TRACT UROTHELIAL CARCINOMA - LIMITED UPDATE 2021 37


273. Bagley, D.H., et al. Ureteroscopic laser treatment of upper urinary tract neoplasms. World J Urol,
2010. 28: 143.
https://pubmed.ncbi.nlm.nih.gov/20229233
274. Mohapatra, A., et al. Importance of long-term follow-up after endoscopic management for upper
tract urothelial carcinoma and factors leading to surgical management. Int Urol Nephrol, 2020.
52: 1465.
https://pubmed.ncbi.nlm.nih.gov/ https://pubmed.ncbi.nlm.nih.gov/32157621

10. CONFLICT OF INTEREST


All members of the Non-Muscle-Invasive Bladder Cancer Guidelines working panel have provided disclosure
statements on all relationships that they have that might be perceived to be a potential source of a conflict of
interest. This information is publically accessible through the European Association of Urology website:
http://uroweb.org/guideline/upper-urinary-tract-urothelial-cell-carcinoma/.
This guidelines document was developed with the financial support of the European Association of
Urology. No external sources of funding and support have been involved. The EAU is a non-profit organisation,
and funding is limited to administrative assistance and travel and meeting expenses. No honoraria or other
reimbursements have been provided.

11. CITATION INFORMATION


The format in which to cite the EAU Guidelines will vary depending on the style guide of the journal in which the
citation appears. Accordingly, the number of authors or whether, for instance, to include the publisher, location,
or an ISBN number may vary.

The compilation of the complete Guidelines should be referenced as:


EAU Guidelines. Edn. presented at the EAU Annual Congress Milan, 2021. ISBN 978-94-92671-13-4.

If a publisher and/or location is required, include:


EAU Guidelines Office, Arnhem, The Netherlands. http://uroweb.org/guidelines/compilations-of-all-guidelines/

References to individual guidelines should be structured in the following way:


Contributors’ names. Title of resource. Publication type. ISBN. Publisher and publisher location, year.

38 UPPER URINARY TRACT UROTHELIAL CARCINOMA - LIMITED UPDATE 2021


EAU - EANM - ESTRO -
ESUR - ISUP - SIOG
Guidelines on
Prostate Cancer
N. Mottet (Chair), P. Cornford (Vice-chair), R.C.N. van den Bergh,
E. Briers, Expert Patient Advocate (European Prostate Cancer
Coalition/Europa UOMO), M. De Santis, S. Gillessen,
J. Grummet, A.M. Henry, T.H. van der Kwast, T.B. Lam,
M.D. Mason, S. O’Hanlon, D.E. Oprea-Lager, G. Ploussard,
H.G. van der Poel, O. Rouvière, I.G. Schoots. D. Tilki, T. Wiegel
Guidelines Associates: T. Van den Broeck, M. Cumberbatch,
A. Farolfi, N. Fossati, G. Gandaglia, N. Grivas, M. Lardas,
M. Liew, L. Moris, P-P.M. Willemse

© European Association of Urology 2021


TABLE OF CONTENTS PAGE
1. INTRODUCTION 10
1.1 Aims and scope 10
1.2 Panel composition 10
1.2.1 Acknowledgement 10
1.3 Available publications 10
1.4 Publication history and summary of changes 10
1.4.1 Publication history 10
1.4.2 Summary of changes 10

2. METHODS 15
2.1 Data identification 15
2.2 Review 16
2.3 Future goals 16

3. EPIDEMIOLOGY AND AETIOLOGY 16


3.1 Epidemiology 16
3.2 Aetiology 16
3.2.1 Family history / hereditary prostate cancer 16
3.2.1.1 Germline mutations and prostate cancer 17
3.2.2 Risk factors 17
3.2.2.1 Metabolic syndrome 17
3.2.2.1.1 Diabetes/metformin 17
3.2.2.1.2 Cholesterol/statins 18
3.2.2.1.3 Obesity 18
3.2.2.2 Dietary factors 18
3.2.2.3 Hormonally active medication 18
3.2.2.3.1 5-alpha-reductase inhibitors (5-ARIs) 18
3.2.2.3.2 Testosterone 18
3.2.2.4 Other potential risk factors 19
3.2.3 Summary of evidence for epidemiology and aetiology 19

4. CLASSIFICATION AND STAGING SYSTEMS 19


4.1 Classification 19
4.2 Gleason score and International Society of Urological Pathology 2014 grade 20
4.3 Prognostic relevance of stratification 21
4.4 Guidelines for classification and staging systems 21

5. DIAGNOSTIC EVALUATION 21
5.1 Screening and early detection 21
5.1.1 Screening 21
5.1.2 Early detection 23
5.1.3 Genetic testing for inherited prostate cancer 24
5.1.4 Guidelines for germline testing* 25
5.1.5 Guidelines for screening and early detection 25
5.2 Clinical diagnosis 25
5.2.1 Digital rectal examination 25
5.2.2 Prostate-specific antigen 25
5.2.2.1 PSA density 26
5.2.2.2 PSA velocity and doubling time 26
5.2.2.3 Free/total PSA ratio 26
5.2.3 Biomarkers in prostate cancer 26
5.2.3.1 Blood based biomarkers: PHI/4K score/IsoPSA 26
5.2.3.2 Urine biomarkers: PCA3/SelectMDX/Mi Prostate score (MiPS)/ExoDX 27
5.2.3.3 Tests to select men for a repeat biopsy 27
5.2.3.4 Guidelines for risk-assessment of asymptomatic men 28
5.2.4 The role of imaging in clinical diagnosis 28
5.2.4.1 Transrectal ultrasound and ultrasound-based techniques 28
5.2.4.2 Multiparametric magnetic resonance imaging 28

2 PROSTATE CANCER - LIMITED UPDATE 2021


5.2.4.2.1 Multiparametric magnetic resonance imaging
performance in detecting PCa 28
5.2.4.2.2 Targeted biopsy improves the detection of ISUP
grade > 2 cancer as compared to systematic biopsy. 28
5.2.4.2.3 MRI-TBx without systematic biopsy reduces the detection
of ISUP grade 1 PCa as compared to systematic biopsy. 29
5.2.4.2.4 The added value of systematic and targeted biopsy 29
5.2.4.2.5 Number of biopsy procedures potentially avoided in
the ‘MR pathway’ 30
5.2.4.2.6 Other considerations 30
5.2.4.2.6.1 Multiparametric magnetic resonance
imaging reproducibility 30
5.2.4.2.6.2 Targeted biopsy accuracy and reproducibility 31
5.2.4.2.6.3 Role of risk-stratification 31
5.2.4.2.6.4 Pre-biopsy MRI and MRI-TBx may induce
an ISUP prostate cancer grade shift, as a
result of improved diagnosis 33
5.2.4.2.7 Summary of evidence and practical considerations on
pre-biopsy MRI and MRI-TBx 33
5.2.4.3 Guidelines for imaging in PCa detection 33
5.2.5 Baseline biopsy 34
5.2.6 Repeat biopsy 34
5.2.6.1 Repeat biopsy after previously negative biopsy 34
5.2.6.2 Saturation biopsy 34
5.2.7 Prostate biopsy procedure 34
5.2.7.1 Sampling sites and number of cores 34
5.2.7.2 Antibiotics prior to biopsy 35
5.2.7.2.1 Transperineal prostate biopsy 35
5.2.7.2.2 Transrectal prostate biopsy 35
5.2.7.3 Summary of evidence and recommendations for performing prostate
biopsy (in line with the Urological Infections Guidelines Panel) 36
5.2.7.4 Local anaesthesia prior to biopsy 37
5.2.7.5 Complications 38
5.2.7.6 Seminal vesicle biopsy 38
5.2.7.7 Transition zone biopsy 38
5.2.8 Pathology of prostate needle biopsies 38
5.2.8.1 Processing 38
5.2.8.2 Microscopy and reporting 38
5.2.8.2.1 Recommended terminology for reporting prostate
biopsies 39
5.2.8.3 Tissue-based prognostic biomarker testing 39
5.2.8.4 Histopathology of radical prostatectomy specimens 40
5.2.8.4.1 Processing of radical prostatectomy specimens 40
5.2.8.4.1.1 Guidelines for processing prostatectomy
specimens 40
5.2.8.4.2 Radical prostatectomy specimen report 40
5.2.8.4.3 ISUP grade in prostatectomy specimens 41
5.2.8.4.4 Definition of extraprostatic extension 41
5.2.8.4.5 PCa volume 42
5.2.8.4.6 Surgical margin status 42
5.3 Diagnosis - Clinical Staging 42
5.3.1 T-staging 42
5.3.1.1 TRUS 42
5.3.1.2 MRI 42
5.3.2 N-staging 43
5.3.2.1 Computed tomography and magnetic resonance imaging 43
5.3.2.2 Risk calculators incorporating MRI findings 43
5.3.2.3 Choline PET/CT 43
5.3.2.4 Prostate-specific membrane antigen-based PET/CT 43
5.3.3 M-staging 44

PROSTATE CANCER - LIMITED UPDATE 2021 3


5.3.3.1 Bone scan 44
5.3.3.2 Fluoride PET and PET/CT, choline PET/CT and MRI 44
5.3.3.3 Prostate-specific membrane antigen-based PET/CT 45
5.3.4 Summary of evidence and practical considerations on initial N/M staging 45
5.3.5 Summary of evidence and guidelines for staging of prostate cancer 46
5.4 Estimating life expectancy and health status 46
5.4.1 Introduction 46
5.4.2 Life expectancy 46
5.4.3 Health status screening 47
5.4.3.1 Co-morbidity 47
5.4.3.2 Nutritional status 47
5.4.3.3 Cognitive function 48
5.4.3.4 Physical function 48
5.4.3.5 Shared decision-making 48
5.4.4 Conclusion 48
5.4.5 Guidelines for evaluating health status and life expectancy 50

6. TREATMENT 50
6.1 Treatment modalities 50
6.1.1 Deferred treatment (active surveillance/watchful waiting) 50
6.1.1.1 Definitions 51
6.1.1.2 Active surveillance 51
6.1.1.3 Watchful Waiting 52
6.1.1.3.1 Outcome of watchful waiting compared with active
treatment 52
6.1.1.4 The ProtecT study 52
6.1.2 Radical prostatectomy 53
6.1.2.1 Introduction 53
6.1.2.2 Pre-operative preparation 53
6.1.2.2.1 Pre-operative patient education 53
6.1.2.3 Surgical techniques 54
6.1.2.3.1 Robotic anterior versus Retzius-sparing dissection 54
6.1.2.3.2 Pelvic lymph node dissection 55
6.1.2.3.3 Sentinel node biopsy analysis 55
6.1.2.3.4 Prostatic anterior fat pad dissection and histologic
analysis 55
6.1.2.3.5 Management of the dorsal venous complex 56
6.1.2.3.6 Nerve-sparing surgery 56
6.1.2.3.7 Lymph-node-positive patients during radical
prostatectomy 56
6.1.2.3.8 Removal of seminal vesicles 56
6.1.2.3.9 Techniques of vesico-urethral anastomosis 56
6.1.2.3.10 Bladder neck management 57
6.1.2.3.11 Urethral length preservation 57
6.1.2.3.12 Cystography prior to catheter removal 57
6.1.2.3.13 Urinary catheter 58
6.1.2.3.14 Use of a pelvic drain 58
6.1.2.4 Acute and chronic complications of surgery 58
6.1.2.4.1 Effect of anterior and posterior reconstruction on
continence 58
6.1.2.4.2 Deep venous thrombosis prophylaxis 59
6.1.2.4.3 Early complications of extended lymph node dissection 59
6.1.3 Radiotherapy 59
6.1.3.1 External beam radiation therapy 59
6.1.3.1.1 Technical aspects: intensity-modulated external-beam
radiotherapy and volumetric arc external-beam
radiotherapy 59
6.1.3.1.2 Dose escalation 60
6.1.3.1.3 Hypofractionation 61
6.1.3.1.4 Neoadjuvant or adjuvant hormone therapy plus
radiotherapy 63

4 PROSTATE CANCER - LIMITED UPDATE 2021


6.1.3.1.5 Combined dose-escalated radiotherapy and androgen-
deprivation therapy 64
6.1.3.2 Proton beam therapy 65
6.1.3.3 Spacer during external beam radiation therapy 65
6.1.3.4 Brachytherapy 65
6.1.3.4.1 Low-dose rate (LDR) brachytherapy 65
6.1.3.4.2 High-dose rate brachytherapy 66
6.1.3.5 Acute side effects of external beam radiotherapy and brachytherapy 66
6.1.4 Hormonal therapy 67
6.1.4.1 Introduction 67
6.1.4.1.1 Different types of hormonal therapy 67
6.1.4.1.1.1 Testosterone-lowering therapy (castration) 67
6.1.4.1.1.1.1 Castration level 67
6.1.4.1.1.1.2 Bilateral orchiectomy 67
6.1.4.1.1.1.3 Oestrogens 67
6.1.4.1.1.1.4 Luteinising-hormone-releasing
hormone agonists 67
6.1.4.1.1.1.5 Luteinising-hormone-releasing
hormone antagonists 67
6.1.4.1.1.1.6 Anti-androgens 68
6.1.4.1.1.1.6.1 Steroidal anti-androgens 68
6.1.4.1.1.1.6.2 Non-steroidal anti-androgens 68
6.1.4.1.1.2 New androgen receptor pathway targetting
agents (ARTA) 68
6.1.4.1.1.2.1 Abiraterone acetate 68
6.1.4.1.1.2.2 Apalutamide, darolutamide,
enzalutamide (alphabetical order) 68
6.1.4.1.1.3 New compounds 69
6.1.4.1.1.3.1 PARP inhibitors 69
6.1.4.1.1.3.2 Immune checkpoint inhibitors 69
6.1.4.1.1.3.3 Protein kinase B (AKT) inhibitors 69
6.1.5 Investigational therapies 69
6.1.5.1 Background 69
6.1.5.2 Cryotherapy 69
6.1.5.3 High-intensity focused ultrasound 70
6.1.5.4 Focal therapy 70
6.1.6 General guidelines for the treatment of prostate cancer 71
6.2 Treatment by disease stages 72
6.2.1 Treatment of low-risk disease 72
6.2.1.1 Active surveillance 72
6.2.1.1.1 Active surveillance - inclusion criteria 72
6.2.1.1.2 Tissue-based prognostic biomarker testing 72
6.2.1.1.3 Imaging for treatment selection 72
6.2.1.1.4 Monitoring during active surveillance 73
6.2.1.1.5 Active Surveillance - when to change strategy 74
6.2.1.2 Alternatives to active surveillance for the treatment of low-risk disease 74
6.2.1.3 Summary of evidence and guidelines for the treatment of low-risk
disease 74
6.2.2 Treatment of intermediate-risk disease 75
6.2.2.1 Active Surveillance 75
6.2.2.2 Surgery 75
6.2.2.3 Radiation therapy 75
6.2.2.3.1 Recommended IMRT for intermediate-risk PCa 75
6.2.2.3.2 Brachytherapy monotherapy 75
6.2.2.4 Other options for the primary treatment of intermediate-risk PCa
(experimental therapies) 75
6.2.2.5 Guidelines for the treatment of intermediate-risk disease 76
6.2.3 Treatment of high-risk localised disease 76
6.2.3.1 Radical prostatectomy 76
6.2.3.1.1 ISUP grade 4–5 76
6.2.3.1.2 Prostate-specific antigen > 20 ng/mL 77

PROSTATE CANCER - LIMITED UPDATE 2021 5


6.2.3.1.3 Radical prostatectomy in cN0 patients who are found to
have pathologically confirmed lymph node invasion (pN1) 77
6.2.3.2 External beam radiation therapy 77
6.2.3.2.1 Recommended external beam radiation therapy
treatment policy for high-risk localised PCa 77
6.2.3.2.2 Lymph node irradiation in cN0 77
6.2.3.2.3 Brachytherapy boost 77
6.2.3.3 Options other than surgery and radiotherapy for the primary
treatment of localised PCa 77
6.2.3.4 Guidelines for radical treatment of high-risk localised disease 78
6.2.4 Treatment of locally advanced PCa 78
6.2.4.1 Surgery 78
6.2.4.2 Radiotherapy for locally advanced PCa 78
6.2.4.3 Treatment of cN1 M0 PCa 78
6.2.4.3.1 Guidelines for the management of cN1 M0 prostate
cancer 80
6.2.4.4 Options other than surgery and radiotherapy for primary treatment 80
6.2.4.4.1 Investigational therapies 80
6.2.4.4.2 Androgen deprivation therapy monotherapy 80
6.2.4.5 Guidelines for radical treatment of locally-advanced disease 80
6.2.5 Adjuvant treatment after radical prostatectomy 80
6.2.5.1 Introduction 80
6.2.5.2 Risk factors for relapse 81
6.2.5.2.1 Biomarker-based risk stratification after radical
prostatectomy 81
6.2.5.3 Immediate (adjuvant) post-operative external irradiation after RP
(cN0 or pN0) 81
6.2.5.4 Comparison of adjuvant- and salvage radiotherapy 82
6.2.5.5 Adjuvant androgen ablation in men with N0 disease 83
6.2.5.6 Adjuvant treatment in pN1 disease 83
6.2.5.6.1 Adjuvant androgen ablation alone 83
6.2.5.6.2 Adjuvant radiotherapy combined with ADT in pN1 disease 83
6.2.5.6.3 Observation of pN1 patients after radical prostatectomy
and extended lymph node dissection 84
6.2.5.7 Guidelines for adjuvant treatment in pN0 and pN1 disease after
radical prostatectomy 84
6.2.5.8 Guidelines for non-curative or palliative treatments in prostate cancer 84
6.2.6 Persistent PSA after radical prostatectomy 85
6.2.6.1 Natural history of persistently elevated PSA after RP 85
6.2.6.2 Imaging in patients with persistently elevated PSA after RP 86
6.2.6.3 Impact of post-operative RT and/or ADT in patients with persistent PSA 87
6.2.6.4 Conclusion 87
6.2.6.5 Recommendations for the management of persistent PSA after
radical prostatectomy 88
6.3 Management of PSA-only recurrence after treatment with curative intent 88
6.3.1 Background 88
6.3.2 Definitions of clinically relevant PSA relapse 88
6.3.3 Natural history of biochemical recurrence 88
6.3.4 The role of imaging in PSA-only recurrence 89
6.3.4.1 Assessment of metastases 89
6.3.4.1.1 Bone scan and abdominopelvic CT 89
6.3.4.1.2 Choline PET/CT 89
6.3.4.1.3 Fluoride PET and PET/CT 89
6.3.4.1.4 Fluciclovine PET/CT 89
6.3.4.1.5 Prostate-specific membrane antigen based PET/CT 90
6.3.4.1.6 Whole-body and axial MRI 90
6.3.4.2 Assessment of local recurrences 90
6.3.4.2.1 Local recurrence after radical prostatectomy 90
6.3.4.2.2 Local recurrence after radiation therapy 91
6.3.4.3 Summary of evidence on imaging in case of biochemical recurrence 91

6 PROSTATE CANCER - LIMITED UPDATE 2021


6.3.4.4 Guidelines for imaging in patients with biochemical recurrence 91
6.3.5 Treatment of PSA-only recurrences 91
6.3.5.1 Treatment of PSA-only recurrences after radical prostatectomy 91
6.3.5.1.1 Salvage radiotherapy for PSA-only recurrence after
radical prostatectomy (cTxcN0M0, without PET/ CT) 91
6.3.5.1.2 Salvage radiotherapy combined with androgen
deprivation therapy (cTxcN0, without PET/CT) 93
6.3.5.1.2.1 Target volume, dose, toxicity 94
6.3.5.1.2.2 Salvage RT with or without ADT
(cTx CN0/1) (with PET/CT) 95
6.3.5.1.2.3 Metastasis-directed therapy for
rcN+ (with PET/CT) 95
6.3.5.1.3 Salvage lymph node dissection 95
6.3.5.1.4 Comparison of adjuvant- and salvage radiotherapy 96
6.3.5.2 Management of PSA failures after radiation therapy 96
6.3.5.2.1 Salvage radical prostatectomy 96
6.3.5.2.1.1 Oncological outcomes 96
6.3.5.2.1.2 Morbidity 97
6.3.5.2.1.3 Summary of salvage radical prostatectomy 97
6.3.5.2.2 Salvage cryoablation of the prostate 97
6.3.5.2.2.1 Oncological outcomes 97
6.3.5.2.2.2 Morbidity 98
6.3.5.2.2.3 Summary of salvage cryoablation of the
prostate 98
6.3.5.2.3 Salvage re-irradiation 98
6.3.5.2.3.1 Salvage brachytherapy for radiotherapy
failure 98
6.3.5.2.3.2 Salvage stereotactic ablative body
radiotherapy for radiotherapy failure 99
6.3.5.2.3.2.1 Oncological outcomes and morbidity 99
6.3.5.2.3.2.2 Morbidity 99
6.3.5.2.3.2.3 Summary of salvage stereotactic
ablative body radiotherapy 99
6.3.5.2.4 Salvage high-intensity focused ultrasound 99
6.3.5.2.4.1 Oncological outcomes 99
6.3.5.2.4.2 Morbidity 100
6.3.5.2.4.3 Summary of salvage high-intensity focused
ultrasound 100
6.3.6 Hormonal therapy for relapsing patients 100
6.3.7 Observation 101
6.3.8 Guidelines for second-line therapy after treatment with curative intent 101
6.4 Treatment: Metastatic prostate cancer 101
6.4.1 Introduction 101
6.4.2 Prognostic factors 101
6.4.3 First-line hormonal treatment 102
6.4.3.1 Non-steroidal anti-androgen monotherapy 102
6.4.3.2 Intermittent versus continuous androgen deprivation therapy 102
6.4.3.3 Immediate versus deferred androgen deprivation therapy 103
6.4.4 Combination therapies 103
6.4.4.1 ‘Complete’ androgen blockade 103
6.4.4.2 Androgen deprivation combined with other agents 103
6.4.4.2.1 Androgen deprivation therapy combined with
chemotherapy 103
6.4.4.2.2 Combination with the new hormonal treatments
(abiraterone, apalutamide, enzalutamide) 104
6.4.5 Treatment selection and patient selection 106
6.4.6 Deferred treatment for metastatic PCa (stage M1) 106
6.4.7 Treatment of the primary tumour in newly diagnosed metastatic disease 106
6.4.8 Metastasis-directed therapy in M1-patients 107
6.4.9 Guidelines for the first-line treatment of metastatic disease 107

PROSTATE CANCER - LIMITED UPDATE 2021 7


6.5 Treatment: Castration-resistant PCa (CRPC) 108
6.5.1 Definition of CRPC 108
6.5.2 Management of mCRPC - general aspects 108
6.5.2.1 Molecular diagnostics 108
6.5.3 Treatment decisions and sequence of available options 108
6.5.4 Non-metastatic CRPC 108
6.5.5 Metastatic CRPC 109
6.5.5.1 Conventional androgen deprivation in CRPC 109
6.5.6 First-line treatment of metastatic CRPC 110
6.5.6.1 Abiraterone 110
6.5.6.2 Enzalutamide 110
6.5.6.3 Docetaxel 111
6.5.6.4 Sipuleucel-T 111
6.5.6.5 Ipatasertib 111
6.5.7 Second-line treatment for mCRPC and sequence 112
6.5.7.1 Cabazitaxel 112
6.5.7.2 Abiraterone acetate after prior docetaxel 113
6.5.7.3 Enzalutamide after docetaxel 113
6.5.7.4 Radium-223 113
6.5.8 Treatment after docetaxel and one line of hormonal treatment for mCRPC 113
6.5.8.1 PARP inhibitors for mCRPC 114
6.5.8.2 Sequencing treatment 114
6.5.8.2.1 ARTA -> ARTA (chemotherapy-naïve patients) 114
6.5.8.2.2 ARTA -> PARP inhibitor/olaparib 115
6.5.8.2.3 Docetaxel for mHSPC -> docetaxel rechallenge 115
6.5.8.2.4 ARTA -> docetaxel or docetaxel -> ARTA followed by
PARP inhibitor 115
6.5.8.2.5 ARTA before or after docetaxel 115
6.5.8.2.6 ARTA –> docetaxel -> cabazitaxel or docetaxel –>
ARTA -> cabazitaxel 115
6.5.9 Prostate-specific membrane antigen (PSMA) therapy 115
6.5.9.1 Background 115
6.5.9.2 PSMA-based therapy 115
6.5.10 Immunotherapy for mCRPC 115
6.5.11 Monitoring of treatment 115
6.5.12 When to change treatment 116
6.5.13 Symptomatic management in metastatic CRPC 116
6.5.13.1 Common complications due to bone metastases 116
6.5.13.2 Preventing skeletal-related events 116
6.5.13.2.1 Bisphosphonates 116
6.5.13.2.2 RANK ligand inhibitors 117
6.5.14 Summary of evidence and guidelines for life-prolonging treatments of
castrate-resistant disease 117
6.5.15 Guidelines for systematic treatments of castrate-resistant disease 118
6.5.16 Guidelines for supportive care of castrate-resistant disease 118
6.5.17 Guideline for non-metastatic castrate-resistant disease 118
6.6 Summary of guidelines for the treatment of prostate cancer 118
6.6.1 General guidelines recommendations for treatment of prostate cancer 119
6.6.2 Guidelines recommendations for the various disease stages 119
6.6.3 Guidelines for metastatic disease, second-line and palliative treatments 122

7. FOLLOW-UP 124
7.1 Follow-up: After local treatment 124
7.1.1 Definition 124
7.1.2 Why follow-up? 124
7.1.3 How to follow-up? 124
7.1.3.1 Prostate-specific antigen monitoring 124
7.1.3.1.1 Active surveillance follow-up 124
7.1.3.1.2 Prostate-specific antigen monitoring after radical
prostatectomy 124

8 PROSTATE CANCER - LIMITED UPDATE 2021


7.1.3.1.3 Prostate-specific antigen monitoring after radiotherapy 124
7.1.3.1.4 Digital rectal examination 125
7.1.3.1.5 Transrectal ultrasound, bone scintigraphy, CT, MRI and
PET/CT 125
7.1.4 How long to follow-up? 125
7.1.5 Summary of evidence and guidelines for follow-up after treatment with curative
intent 125
7.2 Follow-up: During first line hormonal treatment (androgen sensitive period) 125
7.2.1 Introduction 125
7.2.2 Purpose of follow-up 125
7.2.3 General follow-up of men on ADT 126
7.2.3.1 Testosterone monitoring 126
7.2.3.2 Liver function monitoring 126
7.2.3.3 Serum creatinine and haemoglobine 126
7.2.3.4 Monitoring of metabolic complications 126
7.2.3.5 Monitoring bone problems 126
7.2.3.6 Monitoring lifestyle and cognition 127
7.2.4 Methods of follow-up in men on ADT without metastases 127
7.2.4.1 Prostate-specific antigen monitoring 127
7.2.4.2 Imaging 127
7.2.5 Methods of follow-up in men under ADT for metastatic hormone-sensitive PCa 127
7.2.5.1 PSA monitoring 127
7.2.5.2 Imaging as a marker of response in metastatic PCa 127
7.2.6 Guidelines for follow-up during hormonal treatment 128

8. QUALITY OF LIFE OUTCOMES IN PROSTATE CANCER 128


8.1 Introduction 128
8.2 Adverse effects of PCa therapies 128
8.2.1 Surgery 128
8.2.2 Radiotherapy 129
8.2.2.1 Side-effects of external beam radiotherapy 129
8.2.2.2 Side effects from brachytherapy 129
8.2.3 Local primary whole-gland treatments other than surgery or radiotherapy 129
8.2.3.1 Cryosurgery 129
8.2.3.2 High-intensity focused ultrasound 129
8.2.4 Hormonal therapy 129
8.2.4.1 Sexual function 130
8.2.4.2 Hot flushes 130
8.2.4.3 Non-metastatic bone fractures 130
8.2.4.4 Metabolic effects 130
8.2.4.5 Cardiovascular morbidity 131
8.2.4.6 Fatigue 131
8.2.4.7 Neurological side effects 131
8.3 Overall quality of life in men with PCa 131
8.3.1 Long-term (> 12 months) quality of life outcomes in men with localised disease 132
8.3.1.1 Men undergoing local treatments 132
8.3.1.2 Guidelines for quality of life in men undergoing local treatments 133
8.3.2 Improving quality of life in men who have been diagnosed with PCa 133
8.3.2.1 Men undergoing local treatments 133
8.3.2.2 Men undergoing systemic treatments 133
8.3.2.3 Decision regret 134
8.3.2.4 Decision aids in prostate cancer 134
8.3.2.5 Guidelines for quality of life in men undergoing systemic treatments 135

9. REFERENCES 135

10. CONFLICT OF INTEREST 211

11. CITATION INFORMATION 211

PROSTATE CANCER - LIMITED UPDATE 2021 9


1. INTRODUCTION
1.1 Aims and scope
The Prostate Cancer (PCa) Guidelines Panel have prepared this guidelines document to assist medical
professionals in the evidence-based management of PCa.
It must be emphasised that clinical guidelines present the best evidence available to the experts
but following guideline recommendations will not necessarily result in the best outcome. Guidelines can never
replace clinical expertise when making treatment decisions for individual patients, but rather help to focus
decisions - also taking personal values and preferences/individual circumstances of patients into account.
Guidelines are not mandates and do not purport to be a legal standard of care.

1.2 Panel composition


The PCa Guidelines Panel consists of an international multidisciplinary group of urologists, radiation
oncologists, medical oncologists, radiologists, a pathologist, a geriatrician and a patient representative.

All imaging sections in the text have been developed jointly with the European Society of Urogenital Radiology
(ESUR) and the European Association of Nuclear Medicine (EANM). Representatives of the ESUR and the
EANM in the PCa Guidelines Panel are (in alphabetical order): Dr. D. Oprea-Lager, Prof.Dr. O Rouvière and
Dr. I.G. Schoots.
All radiotherapy sections have been developed jointly with the European Society for Radiotherapy
& Oncology (ESTRO). Representatives of ESTRO in the PCa Guidelines Panel are (in alphabetical order):
Prof.Dr. A.M. Henry, Prof.Dr. M.D. Mason and Prof.Dr. T. Wiegel.
The International Society of Urological Pathology is represented by Prof.Dr. T. van der Kwast.
Dr. S. O’Hanlon, consultant geriatrician, representing the International Society of Geriatric Oncology
(SOIG) contributed to the sections addressing life expectancy, health status and quality of life in particular.
Dr. E. Briers, expert Patient Advocate Hasselt-Belgium representing the patient voice as delegated
by the European Prostate Cancer Coalition/Europa UOMO.
All experts involved in the production of this document have submitted potential conflict of interest
statements which can be viewed on the EAU website Uroweb: https://uroweb.org/guideline/prostate-cancer/.

1.2.1 Acknowledgement
The PCa Guidelines Panel gratefully acknowledges the assistance and general guidance provided by
Prof.Dr. M. Bolla, honorary member of the PCa Guidelines Panel.

1.3 Available publications


A quick reference document (Pocket guidelines) is available, both in print and as an app for iOS and Android
devices. These are abridged versions which may require consultation together with the full text version. Several
scientific publications are available [1, 2] as are a number of translations of all versions of the PCa Guidelines.
All documents can be accessed on the EAU website: http://uroweb.org/guideline/prostate-cancer/.

1.4 Publication history and summary of changes


1.4.1 Publication history
The EAU PCa Guidelines were first published in 2001. This 2021 document presents an update of the 2020
PCa Guidelines publication.

1.4.2 Summary of changes


The literature for the complete document has been assessed and updated based upon a review of all
recommendations and creation of appropriate GRADE forms. Evidence summaries and recommendations have
been amended throughout the current document and several new sections have been added.

All chapters of the 2021 PCa Guidelines have been updated. New data have been included in the following
sections, resulting in new sections and added and revised recommendations in:

10 PROSTATE CANCER - LIMITED UPDATE 2021


5.1.4 Guidelines for germline testing*

Recommendations Strength rating


Consider germline testing in men with metastatic PCa. weak
Consider germline testing in men with high-risk PCa who have a family member weak
diagnosed with PCa at age < 60 years.
Consider germline testing in men with multiple family members diagnosed with PCa weak
at age < 60 years or a family member who died from PCa.
Consider germline testing in men with a family history of high-risk germline weak
mutations or a family history of multiple cancers on the same side of the family.
*Genetic counseling is required prior to germline testing.

5.2.7.3 Summary of evidence and recommendations for performing prostate biopsy


(in line with the Urological Infections Guidelines Panel)

Summary of evidence LE
A meta-analysis of seven studies including 1,330 patients showed significantly reduced 1a
infectious complications in patients undergoing transperineal biopsy as compared to
transrectal biopsy.
Meta-analysis of eight RCTs including 1,786 men showed that use of a rectal povidone-iodine 1a
preparation before transrectal biopsy, in addition to antimicrobial prophylaxis, resulted in a
significantly lower rate of infectious complications.
A meta-analysis on eleven studies with 1,753 patients showed significantly reduced infections 1a
after transrectal biopsy when using antimicrobial prophylaxis as compared to placebo/control.

Recommendations Strength rating*


Perform prostate biopsy using the transperineal approach due to the lower risk of Strong
infectious complications.
Use routine surgical disinfection of the perineal skin for transperineal biopsy. Strong
Use rectal cleansing with povidone-iodine in men prior to transrectal prostate biopsy. Strong
Do not use fluoroquinolones for prostate biopsy in line with the European Strong
Commission final decision on EMEA/H/A-31/1452.
Use either target prophylaxis based on rectal swab or stool culture; augmented Weak
prophylaxis (two or more different classes of antibiotics); or alternative antibiotics
(e.g., fosfomycin trometamol, cephalosporin, aminoglycoside) for antibiotic
prophylaxis for transrectal biopsy.
Use a single oral dose of either cefuroxime or cephalexin or cephazolin as antibiotic Weak
prophylaxis for transperineal biopsy. Patients with severe penicillin allergy may be
given sulphamethoxazole.
Ensure that prostate core biopsies from different sites are submitted separately for Strong
processing and pathology reporting.
*Note on strength ratings:
The above strength ratings are explained here due to the major clinical implications of these new recommendations.
Although data showing the lower risk of infection via the transperineal approach is low in certainty, its statistical and
clinical significance warrants its Strong rating. Strong ratings are also given for routine surgical disinfection of skin
in transperineal biopsy and povidone-iodine rectal cleansing in transrectal biopsy as, although quality of data is low,
the clinical benefit is high and practical application simple. A Strong rating is given for avoiding fluoroquinolones in
prostate biopsy due to its legal implications in Europe.

PROSTATE CANCER - LIMITED UPDATE 2021 11


Figure 5.1: Prostate biopsy workflow to reduce infectious complications*

Indication for prostate biopsy?

Transperineal biopsy feasible?

Yes No

Transperineal biopsy - 1st choice (⊕⊕⊝⊝) Transrectal biopsy – 2nd choice (⊕⊕⊝⊝)
with: with:
• perineal cleansing • povidone-iodine rectal preparation
• antibiotic prophylaxis • antibiotic prophylaxis

Fluoroquinolones licensed?

No Yes

1. Targeted prophylaxis: based on rectal Duration of antibiotic prophylaxis ≥ 24 hrs


swab or stool cultures (⊕⊕⊝⊝)
2. Augmented prophylaxis: two or more 1. Targeted prophylaxis (⊕⊕⊝⊝): based
different classes of antibiotics on rectal swab or stool cultures
3. Alternative antibiotics (⊕⊝⊝⊝): 2. Augmented prophylaxis (⊕⊝⊝⊝):
• fosfomycin trometamol (e.g. 3 g before • Fluoroquinolone plus aminoglycoside
and 3 g 24-48 hrs after biopsy) • Fluoroquinolone plus cephalosporin
• cephalosporin (e.g. ceftriaxone 1 g i.m.;
cefixime 400 mg p.o. for 3 days starting 3. Fluoroquinolone prophylaxis
24 hrs before biopsy) (range: ⊕⊝⊝⊝ - ⊕⊕⊝⊝)
• aminoglycoside (e.g. gentamicin 3 mg/kg
i.v.; amikacin 15 mg/kg i.m.)

6.1.6 General guidelines for the treatment of prostate cancer

Recommendations Strength rating


Radiotherapeutic treatment
Offer low-dose rate (LDR) brachytherapy monotherapy to patients with good urinary Strong
function and low- or good prognosis intermediate-risk localised disease.
Offer LDR or high-dose rate brachytherapy boost combined with IMRT including Strong
IGRT to patients with good urinary function and intermediate-risk disease with
adverse features or high-risk disease.

6.2.1.3 Summary of evidence and guidelines for the treatment of low-risk disease

Summary of evidence
Systematic biopsies have been scheduled in AS protocols, the number and frequency of biopsies
varied, there is no approved standard.
Although per-protocol MR scans are increasingly used in AS follow-up no conclusive evidence exist in
terms of their benefit/and whether biopsy may be ommitted based on the imaging findings.
Personalised risk-based approaches will ultimately replace protocol-based management of patients on
AS.

Recommendations Strength rating


Active surveillance (AS)
Selection of patients
Perform a mpMRI before a confirmatory biopsy if no MRI has been performed Strong
before the initial biopsy.

12 PROSTATE CANCER - LIMITED UPDATE 2021


Radiotherapeutic treatment
Offer low-dose rate brachytherapy to patients with low-risk PCa, without a recent Strong
transurethral resection of the prostate and a good International Prostatic Symptom
Score.
Other therapeutic options
Do not offer ADT monotherapy to asymptomatic men not able to receive any local Strong
treatment.

6.2.2.5 Guidelines for the treatment of intermediate-risk disease

Recommendations Strength rating


Active surveillance (AS)
Offer AS to highly selected patients with ISUP grade group 2 disease (i.e. < 10% Weak
pattern 4, PSA < 10 ng/mL, < cT2a, low disease extent on imaging and biopsy)
accepting the potential increased risk of metastatic progression.
Radiotherapeutic treatment
Offer low-dose rate brachytherapy to intermediate-risk patients with ISUP grade 2 Strong
with < 33% of biopsy cores involved, without a recent transurethral resection of the
prostate and with a good International Prostatic Symptom Score.
For intensity-modulated radiotherapy (IMRT) plus image-guided radiotherapy (IGRT), Strong
use a total dose of 76–78 Gy or moderate hypofractionation (60 Gy/20 fx in 4 weeks
or 70 Gy/28 fx in 6 weeks), in combination with short-term androgen deprivation
therapy (ADT) (4 to 6 months).
In patients not willing to undergo ADT, use a total dose of IMRT plus IGRT Weak
(76–78 Gy) or moderate hypofractionation (60 Gy/20 fx in 4 weeks or 70 Gy/28 fx
in 6 weeks) or a combination with brachytherapy.

6.2.3.4 Guidelines for radical treatment of high-risk localised disease

Recommendation Strength rating


Therapeutic options outside surgery and radiotherapy
Only offer ADT monotherapy to those patients unwilling or unable to receive any Strong
form of local treatment if they have a prostate-specific antigen (PSA)-doubling time
< 12 months, and either a PSA > 50 ng/mL or a poorly-differentiated tumour.

6.2.5.7 Guidelines for adjuvant treatment in pN0 and pN1 disease after radical prostatectomy

Recommendation Strength rating


Offer adjuvant intensity-modulated radiation therapy (IMRT) plus image-guided Strong
radiation therapy (IGRT) to high-risk patients (pN0) with ISUP grade group 4–5 and
pT3 ± positive margins.

6.2.6.5 Recommendations for the management of persistent PSA after radical prostatectomy

Recommendation Strength rating


Offer a prostate-specific membrane antigen positron emission tomography (PSMA Weak
PET) scan to men with a persistent PSA > 0.2 ng/mL if the results will influence
subsequent treatment decisions.

6.3.14 Guidelines for second-line therapy after treatment with curative intent

Local salvage treatment Strength rating


Recommendations for biochemical recurrence (BCR) after radical prostatectomy
Offer monitoring, including prostate-specific antigen (PSA), to EAU BCR low-risk Weak
patients.

PROSTATE CANCER - LIMITED UPDATE 2021 13


Offer early salvage intensity-modulated radiotherapy plus image-guided Strong
radiotherapy to men with two consecutive PSA rises.
A negative PET/CT scan should not delay salvage radiotherapy (SRT), if otherwise Strong
indicated.
Do not wait for a PSA threshold before starting treatment. Once the decision for Strong
SRT has been made, SRT (at least 66 Gy) should be given as soon as possible.
Recommendations for BCR after radiotherapy
Offer monitoring, including prostate-specific antigen (PSA), to EAU Low-Risk BCR Weak
patients.
Only offer salvage radical prostatectomy (RP), brachytherapy, high-intensity focused Strong
ultrasound, or cryosurgical ablation to highly selected patients with biopsy proven
local recurrence within a clinical trial setting or well-designed prospective cohort
study undertaken in experienced centres.

6.4.9 Guidelines for the first-line treatment of metastatic disease

Recommendations Strength rating


Discuss combination therapy including ADT plus systemic therapy with all M1 patients. Strong
Do not offer ADT monotherapy to patients whose first presentation is M1 disease Strong
if they have no contraindications for combination therapy and have a sufficient
life expectancy to benefit from combination therapy and are willing to accept the
increased risk of side effects.
Do not offer ADT combined with surgery to M1 patients outside of clinical trials. Strong
Only offer metastasis-directed therapy to M1 patients within a clinical trial setting or Strong
well-designed prospective cohort study.

6.5.14 Summary of evidence and guidelines for life-prolonging treatments of castrate-resistant disease

Recommendations Strength rating


Treat patients with mCRPC with life-prolonging agents. Strong
Offer mCRPC patients somatic and/or germline molecular testing as well as testing Strong
for mismatch repair deficiencies or microsatellite instability.

6.5.15 Guidelines for systematic treatments of castrate-resistant disease

Recommendations Strength rating


Base the choice of treatment on the performance status, symptoms, co-morbidities, Strong
location and extent of disease, genomic profile, patient preference, and on the
previous treatment for hormone-sensitive metastatic PCa (mHSPC) (alphabetical
order: abiraterone, cabazitaxel, docetaxel, enzalutamide, olaparib, radium-223,
sipuleucel-T).
Offer patients with mCRPC who are candidates for cytotoxic therapy and are Strong
chemotherapy naïve docetaxel with 75 mg/m2 every 3 weeks.
Offer patients with mCRPC and progression following docetaxel chemotherapy Strong
further life-prolonging treatment options, which include abiraterone, cabazitaxel,
enzalutamide, radium-223 and olaparib in case of DNA homologous recombination
repair (HRR) alterations.
Base further treatment decisions of mCRPC on performance status, previous Strong
treatments, symptoms, co-morbidities, genomic profile, extent of disease and
patient preference.
Offer abiraterone or enzalutamide to patients previously treated with one or two lines Strong
of chemotherapy.
Avoid sequencing of androgen receptor targeted agents, Weak
Offer chemotherapy to patients previously treated with abiraterone or enzalutamide. Strong
Offer cabazitaxel to patients previously treated with docetaxel. Strong
Recommendations for BCR after radiotherapy
Offer poly(ADP-ribose) polymerase (PARP) inhibitors to pretreated mCRPC patients Strong
with relevant DNA repair gene mutations.

14 PROSTATE CANCER - LIMITED UPDATE 2021


7.2.6 Guidelines for follow-up during hormonal treatment

Recommendations Strength rating


In M1 patients, schedule follow-up at least every 3 to 6 months. Strong
In patients on long-term androgen deprivation therapy (ADT), measure initial bone Strong
mineral density to assess fracture risk.
During follow-up of patients receiving ADT, check PSA and testosterone levels and Strong
monitor patients for symptoms associated with metabolic syndrome as a side effect
of ADT. As a minimum requirement, include a disease-specific history, haemoglobin,
serum creatinine, alkaline phosphatase, lipid profiles and HbA1c level measurements.
When disease progression is suspected, restaging is needed and the subsequent Strong
follow up adapted/individualised.
In M1 patients perform regular imaging (CT and bone scan) even without PSA Weak
progression.
In patients with suspected progression, assess the testosterone level. By definition, Strong
castration- resistant PCa requires a testosterone level < 50 ng/dL (< 1.7 nM/L).

8.3.2.5 Guidelines for quality of life in men undergoing systemic treatments

Recommendations Strength rating


Advise men on ADT to maintain a healthy weight and diet, to stop smoking, Strong
reduce alcohol to < 2 units daily and have yearly screening for diabetes and
hypercholesterolemia. Ensure that calcium and vitamin D meet recommended levels.
Offer anti-resorptive therapy to men on long term ADT with either a BMD T-score of Strong
<-2.5 or with an additional clinical risk factor for fracture or annual bone loss on ADT
is confirmed to exceed 5%.

2. METHODS
2.1 Data identification
For the 2020 PCa Guidelines, new and relevant evidence has been identified, collated and appraised through a
comprehensive review of the GRADE forms [see definition below) and associated recommendation. Changes in
recommendations were only considered on the basis of high level evidence (i.e. systematic reviews with meta-
analysis, randomised controlled trials [RCTs], and prospective comparative studies) published in the English
language. A total of 279 new references were added to the 2021 PCa Guidelines. Additional information can
be found in the general Methodology section of this print and online at the EAU website: https://uroweb.org/
guidelines/policies-and-methodological-documents/.

For each recommendation within the guidelines there is an accompanying online strength rating form, the basis
of which is a modified GRADE methodology [3, 4]. These forms address a number of key elements namely:
1. the overall quality of the evidence which exists for the recommendation, references used in
this text are graded according to a classification system modified from the Oxford Centre for
Evidence-Based Medicine Levels of Evidence [5];
2. the magnitude of the effect (individual or combined effects);
3. the certainty of the results (precision, consistency, heterogeneity and other statistical or
study related factors);
4. the balance between desirable and undesirable outcomes;
5. the impact of patient values and preferences on the intervention;
6. the certainty of those patient values and preferences.

These key elements are the basis which panels use to define the strength rating of each recommendation.
The strength of each recommendation is represented by the words ‘strong’ or ‘weak’ [6]. The strength of each
recommendation is determined by the balance between desirable and undesirable consequences of alternative
management strategies, the quality of the evidence (including certainty of estimates), and nature and variability
of patient values and preferences. The strength rating forms will be available online.

PROSTATE CANCER - LIMITED UPDATE 2021 15


A list of Associations endorsing the EAU Guidelines can also be viewed online at the above address. In addition,
the International Society of Geriatric Oncology (SIOG), the European Society for Radiotherapy & Oncology
(ESTRO), the European Society for Urogenital Radiology (ESUR), the European Association of Nuclear Medicine
(EANM) and the International Society of Urological Pathology (ISUP) have endorsed the PCa Guidelines.

2.2 Review
All radiotherapy sections from the 2021 print were peer-reviewed prior to publication, as were Sections 5.4
(Evaluating life expectancy and health status) and Chapter 8 (Quality of life). Publications ensuing from systematic
reviews have all been peer-reviewed.

2.3 Future goals


Results of ongoing and new systematic reviews will be included in the 2022 update of the PCa Guidelines:

• A systematic review on the deferred treatment with curative intent for localised PCa, explore
heterogeneity of definitions, thresholds and criteria [7];
• A systematic review on progression criteria and quality of life (QoL) of patients diagnosed with PCa;
• A systematic review on the impact of surgeon and hospital caseload volume on oncological and non-
oncological outcomes after radical prostatectomy for non-metastatic prostate cancer [8];
• Evaluation of oncological outcomes and data quality in studies assessing nerve sparing versus non-nerve
sparing radiccal prostatectomy in non-metastatic prostate cancer: a systematic review [9];
• Patient- and tumor-related prognostic factors for post-operative incontinence after radical prostatectomy
for non-metastatic prostate cancer: A systematic review and meta-analysis [10];
• Care pathways for the various stages of PCa management are being developed. These pathways will, in
due time, inform treatment flowcharts and an interactive app.

3. EPIDEMIOLOGY AND AETIOLOGY


3.1 Epidemiology
Prostate cancer is the second most commonly diagnosed cancer in men, with an estimated 1.1 million
diagnoses worldwide in 2012, accounting for 15% of all cancers diagnosed [11]. The frequency of autopsy-
detected PCa is roughly the same worldwide [12]. A systematic review of autopsy studies reported a
prevalence of PCa at age < 30 years of 5% (95% confidence interval [CI]: 3–8%), increasing by an odds ratio
(OR) of 1.7 (1.6–1.8) per decade, to a prevalence of 59% (48–71%) by age > 79 years [13].
The incidence of PCa diagnosis varies widely between different geographical areas, being highest
in Australia/New Zealand and Northern America (age-standardised rates [ASR] per 100,000 of 111.6 and 97.2,
respectively), and in Western and Northern Europe (ASRs of 94.9 and 85, respectively), largely due to the
use of prostate-specific antigen (PSA) testing and the aging population. The incidence is low in Eastern and
South-Central Asia (ASRs of 10.5 and 4.5, respectively). Rates in Eastern and Southern Europe were low but
have shown a steady increase [11, 12]. Incidence and disease stage distibution patters follow (inter)national
organisations‘ recommendations (see Section 5.1) [14].
There is relatively less variation in mortality rates worldwide, although rates are generally high in
populations of African descent (Caribbean: ASR of 29 and Sub-Saharan Africa: ASRs ranging between 19 and
14), intermediate in the USA and very low in Asia (South-Central Asia: ASR of 2.9) [11].

3.2 Aetiology
3.2.1 Family history/hereditary prostate cancer
Family history and ethnic background are associated with an increased PCa incidence suggesting a genetic
predisposition [15, 16]. Only a small subpopulation of men with PCa have true hereditary disease. Hereditary
PCa (HPCa) is associated with a six to seven year earlier disease onset but the disease aggressiveness and
clinical course does not seem to differ in other ways [15, 17].

In a large USA population database, HPCa (in 2.18% of participants) showed a relative risk (RR) of 2.30 for
diagnosis of any PCa, 3.93 for early-onset PCa, 2.21 for lethal PCa, and 2.32 for clinically significant PCa
(csPCa) [18]. These increased risks of HPCa were higher than for familial PCa (> 2 first- or second-degree
relatives with PCa on the same side of the pedigree), or familial syndromes such as hereditary breast and
ovarian cancer and Lynch syndrome. The probability of high-risk PCa at age 65 was 11.4% (vs. a population

16 PROSTATE CANCER - LIMITED UPDATE 2021


risk of 1.4%) in a Swedish population-based study [19].

3.2.1.1 Germline mutations and prostate cancer


Genome-wide association studies have identified more than 100 common susceptibility loci contributing to
the risk for PCa [20-22]. Clinical cohort studies have reported rates of 15% to 17% of germline mutations
independent of stage [23, 24]. Giri et al. studied clinical genetic data from men with PCa unselected for
metastatic disease undergoing multigene testing across the US [23]. The authors found that 15.6% of men
with PCa have pathogenic variants identified in genes tested (BRCA1, BRCA2, HOXB13, MLH1, MSH2, PMS2,
MSH6, EPCAM, ATM, CHEK2, NBN, and TP53), and 10.9% of men have germline pathogenic variants in DNA
repair genes (see Table 5.2). Pathogenic variants were most commonly identified in BRCA2 (4.5%), CHEK2
(2.2%), ATM (1.8%), and BRCA1 (1.1%) [23]. Presence of Gleason 8 or higher was significantly associated with
DNA repair pathogenic variants (OR 1.85 [95% CI: 1.22–2.80], p = 0.004) [23].

Nicolosi and colleagues reported frequency and distribution of positive germline variants in 3,607 unselected
PCa patients and found that 620 (17.2%) had a pathogenic germline variant [24]. The percentage of BRCA1/2
mutations in this study was 6%. Among unselected men with metastatic PCa, an incidence of 11.8% was
found for germline mutations in genes mediating DNA-repair processes [25]. Most mutations were seen in
BRCA2 (5.35%), ATM (1.6%), CHEK2 (1.9%), BRCA1 (0.9%), and PALB2 (0.4%).
Targeted genomic analysis of genes associated with an increased risk of PCa could offer options to
identify families at high risk [26, 27].

Nyberg et al. presented results of a prospective cohort study of male BRCA1 and BRCA2 carriers and their
PCa risk confirming BRCA2 association with aggressive PCa [28]. Mutations in the PCa cluster region of the
BRCA2 gene may increase the PCa risk in particular. Castro and colleagues analysed the outcomes of 2,019
patients with PCa (18 BRCA1 carriers, 61 BRCA2 carriers, and 1,940 non-carriers). Prostate cancers with
germline BRCA1/2 mutations were more frequently associated with ISUP > 4, T3/T4 stage, nodal involvement,
and metastases at diagnosis than PCa in non-carriers [29]. BReast-CAncer susceptibility gene mutation
carriers were reported to have worse outcome when compared to non-carriers after local therapy [30].
In a retrospective study of 313 patients who died of PCa and 486 patients with low-risk localised
PCa, the combined BRCA1/2 and ATM mutation carrier rate was significantly higher in lethal PCa patients
(6.07%) than in localised PCa patients (1.44%) [31].

The Identification of Men With a Genetic Predisposition to ProstAte Cancer (IMPACT) study, which evaluates
targeted PCa screening (annually, biopsy recommended if PSA > 3.0 ng/mL) using PSA in men aged 40–69
years with germline BRCA1/2 mutations, has recently reported interim results [32]. The authors found that after
3 years of screening, BRCA2 mutation carriers were associated with a higher incidence of PCa, a younger
age of diagnosis, and more clinically significant tumours compared with non-carriers. The influence of BRCA1
mutations on PCa remained unclear. No differences in age or tumour characteristics were detected between
BRCA1 carriers and BRCA1 non-carriers. Limitations of the IMPACT study include the lack of multiparametric
magnetic resonance imaging (mpMRI) data and targeted biopsies as it was initiated before that era.

Similarly, Mano et al. reported on an Israeli cohort in which men with BRCA1 and BRCA2 mutations had a
significantly higher incidence of malignant disease. In contrast to findings of the IMPACT study, the rate of PCa
among BRCA1 carriers was more than twice as high (8.6% vs. 3.8%) compared to the general population [33].

3.2.2 Risk factors


A wide variety of exogenous/environmental factors have been discussed as being associated with the risk
of developing PCa or as being aetiologically important for the progression from latent to clinical PCa [34].
Japanese men have a lower PCa risk compared to men from the Western world. However, as Japanese men
move from Japan to California, their risk of PCa increases, approaching that of American men, implying a role
of environmental or dietary factors [35]. However, currently there are no known effective preventative dietary or
pharmacological interventions.

3.2.2.1 Metabolic syndrome


The single components of metabolic syndrome (MetS) hypertension (p = 0.035) and waist circumference
> 102 cm (p = 0.007) have been associated with a significantly greater risk of PCa, but in contrast, having
> 3 components of MetS is associated with a reduced risk (OR: 0.70, 95% CI: 0.60–0.82) [36, 37].

3.2.2.1.1 Diabetes/metformin
On a population level, metformin users (but not other oral hypoglycaemic agents) were found to be at a
decreased risk of PCa diagnosis compared with never-users (adjusted OR: 0.84, 95% CI: 0.74–0.96) [38].

PROSTATE CANCER - LIMITED UPDATE 2021 17


In 540 diabetic participants of the Reduction by Dutasteride of Prostate Cancer Events (REDUCE) study,
metformin use was not significantly associated with PCa and therefore not advised as a preventive measure
(OR: 1.19, p = 0.50) [39]. The ongoing Systemic Therapy in Advancing or Metastatic Prostate Cancer:
Evaluation of Drug Efficacy (STAMPEDE) trial assesses metformin use in advanced PCa (Arm K) [40].

3.2.2.1.2 Cholesterol/statins
A meta-analysis of 14 large prospective studies did not show an association between blood total cholesterol, high-
density lipoprotein cholesterol, low-density lipoprotein cholesterol levels and the risk of either overall PCa or high-
grade PCa [36]. Results from the REDUCE study also did not show a preventive effect of statins on PCa risk [39].

3.2.2.1.3 Obesity
Within the REDUCE study, obesity was associated with lower risk of low-grade PCa in multivariable analyses
(OR: 0.79, p = 0.01), but increased risk of high-grade PCa (OR: 1.28, p = 0.042) [41]. This effect seems mainly
explained by environmental determinants of height/body mass index (BMI) rather than genetically elevated
height or BMI [42].

3.2.2.2 Dietary factors


The association between a wide variety of dietary factors and PCa have been studied (Table 3.1).

Table 3.1: Dietary factors that have been associated with PCa

Alcohol High alcohol intake, but also total abstention from alcohol has been associated with
a higher risk of PCa and PCa-specific mortality [43]. A meta-analysis shows a dose-
response relationship with PCa [44].
Dairy A weak correlation between high intake of protein from dairy products and the risk of
PCa was found [45].
Fat No association between intake of long-chain omega-3 poly-unsaturated fatty acids
and PCa was found [46]. A relation between intake of fried foods and risk of PCa may
exist [47].
Tomatoes A trend towards a favourable effect of tomato intake (mainly cooked) and lycopenes on
(lycopenes/ PCa incidence has been identified in meta-analyses [48, 49]. Randomised controlled
carotenes) trials comparing lycopene with placebo did not identify a significant decrease in the
incidence of PCa [50].
Meat A meta-analysis did not show an association between red meat or processed meat
consumption and PCa [51].
Phytoestrogens Phytoestrogen intake was significantly associated with a reduced risk of PCa in a
meta-analysis [52].
Soy Total soy food intake has been associated with reduced risk of PCa, but also with
(phytoestrogens increased risk of advanced disease [53, 54].
[isoflavones/
coumestans])
Vitamin D A U-shaped association has been observed, with both low- and high vitamin-D
concentrations being associated with an increased risk of PCa, and more strongly for
high-grade disease [54, 55].
Vitamin E/Selenium An inverse association of blood, but mainly nail selenium levels (reflecting long-term
exposure) with aggressive PCa have been found [56, 57]. Selenium and Vitamin E
supplementation were, however, found not to affect PCa incidence [58].

3.2.2.3 Hormonally active medication


3.2.2.3.1 5-alpha-reductase inhibitors (5-ARIs)
Although it seems that 5-ARIs have the potential of preventing or delaying the development of PCa (~25%, for
ISUP grade 1 cancer only), this must be weighed against treatment-related side effects as well as the potential
small increased risk of high-grade PCa [59-61]. None of the available 5-ARIs have been approved by the
European Medicines Agency (EMA) for chemoprevention.

3.2.2.3.2 Testosterone
Hypogonadal men receiving testosterone supplements do not have an increased risk of PCa [62]. A pooled
analysis showed that men with very low concentrations of free testosterone (lowest 10%) have a below average
risk (OR: 0.77) of PCa [63].

18 PROSTATE CANCER - LIMITED UPDATE 2021


3.2.2.4 Other potential risk factors
A significantly higher rate of ISUP > 2 PCas (hazard ratio [HR]: 4.04) was found in men with inflammatory
bowel disease when compared to the general population [64]. Balding was associated with a higher risk
of PCa death [65]. Gonorrhoea was significantly associated with an increased incidence of PCa (OR: 1.31,
95% CI: 1.14–1.52) [66]. Occupational exposure may also play a role, based on a meta-analysis which
revealed that night-shift work is associated with an increased risk (2.8%, p = 0.030) of PCa [67]. Current
cigarette smoking was associated with an increased risk of PCa death (RR: 1.24, 95% CI: 1.18–1.31) and with
aggressive tumour features and worse prognosis, even after cessation [68, 69]. A meta-analysis on Cadmium
(Cd) found a positive association (magnitude of risk unknown due to heterogeneity) between high Cd exposure
and risk of PCa for occupational exposure, but not for non-occupational exposure, potentially due to higher Cd
levels during occupational exposure [70]. Men positive for human papillomavirus-16 may be at increased risk
[71]. Plasma concentration of the estrogenic insecticide chlordecone is associated with an increase in the risk
of PCa (OR: 1.77 for highest tertile of values above the limit of detection) [72].
A number of other factors previously linked to an increased risk of PCa have been disproved
including vasectomy [73] and self-reported acne [74]. There are conflicting data about the use of aspirin or
nonsteroidal anti-inflammatory drugs and the risk of PCa and mortality [75, 76].
Ultraviolet radiation exposure decreased the risk of PCa (HR: 0.91, 95% CI: 0.88–0.95) [77]. A review
found a small but protective association of circumcision status with PCa [78]. Higher ejaculation frequency
(> 21 times a month vs. 4 to 7 times) has been associated with a 20% lower risk of PCa [79].
The associations with PCa identified to date lack evidence for causality. As a consequence there is
no data to suggest effective preventative strategies.

3.2.3 Summary of evidence for epidemiology and aetiology

Summary of evidence LE
Prostate cancer is a major health concern in men, with incidence mainly dependent on age. 3
Genetic factors are associated with risk of (aggressive) PCa. 3
A variety of exogenous/environmental factors have been associated with PCa incidence and prognosis. 3
Selenium or vitamin-E supplements have no beneficial effect in preventing PCa. 2a
In hypogonadal men, testosterone supplements do not increase the risk of PCa. 2
No specific preventive or dietary measures are recommended to reduce the risk of developing PCa. 1a

4. CLASSIFICATION AND STAGING SYSTEMS


4.1 Classification
The objective of a tumour classification system is to combine patients with a similar clinical outcome. This
allows for the design of clinical trials on relatively homogeneous patient populations, the comparison of
clinical and pathological data obtained from different hospitals across the world, and the development of
recommendations for the treatment of these patient populations. Throughout these Guidelines the 2017
Tumour, Node, Metastasis (TNM) classification for staging of PCa (Table 4.1) [80] and the EAU risk group
classification, which is essentially based on D’Amico’s classification system for PCa, are used (Table 4.2) [81].
The latter classification is based on the grouping of patients with a similar risk of biochemical recurrence (BCR)
after radical prostatectomy (RP) or external beam radiotherapy (EBRT). Multiparametric resonance imaging and
targeted biopsy may cause a stage shift in risk classification systems [82].

PROSTATE CANCER - LIMITED UPDATE 2021 19


Table 4.1: Clinical Tumour Node Metastasis (TNM) classification of PCa [80]

T - Primary Tumour (stage based on digital rectal examination [DRE] only)


TX Primary tumour cannot be assessed
T0 No evidence of primary tumour
T1 Clinically inapparent tumour that is not palpable
T1a Tumour incidental histological finding in 5% or less of tissue resected
T1b Tumour incidental histological finding in more than 5% of tissue resected
T1c Tumour identified by needle biopsy (e.g. because of elevated prostate-specific antigen [PSA])
T2 Tumour that is palpable and confined within the prostate
T2a Tumour involves one half of one lobe or less
T2b Tumour involves more than half of one lobe, but not both lobes
T2c Tumour involves both lobes
T3 Tumour extends through the prostatic capsule
T3a Extracapsular extension (unilateral or bilateral)
T3b Tumour invades seminal vesicle(s)
T4 Tumour is fixed or invades adjacent structures other than seminal vesicles: external sphincter, rectum,
levator muscles, and/or pelvic wall
N - Regional (pelvic) Lymph Nodes1
NX Regional lymph nodes cannot be assessed
N0 No regional lymph node metastasis
N1 Regional lymph node metastasis
M - Distant Metastasis2
M0 No distant metastasis
M1 Distant metastasis
M1a Non-regional lymph node(s)
M1b Bone(s)
M1c Other site(s)
1 Metastasis no larger than 0.2 cm can be designated pNmi.
2 When more than one site of metastasis is present, the most advanced category is used. (p)M1c is the most
advanced category.

Clinical T stage only refers to DRE findings; local imaging findings are not considered in the TNM classification.
Pathological staging (pTNM) is based on histopathological tissue assessment and largely parallels the clinical
TNM, except for clinical stage T1 and the T2 substages. Pathological stages pT1a/b/c do not exist and
histopathologically confirmed organ-confined PCas after RP are pathological stage pT2. The current Union for
International Cancer Control (UICC) no longer recognises pT2 substages [80].

4.2 Gleason score and International Society of Urological Pathology 2014 grade
In the original Gleason grading system, 5 Gleason grades (ranging from 1–5) based on histological tumour
architecture were distinguished, but in the 2005 and subsequent 2014 International Society of Urological
Pathology (ISUP) Gleason score (GS) modifications Gleason grades 1 and 2 were eliminated [83, 84]. The
2005 ISUP modified GS of biopsy-detected PCa comprises the Gleason grade of the most extensive (primary)
pattern, plus the second most common (secondary) pattern, if two are present. If one pattern is present, it
needs to be doubled to yield the GS. For three grades, the biopsy GS comprises the most common grade plus
the highest grade, irrespective of its extent. The grade of intraductal carcinoma should also be incorporated in
the GS [85]. In addition to reporting of the carcinoma features for each biopsy, an overall (or global) GS based
on the carcinoma-positive biopsies can be provided. The global GS takes into account the extent of each
grade from all prostate biopsies. The 2014 ISUP endorsed grading system [84, 86] limits the number of PCa
grades, ranging them from 1 to 5 (see Table 4.2), in order to:
1. align the PCa grading with the grading of other carcinomas;
2. eliminate the anomaly that the most highly differentiated PCas have a GS 6;
3. to further define the clinically highly significant distinction between GS 7(3+4) and 7(4+3) PCa [87].

20 PROSTATE CANCER - LIMITED UPDATE 2021


Table 4.2: EAU risk groups for biochemical recurrence of localised and locally advanced prostate cancer

Definition
Low-risk Intermediate-risk High-risk
PSA < 10 ng/mL PSA 10-20 ng/mL PSA > 20 ng/mL any PSA
and GS < 7 (ISUP grade 1) or GS 7 (ISUP grade 2/3) or GS > 7 (ISUP grade 4/5) any GS (any ISUP grade)
and cT1-2a or cT2b or cT2c cT3-4 or cN+
Localised Locally advanced
GS = Gleason score; ISUP = International Society for Urological Pathology; PSA = prostate-specific antigen.

Table 4.3: International Society of Urological Pathology 2014 grade (group) system

Gleason score ISUP grade


2-6 1
7 (3+4) 2
7 (4+3) 3
8 (4+4 or 3+5 or 5+3) 4
9-10 5

4.3 Prognostic relevance of stratification


Comparison of various risk stratification tools using death of disease as outcome parameter revealed that
3-tiered systems, including the EAU and D’Amico risk classification, were inferior to the MSKCC nomograms,
the CAPRA score and the 5-tiered CPG system [88]. Similarly, using death of disease as outcome parameter a
9-tier risk clinical prognostic stage group system, based on age, cT, cN, ISUP grade, percentage positive cores and
PSA as input variables further improved risk stratification in patients treated by surgery or radiotherapy (RT) [89].
A more precise stratification of the clinically heterogeneous subset of intermediate-risk group
patients could provide a better framework for their management. The adoption of the current ISUP grading
system, defining the split-up of GS 7 cancers into ISUP grade 2 (primary Gleason grade 3) and ISUP grade
3 (primary Gleason grade 4) because of their distinct prognostic impact strengthens such a separation of the
intermediate-risk group into a low-intermediate (ISUP grade 2) and high-intermediate (ISUP grade 3) risk group
[86] (see Section 6.2.2).
Emerging clinical data support this distinction between favourable- and unfavourable-risk patient
categories within the intermediate-risk group [87, 90].

4.4 Guidelines for classification and staging systems

Recommendations Strength rating


Use the Tumour, Node, Metastasis (TNM) classification for staging of PCa. Strong
Use the International Society of Urological Pathology (ISUP) 2014 system for grading of PCa. Strong

5. DIAGNOSTIC EVALUATION
5.1 Screening and early detection
5.1.1 Screening
Population or mass screening is defined as the ‘systematic examination of asymptomatic men (at risk)’ and is
usually initiated by health authorities. The co-primary objectives are:
• reduction in mortality due to PCa;
• a maintained QoL as expressed by QoL-adjusted gain in life years (QALYs).

Prostate cancer mortality trends range widely from country to country in the industrialised world [91]. Mortality
due to PCa has decreased in most Western nations but the magnitude of the reduction varies between countries.
Currently, screening for PCa still remains one of the most controversial topics in the urological literature [92].

PROSTATE CANCER - LIMITED UPDATE 2021 21


Initial widespread aggressive screening in USA was associated with a decrease in mortality [93].
In 2012 the US Preventive Services Task Force (USPSTF) released a recommendation against PSA-based
screening [94], which was adopted in the 2013 AUA Guidelines [95] and resulted in a reduction in the use
of PSA for early detection [96]. This reduction in the use of PSA testing was associated with higher rates of
advanced disease at diagnosis (e.g., a 6% increase in the number of patients with metastatic PCa [14, 97-100].
While PCa mortality had decreased for two decades since the introduction of PSA testing [101], the incidence
of advanced disease and, possibly, cancer-related mortality slowly increased from 2008 and accelerated
in 2012 [102]. Moreover, additional evidence suggests a long-term benefit of PSA population screening in
terms of reduction of cancer-specific mortality [103, 104]. However, the temporal relationship between PSA
testing and decreased mortality, as well as a rising mortality following immediately after the USPSTF and AUA
Guidelines recommendation against PSA testing questions the direct causative link between both points.

In 2017 the USPSTF issued an updated statement suggesting that men aged 55–69 should be informed about
the benefits and harms of PSA-based screening as this might be associated with a small survival benefit. The
USPSTF has now upgraded this recommendation to a grade C [105], from a previous grade ‘D’ [105-107]. They
highlighted the fact that the decision to be screened should be an individual one. The grade D recommendation
remains in place for men over 70 years old. This represents a major switch from discouraging PSA-based
screening (grade D) to offering early diagnosis to selected men depending on individual circumstances.
A comparison of systematic and opportunistic screening suggested over-diagnosis and mortality
reduction in the systematic screening group compared to a higher over-diagnosis with a marginal survival
benefit, at best, in the opportunistic screening regimen [108].

A Cochrane review published in 2013 [109], which has since been updated [110], presents the main overview
to date. The findings of the updated publication (based on a literature search until April 3, 2013) are almost
identical to the 2009 review:
• Screening is associated with an increased diagnosis of PCa (RR: 1.3, 95% CI: 1.02–1.65).
• Screening is associated with detection of more localised disease (RR: 1.79, 95% CI: 1.19–2.70) and less
advanced PCa (T3–4, N1, M1) (RR: 0.80, 95% CI: 0.73–0.87).
• From the results of 5 RCTs, randomising more than 341,000 men, no PCa-specific survival benefit was
observed (RR: 1.00, 95% CI: 0.86–1.17). This was the main endpoint in all trials.
• From the results of four available RCTs, no overall survival (OS) benefit was observed (RR: 1.00, 95% CI:
0.96–1.03).
The included studies applied a range of different screening measures and testing intervals in patients who had
ondergone prior PSA testing, to various degrees. None included the use of risk calculators, MRI prior to biopsy
(vs. a single PSA threshold) or AS (as an alternative to RP) which no longer reflects current standard practice.

The diagnostic tool (i.e. biopsy procedure) was not associated with increased mortality within 120 days after
biopsy in screened men as compared to controls in the two largest population-based screening populations
(ERSPC and PLCO), in contrast to a 120-day mortality rate of 1.3% in screened vs. 0.3% in controls,
respectively, in a Canadian population-based screening study [111]. Increased diagnosis has historically led
to over-treatment with associated side effects. However, despite this, the impact on the patient’s overall QoL
is still unclear. Population level screening has never been shown to be detrimental [112-114]. Nevertheless,
all these findings have led to strong advice against systematic population-based screening in most countries,
including those in Europe.

In case screening is considered, a single PSA test is not enough based on the Cluster Randomized Trial
of PSA Testing for Prostate Cancer (CAP) trial. The CAP trial evaluated a single PSA screening vs. controls
not undergoing PSA screening on PCa detection in men aged 50 to 69 years old. The single PSA screening
intervention detected more low-risk PCa cases but had no significant effect on PCa mortality after a median
follow-up of 10 years [115].
Since 2013, the European Randomized Study of Screening for Prostate Cancer (ERSPC) data have
been updated with 16 years of follow-up (see Table 5.1) [116]. The key message is that with extended follow-
up, the mortality reduction remains unchanged (21%, and 29% after non-compliance adjustment). However
the number needed to screen (NNS) and to treat is decreasing, and is now below the NNS observed in breast
cancer trials [116, 117]. Long-term follow-up of the PLCO (Prostate, Lung, Colon, Ovarian cancer screening
trial) showed no survival benefit for screening at a median follow-up of 16.7 years but a significant 17%
increase in Gleason score 2–6 cancers and 11% decrease in Gleason score 8–10 cancers [118].

22 PROSTATE CANCER - LIMITED UPDATE 2021


Table 5.1: Follow-up data from the ERSPC study [116]

Years of follow-up Number needed to screen Number needed to treat


9 1,410 48
11 979 35
13 781 27
16 570 18

5.1.2 Early detection


An individualised risk-adapted strategy for early detection may still be associated with a substantial risk of
over-diagnosis. It is essential to remember that breaking the link between diagnosis and active treatment is the
only way to decrease over-treatment, while still maintaining the potential benefit of individual early diagnosis for
men requesting it [16, 119].

Men at elevated risk of having PCa are those > 50 years [120] or at age > 45 years with a family history of
PCa (either paternal or maternal) [121] or of African descent [122, 123]. Men of African descent are more likely
to be diagnosed with more advanced disease [124] and upgrade was more frequent after prostatectomy as
compared to Caucasian men (49% vs. 26%) [125].
Germline mutations are associated with an increased risk of the development of aggressive PCa, i.e.
BRCA2 [126, 127]. Prostate-specific antigen screening in male BRCA1 and 2 carriers detected more significant
cancers at a younger age compared to non-mutation carriers [32, 33].

Men with a baseline PSA < 1 ng/mL at 40 years and < 2 ng/mL at 60 years are at decreased risk of PCa
metastasis or death from PCa several decades later [128, 129].

The use of DRE alone in the primary care setting had a sensitivity and specificity below 60%, possibly due to
inexperience, and can therefore not be recommended to exclude PCa [130]. Informed men requesting an early
diagnosis should be given a PSA test and undergo a DRE [131]. Prostate-specific antigen measurement and
DRE need to be repeated [115], but the optimal intervals for PSA testing and DRE follow-up are unknown as
they varied between several prospective trials. A risk-adapted strategy might be a consideration, based on the
initial PSA level. This could be every 2 years for those initially at risk, or postponed up to 8 to 10 years in those
not at risk with an initial PSA < 1 ng/mL at 40 years and a PSA < 2 ng/mL at 60 years of age and a negative
family history [132]. An analysis of ERSPC data supports a recommendation for an 8-year screening interval in
men with an initial PSA concentration < 1 μg/L; fewer than 1% of men with an initial PSA concentration < 1 ng/mL
were found to have a concentration above the biopsy threshold of 3 μg/L at 4-year follow-up; the cancer
detection rate by 8 years was close to 1% [133]. The long-term survival and QoL benefits of extended PSA
re-testing (every 8 years) remain to be proven at a population level. Data from the Goteborg arm of the ERSPC
trial suggest that the age at which early diagnosis should be stopped remains controversial, but an individual’s
life expectancy must definitely be taken into account. Men who have less than a 15-year life expectancy are
unlikely to benefit, based on data from the Prostate Cancer Intervention Versus Observation Trial (PIVOT)
and the ERSPC trials. Furthermore, although there is no simple tool to evaluate individual life expectancy;
co-morbidity is at least as important as age. A detailed review can be found in Section 5.4 ‘Estimating life
expectancy and health status’ and in the SIOG Guidelines [134].
Multiple tools are now available to determine the need for a biopsy to establish the diagnosis of a
PCa, including imaging by MRI, if available (see Section 5.2.4).

Urine and serum biomarkers as well as tissue-based biomarkers have been proposed for improving detection
and risk stratification of PCa patients, potentially avoiding unnecessary biopsies. However, further studies are
necessary to validate their efficacy [135]. At present there is too limited data to implement these markers into
routine screening programmes (see Section 5.2.3).

Risk calculators may be useful in helping to determine (on an individual basis) what the potential risk of cancer
may be, thereby reducing the number of unnecessary biopsies. Several tools developed from cohort studies
are available including:
• the PCPT cohort: PCPTRC 2.0 http://myprostatecancerrisk.com/;
• the ERSPC cohort: http://www.prostatecancer-riskcalculator.com/seven-prostate-cancer-risk-calculators;
An updated version was presented in 2017 including prediction of low and high risk now also based on
the ISUP grading system and presence of cribriform growth in histology [136].
• a local Canadian cohort: https://sunnybrook.ca/content/?page=asure-calc (among others).

PROSTATE CANCER - LIMITED UPDATE 2021 23


Since none of these risk calculators has clearly shown superiority, it remains a personal decision as to which
one to use [137]. A comparative analysis showed risk calculators containing MRI to be most predictive,
however, regional modifications may be required before implementation [138].

5.1.3 Genetic testing for inherited prostate cancer


Increasing evidence supports the implementation of genetic counselling and germline testing in early detection
and PCa management. Several commercial screening panels are now available to assess main PCa risk genes
[139]. However, it remains unclear when germline testing should be considered and how this may impact
localised and metastatic disease management. Germline BRCA1 and BRCA2 mutations occur in approximately
0.2% to 0.3% of the general population [140]. It is important to understand the difference between somatic
testing, which is performed on the tumour, and germline testing, which is performed on blood or saliva and
identifies inherited mutations. Genetic counselling is required prior to and after undergoing germline testing.

Germline mutations can drive the development of aggressive PCa. Therefore, the following men with a personal
or family history of PCa or other cancer types arising from DNA repair gene mutations should be considered for
germline testing:
• Men with metastatic PCa;
• Men with high-risk PCa and a family member diagnosed with PCa at age < 60 years;
• Men with multiple family members diagnosed with csPCa at age < 60 years or a family member who died
from PCa cancer;
• Men with a family history of high-risk germline mutations or a family history of multiple cancers on the
same side of the family.

Further research in this field (including not so well-known germline mutations) is needed to develop screening,
early detection and treatment paradigms for mutation carriers and family members.

Table 5.2: Germline mutations in DNA repair genes associated with increased risk of prostate cancer

Gene Location Prostate Cancer risk Findings


BRCA2 13q12.3 - 2.5 to 4.6 [141, 142] • up to 12 % of men with metastatic PCa harbour
-P Ca at 55 years or germline mutations in 16 genes (including BRCA2
under: RR: 8–23 [5.3%]) [25]
[141, 143] • 2% of men with early-onset PCa harbour germline
mutations in the BRCA2 gene [141]
• BRCA2 germline alteration is an independent
predictor of metastases and worse PCa-specific
survival [29, 144]
ATM 11q22.3 RR: 6.3 for metastatic • higher rates of lethal PCa among mutation carriers [31]
prostate [25] • up to 12% of men with metastatic PCa harbour germline
mutations in 16 genes (including ATM [1.6%]) [25]
CHEK2 22q12.1 OR 3.3 [145, 146] • up to 12% of men with metastatic PCa harbour germline
mutations in 16 genes (including CHEK2 [1.9%]) [25]
BRCA1 17q21 RR: 1.8–3.8 at 65 • higher rates of lethal PCa among mutation carriers [31]
years or under • up to 12% of men with metastatic PCa harbour germline
[147, 148] mutations in 16 genes (including BRCA1 [0.9%]) [25]
HOXB13 17q21.2 OR 3.4–7.9 [26, 149] • significantly higher PSA at diagnosis, higher Gleason
score and higher incidence of positive surgical margins
in the radical prostatectomy specimen than non-carriers
[150]
MMR genes 3p21.3 RR: 3.7 [151] • Mutations in MMR genes are responsible for Lynch
MLH1 2p21 syndrome [146]
MSH2 2p16 • MSH2 mutation carriers are more likely to develop PCa
MSH6 7p22.2 than other MMR gene mutation carriers [152]
PMS2
BRCA2 = breast cancer gene 2; ATM = ataxia telangiectasia mutated; CHEK2 = checkpoint kinase 2;
BRCA1 = breast cancer gene 1; HOXB13 = homeobox B13; MMR = mismatch repair; MLH1 = mutL homolog 1;
MSH2 = mutS homolog 2; MSH6 = mutS homolog 6; PMS2 = post-meiotic segregation increased 2;
PCa = prostate cancer; RR = relative risk; OR = odds ratio.

24 PROSTATE CANCER - LIMITED UPDATE 2021


5.1.4 Guidelines for germline testing*

Recommendations Strength rating


Consider germline testing in men with metastatic PCa. weak
Consider germline testing in men with high-risk PCa who have a family member diagnosed weak
with PCa at age < 60 years.
Consider germline testing in men with multiple family members diagnosed with PCa at age weak
< 60 years or a family member who died from PCa.
Consider germline testing in men with a family history of high-risk germline mutations or a weak
family history of multiple cancers on the same side of the family.
*Genetic counseling is required prior to germline testing.

5.1.5 Guidelines for screening and early detection

Recommendations Strength rating


Do not subject men to prostate-specific antigen (PSA) testing without counselling them on Strong
the potential risks and benefits.
Offer an individualised risk-adapted strategy for early detection to a well-informed man and Weak
a life-expectancy of at least 10 to 15 years.
Offer early PSA testing to well-informed men at elevated risk of having PCa: Strong
• men from 50 years of age;
• men from 45 years of age and a family history of PCa;
• men of African descent from 45 years of age;
• men carrying BRCA2 mutations from 40 years of age.
Offer a risk-adapted strategy (based on initial PSA level), with follow-up intervals of 2 years Weak
for those initially at risk:
• men with a PSA level of > 1 ng/mL at 40 years of age;
• men with a PSA level of > 2 ng/mL at 60 years of age;
Postpone follow-up to 8 years in those not at risk.
Stop early diagnosis of PCa based on life expectancy and performance status; men who Strong
have a life-expectancy of < 15 years are unlikely to benefit.

5.2 Clinical diagnosis


Prostate cancer is usually suspected on the basis of DRE and/or PSA levels. Definitive diagnosis depends on
histopathological verification of adenocarcinoma in prostate biopsy cores.

5.2.1 Digital rectal examination


Most PCas are located in the peripheral zone and may be detected by DRE when the volume is > 0.2 mL.
In ~18% of cases, PCa is detected by suspect DRE alone, irrespective of PSA level [153]. A suspect DRE in
patients with a PSA level < 2 ng/mL has a positive predictive value (PPV) of 5–30% [154]. An abnormal DRE is
associated with an increased risk of a higher ISUP grade and is an indication for biopsy [155, 156].

5.2.2 Prostate-specific antigen


The use of PSA as a serum marker has revolutionised PCa diagnosis [157]. Prostate-specific antigen is organ
but not cancer specific; therefore, it may be elevated in benign prostatic hypertrophy (BPH), prostatitis and
other non-malignant conditions. As an independent variable, PSA is a better predictor of cancer than either
DRE or transrectal ultrasound (TRUS) [158].

There are no agreed standards defined for measuring PSA [159]. It is a continuous parameter, with higher
levels indicating greater likelihood of PCa. Many men may harbour PCa despite having low serum PSA [160].
Table 5.3 demonstrates the occurrence of ISUP > grade 2 PCa at low PSA levels, precluding an optimal PSA
threshold for detecting non-palpable but csPCa. The use of nomograms may help in predicting indolent PCa
[161].

PROSTATE CANCER - LIMITED UPDATE 2021 25


Table 5.3: Risk of PCa identified by systemic PCa biopsy in relation to low PSA values [160]

PSA level (ng/mL) Risk of PCa (%) Risk of ISUP grade > 2 PCa (%)
0.0–0.5 6.6 0.8
0.6–1.0 10.1 1.0
1.1–2.0 17.0 2.0
2.1–3.0 23.9 4.6
3.1–4.0 26.9 6.7

5.2.2.1 PSA density


Prostate-specific antigen density is the level of serum PSA divided by the prostate volume. The higher the PSA
density, the more likely it is that the PCa is clinically significant (see Section 6.2.1 - Treatment of low-risk disease).

5.2.2.2 PSA velocity and doubling time


There are two methods of measuring PSA kinetics:
• PSA velocity (PSAV): absolute annual increase in serum PSA (ng/mL/year) [162];
• PSA doubling time (PSA-DT): which measures the exponential increase in serum PSA over time [163].

Prostate-specific antigen velocity and PSA-DT may have a prognostic role in treating PCa but have limited
diagnostic use because of background noise (total prostate volume, and BPH), different intervals between PSA
determinations, and acceleration/deceleration of PSAV and PSA-DT over time [164]. These measurements do
not provide additional information compared with PSA alone [165-168].

Prostate-specific antigen kinetics is mainly exponential, especially during relapse. Two methods have been
described to calculate PSA-DT: the 2 -points and the log-slope method using more PSA measurements
[169]. The latter is considered more accurate as it eliminates inter- and intra-essay variations as well as any
physiological variations and potential imprecise measurements of low values [170, 171]. To obtain a reliable
evaluation, at least 3 PSA measurements must be available, collected four weeks apart as a minimum from the
same clinical laboratory [164]. A minimum increase of between 0.2 and 0.4 ng/mL is required [172].

As PSA-DT values may fluctuate over time in some men, only values obtained over the past 12 months should
be considered to assess disease course [164]. The Memorial Sloan Kettering Cancer Center PSA Doubling
Time calculator is one of the most widely used tools: https://www.mskcc.org/nomograms/prostate/psa_
doubling_time.

5.2.2.3 Free/total PSA ratio


Free/total (f/t) PSA must be used cautiously because it may be adversely affected by several pre-analytical and
clinical factors (e.g., instability of free PSA at 4°C and room temperature, variable assay characteristics, and
concomitant BPH in large prostates) [173]. Prostate cancer was detected in men with a PSA 4–10 ng/mL by
biopsy in 56% of men with f/t PSA < 0.10, but in only 8% with f/t PSA > 0.25 ng/mL [174]. A systematic review
including 14 studies found a pooled sensitivity of 70% in men with a PSA of 4–10 ng/mL [175]. Free/total PSA
is of no clinical use if the total serum PSA is > 10 ng/mL or during follow-up of known PCa. The clinical value of
f/t PSA is limited in light of the new diagnostic pathways incorporating MRI (see Section 5.2.4.2).

5.2.3 Biomarkers in prostate cancer


5.2.3.1 Blood based biomarkers: PHI/4K score/IsoPSA
Several assays measuring a panel of kallikreins in serum or plasma are now commercially available, including
the U.S. Food and Drug Administration (FDA) approved Prostate Health Index (PHI) test (combining free and
total PSA and the [-2]pro-PSA isoform [p2PSA]), and the four kallikrein (4K) score test (measuring free, intact
and total PSA and kallikrein-like peptidase 2 [hK2] in addition to age, DRE and prior biopsy status). Both tests
are intended to reduce the number of unnecessary prostate biopsies in PSA-tested men. A few prospective
multi-centre studies demonstrated that both the PHI and 4K score test out-performed f/t PSA PCa detection,
with an improved prediction of csPCa in men with a PSA between 2–10 ng/mL [176-179]. In a head-to-head
comparison both tests performed equally [180].
In contrast to the 4K score and PHI, which focus on the concentration of PSA isoforms, IsoPSA
utilises a novel technology which focuses on the structure of PSA [181]. Using an aqueous two-phase
solution, it partitions the isoforms of PSA and assesses for structural changes in PSA. In a recent multi-centre
prospective validation in 271 men the assay AUC was 0.784 for high-grade vs. low-grade cancer/benign
histology, which was superior to the AUCs of total PSA and percent free PSA [182].

26 PROSTATE CANCER - LIMITED UPDATE 2021


5.2.3.2 Urine biomarkers: PCA3/SelectMDX/Mi Prostate score (MiPS)/ExoDX
Prostate cancer gene 3 (PCA3) is a prostate-specific, non-coding microRNA (mRNA) biomarker that is
detectable in urine sediments obtained after three strokes of prostatic massage during DRE. The commercially
available Progensa urine test for PCA3 is superior to total and percent-free PSA for the detection of PCa in
men with elevated PSA as it shows significant increases in the area under the receiver-operator characteristic
curve (AUC) for positive biopsies [183-186]. PCA3 score increases with PCa volume, but there are conflicting
data about whether it independently predicts the ISUP grade [187]. Currently, the main indication for the
Progensa test is to determine whether repeat biopsy is needed after an initially negative biopsy, but its clinical
effectiveness for this purpose is uncertain [188]. Wei et al. showed 42% sensitivity at a cut-off of 60 in the
primary biopsy setting with a high specificity (91%) and a PPV of 80% suggesting that the assay may be used
in the primary setting [189].
The SelectMDX test is similarly based on mRNA biomarker isolation from urine. The presence of
HOXC6 and DLX1 mRNA levels is assessed to provide an estimate of the risk of both presence of PCa on
biopsy as well as presence of high-risk cancer [190].
TMPRSS2-ERG fusion, a fusion of the trans-membrane protease serine 2 (TMPRSS2) and the ERG
gene can be detected in 50% of PCas [191]. When detection of TMPRSS2-ERG in urine was added to PCA3
expression and serum PSA (Mi(chigan)Prostate Score [MiPS]), cancer prediction improved [192]. Exosomes
secreted by cancer cells may contain mRNA diagnostic for high-grade PCa [193, 194]. Use of the ExoDx
Prostate IntelliScore urine exosome assay resulted in avoiding 27% of unnecessary biopsies when compared
to standard of care. However, currently, both the MiPS-score and ExoDx assay are considered investigational.

In 6 head-to-head comparison studies of PCA3 and PHI, only Seisen et al. found a significant difference;
PCA3 detected more cancers, but for the detection of significant disease, defined as ISUP grade > 2, more
than three positive cores, or > 50% cancer involvement in any core, PHI proved superior [195]. Russo et al.
suggested in their systematic review that, based on moderate quality data, PHI and the 4K panel had a high
diagnostic accuracy and showed similar performance in predicting the detection of significant disease with a
AUC of 0.82 and 0.81, respectively [196]. However, in the screening population of the ERSPC study the use
of both PCA3 and 4K panel when added to the risk calculator led to an improvement in AUC of less than 0.03
[197]. Based on the available evidence, some biomarkers could help in discriminating between aggressive and
non-aggressive tumours with an additional value compared to the prognostic parameters currently used by
clinicians [198]. However, upfront multiparametric magnetic resonance imaging (mpMRI) is also likely to affect
the utility of above-mentioned biomarkers (see Section 5.2.4).

5.2.3.3 Tests to select men for a repeat biopsy


In men with an elevated risk of PCa with a prior negative biopsy, additional information may be gained by
the Progensa-PCA3 and SelectMDX DRE urine tests, the serum 4Kscore and PHI tests or a tissue-based
epigenetic test (ConfirmMDx). The role of PHI, Progensa PCA3, and SelectMDX in deciding whether to take a
repeat biopsy in men who had a previous negative biopsy is uncertain and probably not cost-effective [188].
The ConfirmMDx test is based on the concept that benign prostatic tissue in the vicinity of a PCa focus shows
distinct epigenetic alterations. In case PCa is missed at biopsy, demonstration of epigenetic changes in the
benign tissue would indicate the presence of carcinoma. The ConfirmMDX test quantifies the methylation
level of promoter regions of three genes (Methylated APC, RASSF1 and GSTP1) in benign prostatic tissue.
A multi-centre study found a NPV of 88% when methylation was absent in all three markers, implying that a
repeat biopsy could be avoided in these men [199]. Given the limited available data and the fact that the role
of mpMRI in tumour detection was not accounted for, no recommendation can be made regarding the routine
application of ConfirmMDX, in particular in the light of current use of mpMRI before biopsy. Table 5.4 presents
an overview of the most commonly available commercial tests.

Table 5.4: Description of additional investigational tests after a negative prostate biopsy*

Name of test Test substrate Molecular FDA approved


Progensa DRE urine lncRNA PCA3 Yes
SelectMDX DRE urine mRNA HOXC6, DLX1 No
PHI Serum Total, free and p2PSA Yes
4Kscore Test Serum/plasma Total, free, intact PSA, hK2 No
ConfirmMDX Benign prostate biopsy Methylated APC, RASSF1 and GSTP1 No
*Isolated high-grade prostatic intraepithelial neoplasia (PIN) in one or two biopsy sites is no longer an indication
for repeat biopsy [200].

PROSTATE CANCER - LIMITED UPDATE 2021 27


5.2.3.4 Guidelines for risk-assessment of asymptomatic men

Recommendations Strength rating


In asymptomatic men with a prostate-specific antigen level between 2–10 ng/mL and a Strong
normal digital rectal examination, use one oft the following tools for biopsy indication:
• risk-calculator;
• imaging;
• an additional serum, urine or tissue-based test. Weak

5.2.4 The role of imaging in clinical diagnosis


5.2.4.1 Transrectal ultrasound and ultrasound-based techniques
Standard TRUS is not reliable at detecting PCa [201] and the diagnostic yield of additional biopsies performed
on hypoechoic lesions is negligible [202]. Prostate HistoScanningTM provided inconsistent results across
studies [203]. New sonographic modalities such as micro-Doppler, sonoelastography contrast-enhanced US
or high-resolution micro-ultrasound provided promising preliminary findings, either alone, or combined with
so-called ‘multiparametric US’. However, these techniques still have limited clinical appllicability due to lack of
standardisation, lack of large-scale evaluation of inter-reader variability and unclear results in transition zones
[204-206].

5.2.4.2 Multiparametric magnetic resonance imaging


5.2.4.2.1 Multiparametric magnetic resonance imaging performance in detecting PCa
Correlation with RP specimens shows that MRI has good sensitivity for the detection and localisation of ISUP
grade > 2 cancers, especially when their diameter is larger than 10 mm [207-209]. This good sensitivity was
further confirmed in patients who underwent template biopsies. In a recent Cochrane meta-analysis which
compared MRI to template biopsies (> 20 cores) in biopsy-naïve and repeat-biopsy settings, MRI had a pooled
sensitivity of 0.91 (95% CI: 0.83–0.95) and a pooled specificity of 0.37 (95% CI: 0.29–0.46) for ISUP grade > 2
cancers [210]. For ISUP grade > 3 cancers, MRI pooled sensitivity and specificity were 0.95 (95% CI: 0.87–
0.99) and 0.35 (95% CI: 0.26–0.46), respectively.
MRI is less sensitive in identifying ISUP grade 1 PCa. It identifies less than 30% of ISUP grade 1
cancers smaller than 0.5 cc identified on RP specimens by histopathology analysis [207]. In series using
template biopsy findings as the reference standard, MRI has a pooled sensitivity of 0.70 (95% CI: 0.59–0.80)
and a pooled specificity of 0.27 (95% CI: 0.19–0.37) for identifying ISUP grade 1 cancers [210].
The probability of detecting malignancy by MRI-identified lesions was standardised first by the
use of a 5-grade Likert score [211], and then by the Prostate Imaging – Reporting and Data System (PI-RADS)
score which has been updated several times since its introduction [212, 213]. In a meta-analysis of 13 studies
involving men with suspected or biopsy-proven PCa, the average PPVs for ISUP grade > 2 cancers of lesions
with a PI-RADSv2 score of 3, 4 and 5 were 12%, 48% and 72% respectively, but with significant heterogeneity
among studies (Table 5.5) [214]. Another retrospective study of 3,449 men with suspected or biopsy-proven
untreated PCa imaged across 26 academic centres found similar results with lesions of PI-RADS v2 scores of
3, 4 and 5 having a PPV for ISUP grade > 2 cancers of 15% (95% CI: 11–19), 39% (95% CI: 34–45) and 72%
(95% CI: 61–82), respectively [215].

Table 5.5: Proportion of malignant MRI lesions by ISUP grade groups and PI-RADS v2 scores [214]

PI-RADS v2 Total number ISUP 1 ISUP 2 ISUP 3 ISUP > 4


of lesions
3 707 14% (9.4–18.7) 9.3% (4.3–14.1) 1.5% (0.05–3) 0.7% (0–1.6)
4 886 21% (13.0–28.9 29.7% (13.9–45.5) 7.7% (3.4–12) 10.8% (5.7–15.9)
5 495 12% (5.3–18.7) 33.5% (8.0–59.0) 15.7% (6.4–25.1) 23% (8.2–37.9)
Intervals in parenthesis are 95% CIs.

 argeted biopsy improves the detection of ISUP grade > 2 cancer as compared to systematic
5.2.4.2.2 T
biopsy.
In pooled data of 25 reports on agreement analysis (head-to-head comparisons) between systematic biopsy
(median number of cores, 8–15) and MRI-targeted biopsies (MRI-TBx; median number of cores, 2–7), the
detection ratio (i.e. the ratio of the detection rates obtained by MRI-TBx alone and by systematic biopsy alone)
was 1.12 (95% CI: 1.02–1.23) for ISUP grade > 2 cancers and 1.20 (95% CI: 1.06–1.36) for ISUP grade > 3
cancers, and therefore in favour of MRI-TBx.

28 PROSTATE CANCER - LIMITED UPDATE 2021


However, the pooled detection ratios for ISUP grade > 2 cancers and ISUP grade > 3 cancers
were 1.44 (95% CI: 1.19–1.75) and 1.64 (95% CI: 1.27–2.11), respectively, in patients with prior negative
systematic biopsies, and only 1.05 (95% CI: 0.95–1.16) and 1.09 (95% CI: 0.94–1.26) in biopsy-naïve patients
[210]. Another meta-analysis of RCTs limited to biopsy-naïve patients with a positive MRI found that MRI-TBx
detected significantly more ISUP grade > 2 cancers than systematic biopsy (risk difference, -0.11 (95% CI: -0.2
to 0.0); p = 0.05), in prospective cohort studies (risk difference, -0.18 (95% CI: -0.24 to -0.11); p < 0.00001),
and in retrospective cohort studies (risk difference, -0.07 (95% CI: -0.12 to -0.02); p = 0.004). There was no
significant increase in the detection rate when systematic biopsies were combined to MRI-TBx [216].
Three prospective multi-centre trials evaluated MRI-TBx in biopsy-naïve patients. In the PRostate
Evaluation for Clinically Important Disease: Sampling Using Image-guidance Or Not? (PRECISION) trial, 500
biopsy-naïve patients were randomised to either MRI-TBx only or TRUS-guided systematic biopsy only. The
detection rate of ISUP grade > 2 cancers was significantly higher in men assigned to MRI-TBx (38%) than in
those assigned to SBx (26%, p = 0.005, detection ratio 1.46) [217]. In the Assessment of Prostate MRI Before
Prostate Biopsies (MRI-FIRST) trial, 251 biopsy-naïve patients underwent TRUS-guided systematic biopsy by
an operator who was blinded to MRI findings, and MRI-TBx by another operator. Magnetic resonance Imaging-
TBx detected ISUP grade > 2 cancers in a higher percentage of patients but the difference was not significant
(32.3% vs. 29.9%, p = 0.38; detection ratio: 1.08) [202]. However, MRI-TBx detected significantly more ISUP
grade > 3 cancers than systematic biopsy (19.9% vs. 15.1%, p = 0.0095; detection ratio: 1.32). A similar
trend for improved detection of ISUP grade > 3 cancers by MRI-TBx was observed in the Cochrane analysis;
however, it was not statistically significant (detection ratio 1.11 [0.88–1.40]) [210]. The Met Prostaat MRI Meer
Mans (4M) study included 626 biopsy-naïve patients; all patients underwent systematic biopsy, and those with
a positive MRI (PI-RADSv2 score of 3–5, 51%) underwent additional in-bore MRI-TBx. The results were close
to those of the MRI-FIRST trial with a detection ratio for ISUP grade > 2 cancers of 1.09 (detection rate: 25%
for MRI-TBx vs. 23% for systematic biopsy) [218]. However, in this study, MRI-TBx and systematic biopsy
detected an equal number of ISUP grade > 3 cancers (11% vs. 12%; detection ratio: 0.92).
The Target Biopsy Techniques Based on Magnetic Resonance Imaging in the Diagnosis of Prostate
Cancer in Patients with Prior Negative Biopsies (FUTURE) randomised trial compared three techniques of
MRI-TBx in the repeat-biopsy setting [219]. In the subgroup of 152 patients who underwent both MRI-TBx and
systematic biopsy, MRI-TBx detected significantly more ISUP grade > 2 cancers than systematic biopsy (34%
vs. 16%; p < 0.001, detection ratio of 2.1), which is a finding consistent with the Cochrane agreement analysis
(detection ratio: 1.44). An ISUP grade > 2 cancer would have been missed in only 1.3% (2/152) of patients,
had systematic biopsy been omitted [220]. These findings support that MRI-TBx significantly out-performs
systematic biopsy for the detection of ISUP grade > 2 in the repeat-biopsy setting. In biopsy-naïve patients, the
difference appears to be less marked but remains in favour of MRI-TBx.

5.2.4.2.3 M
 RI-TBx without systematic biopsy reduces the detection of ISUP grade 1 PCa as compared to
systematic biopsy.
In pooled data of 25 head-to-head comparisons between systematic biopsy and MRI-TBx, the detection ratio
for ISUP grade 1 cancers was 0.62 (95% CI: 0.44–0.88) in patients with prior negative biopsy and 0.63 (95%
CI: 0.54–0.74) in biopsy-naïve patients [210]. In the PRECISION and 4M trials, the detection rate of ISUP grade 1
patients was significantly lower in the MRI-TBx group as compared to systematic biopsy (9% vs. 22%,
p < 0.001, detection ratio of 0.41 for PRECISION; 14% vs. 25%, p < 0.001, detection ratio of 0.56 for 4M) [217,
218]. In the MRI-FIRST trial, MRI-TBx detected significantly fewer patients with clinically insignificant PCa
(defined as ISUP grade 1 and maximum cancer core length < 6 mm) than systematic biopsy (5.6% vs. 19.5%,
p < 0.0001, detection ratio of 0.29) [202]. Consequently, MRI-TBx without systematic biopsy significantly
reduces over-diagnosis of low-risk disease, as compared to systematic biopsy.

5.2.4.2.4 The added value of systematic and targeted biopsy


Magnetic resonance imaging-targeted biopsies can be used in two different diagnostic pathways: 1) the
‘combined pathway’, in which patients with a positive MRI undergo combined systematic and targeted biopsy,
and patients with a negative MRI undergo systematic biopsy; 2) the ‘MRI pathway’, in which patients with a
positive MRI undergo only MRI-TBx, and patients with a negative MRI who are not biopsied at all.
Choosing between these pathways depends not only on the detection rates obtained by the
two biopsy techniques, but also on whether or not they detect the same patients. Many studies evaluated
combined systematic and targeted biopsy in the same patients and could therefore assess the absolute added
value of each technique (i.e. the percentage of patients diagnosed by only one biopsy technique). Data from
the Cochrane meta-analysis of these studies and from the MRI-FIRST and 4M trials suggest that the absolute
added value of MRI-TBx for detecting ISUP grade > 2 cancers is higher than that of systematic biopsy (see
Table 5.6).

PROSTATE CANCER - LIMITED UPDATE 2021 29


 bsolute added values of targeted and systematic biopsies for ISUP grade > 2 and > 3 cancer
Table 5.6: A
detection

ISUP > 2 ISUP > 3


ISUP grade Cochrane MRI-FIRST 4M trial Cochrane MRI-FIRST 4M trial
meta- trial* [202] [218] meta- trial* [202] [218]
analysis* analysis*
[210] [210]
Added value of 6.3% 7.6% 7.0% (ND) 4.7% 6.0% 3.2% (ND)
MRI-TBx (4.8-8.2) (4.6-11.6) (3.5-6.3) (3.4-9.7)
Added value 4.3% 5.2% 5.0% (ND) 2.8% 1.2% 4.1% (ND)
Biopsy-naïve of systematic (2.6-6.9) (2.8-8.7) (1.7-4.8) (0.2-3.5)
biopsy
Overall 27.7% 37.5% 30% (ND) 15.5% 21.1% 15% (ND)
prevalence (23.7-32.6) (31.4-43.8) (12.6-19.5) (16.2-26.7)
Added value of 9.6% - - 6.3% - -
MRI-TBx (7.7-11.8) (5.2-7.7)
Added value 2.3% - - 1.1% - -
Prior negative
of systematic (1.2-4.5) (0.5-2.6)
biopsy
biopsy
Overall 22.8% - - 12.6% - -
prevalence (20.0-26.2) (10.5-15.6)
*Intervals in parenthesis are 95% CI.
The absolute added value of a given biopsy technique is defined by the percentage of patients of the entire
cohort diagnosed only by this biopsy technique.
ISUP = International Society for Urological Pathology (grade); MRI-TBx = magnetic resonance imaging-targeted
biopsies; ND = not defined.

In Table 5.6, the absolute added values refer to the percentage of patients in the entire cohort; if the cancer
prevalence is taken into account, the ‘relative’ percentage of additional detected PCa can be computed.
Adding MRI-TBx to systematic biopsy in biopsy-naïve patients increases the number of ISUP grade > 2 and
grade > 3 PCa by approximately 20% and 30%, respectively. In the repeat-biopsy setting, adding MRI-TBx
increases detection of ISUP grade > 2 and grade > 3 PCa by approximately 40% and 50%, respectively.
Omitting systematic biopsy in biopsy-naïve patients would miss approximately 16% of ISUP grade > 2 PCa
and 18% of ISUP grade > 3 PCa. In the repeat-biopsy setting, it would miss approximately 10% of ISUP grade
> 2 PCa and 9% of ISUP grade > 3 PCa.

5.2.4.2.5 Number of biopsy procedures potentially avoided in the ‘MR pathway’


The diagnostic yield and number of biopsy procedures potentially avoided by the ‘MR pathway’ depends on
the Likert/PI-RADS threshold used to define positive MRI. In pooled studies on biopsy-naïve patients and
patients with prior negative biopsies, a Likert/PI-RADS threshold of > 3 would have avoided 30% (95% CI:
23–38) of all biopsy procedures while missing 11% (95% CI: 6–18) of all detected ISUP grade > 2 cancers
(relative percentage) [210]. Increasing the threshold to > 4 would have avoided 59% (95% CI: 43–78) of all
biopsy procedures while missing 28% (95% CI: 14–48) of all detected ISUP grade > 2 cancers [210]. Of note,
the percentages of negative MRI (Likert/PI-RADS score < 2) in the MRI-FIRST, PRECISION and 4M trials were
21.1%, 28.9% and 49%, respectively [202, 217, 218].

5.2.4.2.6 Other considerations


5.2.4.2.6.1 Multiparametric magnetic resonance imaging reproducibility
Despite the use of the PI-RADSv2 scoring system [212], MRI inter-reader reproducibility remains moderate at
best which currently limits its broad use by non-dedicated radiologists [221]. However, significant improvement
in the accuracy of MRI and MRI-TBx can be observed over time, both in academic and community hospitals,
especially after implementation of PI-RADSv2 scoring and multidisciplinary meetings using pathological
correlation and feedback [222-225]. An updated version of the PI-RADS score (PI-RADSv2.1) has been
recently published to improve reader reproducibility but it has not yet been fully evaluated [213]. It is still too
early to predict whether quantitative approaches and computer-aided diagnostic systems will improve the
characterisation of lesions seen at MRI [226-228].

30 PROSTATE CANCER - LIMITED UPDATE 2021


5.2.4.2.6.2 Targeted biopsy accuracy and reproducibility
Clinically significant PCa not detected by the ‘MRI pathway’ can be missed because of MRI failure (invisible
cancer or reader’s misinterpretation) or because of targeting failure (target missed or undersampled by MRI-
TBx). In two retrospective studies of 211 and 116 patients with a unilateral MRI lesion, targeted biopsy alone
detected 73.5–85.5% of all csPCa (ISUP grade > 2); combining MRI-TBx with systematic biopsy of the lobe
with the MRI lesion detected 96–96.4% of all csPCas and combined targeted and systematic biopsy of
the contralateral lobe only identified 81.6–92.7% of csPCas [229, 230]. The difference may reflect targeting
errors leading to undersampling of the tumour. The accuracy of MRI-TBx is indeed substantially impacted
by the experience of the biopsy operator [221]. Increasing the number of cores taken per target may partially
compensate for guiding imprecision. In a retrospective study of 479 patients who underwent MRI-TBx with 4
cores per target that were sequentially labelled, the first 3 cores detected 95.1% of the csPCas detected by
the 4-core strategy [231]. In two other retrospective studies of 330 and 744 patients who underwent MRI-TBx
with up to 5 cores per target, the one-core and 3-core sampling strategies detected respectively 63–75% and
90–93% of the ISUP grade > 2 PCa detected by the 5-core strategy [232, 233]. These percentages are likely to
be influenced by the lesion size and location, the prostate volume or the operator’s experience, but no study
has quantified the impact of these factors yet.

5.2.4.2.6.3 Role of risk-stratification


Using risk-stratification to avoid biopsy procedures
Prostate-specific antigen density (PSAD) may help refine the risk of csPCa in patients undergoing MRI as
PSAD and the PI-RADS score are significant independent predictors of csPCa at biopsy [234, 235]. In a
meta-analysis of 8 studies, pooled MRI NPV for ISUP grade > 2 cancer was 84.4% (95% CI: 81.3–87.2) in the
wole cohort, 82.7% (95% CI: 80.5–84.7) in biopsy-naïve men and 88.2% (95% CI: 85–91.1) in men with prior
negative biopsies. In the subgroup of patients with PSAD < 0.15 ng/ml, NPV increased to respectively 90.4%
(95% CI: 86.8–93.4), 88.7% (95% CI: 83.1–93.3) and 94.1% (95% CI: 90.9–96.6) [236]. In contrast, the risk of
csPCa is as high as 27–40% in patients with negative MRI and PSAD > 0.15–0.20 ng/mL/cc [218, 235, 237-
240] (Table 5.7).
Combining MRI findings with the PCA3 score may also improve risk stratification [241]. Several
groups have developed comprehensive risk calculators which combine MRI findings with simple clinical data
as a tool to predict subsequent biopsy results [242]. At external validation, they tended to outperform risk
calculators not incorporating MRI findings (ERSPC and PBCG) with good discriminative power (as measured
by the AUC). However, they also tended to be miscalibrated with under- or over-prediction of the risk of ISUP
grade > 2 cancer [243, 244]. In one study that externally assessed four risk calculators combining MRI findings
and clinical data, only two demonstrated a distinct net benefit when a risk of false-negative prediction of 15%
was accepted. The others were harmful for this risk level, as compared to the ‘biopsy all’ strategy [243]. This
illustrates the prevalence-dependence of risk models. Recalibrations taking into account the local prevalence
are possible, but this approach is difficult in routine clinical practice as the local prevalence is difficult to
estimate and may change over time.

Using risk-stratification to avoid MRI and biopsy procedures


A retrospective analysis including 200 men from a prospective database of patients who underwent MRI and
combined systematic and targeted biopsy showed that upfront use of the Rotterdam Prostate Cancer Risk
Calculator (RPCRC) would have avoided MRI and biopsy in 73 men (37%). Of these 73 men, 10 had ISUP
grade 1 cancer and 4 had ISUP grade > 2 cancer [245]. A prospective multi-centre study evaluated several
diagnostic pathways in 545 biopsy-naïve men who underwent MRI and systematic and targeted biopsy. Using
a PHI threshold of > 30 to perform MRI and biopsy would have avoided MRI and biopsy in 25% of men at the
cost of missing 8% of the ISUP grade > 2 cancers [246]. Another prospective multi-centre trial including 532
men (with or without history of prostate biopsy) showed that using a threshold of > 10% for the Stockholm3
test to perform MRI and biopsy would have avoided MRI and biopsy in 38% of men at the cost of missing 8%
of ISUP grade > 2 cancers [247].

PROSTATE CANCER - LIMITED UPDATE 2021 31


Table 5.7: Impact of the PSA density on csPCa detection rates in patients with negative MRI findings

Study Study design Population Biopsy protocol csPCa csPCa detection rate
definition
Washino, Retrosp. n = 288 SBx (14 cores) + ISUP > 2 or Whole cohort (prevalence):
et al. 2017 Single-centre Biopsy naive cognitive TBx MCCL > 4 mm 49%
[235]
PI-RADS 1-2:
0% if PSAD < 0.15,
20% if PSAD = 0.15-0.29,
30% if PSAD > 0.3
Distler, Retrosp. n = 1,040 TTP (24 cores) + ISUP > 2 Whole cohort (prevalence):
et al. 2017 analysis of Biopsy fusion TBx 43%
[234] prospective naive + prior
database negative PI-RADS 1-2 / Whole
Single-centre biopsy cohort: 11% if PSAD < 0.15,
33% if PSAD > 0.15

PI-RADS 1-2 / prior


negative biopsy:
7% if PSAD < 0.15,
27% if PSAD > 0.15
Hansen, Retrosp. n = 514 TTP (24 cores) + ISUP > 2 Whole cohort (prevalence):
et al. 2017 Single-centre Prior negative fusion TBx 31%
[237] biopsy or AS
for ISUP 1 Likert 1-2:
PCa 9% if PSAD < 0.10,
9% if PSAD < 0.2,
29% if PSAD > 0.2
Hansen, Retrosp. n = 807 TTP + cognitive ISUP > 2 Whole cohort (prevalence):
et al. 2018 Multi-centre Biopsy naive or fusion TBx 49%
[238]
PI-RADS 1-2:
10% if PSAD < 0.10,
21% if PSAD = 0.1-0.2,
33% if PSAD > 0.2
Oishi, Retrosp. n = 135 SBx (12 cores) ISUP > 2 Whole cohort (prevalence):
et al. 2019 analysis of Biopsy 18%
[239] prospective naive + prior
database negative PSAD < 0.10: 6% (overall
Single-centre biopsy + AS pop), 15% (biopsy naïve),
+ Restaging 0% (prior negative biopsy)
Only pts with PSAD < 0.15: 10% (overall
negative MRI pop), 20% (biopsy naïve),
(PI-RADS < 2) 0% (prior negative biopsy)
PSAD > 0.15: 40% (overall
pop), 29% (biopsy naïve),
29% (prior negative biopsy)
Boesen, Retrosp. n = 808 SBx (10 cores) + ISUP > 2 Whole cohort (prevalence):
et al. 2019 analysis of Biopsy-naive fusion TBx 35%
[240] prospective
database PI-RADS 1-2:
Single-centre 3% if PSAD < 0.10,
5% < 0.15,
5% if PSAD < 0.2,
32% if PSAD > 0.2

32 PROSTATE CANCER - LIMITED UPDATE 2021


Van der Prospective n = 626 TTP Bx (median ISUP > 2 Whole cohort (prevalence):
Leest, et al. Multi-centre Biopsy naive 24 cores) + 30%
2019 [218] fusion TBx
PI-RADS 1-2:
1.3% if PSAD < 0.10,
2% if PSAD < 0.15
csPCa = clinically significant prostate cancer; Retrosp. = retrospective; n = number of patients;
SBx = systematic transrectal biopsy; TBx = targeted biopsy; TTP = transperineal template biopsy;
PSAD = PSA density; ISUP = international Society of Urological Pathology; MCCL = maximum cancer core length.

5.2.4.2.6.4 Pre-biopsy MRI and MRI-TBx may induce an ISUP prostate cancer grade shift, as a result of
improved diagnosis
Magnetic resonance imaging findings are significant predictors of adverse pathology features on prostatectomy
specimens, and of survival-free BCR after RP or RT [82, 248-250]. In addition, tumours visible on MRI are
enriched in molecular hallmarks of aggressivity, as compared to invisible lesions [251]. Thus, MRI does identify
aggressive tumours.
Nonetheless, improved targeting obtained by MRI-TBx can artificially inflate the ISUP grade of
the tumours by focusing at the areas of high-grade cancer. As a result, ISUP grade > 2 cancers detected by
MRI-TBx are, on average, of better prognosis than those detected by the classical diagnostic pathway. This
is illustrated in a retrospective series of 1,345 patients treated by RP which showed that, in all risk groups,
patients diagnosed by MRI-TBx had better BCR-free survival than those diagnosed by systematic biopsy
only [82]. To mitigate this grade shift, in case of targeted biopsies, the 2019 ISUP consensus conference
recommended using an aggregated ISUP grade summarizing the results of all biopsy cores from the same MR
lesion, rather than using the result from the core with the highest ISUP grade [85]. When long-term follow-up
of patients who underwent MRI-TBx is available, a revision of the risk-groups definition will become necessary.
In the meantime, results of MRI-TBx must be interpreted in the context of this potential grade shift as an
increasing proportion of patients diagnosed with ISUP grade 2 cancers by MRI-TBx might be eligible for AS
[252].

5.2.4.2.7 Summary of evidence and practical considerations on pre-biopsy MRI and MRI-TBx
Pre-biopsy MRI and MRI-TBx substantially improve the detection of ISUP grade > 2 PCa. This improvement is
most notable in the repeat-biopsy setting, with marginal added value for systematic biopsies. It is less marked
in biopsy-naïve patients in whom systematic biopsy retains a higher added value, at least for the detection of
ISUP grade 2 cancers. MRI-TBx also detects significantly less ISUP grade 1 cancers than systematic biopsies.
The ‘MRI pathway’ is appealing since it could decrease the number of biopsy procedures, reduce
the detection of low-grade PCa while maintaining (or even improving) the detection of csPCa, as compared
to systematic biopsy. However, some cavaets need pointing out. First, MRI findings must be interpreted in
the light of the a priori risk of csPCa. Risk stratification combining clinical data, MRI findings and (maybe)
other biomarkers will help, in the future, defining those patients that can safely avoid biopsy. Second, without
standardisation of MRI interpretation and of MRI-TBx technique the ‘MR pathway’ may lead to suboptimal
care outside large-volume (expert) centres. Indeed, the inter-reader reproducibility of MRI is moderate at
best. Current biopsy-targeting methods remain imprecise and their accuracy is substantially impacted by the
operator’s experience. As a result, 3 to 5 biopsy cores per target may be needed to reduce the risk of missing
or undersampling the lesion, even with US/MR fusion systems. Third, the use of pre-biopsy MRI may induce
grade shift, even with the use of an aggregated ISUP grade for each MR lesion targeted at biopsy. Clinicians
must interpret MRI-TBx results in the context of this potential grade shift. A revision of the definitions of the risk
groups will be needed in the future to take into account wider use of MRI and MRI-TBx.
Finally, it must be emphasised that the ‘MR pathway’ has only been evaluated in patients in whom
the risk of csPCa was judged high enough to deserve biopsy. Pre-biopsy MRI must not be used in patients
who do not have an indication for prostate biopsy based on their family history or clinical and biochemical data.
Because of its low specificity, MRI in very low-risk patients would result in an inflation of false-positive findings
and subsequent unnecessary biopsies.

5.2.4.3 Guidelines for imaging in PCa detection

Introductory statement LE
Systematic biopsy is an acceptable approach in case magnetic resonance imaging (MRI) is 3
unavailable.

PROSTATE CANCER - LIMITED UPDATE 2021 33


Recommendations for all patients Strength rating
Do not use multiparametric magnetic resonance imaging (mpMRI) as an initial screening tool. Strong
Adhere to PI-RADS guidelines for mpMRI acquisition and interpretation and evaluate Strong
mpMRI results in multidisciplinary meetings with pathological feedback.

Recommendations in biopsy naïve patients Strength rating


Perform mpMRI before prostate biopsy. Strong
When mpMRI is positive (i.e. PI-RADS > 3), combine targeted and systematic biopsy. Strong
When mpMRI is negative (i.e., PI-RADS < 2), and clinical suspicion of PCa is low, omit Weak
biopsy based on shared decision-making with the patient.

Recommendations in patients with prior negative biopsy Strength rating


Perform mpMRI before prostate biopsy. Strong
When mpMRI is positive (i.e. PI-RADS > 3), perform targeted biopsy only. Weak
When mpMRI is negative (i.e., PI-RADS < 2), and clinical suspicion of PCa is high, perform Strong
systematic biopsy based on shared share decision-making with the patient.

5.2.5 Baseline biopsy


The need for prostate biopsy is based on PSA level, other biomarkers and/or suspicious DRE and/or imaging
(see Section 5.2.4). Age, potential co-morbidity and therapeutic consequences should also be considered and
discussed beforehand [253]. Risk stratification is a potential tool for reducing unnecessary biopsies [253].
Limited PSA elevation alone should not prompt immediate biopsy. Prostate-specific antigen
level should be verified after a few weeks, in the same laboratory using the same assay under standardised
conditions (i.e. no ejaculation, manipulations, and urinary tract infections [UTIs]) [254, 255]. Empiric use of
antibiotics in an asymptomatic patient in order to lower the PSA should not be undertaken [256].
Ultrasound (US)-guided biopsy is now the standard of care. Prostate biopsy is performed by either
the transrectal or transperineal approach. Cancer detection rates, when performed without prior imaging with
MRI, are comparable between the two approaches [257], however, some evidence suggests reduced infection
risk with the transperineal route (see Section 5.2.6.4) [258, 259].
Transurethral resection of the prostate (TURP) should not be used as a tool for cancer detection [260].

5.2.6 Repeat biopsy


5.2.6.1 Repeat biopsy after previously negative biopsy
The indications for repeat biopsy are:
• rising and/or persistently elevated PSA (see Table 5.3 for risk estimates);
• suspicious DRE, 5–30% PCa risk [153, 154];
• intraductal carcinoma as a solitary finding, > 90% risk of associated high-grade PCa [261];
• positive mpMRI findings (see Section 5.2.4.2).

The recommendation to perform a repeat biopsy after a diagnosis of atypical small acinar proliferation and
extensive high-grade PIN is based on earlier studies on systematic biopsies using a 6–10 core biopsy protocol.
In a contemporary series of biopsies the likelihood of finding a csPCa after follow-up biopsy after a diagnosis
of atypical small acinar proliferation was only 6% [262].
The added value of other biomarkers remains unclear (see Sections 5.2.3.1 and 5.2.3.2).

5.2.6.2 Saturation biopsy


The incidence of PCa detected by saturation repeat biopsy (> 20 cores) is 30–43% and depends on the
number of cores sampled during earlier biopsies [263]. Saturation biopsy may be performed with the
transperineal technique, which detects an additional 38% of PCa. The rate of urinary retention varies
substantially from 1.2% to 10% [264-267].

5.2.7 Prostate biopsy procedure


5.2.7.1 Sampling sites and number of cores
On baseline biopsies, where no prior imaging with mpMRI has been performed, or where mpMRI has not
shown any suspicious lesion, the sample sites should be bilateral from apex to base, as far posterior and lateral
as possible in the peripheral gland. Additional cores should be obtained from suspect areas identified by DRE;
suspect areas on TRUS might be consideration for additional biopsies. Sextant biopsy is no longer considered
adequate. At least 8 systematic biopsies are recommended in prostates with a size of about 30 cc [268]. Ten to

34 PROSTATE CANCER - LIMITED UPDATE 2021


12 core biopsies are recommended in larger prostates, with > 12 cores not being significantly more conclusive
[269, 270].
Eighteen to 24 template cores may be taken in transperineal biopsy with acceptable morbidity, but it
is unknown whether this number improves detection of significant cancer [271-273]. As per transrectal biopsy,
for maximal detection of significant cancer, cores should be directed towards the peripheral zone posteriorly
and laterally, but in transperineal biopsy can also more easily be directed to the anterior horns of the peripheral
zone as well. In the setting of a positive MRI with targeted biopsy cores being taken, the addition of template
cores may increase the detection of significant cancer slightly, but also increases the detection of insignificant
cancer [274, 275]. The optimal number of template cores in this setting is unknown.

Where mpMRI has shown a suspicious lesion MR-TBx can be obtained through cognitive guidance, US/MR
fusion software or direct in-bore guidance. Current literature, including systematic reviews and meta-analyses,
does not show a clear superiority of one image-guided technique over another [219, 276-279]. However,
regarding approach, the only systematic review and meta-analysis comparing MRI-targeted transrectal biopsy
to MRI-targeted transperineal biopsy, analysing 8 studies, showed a higher sensitivity for detection of csPCa
when the transperineal approach was used (86% vs. 73%) [280]. This benefit was especially pronounced for
anterior tumours. Multiple (3–5) cores should be taken from each lesion (see Section 5.2.4.2.7.2).

5.2.7.2 Antibiotics prior to biopsy


5.2.7.2.1 Transperineal prostate biopsy
A total of seven randomised studies including 1,330 patients compared the impact of biopsy route on
infectious complications. Infectious complications were significantly higher following transrectal biopsy (37
events among 657 men) compared to transperineal biopsy (22 events among 673 men) (RR: 1.81 [range
1.09–3.00], 95% CI) [281-288]. In addition, a systematic review including 165 studies with 162,577 patients
described sepsis rates of 0.1% and 0.9% for transperineal and transrectal biopsies, respectively [289]. Finally,
a population-based study from the UK (n = 73,630) showed lower re-admission rates for sepsis in patients
who had transperineal vs. transrectal biopsies (1.0% vs. 1.4%, respectively) [290]. The available evidence
demonstrates that the transrectal approach should be abandoned in favour of the transperineal approach
despite any possible logistical challenges.

To date, no RCT has been published investigating different antibiotic prophylaxis regimens for transperineal
prostate biopsy. However, as it is a clean procedure that avoids rectal flora, quinolones or other antibiotics
to cover rectal flora may not be necessary. A single dose of cephalosporin only to cover skin commensals
has been shown to be sufficient in multiple single cohort series [267, 291]. Prior negative mid-stream urine
(MSU) test and routine surgical disinfecting preparation of the perineal skin are mandatory. In one of the
largest studies to date, 1,287 patients underwent transperineal biopsy under local anaesthesia only [292].
Antibiotic prophylaxis consisted of a single oral dose of either cefuroxime or cephalexin. Patients with cardiac
valve replacements received amoxycillin and gentamicin, and those with severe penicillin allergy received
sulphamethoxazole. No quinolones were used. Only one patient developed a UTI with positive urine culture and
there was no urosepsis requiring hospitalisation.
In another study of 577 consecutive patients undergoing transperineal biopsy using single dose IV
cephazolin prophylaxis, one patient (0.2%) suffered prostatitis not requiring hospitalisation [267]. There were
no incidences of sepsis. In a further study of 485 patients using only cephazolin, 4 patients (0.8%) suffered
infectious complications [293].

5.2.7.2.2 Transrectal prostate biopsy


Meta-analysis of eight RCTs including 1,786 men showed that use of a rectal povidone-iodine preparation
before biopsy, in addition to antimicrobial prophylaxis, resulted in a significantly lower rate of infectious
complications (RR [95% CIs] 0.55 [0.41–0.72]) [288, 294-299]. Single RCTs showed no evidence of benefit for
perineal skin disinfection [300], but reported an advantage for rectal povidone-iodine preparation before biopsy
compared to after biopsy [301].
A meta-analysis of four RCTs including 671 men evaluated the use of rectal preparation by enema
before transrectal biopsy. No significant advantage was found regarding infectious complications (RR [95%
CIs] 0.96 [0.64–1.54]) [288, 302-304].
A meta-analysis of 26 RCTs with 3,857 patients found no evidence that use of peri-prostatic
injection of local anaesthesia resulted in more infectious complications than no injection (RR [95% CIs]
1.07 [0.77-1.48]) [288]. A meta-analysis of 9 RCTs including 2,230 patients found that extended biopsy
templates showed comparable infectious complications to standard templates (RR [95% CIs] 0.80 [0.53-1.22])
[288]. Additional meta-analyses found no difference in infections complications regarding needle guide type
(disposable vs. reusable), needle type (coaxial vs. non-coaxial), needle size (large vs. small), and number of
injections for peri-prostatic nerve block (standard vs. extended) [288].

PROSTATE CANCER - LIMITED UPDATE 2021 35


A meta-analysis of eleven studies with 1,753 patients showed significantly reduced infections after transrectal
prostate biopsy when using antimicrobial prophylaxis as compared to placebo/control (RR [95% CI] 0.56 [0.40–
0.77]) [305].
Fluoroquinolones have been traditionally used for antibiotic prophylaxis in this setting; however,
overuse and misuse of fluoroquinolones has resulted in an increase in fluoroquinolone resistance. In
addition, the European Commission has implemented stringent regulatory conditions regarding the use of
fluoroquinolones resulting in the suspension of the indication for peri-operative antibiotic prophylaxis including
prostate biopsy [306].
A systematic review and meta-analysis on antibiotic prophylaxis for the prevention of infectious
complications following prostate biopsy concluded that in countries where fluoroquinolones are allowed
as antibiotic prophylaxis, a minimum of a full one-day administration, as well as targeted therapy in case
of fluoroquinolone resistance, or augmented prophylaxis (combination of two or more different classes of
antibiotics) is recommended [305]. In countries where use of fluoroquinolones are suspended, cephalosporins
or aminoglycosides can be used as individual agents with comparable infectious complications based on a
meta-analysis of two RCTs [305]. A meta-analysis of three RCTs reported that fosfomycin trometamol was
superior to fluoroquinolones (RR [95% CI] 0.49 [0.27–0.87]) [305], but routine general use should be critically
assessed due to the relevant infectious complications reported in non-randomised studies [307]. Another
possibility is the use of augmented prophylaxis without fluoroquinolones, although no standard combination
has been established to date. Finally, targeted prophylaxis based on rectal swap/stool culture is plausible, but
no RCTs are available on non-fluoroquinolones. See figure 5.1 for prostate biopsy workflow to reduce infections
complications.

Based on a meta-analysis, suggested antimicrobial prophylaxis before transrectal biopsy may consist of:
1. Targeted prophylaxis - based on rectal swab or stool culture.
2. Augmented prophylaxis - two or more different classes of antibiotics (of note: this option is against
antibiotic stewardship programmes).
3. Alternative antibiotics:
• fosfomycin trometamol (e.g., 3 g before and 3 g 24–48 hrs. after biopsy);
• cephalosporin (e.g., ceftriaxone 1 g i.m; cefixime 400 mg p.o for 3 days starting 24 hrs. before biopsy)
aminoglycoside (e.g., gentamicin 3 mg/kg i.v.; amikacin 15 mg/kg i.m).

5.2.7.3 Summary of evidence and recommendations for performing prostate biopsy


(in line with the Urological Infections Guidelines Panel)

Summary of evidence LE
A meta-analysis of seven studies including 1,330 patients showed significantly reduced infectious 1a
complications in patients undergoing transperineal biopsy as compared to transrectal biopsy.
Meta-analysis of eight RCTs including 1,786 men showed that use of a rectal povidone-iodine 1a
preparation before transrectal biopsy, in addition to antimicrobial prophylaxis, resulted in a significantly
lower rate of infectious complications.
A meta-analysis on eleven studies with 1,753 patients showed significantly reduced infections after 1a
transrectal biopsy when using antimicrobial prophylaxis as compared to placebo/control.

Recommendations Strength rating*


Perform prostate biopsy using the transperineal approach due to the lower risk of infectious Strong
complications.
Use routine surgical disinfection of the perineal skin for transperineal biopsy. Strong
Use rectal cleansing with povidone-iodine in men prior to transrectal prostate biopsy. Strong
Do not use fluoroquinolones for prostate biopsy in line with the European Commission final Strong
decision on EMEA/H/A-31/1452.
Use either target prophylaxis based on rectal swab or stool culture; augmented prophylaxis Weak
(two or more different classes of antibiotics); or alternative antibiotics (e.g., fosfomycin
trometamol, cephalosporin, aminoglycoside) for antibiotic prophylaxis for transrectal biopsy.
Use a single oral dose of either cefuroxime or cephalexin or cephazolin as antibiotic Weak
prophylaxis for transperineal biopsy. Patients with severe penicillin allergy may be given
sulphamethoxazole.
Ensure that prostate core biopsies from different sites are submitted separately for Strong
processing and pathology reporting.

36 PROSTATE CANCER - LIMITED UPDATE 2021


Figure 5.1: Prostate biopsy workflow to reduce infectious complications*

Indication for prostate biopsy?

Transperineal biopsy feasible?

Yes No

Transperineal biopsy - 1st choice (⊕⊕⊝⊝) Transrectal biopsy – 2nd choice (⊕⊕⊝⊝)
with: with:
• perineal cleansing • povidone-iodine rectal preparation
• antibiotic prophylaxis • antibiotic prophylaxis

Fluoroquinolones licensed?

No Yes

1. Targeted prophylaxis: based on rectal Duration of antibiotic prophylaxis ≥ 24 hrs


swab or stool cultures (⊕⊕⊝⊝)
2. Augmented prophylaxis: two or more 1. Targeted prophylaxis (⊕⊕⊝⊝): based
different classes of antibiotics on rectal swab or stool cultures
3. Alternative antibiotics (⊕⊝⊝⊝): 2. Augmented prophylaxis (⊕⊝⊝⊝):
• fosfomycin trometamol (e.g. 3 g before • Fluoroquinolone plus aminoglycoside
and 3 g 24-48 hrs after biopsy) • Fluoroquinolone plus cephalosporin
• cephalosporin (e.g. ceftriaxone 1 g i.m.;
cefixime 400 mg p.o. for 3 days starting 3. Fluoroquinolone prophylaxis
24 hrs before biopsy) (range: ⊕⊝⊝⊝ - ⊕⊕⊝⊝)
• aminoglycoside (e.g. gentamicin 3 mg/kg
i.v.; amikacin 15 mg/kg i.m.)

GRADE Working Group grades of evidence.


• High certainty: (⊕⊕⊕⊕) very confident that the true effect lies close to that of the estimate of the effect.
• M oderate certainty: (⊕⊕⊕) moderately confident in the effect estimate: the true effect is likely to be close to
the estimate of the effect, but there is a possibility that it is substantially different.
• L ow certainty: (⊕⊕) confidence in the effect estimate is limited: the true effect may be substantially
different from the estimate of the effect.
• V ery low certainty: (⊕) very little confidence in the effect estimate: the true effect is likely to be
substantially different from the estimate of effect.
*Figure adapted from Pilatz et al., [308] with permission from Elsevier.

*Note on strength ratings:


The above strength ratings are explained here due to the major clinical implications of these new
recommendations. Although data showing the lower risk of infection via the transperineal approach is low
in certainty, its statistical and clinical significance warrants its Strong rating. Strong ratings are also given for
routine surgical disinfection of skin in transperineal biopsy and povidone-iodine rectal cleansing in transrectal
biopsy as, although quality of data is low, the clinical benefit is high and practical application simple. A ‘Strong’
rating is given for avoiding fluoroquinolones in prostate biopsy due to its legal implications in Europe.

5.2.7.4 Local anaesthesia prior to biopsy


Ultrasound-guided peri-prostatic block is recommended [309]. It is not important whether the depot is apical or
basal. Intra-rectal instillation of local anaesthesia is inferior to peri-prostatic infiltration [310]. Local anaesthesia
can also be used effectively for mpMRI-targeted and systemic transperineal biopsy [311]. Patients are placed
in the lithotomy position. Bupivacaine is injected into the perineal skin and subcutaneous tissues, followed
two minutes later by a peri-prostatic block. A systematic review evaluating pain in 3 studies comparing
transperineal vs. transrectal biopsies found that the transperineal approach significantly increased patient pain

PROSTATE CANCER - LIMITED UPDATE 2021 37


(RR: 1.83 [1.27–2.65]) [312]. In a randomised comparison a combination of peri-prostatic and pudendal block
anaesthesia reduced pain during transperineal biopsies compared to peri-prostatic anaesthesia only [313].
Targeted biopsies can then be taken via a brachytherapy grid or a freehand needle-guiding device under local
infiltration anaesthesia [311, 314, 315].

5.2.7.5 Complications
Complications of TRUS biopsy are listed in Table 5.8 [316]. Mortality after prostate biopsy is extremely rare and
most are consequences of sepsis [111]. Low-dose aspirin is no longer an absolute contraindication [317]. A
systematic review found favourable infection rates for transperineal compared to TRUS biopsies with similar
rates of haematuria, haematospermia and urinary retention [318]. A meta-analysis of 4,280 men randomised
between transperineal vs. TRUS biopsies in 13 studies found no significant differences in complication rates,
however, data on sepsis compared only 497 men undergoing TRUS biopsy to 474 having transperineal biopsy.
The transperineal approach required more (local) anaesthesia [257].

Table 5.8: Percentage of complications per TRUS biopsy session, irrespective of the number of cores

Complications Percentage of patients affected


Haematospermia 37.4
Haematuria > 1 day 14.5
Rectal bleeding < 2 days 2.2
Prostatitis 1.0
Fever > 38.5°C 0.8
Epididymitis 0.7
Rectal bleeding > 2 days +/− surgical intervention 0.7
Urinary retention 0.2
Other complications requiring hospitalisation 0.3

5.2.7.6 Seminal vesicle biopsy


Indications for seminal vesicle (SV) (staging) biopsies are poorly defined. At a PSA of > 15 ng/mL, the odds
of tumour involvement are 20–25% [319]. A SV staging biopsy is only useful if it has a decisive impact
on treatment, such as ruling out radical tumour resection or for potential subsequent RT. Its added value
compared with mpMRI is questionable.

5.2.7.7 Transition zone biopsy


Transition zone sampling during baseline biopsies has a low detection rate and should be limited to MRI-
detected lesions or repeat biopsies [320].

5.2.8 Pathology of prostate needle biopsies


5.2.8.1 Processing
Prostate core biopsies from different sites are processed separately. Before processing, the number and length
of the cores are recorded. The length of biopsy tissue significantly correlates with the PCa detection rate
[321]. To achieve optimal flattening and alignment, a maximum of three cores should be embedded per tissue
cassette, and sponges or paper used to keep the cores stretched and flat [322, 323]. To optimise detection of
small lesions, paraffin blocks should be cut at three levels and intervening unstained sections may be kept for
immunohistochemistry (IHC) [320].

5.2.8.2 Microscopy and reporting


Diagnosis of PCa is based on histology. The diagnostic criteria include features pathognomonic of cancer,
major and minor features favouring cancer and features against cancer. Ancillary staining and additional
(deeper) sections should be considered if a suspect lesion is identified [324-326]. Diagnostic uncertainty
is resolved by intradepartmental or external consultation [324]. Section 5.2.8.3 lists the recommended
terminology for reporting prostate biopsies [322]. Type and subtype of PCa should be reported such as for
instance acinar adenocarcinoma (> 95% of PCa), ductal adenocarcinoma (< 5%) and poorly differentiated
small or large cell neuroendocrine carcinoma (< 1%), even if representing a small proportion of the PCa. The
distinct aggressive nature of ductal adenocarcinoma and small/large cell neuroendocrine carcinoma should be
commented upon in the pathology report [322].

38 PROSTATE CANCER - LIMITED UPDATE 2021


5.2.8.2.1 Recommended terminology for reporting prostate biopsies [255]

Recommendations Strength rating


Benign/negative for malignancy; if appropriate, include a description. Strong
Active inflammation.
Granulomatous inflammation.
High-grade prostatic intraepithelial neoplasia (PIN).
High-grade PIN with atypical glands, suspicious for adenocarcinoma (PINATYP).
Focus of atypical glands/lesion suspicious for adenocarcinoma/atypical small acinar
proliferation, suspicious for cancer.
Adenocarcinoma, provide type and subtype.
Intraductal carcinoma.

Each biopsy site should be reported individually, including its location (in accordance with the sampling site)
and histopathological findings, which include the histological type and the ISUP 2014 grade [84]. For mpMRI
targeted biopsies consisting of multiple cores per target the aggregated ISUP grade and percentage of high-
grade carcinoma should be reported per targeted lesion [85]. If the targeted biopsies are negative, presence
of specific benign pathology should be mentioned, such as dense inflammation, fibromuscular hyperplasia
or granulomatous inflammation [327]. A global ISUP grade comprising all systematic (non-targeted) biopsies
is also reported (see Section 4.2). The global ISUP grade takes into account all systemic biopsies positive for
carcinoma, by estimating the total extent of each Gleason grade present. For instance, if three biopsy sites
are entirely composed of Gleason grade 3 and one biopsy site of Gleason grade 4 only, the global ISUP grade
would be 2 (i.e. GS 7[3+4]) or 3 (i.e. GS 7[4+3]), dependent on whether the extent of Gleason grade 3 exceeds
that of Gleason grade 4, whereas the worse grade would be ISUP grade 4 (i.e. GS 8[4+4]). Recent publications
demonstrated that global ISUP grade is somewhat superior in predicting prostatectomy ISUP grade [328] and
BCR [329].
Lymphovascular invasion (LVI) and extraprostatic extension (EPE) must each be reported, if
identified. More recently, expansile cribriform pattern of PCa as well as intraductal carcinoma in biopsies were
identified as independent prognosticators of metastatic disease [330] and PCa-specific survival [331] and
presence of both adverse pathologies should be reported [85].
The proportion of systematic (non-targeted) carcinoma-positive cores as well as the extent of
tumour involvement per biopsy core correlate with the ISUP grade, tumour volume, surgical margins and
pathologic stage in RP specimens and predict BCR, post-prostatectomy progression and RT failure. These
parameters are included in nomograms created to predict pathologic stage and SV invasion after RP and RT
failure [332-334]. A pathology report should therefore provide both the proportion of carcinoma-positive cores
and the extent of cancer involvement for each core. The length in mm and percentage of carcinoma in the
biopsy have equal prognostic impact [335]. An extent of > 50% of adenocarcinoma in a single core is used in
some AS protocols as a cut off [336] triggering immediate treatment vs. AS in patients with ISUP grade 1.

A prostate biopsy that does not contain glandular tissue should be reported as diagnostically inadequate.
Mandatory elements to be reported for a carcinoma-positive prostate biopsy are:
• type of carcinoma;
• primary and secondary/worst Gleason grade (per biopsy site and global);
• percentage high-grade carcinoma (global);
• extent of carcinoma (in mm or percentage) (per biopsy site);
• if present: EPE, SV invasion, LVI, intraductal carcinoma/cribriform pattern, peri-neural invasion;
• ISUP grade (global);
• In targeted biopsies aggregate ISUP grade and percentage high-grade carcinoma per targeted site;
• In carcinoma-negative targeted biopsy report specific benign pathology, e.g., fibromuscular hyperplasia
or granulomatous inflammation, if present [85];

5.2.8.3 Tissue-based prognostic biomarker testing


After a comprehensive literature review and several panel discussions an ASCO-EAU-AUA multidisciplinary
expert panel made recommendations regarding the use of tissue-based PCa biomarkers. The
recommendations were limited to 5 commercially available tests (Oncotype Dx®, Prolaris®, Decipher®, Decipher
PORTOS and ProMark®) with extensive validation in large retrospective studies and evidence that their test
results might actually impact clinical decision-taking [337].

PROSTATE CANCER - LIMITED UPDATE 2021 39


The selected commercially available tests significantly improved the prognostic accuracy of clinical
multivariable models for identifying men who would benefit of AS and those with csPCa requiring curative
treatment, as well as for guidance of patient management after RP. In addition, a few studies showed that
tissue biomarker tests and MRI findings independently improved the detection of clinically significant cancer
in an AS setting, but it remains unclear which men would benefit of both tests. Since the long-term impact
of the use of these commercially available tests on oncological outcome remains unproven and prospective
trials are largely lacking, the Panel concluded that these tests should not be offered routinely but only in
subsets of patients where the test result provides clinically actionable information, such as for instance in men
with favourable intermediate-risk PCa who might opt for AS or men with unfavourable intermediate-risk PCa
scheduled for RT to decide on treatment intensification with hormonal therapy (HT).

5.2.8.4 Histopathology of radical prostatectomy specimens


5.2.8.4.1 Processing of radical prostatectomy specimens
Histopathological examination of RP specimens describes the pathological stage, histopathological type,
grade and surgical margins of PCa. It is recommended that RP specimens are totally embedded to enable
assessment of cancer location, multifocality and heterogeneity. For cost-effectiveness, partial embedding may
also be considered, particularly for prostates > 60 g. The most widely accepted method includes complete
embedding of the posterior prostate and a single mid-anterior left and right section. Compared with total
embedding, partial embedding detected 98% of PCa with an ISUP grade > 2 with accurate staging in 96% of
cases [338].

The entire RP specimen should be inked upon receipt in the laboratory to demonstrate the surgical margins.
Specimens are fixed by immersion in buffered formalin for at least 24 hours, preferably before slicing. Fixation
can be enhanced by injecting formalin which provides more homogeneous fixation and sectioning after
24 hours [339]. After fixation, the apex and the base (bladder neck) are removed and cut into (para)sagittal
or radial sections; the shave method is not recommended [83]. The remainder of the specimen is cut in
transverse, 3–4 mm sections, perpendicular to the long axis of the urethra. The resultant tissue slices can
be embedded and processed as whole-mounts or after quadrant sectioning. Whole-mounts provide better
topographic visualisation, faster histopathological examination and better correlation with pre-operative
imaging, although they are more time-consuming and require specialist handling. For routine sectioning, the
advantages of whole mounts do not outweigh their disadvantages.

5.2.8.4.1.1 Guidelines for processing prostatectomy specimens

Recommendations Strength rating


Ensure total embedding, by conventional (quadrant) or whole-mount sectioning. Strong
Ink the entire surface before cutting, to evaluate the surgical margin. Strong
Examine the apex and base separately, using the cone method with sagittal or radial Strong
sectioning.

5.2.8.4.2 Radical prostatectomy specimen report


The pathology report provides essential information on the prognostic characteristics relevant for clinical
decision-making (Table 5.9). As a result of the complex information to be provided for each RP specimen, the
use of synoptic(-like) or checklist reporting is recommended (Table 5.10). Synoptic reporting results in more
transparent and complete pathology reporting [340].

Table 5.9: Mandatory elements provided by the pathology report

Histopathological type: > 95% of PCa represents conventional (acinar) adenocarcinoma.


Grading according to ISUP grade (or not applicable if therapy-related changes).
Presence of intraductal and/or cribriform carcinoma.
Tumour (sub)staging and surgical margin status: location and extent of EPE, presence of bladder neck
invasion, laterality of EPE or SV invasion, location and extent of positive surgical margins.
Additional information may be provided on multifocality, and diameter/volume and zonal location of the
dominant tumour.

40 PROSTATE CANCER - LIMITED UPDATE 2021


Table 5.10: Example checklist: reporting of prostatectomy specimens

Histopathological type
Type of carcinoma, e.g. conventional acinar, or ductal
Histological grade
Primary (predominant) Gleason grade
Secondary Gleason grade
Tertiary Gleason grade (if applicable)
Global ISUP grade
Approximate percentage of Gleason grade 4 or 5
Tumour quantitation (optional)
Percentage of prostate involved
Size/volume of dominant tumour nodule
Pathological staging (pTNM)
If extraprostatic extension is present:
• indicate whether it is focal or extensive*;
• specify sites;
• indicate whether there is seminal vesicle invasion.
If applicable, regional lymph nodes:
• location;
• number of nodes retrieved;
• number of nodes involved.
Surgical margins
If carcinoma is present at the margin:
• specify sites.
Other
Presence of lymphovascular/angio-invasion
Location of dominant tumour
Presence of intraductal carcinoma/cribriform architecture
*Focal is defined as carcinoma glands extending less than 1 high-power field (HPF) in the extraprostatic fat in
1 or 2 sections and extensive is extension beyond or in more sections.

5.2.8.4.3 ISUP grade in prostatectomy specimens


Grading of conventional prostatic adenocarcinoma using the (ISUP 2014 modified) Gleason system is the
strongest prognostic factor for clinical behaviour and treatment response [84]. The ISUP grade is incorporated
in nomograms that predict disease-specific survival (DSS) after prostatectomy [341].

The ISUP grade is based on the sum of the most and second-most dominant (in terms of volume) Gleason
grade. ISUP grade 1 is GS 6. ISUP grades 2 and 3 represent carcinomas constituted of Gleason grade 3 and
4 components, with ISUP grade 2 when 50% of the carcinoma, or more, is Gleason grade 3 and ISUP grade 3
when the grade 4 component represents more than 50% of the carcinoma. In a carcinoma almost entirely
composed of Gleason grade 3 the presence of a minor (< 5%) Gleason pattern 4 component is not included in
the Gleason score (ISUP grade 1), but its presence is commented upon.
ISUP grade 4 is largely composed of Gleason grade 4 and ISUP grade 5 of a combination of
Gleason grade 4 and 5 or only Gleason grade 5. A global ISUP grade is given for multiple tumours, but a
separate tumour focus with a higher ISUP grade should also be mentioned. Tertiary Gleason grade 5, if > 5%
of the PCa volume, is an unfavourable prognostic indicator for BCR and should be incorporated in the ISUP
grade. If less than 5% its presence should be mentioned in the report [85, 342].

5.2.8.4.4 Definition of extraprostatic extension


Extraprostatic extension is defined as carcinoma mixed with peri-prostatic adipose tissue, or tissue that
extends beyond the prostate gland boundaries (e.g., neurovascular bundle, anterior prostate). Microscopic
bladder neck invasion is considered EPE. It is useful to report the location and extent of EPE because the latter
is related to recurrence risk [343].
There are no internationally accepted definitions of focal or microscopic, vs. non-focal or extensive
EPE. Some describe focal as a few glands [344] or < 1 HPF in one or at most two sections [345] whereas
others measure the depth of extent in millimetres [346].

PROSTATE CANCER - LIMITED UPDATE 2021 41


At the apex of the prostate, tumour mixed with skeletal muscle does not constitute EPE. In the
bladder neck, microscopic invasion of smooth muscle fibres is not equated to bladder wall invasion, i.e. not
as pT4, because it does not carry independent prognostic significance for PCa recurrence and should be
recorded as EPE (pT3a) [347, 348]. Stage pT4 is only assigned when the tumour invades the bladder muscle
wall as determined macroscopically [349].

5.2.8.4.5 PCa volume


The independent prognostic value of PCa volume in RP specimens has not been established [345, 350-353].
Nevertheless, a cut-off of 0.5 mL is traditionally used to distinguish insignificant from clinically relevant cancer
[350]. Improvement in prostatic radio-imaging allows more accurate pre-operative measurement of cancer
volume. It is recommended that at least the diameter/volume of the dominant tumour nodule should be
assessed, or a rough estimate of the percentage of cancer tissue provided [354].

5.2.8.4.6 Surgical margin status


Surgical margin is an independent risk factor for BCR. Margin status is positive if tumour cells are in contact
with the ink on the specimen surface. Margin status is negative if tumour cells are close to the inked surface
[351] or at the surface of the tissue lacking ink. In tissues that have severe crush artefacts, it may not be
possible to determine margin status [355].
Surgical margin is separate from pathological stage, and a positive margin is not evidence of EPE
[356]. There is insufficient evidence to prove a relationship between margin extent and recurrence risk [345].
However, some indication must be given of the multifocality and extent of margin positivity, such as the linear
extent in mm of involvement: focal, < 1 mm vs. extensive, > 1 mm [357], or number of blocks with positive
margin involvement. Gleason score at the positive margin was found to correlate with outcome, and should be
reported [313].

5.3 Diagnosis - Clinical Staging


5.3.1 T-staging
The cT category used in the risk table only refers to the DRE finding. The imaging parameters and biopsy
results for local staging are, so far, not part of the risk category stratification [358].

5.3.1.1 TRUS
Transrectal ultrasound is no more accurate at predicting organ-confined disease than DRE [359]. Some single-
centre studies reported good results in local staging using 3D TRUS or colour Doppler but these good results
were not confirmed by large-scale studies [360, 361].

5.3.1.2 MRI
T2-weighted imaging remains the most useful method for local staging on MRI. At 1.5 Tesla, MRI has good
specificity but low sensitivity for detecting T3 stages. Pooled data from a meta-analysis showed a sensitivity
and specificity of 0.57 (95% CI: 0.49–0.64) and 0.91 (95% CI: 0.88–0.93), 0.58 (95% CI: 0.47–0.68) and 0.96
(95% CI: 0.95–0.97), and 0.61 (95% CI: 0.54–0.67) and 0.88 (95% CI: 0.85–0.91), for EPE, SVI, and overall
stage T3 assessement, respectively [362]. Magnetic resonance imaging cannot detect microscopic EPE. Its
sensitivity increases with the radius of extension within peri-prostatic fat. In one study, the EPE detection rate
increased from 14 to 100% when the radius of extension increased from < 1 mm to > 3 mm [363]. In another
study, MRI sensitivity, specificity and accuracy for detecting pT3 stage were 40%, 95% and 76%, respectively,
for focal (i.e. microscopic) EPE, and 62%, 95% and 88% for extensive EPE [364].
The use of high field strength (3 Tesla) or functional imaging in addition to T2-weighted imaging
improves sensitivity for EPE or SVI detection [362] but the experience of the reader remains of paramount
importance [365] and the inter-reader agreement remains moderate with kappa values ranging from 0.41 to
0.68 [366].
Several series suggested that MRI findings could improve the prediction of the pathological stage
when combined with clinical and biopsy data [248]. However, all these studies were based on cohorts of men
diagnosed with systematic biopsy and their generalisability in the targeted biopsy setting is questionable. For
risk calculators predicting EPE and SVI based on MRI findings, clinical data and combined systematic and
targeted biopsy results have been recently reported [367], but they have not undergone external validation yet.
Given its low sensitivity for focal (microscopic) EPE, MRI is not recommended for local staging in
low-risk patients [368-370]. However, MRI can still be useful for treatment planning.

42 PROSTATE CANCER - LIMITED UPDATE 2021


5.3.2 N-staging
5.3.2.1 Computed tomography and magnetic resonance imaging
Abdominal CT and T1-T2-weighted MRI indirectly assess nodal invasion by using LN diameter and
morphology. However, the size of non-metastatic LNs varies widely and may overlap the size of LN metastases.
Usually, LNs with a short axis > 8 mm in the pelvis and > 10 mm outside the pelvis are considered malignant.
Decreasing these thresholds improves sensitivity but decreases specificity. As a result, the ideal size threshold
remains unclear [371, 372]. Computed tomography (CT) and MRI sensitivity is less than 40% [373, 374]. Among
4,264 patients, 654 (15.3%) of whom had positive LNs at LND, CT was positive in only 105 patients (2.5%)
[371]. In a multi-centre database of 1,091 patients who underwent pelvic LN dissection, CT sensitivity and
specificity were 8.8% and 98%, respectively [375]. Detection of microscopic LN invasion by CT is < 1% in
patients with ISUP grade < 4 cancer, PSA < 20 ng/mL, or localised disease [376-378].
Diffusion-weighted MRI (DW-MRI) may detect metastases in normal-sized nodes, but a negative
DW-MRI cannot rule out the presence of LN metastases [372, 379].

5.3.2.2 Risk calculators incorporating MRI findings


Because CT and MRI lack sensitivity for direct detection of positive LNs, nomograms combining clinical and
biopsy findings (such as the nomograms developed by the Memorial Sloan Cancer Center (MSKCC) [380],
by Briganti et al. or by Gandaglia et al. [381, 382] have been used to identify patients at higher risk of LN
invasion who should be considered for LN dissection. Although these nomograms are associated with good
performance, they have been developed using systematic biopsy findings and may therefore not be sensitive to
patients diagnosed with combined MRI-TBx and systematic biopsy.
Two models incorporating MRI-TBx biopsy findings and MRI-derived findings recently underwent
external validation [367, 383]. One model tested on an external cohort of 187 patients treated by RP and
extended LN dissection showed a prevalence of LN invasion of 13.9% (vs. 16.9% in the development cohort).
The C-index was 0.73 (vs. 0.81 in the development cohort); at calibration analysis, the model tended to
overpredict the actual risk [383]. Another model was validated in an external multi-centre cohort of 487 patients
with a prevalence of 8% of LN invasion (vs. 12.5% in the development cohort).The AUC was 0.79 (vs. 0.81
in the development cohort). Using a risk cut-off of 7% would have avoided LN dissection in 56% of patients,
while missing LN invasion in 13 (34%) of the 38 patients with LN invasion [367].

5.3.2.3 Choline PET/CT


In a meta-analysis of 609 patients, pooled sensitivity and specificity of choline PET/CT for pelvic LN
metastases were 62% (95% CI: 51–66%) and 92% (95% CI: 89–94%), respectively [384]. In a prospective
trial of 75 patients at intermediate risk of nodal involvement (10–35%), the sensitivity was only 8.2% at
region-based analysis and 18.9% at patient-based analysis, which is too low to be of clinical value [385]. The
sensitivity of choline PET/CT increases to 50% in patients at high risk and to 71% in patients at very high risk,
in both cases out-performing contrast-enhanced CT [386]. However, comparisons between choline PET/CT
and DW-MRI yielded contradictory results, with PET/CT sensitivity found to be superior [387], similar [388, 389]
or inferior [385] than that of DW-MRI.
Due of its low sensitivity, choline PET/CT does not reach clinically acceptable diagnostic accuracy
for detection of LN metastases, or to rule out a nodal dissection based on risk factors or nomograms (see
Section 6.3.4.1.2).

5.3.2.4 Prostate-specific membrane antigen-based PET/CT


Prostate-specific membrane antigen (PSMA) positron-emission tomography (PET)/CT uses several different
radiopharmaceuticals; the majority of published studies used 68Ga-labelling for PSMA PET imaging, but some
used 18F-labelling. At present there are no conclusive data about comparison of such tracers, with additional
new radiotracers being developed. 68Ga-labelled or 18F-labelled radiotracers for PSMA PET/CT imaging provide
good contrast-to-noise ratio, thereby improving the detectability of lesions. Prostate-specific membrane
antigen is also an attractive target because of its specificity for prostate tissue, even if PSMA expression in
other non-prostatic malignancies or benign conditions may cause incidental false-positive findings [390-394].

A prospective, multi-centre study addressed the use of 68Ga-PSMA PET/CT in patients with newly diagnosed
PCa and negative bone scan findings. In 103 eligible patients at increased risk for metastatic LNs prior to
surgery, 97 extended pelvic lymph-node dissections (ePLND) were performed, resulting in the identification of
85 LN metastases in 41 patients (42.3%). Positron-emission tomography was positive in 17 patients, resulting
in a per-patient-based sensitivity and specificity of 41.5% (95% CI: 26.7–57.8) and 90.9% (95% CI: 79.3–96.6),
respectively. A treatment change occurred in 12.6% of patients [395]. Another prospective multi-centre trial
investigated the diagnostic accuracy of 18F-DCFPyL ([2-(3-(1-carboxy-5-[(6-18F-fluoropyridine-3-carbonyl)-
amino]-pentyl)-ureido)-pentanedioic acid] PET/CT for LN staging in 117 patients with primary PCa, prior to

PROSTATE CANCER - LIMITED UPDATE 2021 43


robot-assisted radical prostatectomy (RARP) with ePLND. 18F-DCFPyL PET/CT showed a high specificity
(94.0%; CI: 86.9–97.5%), and a limited sensitivity (41.2%,CI: 19.4–66.5%) for the detection of pelvic LN
metastases [396]. This suggests that current PSMA-based PET/CT imaging cannot yet replace diagnostic
ePLND.
Prostate-specific antigen may be a predictor of a positive PSMA PET/CT. In the primary staging
cohort from a meta-analysis [397], however, only 4 studies reported PSMA PET-positivity based on the PSA
value, offering no robust estimates of positivity. The tracer uptake is also influenced by the ISUP grade and the
PSA level. In a series of 90 patients with primary PCa, tumours with an ISUP grade between 1 and 3 showed
significantly lower tracer uptake than tumours with an ISUP grade > 4. Similarly, patients with PSA levels
> 10 ng/mL showed significantly higher uptake than those with PSA levels < 10 ng/mL [398].

Comparison between PSMA PET/CT and MRI was recently performed in a systematic review and meta-
analysis including 13 studies (n = 1,597) [399]. 68Ga-PSMA was found to have a higher sensitivity and a
comparable specificity for staging pre-operative LN metastases in intermediate- and high-risk PCa. The pooled
sensitivity and specificity of 68Ga-PSMA PET were 0.65 (95% CI: 0.49–0.79) and 0.94 (95% CI: 0.88–0.97),
respectively, while the corresponding values of MRI were 0.41 (95% CI: 0.26–0.57) and 0.92 (95% CI:
0.86–0.95). 68Ga-PSMA PET was potentially a more effective and appropriate imaging modality to predict LN
metastasis prior to surgery as indicated by the area under the symmetric receiver-operating characteristic
(SROC) curve. Another prospective trial reported superior sensitivity of PSMA PET/CT as compared to MRI for
nodal staging of 36 high-risk PCa patients [400].

PSMA PET/CT has a good sensitivity and specificity for LN involvement, possibly impacting clinical decision-
making. In a recent review and meta-analysis including 37 articles, a subgroup analysis was perfomed in
patients undergoing PSMA PET/CT for primary staging. On a per-patient-based analysis, the sensitivity and
specificity of 68Ga-PSMA PET were 77% and 97%, respectively, after eLND at the time of RP. On a per-lesion-
based analysis, sensitivity and specificity were 75% and 99%, respectively [397].

In summary, PSMA PET/CT is more appropriate in N-staging as compared to MRI, abdominal contrast-
enhanced CT or choline PET/CT; however, small LN metastases, under the spatial resolution of PET (~5 mm),
may still be missed.

5.3.3 M-staging
5.3.3.1 Bone scan
99mTc-Bone scan has been the most widely used method for evaluating bone metastases of PCa. A meta-

analysis showed combined sensitivity and specificity of 79% (95% CI: 73–83%) and 82% (95% CI: 78–85%)
at patient level and 59% (95% CI: 55–63%) and 75% (95% CI: 71–79%) at lesion level [401]. Bone scan
diagnostic yield is significantly influenced by the PSA level, the clinical stage and the tumour ISUP grade and
these three factors were the only independent predictors of bone scan positivity in a study of 853 patients [402].
The mean bone scan positivity rate in 23 different series was 2.3% in patients with PSA levels < 10 ng/ mL,
5.3% in patients with PSA levels between 10.1 and 19.9 ng/mL and 16.2% in patients with PSA levels of
20.0–49.9 ng/mL. It was 6.4% in men with organ-confined cancer and 49.5% in men with locally advanced
cancers. Detection rates were 5.6% and 29.9% for ISUP grade 2 and > 3, respectively [371]. In two studies,
a major Gleason pattern of 4 was found to be a significant predictor of positive bone scan [403, 404]. Bone
scanning should be performed in symptomatic patients, independent of PSA level, ISUP grade or clinical stage
[371].

5.3.3.2 Fluoride PET and PET/CT, choline PET/CT and MRI


18F-sodium fluoride (18F-NaF) PET or PET/CT was reported to have similar specificity and superior sensitivity

to bone scintigraphy for detecting bone metastases in patients with newly diagnosed high-risk PCa [405,
406]. However, in a prospective study 18F-NaF PET showed no added value over bone scintigraphy in patients
with newly diagnosed intermediate- or high-risk PCa and negative bone scintigraphy results [407]. Recently,
the interobserver agreement for the detection of bone metastases and the accuracy of 18F-NaF PET/CT in
the diagnosis of bone metastases were investigated. Bone metastases were identified in 211 out of 219
patients with an excellent interobserver agreement, demonstrating that 18F-NaF PET/CT is a robust tool for the
detection of osteoblastic lesions in patients with PCa [408].
It remains unclear whether choline PET/CT is more sensitive than bone scan but it has higher
specificity with fewer indeterminate bone lesions [384, 409, 410].
Diffusion-weighted whole-body and axial MRI are more sensitive than bone scan and targeted
conventional radiography in detecting bone metastases in high-risk PCa [411, 412]. Whole-body MRI is also
more sensitive and specific than combined bone scan, targeted radiography and abdominopelvic CT [413].

44 PROSTATE CANCER - LIMITED UPDATE 2021


A meta-analysis found that MRI is more sensitive than choline PET/CT and bone scan for detecting bone
metastases on a per-patient basis, although choline PET/CT had the highest specificity [401].
It is of note that choline PET/CT and DW-MRI can also detect visceral and nodal metastases. Bone
scan and 18F-NaF PET/CT only assess the presence of bone metastases.

5.3.3.3 Prostate-specific membrane antigen-based PET/CT


A systematic review including 12 studies (n = 322) reported high variation in 68Ga-PSMA PET/CT sensitivity
for initial staging (range 33–99% median sensitivity on per-lesion analysis 33–92%, and on per-patient
analysis 66–91%), with good specificity (per-lesion 82–100%, and per-patient 67–99%), with most studies
demonstrating increased detection rates with respect to conventional imaging modalities (bone scan and CT)
[414]. Table 5.11 reports the data of the 5 studies including histopathologic correlation.

Table 5.11: PSMA PET/CT results in primary staging alone [414]

Study Sensitivity Specificity PPV NPV


(per lesion) (per lesion) (per lesion) (per lesion)
Budaus 33% 100% 100% 69%
Herlemann 84% 82% 84% 82%
Van Leeuwen 58% 100% 94% 98%
Maurer 74% 99% 95% 94%
Rahbar 92% 92% 96% 85%
NPV = negative predictive value; PPV = positive predictive value.

One prospective multi-centre study evaluated changes in planned management before and after PSMA PET/CT
in 108 intermediate- and high-risk patients referred for primary staging. As compared to conventional staging,
additional LNs and bone/visceral metastases were detected in 25% and 6% of patients, respectively [415];
management changes occurred in 21% of patients. A recent retrospective review investigated the risk of
metastases identified by 68Ga-PSMA at initial staging in 1,253 patients (high-risk disease in 49.7%) [416].
Metastatic disease was identified by PSMA PET/CT in 12.1% of men, including 8.2% with a PSA level of
< 10 ng/mL and 43% with a PSA level of > 20 ng/mL. Lymph node metastases were suspected in 107 men,
with 47.7% outside the boundaries of an ePLND. Bone metastases were identified in 4.7%. In men with
intermediate-risk PCa metastases were identified in 5.2%, compared to 19.9% with high-risk disease.
In the PSMA PET/CT prospective multi-centre study in patients with high-risk PCa before curative-
intent surgery or RT (proPSMA), 302 patients were randomly assigned to conventional imaging with CT and
bone scintigraphy or 68Ga-PSMA-11 PET/CT. The primary outcome focused on the accuracy of first-line
imaging for the identification of pelvic LN or distant metastases, using a predefined reference standard
consisting of histopathology, imaging, and biochemistry at 6-month follow-up. Accuracy of 68Ga-PSMA PET/CT
was 27% (95% CI: 23–31) higher than that of CT and bone scintigraphy (92% [88–95] vs. 65% [60–69];
p < 0.0001). Conventional imaging had a lower sensitivity (38% [24–52] vs. 85% [74–96]) and specificity (91%
[85–97] vs. 98% [95–100]) than PSMA PET/CT. Furthermore, 68Ga-PSMA PET/CT scan prompted management
change more frequently as compared to conventional imaging (41 [28%] men [21–36] vs. 23 [15%] men [10–22],
p = 0.08), with less equivocal findings (7% [4–13] vs. 23% [17–31]) and lower radiation exposure (8.4 mSv vs.
19.2 mSv; p < 0.001) [417].

5.3.4 Summary of evidence and practical considerations on initial N/M staging


The field of non-invasive N- and M-staging of PCa patients is evolving very rapidly. Evidence shows that
choline PET/CT, PSMA PET/CT and MRI provide a more sensitive detection of LN- and bone metastases
than the classical work-up with bone scan and abdominopelvic CT. In view of the evidence offered by the
randomised, multi-centre proPSMA trial [417], replacing bone scan and abdominopelvic CT by more sensitive
imaging modalities may be a consideration in patients with high-risk PCa undergoing initial staging. However, in
absence of prospective studies demonstrating survival benefit, caution must be used when taking therapeutic
decisions [418]. The prognosis and ideal management of patients diagnosed as metastatic by these more
sensitive tests is unknown. In particular, it is unclear whether patients with metastases detectable only
with PET/CT or MRI should be managed using systemic therapies, or whether they should be subjected to
aggressive local and metastases-directed therapies [419].
Results from RCTs evaluating the management and outcome of patients with (and without)
metastases detected by choline PET/CT, PSMA PET/CT and MRI are awaited before a decision can be made to
treat patients based on the results of these tests [420].

PROSTATE CANCER - LIMITED UPDATE 2021 45


5.3.5 Summary of evidence and guidelines for staging of prostate cancer

Summary of evidence LE
PSMA PET/CT is more accurate for staging than CT and bone scan but to date no outcome data exist 1b
to inform subsequent management.

Any risk group staging Strength rating


Use pre-biopsy MRI for local staging information. Weak
Low-risk localised disease
Do not use additional imaging for staging purposes. Strong
Intermediate-risk disease
In ISUP grade > 3, include at least cross-sectional abdominopelvic imaging and a bone- Weak
scan for metastatic screening.
High-risk localised disease/locally advanced disease
Perform metastatic screening including at least cross-sectional abdominopelvic imaging Strong
and a bone-scan.

5.4 Estimating life expectancy and health status


5.4.1 Introduction
Evaluation of life expectancy and health status is important in clinical decision-making for screening, diagnosis,
and treatment of PCa. Prostate cancer is common in older men (median age 68) and diagnoses in men > 65 will
result in a 70% increase in annual diagnosis by 2030 in Europe and the USA [421, 422].
Active treatment mostly benefits patients with intermediate- or high-risk PCa and longest expected
survival. In localised disease, over 10 years life expectancy is considered mandatory for any benefit from local
treatment and an improvement in CSS may take longer to become apparent. Older age and worse baseline
health status have been associated with a smaller benefit in PCa-specific mortality (PCSM) and life expectancy
of surgery vs. active surveillance (AS) [423]. Although in a RCT the benefit of surgery with respect to death from
PCa was largest in men < 65 years of age (RR: 0.45), RP was associated with a reduced risk of metastases and
use of androgen deprivation therapy (ADT) among older men (RR: 0.68 and 0.60, respectively) [424]. External
beam radiotherapy shows similar cancer control regardless of age, assuming a dose of > 72 Gy when using
intensity-modulated or image-guided RT [425].
Older men have a higher incidence of PCa and may be under-treated despite the high overall
mortality rates [426, 427]. Of all PCa-related deaths 71% occur in men aged > 75 years [428], probably due
to the higher incidence of advanced disease and death from PCa despite higher death rates from competing
causes [429-431]. In the USA, only 41% of patients aged > 75 years with intermediate- and high-risk disease
received curative treatment compared to 88% aged 65–74 [432].

5.4.2 Life expectancy


Life expectancy tables for European men are available online: https://ec.europa.eu/eurostat/web/products-
datasets/-/tps00205. Survival may be variable and therefore estimates of survival must be individualised. Gait
speed is a good single predictive method of life expectancy (from a standing start, at usual pace, generally over
6 meters). For men at age 75, 10-year survival ranged from 19% < 0.4 m/s to 87%, for > 1.4 m/s [433].

46 PROSTATE CANCER - LIMITED UPDATE 2021


Figure 5.2: Predicted Median Life Expectancy by Age and Gait Speed for males* [433]

*Figure reproduced with permission of the publisher, from Studenski S, et al. JAMA 2011 305(1)50.

5.4.3 Health status screening


Heterogeneity increases with advancing age, so it is important to use measures other than just age or
performance status (PS) when considering treatment options. The International SIOG PCa Working Group
recommends that treatment for adults over 70 years of age should be based on a systematic evaluation of
health status using the G8 (Geriatric 8) screening tool (see Table 5.12) [134]. This tool helps to discriminate
between those who are fit and those with frailty, a syndrome of reduced ability to respond to stressors.
Patients with frailty have a higher risk of mortality and negative side effects of cancer treatment [434]. Healthy
patients with a G8 score > 14 or vulnerable patients with reversible impairment after resolution of their geriatric
problems should receive the same treatment as younger patients. Frail patients with irreversible impairment
should receive adapted treatment. Patients who are too ill should receive only palliative treatment (see Figure
5.3) [134]. Patients with a G8 score < 14 should undergo a comprehensive geriatric assessment (CGA) as this
score is associated with 3-year mortality. A CGA is a multi-domain assessment that includes co-morbidity,
nutritional status, cognitive and physical function, and social supports to determine if impairments are
reversible [435]. A systematic review of the effect of geriatric evaluation for older cancer patients showed
improved treatment tolerance and completion [436].
The Clinical Frailty Scale (CFS) is another screening tool for frailty (see Table 5.4.2) [437]. Although
not frequently used in the cancer setting, it is considered to be a common language for expressing degree of
frailty. The scale runs from 1 to 9, with higher scores indicating increasing frailty. Patients with a higher CFS
score have a higher 30-day mortality after surgery and are less likely to be discharged home [438].
It is important to use a validated tool to identify frailty, such as the G8 or CFS, as clinical judgement
has been shown to be poorly predictive of frailty in older patients with cancer [439].

5.4.3.1 Co-morbidity
Co-morbidity is a major predictor of non-cancer-specific death in localised PCa treated with RP and is more
important than age [440, 441]. Ten years after not receiving active treatment for PCa, most men with a high
co-morbidity score had died from competing causes, irrespective of age or tumour aggressiveness [440].
Measures for co-morbidity include: Cumulative Illness Score Rating-Geriatrics (CISR-G) [442, 443] (Table 5.13)
and Charlson Co-morbidity Index (CCI) [444].

5.4.3.2 Nutritional status


Malnutrition can be estimated from body weight during the previous 3 months (good nutritional status < 5%
weight loss; risk of malnutrition: 5–10% weight loss; severe malnutrition: > 10% weight loss) [445].

PROSTATE CANCER - LIMITED UPDATE 2021 47


5.4.3.3 Cognitive function
Cognitive impairment can be screened for using the mini-COG (https://mini-cog.com/) which consists of
three-word recall and a clock-drawing test and can be completed within 5 minutes. A score of < 3/5 indicates
the need to refer the patient for full cognitive assessment. Patients with any form of cognitive impairment
(e.g., Alzheimer’s or vascular dementia) may need a capacity assessment of their ability to make an informed
decision, which is an increasingly important factor in health status assessment [446-448]. Cognitive impairment
also predicts risk of delirium, which is important for patients undergoing surgery [449].

5.4.3.4 Physical function


Measures for overall physical functioning include: Karnofsky score and ECOG scores [450]. Measures for
dependence in daily activities include: Activities of Daily Living (ADL; basic activities) and Instrumental Activities
of Daily Living (IADL; activities requiring higher cognition and judgement) [451-453].

5.4.3.5 Shared decision-making


The patient’s own values and preferences should be taken into account as well as the above factors. A shared
decision-making process also involves anticipated changes to quality of life (QoL) functional ability, and a
patient’s hopes, worries and expectations about the future [454]. Particularly in older and frail patients, these
aspects should be given equal importance to disease characteristics during the decision-making process [455].
Older patients may also wish to involve family members, and this is particularly important where cognitive
impairment exists.

5.4.4 Conclusion
Individual life expectancy, health status, frailty, and co-morbidity, not only age, should be central in clinical
decisions on screening, diagnostics, and treatment for PCa. A life expectancy of 10 years is most commonly
used as a threshold for benefit of local treatment. Older men may be undertreated. Patients aged 70 years of
age or older who have frailty should receive a comprehensive geriatric assessment. Resolution of impairments
in vulnerable men allows a similar urological approach as in fit patients.

Table 5.12: G8 screening tool (adapted from [456])

Items Possible responses (score)


A Has food intake declined over the past 3 months 0 = severe decrease in food intake
due to loss of appetite, digestive problems, 1 = moderate decrease in food intake
chewing, or swallowing difficulties? 2 = no decrease in food intake
B Weight loss during the last 3 months? 0 = weight loss > 3 kg
1 = does not know
2 = weight loss between 1 and 3 kg
3 = no weight loss
C Mobility? 0 = bed or chair bound
1 = able to get out of bed/chair but does not go out
2 = goes out
D Neuropsychological problems? 0 = severe dementia or depression
1 = mild dementia
2 = no psychological problems
E BMI? (weight in kg)/(height in m2) 0 = BMI < 19
1 = BMI 19 to < 21
2 = BMI 21 to < 23
3 = BMI > 23
F Takes more than three prescription drugs per 0 = yes
day? 1 = no
G In comparison with other people of the same 0.0 = not as good
age, how does the patient consider his/her health 0.5 = does not know
status? 1.0 = as good
2.0 = better
H Age 0 = > 85
1 = 80-85
2 = < 80
Total score 0-17

48 PROSTATE CANCER - LIMITED UPDATE 2021


Figure 5.3: Decision tree for health status screening (men > 70 years)** [134]

Screening by G8 and mini-COG


TM*

G8 score > 14/17


G8 score ≤ 14/17
no geriatric
a full geriatric evaluation is
evaluation is
mandatory
needed

- Abnormal ADL: 1 or 2 - Abnormal ADL: > 2


- Weight loss 5-10% - Weight loss > 10%
- Comorbidities CISR-G - Comorbidities CISR-G
grades 1-2 grades 3-4

Geriatric assessment
then geriatric intervention

Group 1 Group 2 Group 3


Fit Vulnerable Frail

Mini-COGTM = Mini-COGTM cognitive test; ADLs = activities of daily living; CIRS-G = Cumulative Illness


Rating Score - Geriatrics; CGA = comprehensive geriatric assessment.
*F or Mini-COGTM, a cut-off point of < 3/5 indicates a need to refer the patient for full evaluation of potential
dementia.
**Reproduced with permission of Elsevier, from Boyle H.J., et al. Eur J Cancer 2019:116; 116 [134].

Figure 5.4: The Clinical Frailty Scale version 2.0 [437]*

CLINICAL FRAILTY SCALE 6 LIVING


WITH
People who need help with all outside
activities and with keeping house.
MODERATE Inside, they often have problems with
stairs and need help with bathing and
1 VERY People who are robust, active, energetic
FIT and motivated. They tend to exercise
FRAILTY might need minimal assistance (cuing,
standby) with dressing.
regularly and are among the fittest for
their age.
7 LIVING
WITH
Completely dependent for personal
care, from whatever cause (physical or
2 FIT People who have no active disease
symptoms but are less fit than category
SEVERE
FRAILTY
cognitive). Even so, they seem stable
and not at high risk of dying (within ~6
1. Often, they exercise or are very active months).
occasionally, e.g., seasonally.
8 LIVING Completely dependent for personal care

3 MANAGING People whose medical problems are


WELL well controlled, even if occasionally
WITH VERY
SEVERE
and approaching end of life. Typically,
they could not recover even from a
symptomatic, but often are not FRAILTY minor illness.
regularly active beyond routine walking.
9 TERMINALLY Approaching the end of life. This
4 LIVING
WITH
Previously “vulnerable,” this category
marks early transition from complete
ILL category applies to people with a life
expectancy <6 months, who are not
VERY MILD independence. While not dependent on otherwise living with severe frailty.
(Many terminally ill people can still
FRAILTY others for daily help, often symptoms
exercise until very close to death.)
limit activities. A common complaint
is being “slowed up” and/or being tired
during the day. SCORING FRAILTY IN PEOPLE WITH DEMENTIA
The degree of frailty generally In moderate dementia, recent memory is
5 LIVING
WITH
People who often have more evident
slowing, and need help with high
corresponds to the degree of
dementia. Common symptoms in
very impaired, even though they seemingly
can remember their past life events well.
mild dementia include forgetting They can do personal care with prompting.
MILD order instrumental activities of daily
the details of a recent event, though In severe dementia, they cannot do
FRAILTY living (finances, transportation, heavy still remembering the event itself, personal care without help.
housework). Typically, mild frailty repeating the same question/story
In very severe dementia they are often
progressively impairs shopping and and social withdrawal.
bedfast. Many are virtually mute.
walking outside alone, meal preparation,
medications and begins to restrict light Clinical Frailty Scale ©2005–2020 Rockwood,
housework. Version 2.0 (EN). All rights reserved. For permission:
www.geriatricmedicineresearch.ca
Rockwood K et al. A global clinical measure of fitness
www.geriatricmedicineresearch.ca and frailty in elderly people. CMAJ 2005;173:489–495.

*Permission to reproduce the CFS was granted by the copyright holder.

PROSTATE CANCER - LIMITED UPDATE 2021 49


Table 5.13: Cumulative Illness Score Rating-Geriatrics (CISR-G)

1 Cardiac (heart only)


2 Hypertension (rating is based on severity; affected systems are rated separately)
3 Vascular (blood, blood vessels and cells, marrow, spleen, lymphatics)
4 Respiratory (lungs, bronchi, trachea below the larynx)
5 ENT (eye, ear, nose, throat, larynx)
6 Upper GI (esophagus, stomach, duodenum. Biliar and parcreatic trees; do not include diabetes)
7 Lower GI (intestines, hernias)
8 Hepatic (liver only)
9 Renal (kidneys only)
10 Other GU (ureters, bladder, urethra, prostate, genitals)
11 Musculo-Skeletal-Integumentary (muscles, bone, skin)
12 Neurological (brain, spinal cord, nerves; do not include dementia)
13 Endocrine-Metabolic (includes diabetes, diffuse infections, infections, toxicity)
14 Psychiatric/Behavioural (includes dementia, depression, anxiety, agitation, psychosis)
All body systems are scores on a 0 - 4 scale.
- 0: No problem affecting that system.
- 1: Current mild problem or past significant problem.
- 2: Moderate disability or morbidity and/or requires first line therapy.
- 3: S
 evere problem and/or constant and significant disability and/or hard to control chronic
problems.
- 4: E
 xtremely severe problem and/or immediate treatment required and/or organ failure and/or
severe functional impairment.
Total score 0-56

5.4.5 Guidelines for evaluating health status and life expectancy

Recommendations Strength rating


Use individual life expectancy, health status, and co-morbidity in PCa management. Strong
Use the Geriatric-8, mini-COG and Clinical Frailty Scale tools for health status screening. Strong
Perform a full specialist geriatric evaluation in patients with a G8 score < 14. Strong
Consider standard treatment in vulnerable patients with reversible impairments (after Weak
resolution of geriatric problems) similar to fit patients, if life expectancy is > 10 years.
Offer adapted treatment to patients with irreversible impairment. Weak
Offer symptom-directed therapy alone to frail patients. Strong

6. TREATMENT
This chapter reviews the available treatment modalities, followed by separate sections addressing treatment
for the various disease stages.

6.1 Treatment modalities


6.1.1 Deferred treatment (active surveillance/watchful waiting)
In localised disease a life expectancy of at least 10 years is considered mandatory for any benefit from local
treatment. Data are available on patients who did not undergo local treatment with up to 25 years of follow-
up, with endpoints of OS and CSS. Several series have shown a consistent CSS rate of 82–87% at 10 years
[457-462], and 80–95% for T1/T2 and ISUP grade < 2 PCas [463]. In three studies with data beyond 15 years,
the DSS was 80%, 79% and 58% [459, 461, 462], and two reported 20-year CSS rates of 57% and 32%,
respectively [459, 461]. In addition, many patients classified as ISUP grade 1 would now be classified as ISUP
grade 2–3 based on the 2005 Gleason classification, suggesting that the above-mentioned results should be
considered as minimal. Patients with well-, moderately- and poorly-differentiated tumours had 10-year CSS
rates of 91%, 90% and 74%, respectively, correlating with data from the pooled analysis [463]. Observation
was most effective in men aged 65–75 years with low-risk PCa [464].

50 PROSTATE CANCER - LIMITED UPDATE 2021


Co-morbidity is more important than age in predicting life expectancy in men with PCa. Increasing co-morbidity
greatly increases the risk of dying from non-PCa-related causes and for those men with a short life expectancy.
In an analysis of 19,639 patients aged > 65 years who were not given curative treatment, most men with a CCI
score > 2 had died from competing causes at 10 years follow-up regardless of their age at time of diagnosis.
Tumour aggressiveness had little impact on OS suggesting that patients could have been spared biopsy
and diagnosis of cancer. Men with a CCI score < 1 had a low risk of death at 10 years, especially for well-
or moderately-differentiated lesions [440]. This highlights the importance of assessing co-morbidity before
considering a biopsy.
In screening-detected localised PCa the lead-time bias is likely to be greater. Mortality from
untreated screen-detected PCa in patients with ISUP grade 1–2 might be as low as 7% at 15 years follow-
up [465]. Consequently, approximately 45% of men with PSA-detected PCa are suitable for close follow-up
through a robust surveillance programme. There are two distinct strategies for conservative management that
aim to reduce over-treatment: AS and Watchful waiting (WW) (Table 6.1.1).

6.1.1.1 Definitions
Active surveillance aims to avoid unnecessary treatment in men with clinically localised PCa who do not require
immediate treatment, but at the same time achieve the correct timing for curative treatment in those who
eventually do [466]. Patients remain under close surveillance through structured surveillance programmes with
regular follow-up consisting of PSA testing, clinical examination, mpMRI imaging and repeat prostate biopsies,
with curative treatment being prompted by pre-defined thresholds indicative of potentially life-threatening
disease which is still potentially curable, while considering individual life expectancy.

Watchful waiting refers to conservative management for patients deemed unsuitable for curative treatment
from the outset, and patients are clinically ‘watched’ for the development of local or systemic progression
with (imminent) disease-related complaints, at which stage they are then treated palliatively according to their
symptoms in order to maintain QoL.

Table 6.1.1: Definitions of active surveillance and watchful waiting [465]

Active surveillance Watchful waiting


Treatment intent Curative Palliative
Follow-up Pre-defined schedule Patient-specific
Assessment/markers used DRE, PSA, mpMRI, re-biopsy Not pre-defined, but dependent
on development of symptoms of
progression
Life expectancy > 10 years < 10 years
Aim Minimise treatment-related toxicity Minimise treatment-related toxicity
without compromising survival
Comments Low-risk patients Can apply to patients with all stages
DRE = digital rectal examination; PSA = prostate-specific antigen; mpMRI = multiparametric magnetic
resonance imaging.

6.1.1.2 Active surveillance


No formal RCT is available comparing this modality to standard treatment. The Prostate Testing for Cancer and
Treatment (ProtecT) trial is discussed later as it is not a formal AS strategy but rather active monitoring (AM),
which is a significantly less stringent surveillance strategy in terms of clinical follow-up, imaging and repeat
biopsies [467].

Several cohorts have investigated AS in organ-confined disease, the findings of which were summarised in a
systematic review [468]. More recently, the largest prospective series of men with low-risk PCa managed by AS
was published [469]. Table 6.1.2 summarises the results of selective AS cohorts. It is clear that the long-term
OS and CSS of patients on AS are extremely good. However, more than one-third of patients are ‘reclassified’
during follow-up, most of whom undergo curative treatment due to disease upgrading, increase in disease
extent, disease stage, progression or patient preference. There is considerable variation and heterogeneity
between studies regarding patient selection and eligibility, follow-up policies (including frequency and type
of imaging such as mpMRI imaging, type and frequency of repeat prostate biopsies, such as MRI-targeted
biopsies or transperineal template biopsies, use of PSA kinetics and density, and frequency of clinical follow-
up), when active treatment should be instigated (i.e. reclassification criteria) and which outcome measures
should be prioritised [466]. These will be discussed further in section 6.2.1.

PROSTATE CANCER - LIMITED UPDATE 2021 51


Table 6.1.2: Active surveillance in screening-detected prostate cancer

Studies N Median FU (mo) pT3 in RP patients* 10-year 10-year


OS (%) CSS (%)
Van As, et al. 2008 [470] 326 22 8/18 (44%) 98 100
Carter, et al. 2007 [471] 407 41 10/49 (20%) 98 100
Adamy, et al. 2011 [472] 533-1,000 48 4/24 (17%) 90 99
Soloway, et al. 2010 [473] 99 45 0/2 100 100
Roemeling, et al. 2007 [474] 278 41 - 89 100
Khatami, et al. 2007 [475] 270 63 - n.r. 100
Klotz, et al. 2015 [476] 993 77 - 85 98.1
Tosoian, et al. 2015 [469] 1,818 60 - 93 99.9
Total 4,724-5,191 46.5 - 93 99
* Patients receiving active therapy following initial active surveillance.
CSS = cancer-specific survival; FU = follow-up; mo = months; n = number of patients; n.r. = not reported;
OS = overall survival; RP = radical prostatectomy.

6.1.1.3 Watchful Waiting


6.1.1.3.1 Outcome of watchful waiting compared with active treatment
The SPCG-4 study was a RCT from the pre-PSA era, randomising patients to either WW or RP (Table 6.1.3)
[477]. The study found RP to provide superior CSS, OS and progression-free survival (PFS) compared to
WW at a median follow-up of 23.6 years (range 3 weeks–28 years). The PIVOT trial was a RCT conducted in
the early PSA era and made a similar comparison between RP vs. observation in 731 men (50% with non-
palpable disease) [478] but in contrast to the SPCG-4, it found little to no benefit of RP (cumulative incidence
of all-cause death, RP vs. observation: 68% vs. 73%; RR: 0.92, 95% CI: 0.84–1.01) within a median follow-
up period of 18.6 years (interquartile range, 16.6 to 20 years). Exploratory subgroup analysis showed that
the borderline benefit from RP was most marked for intermediate-risk disease (RR: 0.84, 95% CI: 0.73–0.98)
but there was no benefit in patients with low- or high-risk disease. Overall, no adverse effects on HRQoL
and psychological well-being was apparent in the first years [479]. However, one of the criticisms of the
PIVOT trial is the relatively high overall mortality rate in the WW group (73% at a median follow-up period of
18.6 years (interquartile range, 16.6 to 20.0 years) compared with more contemporary series.

Table 6.1.3: Outcome of SPCG-4 at a median follow-up of 23.6 years [477]

RP (n = 348) (%) Watchful waiting Relative risk p-value


(n = 348) (%) (95% CI)
Disease-specific mortality 19.6 31.3 0.55 (0.41–0.74) < 0.001
Overall mortality 71.9 83.8 0.74 (0.62–0.87) < 0.001
Metastatic progression 26.6 43.3 0.54 (0.42–0.70) < 0.001
CI = confidence interval; RP = radical prostatectomy.

6.1.1.4 The ProtecT study


The ProtecT trial randomised 1,643 patients into three arms: active treatment with either RP or EBRT, and
AM [467]. In this AM schedule patients with a PSA rise of more than 50% in 12 months underwent a repeat
biopsy, but none had systematic repeat biopsies. Fifty-six percent of patients had low-risk disease, with
90% having a PSA < 10 ng/mL, 77% ISUP grade 1 (20% ISUP grade 2–3), and 76% T1c, while the other
patients had mainly intermediate-risk disease. After 10 years of follow-up, CSS was the same between
those actively treated and those on AM (99% and 98.8%, respectively), as was the OS. Only metastatic
progression differed (6% in the AM group as compared to 2.6% in the treated group). The key finding was
that AM was as effective as active treatment at 10 years, at a cost of increased progression and double the
metastatic risk. Metastases remained rare (6%), but more frequent than seen with AS protocols; probably
driven by differences in intensity of monitoring and patient selection. It is important to note that the AM arm
in ProtecT represented an intermediate approach between contemporary AS protocols and WW in terms of
a monitoring strategy based almost entirely on PSA measurements alone; there was no use of mpMRI scan,
either at recruitment or during the monitoring period, nor was there any protocol-mandated repeat prostate
biopsies at regular intervals. In addition, approximately 40% of randomised patients had intermediate-risk
disease. Nevertheless, the ProtecT study has reinforced the role of deferred active treatment (i.e. either AS or
some form of initial AM) as a feasible alternative to active curative interventions in patients with low-grade and

52 PROSTATE CANCER - LIMITED UPDATE 2021


low-stage disease. Beyond 10 years, no data is available, as yet, although AS is likely to give more reassurance
especially in younger men, based on more accurate risk stratification at recruitment and more stringent criteria
regarding follow-up, imaging, repeat biopsy and reclassification. Individual life expectancy must be evaluated
before considering any active treatment in low-risk patients and in those with up to 10 years’ individual life
expectancy.

6.1.2 Radical prostatectomy


6.1.2.1 Introduction
The goal of RP by any approach is the eradication of cancer while, whenever possible, preserving pelvic
organ function [480]. The procedure involves removing the entire prostate with its capsule intact and SVs,
followed by vesico-urethral anastomosis. Surgical approaches have expanded from perineal and retropubic
open approaches to laparoscopic and robotic-assisted techniques; anastomoses have evolved from Vest
approximation sutures to continuous suture watertight anastomoses under direct vision and mapping of the
anatomy of the dorsal venous complex (DVC) and cavernous nerves has led to excellent visualisation and
potential for preservation of erectile function [481]. The main results from multi-centre RCTs involving RP are
summarised in Table 6.1.4.

Table 6.1.4: Oncological results of radical prostatectomy in organ-confined disease in RCTs

Study Acronym Population Treatment Median Risk category CSS (%)


period FU (mo)
Bill-Axelson, et al. SPCG-4 Pre-PSA era 1989-1999 283 Low risk and 80.4
2018 [477] Intermediate risk (at 23 yr.)
Wilt, et al. PIVOT Early years of 1994-2002 152 Low risk 95.9
2017 [482] PSA testing Intermediate risk 91.5
(at 19.5 yr.)
Hamdy, et al. ProtecT Screened 1999-2009 120 Mainly low- and 99
2016 [467] population intermediate risk (at 10 yr.)
CSS = cancer-specific survival; FU = follow-up; mo = months; PSA = prostate-specific antigen; yr. = year.

6.1.2.2 Pre-operative preparation


6.1.2.2.1 Pre-operative patient education
As before any surgery appropriate education and patient consent is mandatory prior to RP. Peri-operative
education has been shown to improve long-term patient satisfaction following RP [483]. Augmentation of
standard verbal and written educational materials such as use of interactive multimedia tools [484, 485] and
pre-operative patient-specific 3D printed prostate models has been shown to improve patient understanding
and satisfaction and should be considered to optimise patient-centred care [486].

Pre-operative pelvic floor exercises


Although many patients who have undergone RP will experience a return to urinary continence [487], temporary
urinary incontinence is common early after surgery, reducing QoL. Pre-operative pelvic floor exercises (PFE)
with, or without, biofeedback have been used with the aim of reducing this early post-operative incontinence.
A systematic review and meta-analysis of the effect of pre-RP PFE on post-operative urinary incontinence
showed a significant improvement in incontinence rates at 3 months post-operatively with an OR of 0.64
(p = 0.005), but not at 1 month or 6 months [488]. Pre-operative PFE may therefore provide some benefit,
however the analysis was hampered by the variety of PFE regimens and a lack of consensus on the definition
of incontinence.

Prophylactic antibiotics
Prophylactic antibiotics should be used; however no high-level evidence is available to recommend specific
prophylactic antibiotics prior to RP (See EAU Urological Infections Guidelines). In addition, as the susceptibility
of bacterial pathogens and antibiotic availability varies worldwide, any use of prophylactic antibiotics should
adhere to local guidelines.

Neoadjuvant androgen deprivation therapy


Several RCTs have analysed the impact of neoadjuvant ADT before RP, most of these using a 3-month
period. The main findings were summarised in a Cochrane review [489]. Neoadjuvant ADT is associated with
a decreased rate of pT3 (downstaging), decreased positive margins, and a lower incidence of positive LNs.
These benefits are greater with increased treatment duration (up to 8 months). However, since neither the PSA

PROSTATE CANCER - LIMITED UPDATE 2021 53


relapse-free survival nor CSS were shown to improve, neoadjuvant ADT should not be considered as standard
clinical practice. One recent RCT compared neoadjuvant luteinising hormone-releasing hormone (LHRH) alone
vs. LHRH plus abiraterone plus prednisone prior to RP in 65 localised high-risk PCa patients [490]. Patients in
the combination arm were found to have both significantly lower tumour volume and significantly lower BCR at
> 4 years follow-up (p = 0.0014). Further supportive evidence is required before recommending combination
neoadjuvant therapy including abiraterone prior to RP.

6.1.2.3 Surgical techniques


Prostatectomy can be performed by open-, laparoscopic- or robot-assisted (RARP) approaches. The initial
open technique of RP described by Young in 1904 was via the perineum [481] but suffered from a lack of
access to pelvic LNs. If lymphadenectomy is required during perineal RP it must be done via a separate
open retropubic (RRP) or laparoscopic approach. The open retropubic approach was popularised by Walsh
in 1982 following his anatomical description of the DVC, enabling its early control and of the cavernous
nerves, permitting a bilateral nerve-sparing procedure [491]. This led to the demise in popularity of perineal
RP and eventually to the first laparoscopic RP reported in 1997 using retropubic principles, but performed
transperitoneally [492]. The initial 9 cases averaged 9.4 hours, an indication of the significant technical and
ergonomic difficulties of the technique. Most recently, RARP was introduced using the da Vinci Surgical
System® by Binder in 2002 [493]. This technology combined the minimally-invasive advantages of laparoscopic
RP with improved surgeon ergonomics and greater technical ease of suture reconstruction of the vesico-urethral
anastomosis and has now become the preferred minimally-invasive approach, when available.

In a randomised phase III trial, RARP was shown to have reduced admission times and blood loss, but not
earlier (12 weeks) functional or oncological outcomes compared to open RP [494]. An updated analysis
with follow-up at 24 months did not reveal any significant differences in functional outcomes between the
approaches [495]. Increased surgical experience has lowered the complication rates of RP and improved
cancer cure [461-464]. Lower rates of positive surgical margins for high-volume surgeons suggest that
experience and careful attention to surgical details, can improve cancer control with RP [496-498]. There is
a lack of studies comparing the different surgical modalities for these longer-term outcomes [460, 479, 482,
499]. A systematic review and meta-analysis of non-RCTs demonstrated that RARP had lower peri-operative
morbidity and a reduced risk of positive surgical margins compared with laparoscopic prostatectomy (LRP),
although there was considerable methodological uncertainty [500]. There was no evidence of differences
in urinary incontinence at 12 months and there was insufficient evidence to draw conclusions based on
differences in cancer-related, patient-driven or erectile dysfunction (ED) outcomes. Another systematic review
and meta-analysis included two small RCTs comparing RARP vs. LRP [501]. The results suggested higher rates
of return of erectile function (RR: 1.51, 95% CI: 1.19–1.92) and return to continence function (RR: 1.14, 95% CI:
1.04–1.24) in the RARP group. However, a recent Cochrane review comparing either RARP or LRP vs. open RP
included two RCTs and found no significant differences between the comparisons for oncological-, urinary- and
sexual function outcomes, although RARP and LRP both resulted in statistically significant improvements in
duration of hospital stay and blood transfusion rates over open RP [502]. Therefore, no surgical approach can
be recommended over another.

Outcome after prostatectomy has been shown to be dependent on both surgeon [503] as well as hospital
volume [504]. Although various volume criteria have been set worldwide, the level of evidence is insufficient to
pinpoint a specific lower volume limit.

6.1.2.3.1 Robotic anterior versus Retzius-sparing dissection


Robot-assisted RP has typically been performed via the anterior approach, first dropping the bladder to expose
the space of Retzius. However, the posterior approach (Retzius-sparing [RS-RARP]) has been used to minimise
injury to support structures surrounding the prostate.
Galfano et al., first described RS-RARP in 2010 [505]. This approach commences dissection
posteriorly at the pouch of Douglas, first dissecting the SVs and progressing caudally behind the prostate.
All of the anterior support structures are avoided, giving rise to the hypothetical mechanism for improved
early post-operative continence. Retzius-sparing-RARP thus offers the same potential advantage as the open
perineal approach, but without disturbance of the perineal musculature.

Retzius-sparing-RARP has been recently investigated in RCTs leading to four systematic reviews and
meta-analyses [506-508] including a 2020 Cochrane systematic review [509] and a large propensity
score matched analysis [510]. The Cochrane review used the most rigorous methodology and analysed
5 RCTs with 502 patients. It found with moderate certainty that RS-RARP improved continence at 1 week
post catheter removal compared to standard RARP (RR: 1.74). Continence may also be improved at 3 months

54 PROSTATE CANCER - LIMITED UPDATE 2021


post-operatively (RR: 1.33), but this was based on low-certainty data. Continence outcomes appeared to
equalise by 12 months (RR: 1.01). These findings matched those of the other systematic reviews. However, a
significant concern was that RS-RARP appears to increase the risk of positive margins (RR: 1.95) but this was
also low-certainty evidence. A single-surgeon propensity score matched analysis of 1,863 patients reached
the same conclusion as the systematic reviews regarding earlier return to continence but did not show data on
margin status [510].
Based on these data, recommendations cannot be made for one technique over another. However,
the trade-offs between the risks of a positive margin vs. earlier continence recovery should be discussed with
prospective patients. Furthermore, no high level evidence is available on high-risk disease with some concerns
that RS-RARP may confer an increased positive margin rate based on pT3 results. In addition, RS-RARP may
be more technically challenging in various scenarios such as anterior tumours, post-TURP, a grossly enlarged
gland, or a bulky median lobe [511].

6.1.2.3.2 Pelvic lymph node dissection


A recent systematic review demonstrated that performing PLND during RP failed to improve oncological
outcomes, including survival [512]. Moreover, a RCT failed to show a benefit of an extended approach vs.
a limited PLND on early oncologic outcomes [513]. However, it is generally accepted that eLND provides
important information for staging and prognosis which cannot be matched by any other currently available
procedure [512].

Extended LND includes removal of the nodes overlying the external iliac artery and vein, the nodes within the
obturator fossa located cranially and caudally to the obturator nerve, and the nodes medial and lateral to the
internal iliac artery. With this template, 94% of patients are correctly staged [514].

The individual risk of patients harbouring positive LNs can be estimated based on validated nomograms. The
Briganti nomogram [381, 382], the Roach formula [515] or the Partin and MSKCC nomograms [516] have shown
similar diagnostic accuracy in predicting LN invasion. These nomograms have all been developed in the pre-
MRI setting based on systematic random biopsy. A risk of nodal metastases over 5% can be used to identify
candidates for nodal sampling by eLND during RP [517-519].

An updated nomogram has been externally validated in men diagnosed based on mpMRI followed by MRI-
targeted biopsy [382]. Based on this nomogram patients can be spared an ePLND if their risk of nodal
involvement is less than 7%; which would result in missing only 1.5% of patients with nodal invasion [382,
520]. This 7% cutoff is comparable to the 5% cutoff of the Briganti nomogram in patients diagnosed by
systematic random biopsy alone. Therefore, this novel nomogram and a 7% threshold should be used after
MRI-targeted biopsy to identify candidates for eLND [521].

6.1.2.3.3 Sentinel node biopsy analysis


The rationale for a sentinel node biopsy (SNB) is based on the concept that a sentinel node is the first
to be involved by migrating tumour cells. Therefore, when this node is negative it is possible to avoid an
ePLND. There is heterogeneity and variation in techniques in relation to SNB (e.g. the optimal tracer) but a
multidisciplinary collaborative endeavour attempted to standardise definitions, thresholds and strategies in
relation to techniques of SNB using consensus methods [522].
Intraprostatic injections of indocyanine green (ICG) have been used to visualise prostate-related
LNs during lymphadenectomy. In a randomised comparison, Harke et al., found more cancer containing LNs in
men that underwent a LN dissection guided by ICG but no difference in BCR at 22.9 month follow-up [523]. A
systematic review showed a sensitivity of 95.2% and NPV of 98.0% for SNB in detecting men with metastases
at eLND [524]. However, there is still insufficient high-quality evidence supporting oncological effectiveness
of SNB for nodal staging. Sentinel node biopsy is therefore still considered as an experimental nodal staging
procedure.

6.1.2.3.4 Prostatic anterior fat pad dissection and histologic analysis


Several multi-centre and large single-centre series have shown the presence of lymphoid tissue within the
fat pad anterior to the endopelvic fascia; the prostatic anterior fat pad (PAFP) [525-531]. This lymphoid tissue
is present in 5.5–10.6% of cases and contains metastatic PCa in up to 1.3% of intermediate- and high-risk
patients.
When positive, the PAFP is often the only site of LN metastasis. The PAFP is therefore a rare but
recognised route of spread of disease. Unlike PLND, there is no morbidity associated with removal of the PAFP.
The PAFP is always removed at RP for exposure of the endopelvic fascia and should be sent for histologic
analysis as per all removed tissue.

PROSTATE CANCER - LIMITED UPDATE 2021 55


6.1.2.3.5 Management of the dorsal venous complex
Since the description of the anatomical open RP by Walsh and Donker in the 1980s, various methods of
controlling bleeding from the DVC have been proposed to optimise visualisation [491]. In the open setting,
blood loss and transfusion rates have been found to be significantly reduced when ligating the DVC prior to
transection [532]. However, concerns have been raised regarding the effect of prior DVC ligation on apical
margin positivity and continence recovery due to the proximity of the DVC to both the prostatic apex and the
urethral sphincter muscle fibres. In the robotic-assisted laparoscopic technique, due to the increased pressure
of pneumoperitoneum, whether prior DVC ligation was used or not, blood loss was not found to be significantly
different in one study [533]. In another study, mean blood loss was significantly less with prior DVC ligation
(184 vs. 176 mL, p = 0.033), however it is debatable whether this was clinically significant [534]. The positive
apical margin rate was not different, however, the latter study showed earlier return to full continence at
5 months post-operatively in the no prior DVC ligation group (61% vs. 40%, p < 0.01).

Ligation of the DVC can be performed with standard suture or using a vascular stapler. One study found
significantly reduced blood loss (494 mL vs. 288 mL) and improved apical margin status (13% vs. 2%) when
using the stapler [535].
Given the relatively small differences in outcomes, the surgeon’s choice to ligate prior to transection
or not, or whether to use sutures or a stapler, will depend on their familiarity with the technique and the
equipment available.

6.1.2.3.6 Nerve-sparing surgery


During prostatectomy, preservation of the neurovascular bundles with parasympathetic nerve branches of the
pelvic plexus may spare erectile function [536, 537].
Although age and pre-operative function may remain the most important predictors for post-
operative erectile function, nerve-sparing has also been associated with improved continence outcomes and
may therefore still be relevant for men with poor erectile function [538, 539]. The association with continence
may be mainly due to the dissection technique used during nerve-sparing surgery, and not due to the
preservation of the nerve bundles themselves [538].
Extra-, inter-, and intra-fascial dissection planes can be planned, with those closer to the prostate
and performed bilaterally associated with superior (early) functional outcomes [540-543]. Furthermore, many
different techniques are propagated such as retrograde approach after anterior release (vs. antegrade), and
athermal and traction-free handling of bundles [544-546]. Nerve-sparing does not compromise cancer control if
patients are carefully selected depending on tumour location, size and grade [547-549].

6.1.2.3.7 Lymph-node-positive patients during radical prostatectomy


Although no RCTs are available, data from prospective cohort studies comparing survival of pN+ patients
(as defined following pathological examination after RP) support that RP may have a survival benefit over
abandonment of RP in node-positive cases [550]. As a consequence there is no role for performing frozen
section of suspicious LNs.

6.1.2.3.8 Removal of seminal vesicles


The more aggressive forms of PCa may spread directly into the SVs. For oncological clearance, the SVs have
traditionally been removed intact with the prostate specimen [551]. However, in some patients the tips of the
SVs can be challenging to dissect free. Furthermore, the cavernous nerves run past the SV tips such that
indiscriminate dissection of the SV tips could potentially lead to ED [552]. However, a RCT comparing nerve-
sparing RP with and without a SV-sparing approach found no difference in margin status, PSA recurrence,
continence or erectile function outcomes. Another study of 71 consecutive RPs showed no cancer in any of
the distal 1 cm of SVs, even in 12 patients with SV invasion [553]. Whilst complete SV removal should be the
default, preservation of the SV tips may be considered in cases of low risk of involvement.

6.1.2.3.9 Techniques of vesico-urethral anastomosis


Following prostate removal, the bladder neck is anastomosed to the membranous urethra. The objective is to
create a precisely aligned, watertight, tension-free, and stricture-free anastomosis that preserves the integrity
of the intrinsic sphincter mechanism. Several methods have been described, based on the direct or indirect
approach, the type of suture (i.e. barbed vs. non-barbed/monofilament), and variation in suturing technique
(e.g., continuous vs. interrupted, or single-needle vs. double-needle running suture). The direct vesico-urethral
anastomosis, which involves the construction of a primary end-to-end inter-mucosal anastomosis of the
bladder neck to the membranous urethra by using 6 interrupted sutures placed circumferentially, has become
the standard method of reconstruction for open RP [554].

56 PROSTATE CANCER - LIMITED UPDATE 2021


The development of laparoscopic- and robotic-assisted techniques to perform RP have facilitated the
introduction of new suturing techniques for the anastomosis. A systematic review and meta-analysis [555]
compared unidirectional barbed suture vs. conventional non-barbed suture for vesico-urethral anastomosis
during robotic-assisted laparoscopic prostatectomy (RALP). The review included 3 RCTs and found significantly
reduced anastomosis time, operative time and posterior reconstruction time in favour of the unidirectional
barbed suture technique, but there were no differences in post-operative leak rate, length of catheterisation
and continence rate. However, no definitive conclusions could be drawn due to the relatively low quality of
the data. In regard to suturing technique, a systematic review and meta-analysis compared continuous vs.
interrupted suturing for vesico-urethral anastomosis during RP [556]. The study included only one RCT with
60 patients [557]. Although the review found slight advantages for continuous suturing over interrupted suturing
in terms of catheterisation time, anastomosis time and rate of extravasation, the overall quality of evidence
was low and no clear recommendations were possible. A recent RCT [558] compared the technique of suturing
using a single absorbable running suture vs. a double-needle single-knot running suture (i.e. Van Velthoven
technique) in laparoscopic RP [559]. The study found slightly reduced anastomosis time with the single running
suture technique, but anastomotic leak, stricture, and continence rates were similar.

Overall, although there are a variety of approaches, methods and techniques for performing the vesico-urethral
anastomosis, no clear recommendations are possible due to the lack of high-certainty evidence. In practice,
the chosen method should be based on surgeon experience and individual preference [554-565].

6.1.2.3.10 Bladder neck management


Bladder neck mucosal eversion
Some surgeons perform mucosal eversion of the bladder neck as its own step in open RP with the aim of
securing a mucosa-to-mucosa vesico-urethral anastomosis and avoiding anastomotic stricture. Whilst bringing
bladder and urethral mucosa together by the everted bladder mucosa covering the bladder muscle layer, this
step may actually delay healing of the muscle layers. An alternative is to simply ensure bladder mucosa is
included in the full thickness anastomotic sutures. A non-randomised study of 211 patients with and without
bladder neck mucosal eversion showed no significant difference in anastomotic stricture rate [566]. The
strongest predictor of anastomotic stricture in RP is current cigarette smoking [567].

Bladder neck preservation


Whilst the majority of urinary continence is maintained by the external urethral sphincter at the membranous
urethra (see below), a minor component is contributed by the internal lissosphincter at the bladder neck [568].
Preservation of the bladder neck has therefore been proposed to improve continence recovery post-RP. A RCT
assessing continence recovery at 12 months and 4 years showed improved objective and subjective urinary
continence in both the short- and long term without any adverse effect on oncological outcome [569]. These
findings were confirmed by a systematic review [570]. However, concern remains regarding margin status for
cancers located at the prostate base.
A systematic review addressing site-specific margin status found a mean base-specific positive
margin rate of 4.9% with bladder neck preservation vs. only 1.9% without [568]. This study was inconclusive,
but it would be sensible to exercise caution when considering bladder neck preservation if significant cancer is
known to be at the prostate base. Bladder neck preservation should be performed routinely when the cancer
is distant from the base. However, bladder neck preservation cannot be performed in the presence of a large
median lobe or a previous TURP.

6.1.2.3.11 Urethral length preservation


The membranous urethra sits immediately distal to the prostatic apex and is chiefly responsible, along with its
surrounding pelvic floor support structures, for urinary continence. It consists of the external rhabdosphincter
which surrounds an inner layer of smooth muscle. Using pre-operative MRI, the length of membranous urethra
has been shown to vary widely. A systematic review and meta-analysis has found that every extra millimetre
of membranous urethral length seen on MRI pre-operatively improves early return to continence post-RP
[571]. Therefore, it is likely that preservation of as much urethral length as possible during RP will maximise
the chance of early return to continence. It may also be useful to measure urethral length pre-operatively to
facilitate councelling of patients on their relative likelihood of early post-operative continence.

6.1.2.3.12 Cystography prior to catheter removal


Cystography may be used prior to catheter removal to check for a substantial anastomotic leak. If such a
leak is found, catheter removal may then be deferred to allow further healing and sealing of the anastomosis.
However, small comparative studies suggest that a cystogram to assess anastomotic leakage is not indicated
as standard of care before catheter removal 8 to 10 days after surgery [572]. If a cystogram is used, men with

PROSTATE CANCER - LIMITED UPDATE 2021 57


LUTS, large prostates, previous TURP or bladder neck reconstruction, may benefit as these factors have been
associated with leakage [573, 574]. Contrast-enhanced transrectal US is an alternative [575].

6.1.2.3.13 Urinary catheter


A urinary catheter is routinely placed during RP to enable bladder rest and drainage of urine while the
vesicourethral anastomosis heals. Compared to a traditional catheter duration of around 1 week, some centres
remove the transurethral catheter early (post-operative day 2–3), usually after thorough anastomosis with
posterior reconstruction or in patients selected peri-operatively on the basis of anastomosis quality [576-579].
No higher complication rates were found. Although shorter catheterisation has been associated with more
favourable short-term functional outcomes, no differences in long-term function were found [580]. One RCT has
shown no difference in rate of UTI following indwelling catheter (IDC) removal whether prophylactic ciprofloxacin
was given prior to IDC removal or not, suggesting antibiotics should not be given at catheter removal [581].
As an alternative to transurethral catheterisation, suprapubic catheter insertion during RP has been
suggested. Some reports suggest less bother regarding post-operative hygiene and pain [582-586], while
others did not find any differences [587, 588]. No impact on long-term functional outcomes were seen.

6.1.2.3.14 Use of a pelvic drain


A pelvic drain has traditionally been used in RP for potential drainage of urine leaking from the vesico-urethral
anastomosis, blood, or lymphatic fluid when a PLND has been performed. Two RCTs in the robotic-assisted
laparoscopic setting have been performed [589, 590]. Patients with urine leak at intra-operative anastomosis
watertight testing were excluded. Both trials showed non-inferiority in complication rates when no drain was
used. When the anastomosis is found to be watertight intra-operatively, it is reasonable to avoid inserting a
pelvic drain. There is no evidence to guide usage of a pelvic drain in PLND.

6.1.2.4 Acute and chronic complications of surgery


Post-operative incontinence and ED are common problems following surgery for PCa. A key consideration is
whether these problems are reduced by using newer techniques such as RALP. Recent systematic reviews
have documented complication rates after RALP [500, 591-594], and can be compared with contemporaneous
reports after radical RRP [595]. Recently, a prospective controlled non-RCT of patients undergoing RP in
14 centres using RALP or RRP was published [596]. At 12 months after RALP, 21.3% of patients were
incontinent, as were 20.2% after RRP. The adjusted OR was 1.08 (95% CI: 0.87–1.34). Erectile dysfunction was
observed in 70.4% after RALP and 74.7% after RRP. The adjusted OR was 0.81 (95% CI: 0.66–0.98) [596].
A RCT comparing RALP and RRP reported outcomes at 12 weeks in 326 patients and functional
outcomes at 2 years [494]. Urinary function scores did not differ significantly between RRP vs. RALP at 6 and
12 weeks post-surgery (74–50 vs. 71–10, p = 0.09; 83–80 vs. 82–50, p = 0.48), with comparable outcomes
for sexual function scores (30–70 vs. 32–70, p = 0.45; 35–00 vs. 38–90, p = 0.18). In the RRP group 14
(9%) patients had post-operative complications vs. 6 (4%) in the RALP group. The intra-and peri-operative
complications of retropubic RP and RALP are listed in Table 6.1.5. The early use of phosphodiesterase-5
(PDE5) inhibitors in penile rehabilitation remains controversial resulting in a lack of clear recommendations (see
Section 8.3.2).

6.1.2.4.1 Effect of anterior and posterior reconstruction on continence


Preservation of integrity of the external urethral sphincter is critical for continence post-RP. Less clear is the
effect of reconstruction of surrounding support structures to return to continence. Several small RCTs have
been conducted, however, pooling analyses is hampered by variation in the definitions of incontinence and
surgical approach, such as open vs. robotic and intraperitoneal vs. extraperitoneal. In addition, techniques
used to perform both anterior suspension or reconstruction and posterior reconstruction are varied. For
example, anterior suspension is performed either through periosteum of the pubis or the combination of ligated
DVC and puboprostatic ligaments (PPL). Posterior reconstruction from rhabdosphincter is described to either
Denonvilliers fascia posterior to bladder or to posterior bladder wall itself.
Two trials assessing posterior reconstruction in RALRP found no significant improvement in return
to continence [597, 598]. A third trial using posterior bladder wall for reconstruction showed only an earlier
return to 1 pad per day (median 18 vs. 30 days, p = 0.024) [599]. When combining both anterior and posterior
reconstruction, where for anterior reconstruction the PPL were sutured to the anterior bladder neck, another
RCT found no improvement compared to a standard anastomosis with no reconstruction [600].

Four RCTs including anterior suspension have also shown conflicting results. Anterior suspension alone
through the pubic periosteum, in the setting of extraperitoneal RALRP, showed no advantage [601]. However,
when combined with posterior reconstruction in RRP, one RCT showed significant improvement in return to
continence at one month (7.1% vs. 26.5%, p = 0.047) and 3 months (15.4% vs. 45.2%, p = 0.016), but not

58 PROSTATE CANCER - LIMITED UPDATE 2021


at 6 months (57.9% vs. 65.4%, p = 0.609) [602]. Another anterior plus posterior reconstruction RCT using the
Advanced Reconstruction of VesicoUrethral Support (ARVUS) technique and the strict definition of continence
of ‘no pads‘, showed statistically significant improvement in continence at 2 weeks (43.8% vs. 11.8%), 4 weeks
(62.5% vs. 14.7%), 8 weeks (68.8% vs. 20.6%), 6 months (75% vs. 44.1%) and 12 months (86.7% vs. 61.3%),
when compared to standard posterior Rocco reconstruction [603]. Anterior suspension alone through the DVC
and PPL combined without posterior construction in the setting of RRP has shown improvement in continence
at one month (20% vs. 53%, p = 0.029), 3 months (47% vs. 73%, p = 0.034) and 6 months (83% vs. 100%,
p = 0.02), but not at 12 months (97% vs. 100%, p = 0.313) [604]. Together, these results suggest a possible
earlier return to continence, but no long-term difference.

As there is conflicting evidence on the effect of anterior and/or posterior reconstruction on return to continence
post-RP, no recommendations can be made. However, no studies showed an increase in adverse oncologic
outcome or complications with reconstruction.

6.1.2.4.2 Deep venous thrombosis prophylaxis


For EAU Guidelines recommendations on post-RP deep venous thrombosis prophylaxis, please see the
Thromboprophylaxis Guidelines Section 3.1.6 [605]. However these recommendations should be adapted
based on national recommendations, when available.

Table 6.1.5: Intra-and peri-operative complications of retropubic RP and RALP (Adapted from [500])

Predicted probability of event RALP (%) Laparoscopic RP (%) RRP (%)


Bladder neck contracture 1.0 2.1 4.9
Anastomotic leak 1.0 4.4 3.3
Infection 0.8 1.1 4.8
Organ injury 0.4 2.9 0.8
Ileus 1.1 2.4 0.3
Deep-vein thrombosis 0.6 0.2 1.4
Predicted rates of event RALP (%) Laparoscopic RP (%) RRP (%)
Clavien I 2.1 4.1 4.2
Clavien II 3.9 7.2 17.5
Clavien IIIa 0.5 2.3 1.8
Clavien IIIb 0.9 3.6 2.5
Clavien Iva 0.6 0.8 2.1
Clavien V < 0.1 0.2 0.2
RALP = robot-assisted laparoscopic prostatectomy; RP = radical prostatectomy; RRP = radical retropubic
prostatectomy.

6.1.2.4.3 Early complications of extended lymph node dissection


Pelvic eLND increases morbidity in the treatment of PCa [512]. Overall complication rates of 19.8% vs.
8.2% were noted for eLND vs. limited LND, respectively, with lymphoceles (10.3% vs. 4.6%) being the most
common adverse event. Other authors have reported more acceptable complication rates [606]. Similar rates of
lymphoceles have been observed in RALP series; however, in one subgroup analysis lymphoceles were more
common with the extraperitoneal approach (19%) vs. the transperitoneal approach (0%) [607, 608]. Briganti
et al. [609] also showed more complications after extended compared to limited LND. Twenty percent of men
suffer a complication of some sort after eLND. Thromboembolic events occur in less than 1% of cases.

6.1.3 Radiotherapy
Intensity-modulated radiotherapy (IMRT) with image-guided radiotherapy (IGRT) is currently widely recognised
as the best available approach for EBRT.

6.1.3.1 External beam radiation therapy


6.1.3.1.1 Technical aspects: intensity-modulated external-beam radiotherapy and volumetric arc external-beam
radiotherapy
Intensity-modulated external-beam radiotherapy and volumetric arc external-beam radiotherapy (VMAT)
employ dynamic multileaf collimators, which automatically and continuously adapt to the contours of the target
volume seen by each beam. Viani et al., show significantly reduced acute and late grade > 2 GU and GI toxicity
in favour of IMRT, while BCR-free rates did not differ significantly when comparing IMRT with three-dimensional

PROSTATE CANCER - LIMITED UPDATE 2021 59


conformal radiation therapy (3D-CRT) in a RCT including 215 patients [610]. A meta-analysis by Yu et al.,
(23 studies, 9,556 patients) concluded that IMRT significantly decreases the occurrence of grade 2–4 acute GI
toxicity, late GI toxicity and late rectal bleeding, and achieves better PSA relapse-free survival in comparison
with 3D-CRT. Intensity-modulated external-beam radiotherapy and 3D-CRT show comparable acute rectal
toxicity, late GU toxicity and OS, while IMRT slightly increases the morbidity of acute GU toxicity [611]. Wortel
et al., concluded that, as compared to 3D-CRT, image-guided IMRT was associated with significantly reduced
late GI toxicity whereas GU toxicities remained comparable (242 IMRT patients vs. 189 3D-CRT patients) [612].
Finally, Zapatero et al., found, based on 733 consecutive patients (295 IMRT vs. 438 3D-CRT), that compared
with 3D-CRT, high-dose IMRT/IGRT is associated with a lower rate of late urinary complications despite a
higher radiation dose [613]. In conclusion, IMRT plus IGRT remain the standard of care for the treatment of
PCa.

The advantage of VMAT over IMRT is shorter treatment times, generally two to three minutes. Both techniques
allow for a more complex distribution of the dose to be delivered and provide concave isodose curves, which
are particularly useful as a means of sparing the rectum. Radiotherapy treatment planning for IMRT and VMAT
differs from that used in conventional EBRT, requiring a computer system capable of ‘inverse planning’ and
the appropriate physics expertise. Treatment plans must conform to pre-specified dose constraints to critical
organs at risk of normal tissue damage and a formal quality assurance process should be routine.
With dose escalation using IMRT/VMAT organ movement becomes a critical issue in terms of both
tumour control and treatment toxicity. Techniques will therefore combine IMRT/VMAT with some form of IGRT
in which organ movement can be visualised and corrected for in real time, although the optimum means of
achieving this is still unclear [614]. Tomotherapy is another technique for the delivery of IMRT, using a linear
accelerator mounted on a ring gantry that rotates as the patient is delivered through the centre of the ring,
analogous to spiral CT scanning.

6.1.3.1.2 Dose escalation


Local control is a critical issue for the outcome of radiotherapy of PCa. It has been shown that local failure
due to insufficient total dose is prognostic for death from PCa as a second wave of metastases is seen later
on [615]. Several RCTs have shown that dose escalation (range 74–80 Gy) has a significant impact on 10-year
biochemical relapse as well as metastases and disease-specific mortality [616-623]. These trials have generally
included patients from several risk groups, and the use of neoadjuvant/adjuvant HT has varied (see Table
6.1.6). The best evidence of an OS benefit in patients with intermediate- or high-risk PCa, but not with low-risk
PCa, derives from a non-randomised but well conducted propensity-matched retrospective analysis of the U.S.
National Cancer Database including a total of 42,481 patients [624]. In everyday practice, a minimum dose
of > 74 Gy is recommended for EBRT plus HT, with no different recommendations according to the patient’s
risk group. If IMRT and IGRT are used for dose escalation, rates of severe late side effects (> grade 3) for the
rectum are 2–3% and for the GU tract 2–5% [618, 621, 625-638].

Table 6.1.6: Randomised trials of dose escalation in localised PCa

Trial n PCa condition Radiotherapy Follow-up Outcome Results


Dose (median)
MD Anderson 301 T1-T3, N0, M0, 70 vs.78 Gy 15 yr. DM, DSM, All patients:
study 2011 PSA 10 ng/mL FFF 18.9% FFF at 70 Gy
[623] vs. 12% FFF at 78 Gy
PSA > 10 ng/mL (p = 0.042)
3.4% DM at 70 Gy
1.1% DM at 78 Gy
(p = 0.018)
6.2% DSM at 70 Gy
3.2% DSM at 78 Gy
(p = 0.043)
No difference in OS
(p > 0.05)

60 PROSTATE CANCER - LIMITED UPDATE 2021


PROG 95-09 393 T1b-T2b 70.2 vs.79.2 Gy 8.9 yr. 10-yr. All patients:
2010 [617] PSA 15 ng/mL including ASTRO BCF 32% BF at 70.2 Gy
75% proton boost 17% BF at 79.2 Gy
19.8 vs. 28.8 Gy (p < 0.0001)
Low-risk patients:
28% BF at 70.2 Gy
7% BF at 79.2 Gy
(p < 0.0001)
MRC RT01 843 T1b-T3a, N0, M0 64 vs. 74 Gy 10 yr. BFS, OS 43% BFS at 64 Gy
2014 [622] PSA < 50 ng/mL 55% BFS at 74 Gy
neoadjuvant HT (p = 0.0003) 71%
OS both groups
(p = 0.96)
Dutch 664 T1b-T4 143 pts. 68 vs. 78 Gy 110 mo. Freedom 43% FFF at 68 Gy
randomised with (neo) biochemical 49% FFF at 78 Gy
phase III trial adjuvant HT (Phoenix) (p = 0.045)
2014 [621] and/or clinical
failure at
10 yr.
GETUG 06 306 T1b-T3a, N0, M0 70 vs. 80 Gy 61 mo. BCF (ASTRO) 39% BF at 70 Gy
2011 [620] PSA < 50 ng/mL 28% BF at 80 Gy
RTOG 0126 1,532 T1b-T2b 70.2 vs. 79.2 Gy 100 mo. OS, DM, BCF 75% OS at 70.2 Gy
2018 [616] ISUP grade 1 (ASTRO) 76% OS at 79.2 Gy
+ PSA 10-20 6% DM at 70.2 Gy
ng/mL or ISUP 4% DM at 79.2 Gy
grade 2/3 + (p = 0.05)
PSA < 15 ng/mL 47% BCF at 70.2 Gy
31% BCF at 79.2 Gy
(p < 0.001;
Phoenix, p < 0.001)
BCF = biochemical failure; BFS = biochemical progression-free survival; DM = distant metastases;
DSM = disease specific mortality; FFF = freedom from biochemical or clinical failure; HT = hormone therapy;
mo. = months; n = number of patients; OS = overall survival; PSA = prostate-specific antigen; yr. = year.

6.1.3.1.3 Hypofractionation
Fractionated RT utilises differences in the DNA repair capacity of normal and tumour tissue and slowly
proliferating cells are very sensitive to an increased dose per fraction [639]. A meta-analysis of 25 studies
including > 14,000 patients concluded that since PCa has a slow proliferation rate, hypofractionated RT
could be more effective than conventional fractions of 1.8–2 Gy [640]. Hypofractionation (HFX) has the added
advantage of being more convenient for the patient at lower cost.
Moderate HFX is defined as RT with 2.5–3.4 Gy/fx. Several studies report on moderate HFX applied
in various techniques also including HT in part [641-651]. A systematic review concluded that studies on
moderate HFX (2.5–3.4 Gy/fx) delivered with conventional 3D-CRT/IMRT have sufficient follow-up to support
the safety of this therapy but long-term efficacy data are still lacking [650]. These results were confirmed by
a recent Cochrane review on moderate HFX for clinically localised PCa [652]. Eleven studies were included
(n = 8,278) with a median follow-up of 72 months showing little or no difference in PCa-specific survival
(HR: 1.00). Based on 4 studies (n = 3,848), hypofractionation probably makes little or no difference to late
radiation GU toxicity (RR: 1.05) or GI toxicity (RR: 1.1), but this conclusion is based on relatively short follow-
up, and 10 to 15-year data will be required to confirm these findings [652].

Moderate HFX should only be done by experienced teams using high-quality EBRT using IGRT and IMRT and
published phase III protocols should be adhered to (see Table 6.1.7 below).

PROSTATE CANCER - LIMITED UPDATE 2021 61


Table 6.1.7: Major phase III randomised trials of moderate hypofractionation for primary treatment

Study/ n Risk, ISUP ADT RT Regimen BED, Gy Median Outcome


Author grade, or FU, mo
NCCN
Lee, et al. 550 low risk None 70 Gy/28 fx 80 70 5 yr. DFS 86.3%
2016 [645] 542 73.8 Gy/41 fx 69.6 (n.s.)
5 yr. DFS 85.3%
Dearnaley, 1077/19 fx 15% low 3-6 mo. 57 Gy/19 fx 73.3 62 5 yr. BCDF
et al. CHHiP 1074/20 fx 73% before 60 Gy/20 fx 77.1 85.9% (19 fx)
2012 [641] 1065/37 fx intermediate and 74 Gy/37 fx 74 90.6% (20 fx)
and 2016 12% high during 88.3% (37 fx)
[646] EBRT
De Vries, 403 30% ISUP None 64.6 Gy/19 fx 90.4 89 8-yr. OS 80.8%
et al. 2020 392 grade 1 78 Gy/39 fx 78 vs. 77.6%
[647, 653] 45% ISUP (p = 0.17)
grade 2-3,
25% ISUP 8 yr. TF 24.4%
grade 4-5 vs. 26.3%
Catton, et al. 608 intermediate None 60 Gy/20 fx 77.1 72 5 yr. BCDF both
2017 [649] risk arms 85%
53% T1c HR: 0.96 (n.s)
46% T2a-c
598 9% ISUP 78 Gy/39 fx 78
grade 1
63% ISUP
grade 2
28% ISUP
grade 3
ADT = androgen deprivation therapy; BCDF = biochemical or clinical disease failure; BED = biologically
equivalent dose, calculated to be equivalent in 2 Gy fractions using an α/ß of 1.5 Gy; DFS = disease-
free survival; EBRT = external beam radiotherapy; FU = follow-up; fx = fractions; HR = hazard ratio;
ISUP = International Society of Urological Pathology; mo. = month; n = number of patients; NCCN = National
Comprehensive Cancer Network; n.s. = not significant; TF = treatment failure; yr. = year.

Ultra-HFX has been defined as radiotherapy with > 3.4 Gy per fraction [651]. It requires IGRT and stereotactic
body radiotherapy (SBRT). Table 6.1.8 provides an overview of selected studies. Short-term biochemical
control is comparable to conventional fractionation. However, there are concerns about high-grade GU
and rectal toxicity and long-term side effects may not all be known yet [650, 654, 655]. In the HYPO-RT-PC
randomised trial by Widmark et al. (n = 1,200), no difference in failure-free survival was seen for conventional or
ultra-HFX but acute Grade > 2 GU toxicity was 23% vs. 28% (p = 0.057), favouring conventional fractionation.
There were no significant differences in long-term toxicity [656]. A systematic review by Jackson et al. included
38 studies with 6,116 patients who received RT with < 10 fractions and > 5 Gy per fraction. Five and 7-year
BRFS rates were 95.3% and 93.7%, respectively, and estimated late grade > 3 GU and GI toxicity rates were
2.0% and 1.1%, respectively [657]. The authors conclude that there is sufficient evidence to support SBRT
as a standard treatment option for localised PCa, even though the median follow-up in this review was only
39 months and it included at least one trial (HYPO-RT-PC) which used ‘conventional’ IMRT/VMAT for ultra-
HFX. In their review on SBRT, Cushman and co-workers evaluated 14 trials, including 2,038 patients and
concluded that despite a lack of long-term follow-up and the heterogeneity of the available evidence, prostate
SBRT affords appropriate biochemical control with few high-grade toxicities [658]. In the Intensity-modulated
fractionated radiotherapy vs. stereotactic body radiotherapy for PCa (PACE-B) trial, acute grade > 2 GU or GI
toxicities did not differ significantly between conventional fractionation and ultra-HFX [659]. Therefore, it seems
prudent to restrict extreme HFX to prospective clinical trials and to inform patients on the uncertainties of the
long-term outcome.

62 PROSTATE CANCER - LIMITED UPDATE 2021


Table 6.1.8: Selected trials on ultra-hypofractionation for intact localised PCa

Reference n med FU Risk-Group Regimen (TD/fx) Outcome


(mo)
Widmark et al. 1,200 60 89% intermediate 78 Gy / 39 fx, 8 w FFS at 5 yr.
2019 11% high 42.7 Gy / 7 fx, 2.5 w 84% in both arms
HYPO-RT-PC No SBRT
[656]
Brand et al. 2019 847 variable 8% low 78 Gy / 39 fx, 8 w Grade > 2 acute GI
PACE-B [659] 92% intermediate 36.25 Gy / 5 fx, 1-2 w 12% vs. 10%, p = 0.38
SBRT Grade > 2 acute GU
27% vs. 23%, p = 0.16
FFS = failure-free survival; FU = follow-up; fx = number fractions; GI = gastrointestinal toxicity;
GU = genitourinary toxicity; mo. = months; n = number of patients; TD = total dose; SBRT = stereotactic body
radiotherapy; w = weeks, yr. = years.

6.1.3.1.4 Neoadjuvant or adjuvant hormone therapy plus radiotherapy


The combination of RT with LHRH ADT has definitively proven its superiority compared with RT alone followed
by deferred ADT on relapse, as shown by phase III RCTs [660-664] (Table 6.1.9). The main message is that for
intermediate-risk disease a short duration of around 6 months is optimal while a longer one, around 3 years, is
needed for high-risk patients.
A meta-analysis based on individual patient data from two RCTs (RTOG 9413 and Ottawa 0101)
has compared neoadjuvant/concomitant vs. adjuvant ADT in conjunction with prostate RT and reported
superior PFS with adjuvant ADT [665]. This is an important observation, which should influence future clinical
trial design and evaluation of outcomes. However, there are differences between the two trials in patient
characteristics, exact scheduling of neoadjuvant +/- concomitant ADT, hormonal preparation, and RT schedule.
At present, either neoadjuvant or adjuvant ADT remain acceptable options for patients requiring short-term
ADT in conjunction with EBRT.

Table 6.1.9: Selected studies of use and duration of ADT in combination with RT for PCa

Trial TNM stage n Trial ADT RT Effect on OS


RTOG 85-31 T3 or N1 M0 977 EBRT ± ADT Orchiectomy or 65-70 Gy Significant benefit for
2005 [661] LHRH agonist combined treatment
15% RP (p = 0.002) seems to
be mostly caused by
patients with ISUP
grade 2-5
RTOG 94-13 T1c-4 N0-1 1,292 ADT timing 2 mo. Whole pelvic No significant difference
2007 [666] M0 comparison neoadjuvant RT vs. between neoadjuvant
plus prostate plus concomitant vs.
concomitant vs. only; 70.2 Gy adjuvant androgen
4 mo. adjuvant suppression therapy
suppression groups (interaction
suspected)
RTOG 86-10 T2-4 N0-1 456 EBRT ± ADT Goserelin plus 65-70 Gy RT No significant difference
2008 [662] flutamide 2 mo. at 10 yr.
before, plus
concomitant
therapy
D’Amico AV, T2 N0 M0 206 EBRT ± ADT LHRH agonist 70 Gy Significant benefit (HR:
et al. 2008 (localised plus flutamide 3D-CRT 0·55, 95% CI: 0.34-0.90,
[663] unfavourable for 6 mo. p = 0.01) that may
risk) pertain only to men
with no or minimal
co-morbidity

PROSTATE CANCER - LIMITED UPDATE 2021 63


RTOG 92-02 T2c-4 N0-1 1554 Short vs. LHRH agonist 65-70 Gy p = 0.73,
2008 [667] M0 prolonged given for 2 yr. p = 0.36 overall;
ADT as adjuvant significant benefit
after 4 mo. as (p = 0.044) (p = 0.0061)
neoadjuvant in subset with ISUP
grade 4-5
EORTC T1c-2ab N1 970 Short vs. LHRH agonist 70 Gy Better result with 3 yr.
22961 2009 M0, T2c-4 prolonged for 6 mo. vs. 3D-CRT treatment than with 6 mo.
[668] N0-1 M0 ADT 3 yr. (3.8% improvement in
survival at 5 yr.)
EORTC T1-2 poorly 415 EBRT ± ADT LHRH agonist 70 Gy RT Significant benefit at
22863 2010 differentiated for 3 yr. 10 yr. for combined
[660] and M0, or (adjuvant) treatment (HR: 0.60,
T3-4 N0-1 M0 95% CI: 0.45-0.80,
p = 0.0004).
TROG T2b-4 N0 M0 802 Neoadjuvant Goserelin 66 Gy No significant difference
96-01 2011 ADT plus flutamide 3D-CRT in OS reported; benefit
[664] Duration 3 or 6 mo. in PCa-specific survival
before, plus (HR: 0.56, 95% CI:
concomitant 0.32-0.98, p = 0.04)
suppression (10 yr.: HR: 0.84,
0.65-1.08, p = 0.18)
RTOG 99-10 intermediate 1,579 Short vs. LHRH agonist 70.2 Gy 67 vs. 68%, p = 0.62,
2015 [669] risk prolonged 8 + 8 vs. 8 + 28 2D/3D confirms 8 + 8 wk.
94% T1-T2, ADT wk. LHRH as a standard
6% T3-4
ADT = androgen deprivation therapy; CI = confidence interval; EBRT = external beam radiotherapy in standard
fractionation; HR = hazard ratio; LHRH = luteinising hormone-releasing hormone; mo. = months; n = number of
patients; OS = overall survival; RP = radical prostatectomy; RT = radiotherapy; wk = week; yr. = year.

The question of the added value of EBRT combined with ADT has been clarified with 3 RCTs. All showed a
clear benefit of adding EBRT to long-term ADT (see Table 6.1.10).

Table 6.1.10: Selected studies of ADT in combination with, or without, RT for PCa

Trial TNM stage n Trial design ADT RT


Effect on OS
SPCG-7/ T1b-2 WHO 875 ADT ± EBRT LHRH agonist 34% (95% CI: 29-39%)
70 Gy
SFUO-3 Grade 1-3, for 3 mo. plus vs. 17% (95% CI: 13-22%
3D-CRT
2016 [670] T3 N0 M0 continuous CSM at 12 (15) yr. favouring
vs. no RT
flutamide combined treatment
(p < 0.0001 for 15-yr.
results)
NCIC CTG PR.3/MRC
PRO7/NCIC T3-4 (88%), 1,205 ADT ± EBRT Continuous 65-70 Gy 10-yr. OS = 49% vs. 55%
2011 [671] PSA > 20 ng/mL LHRH agonist 3D-CRT favouring combined
and 2015 (64%), ISUP vs. no RT treatment HR: 0.7,
[672] grade 4-5 (36%) p < 0.001)
N0 M0
Sargos, et al. T3-4 N0 M0 273 ADT ± EBRT LHRH agonist 70 Gy Significant reduction of
2020 [673] for 3 yr. 3D-CRT clinical progression; 5-yr.
vs. no RT OS 71.4% vs. 71.5%
ADT = androgen deprivation therapy; CSM = cancer-specific mortality; EBRT = external beam radiotherapy;
HR = hazard ratio; LHRH = luteinising hormone-releasing hormone; mo. = months; n = number of patients;
OS = overall survival; RT = radiotherapy; 3D-CRT = three-dimensional conformal radiotherapy.

6.1.3.1.5 Combined dose-escalated radiotherapy and androgen-deprivation therapy


Zelefsky et al., reported a retrospective analysis comprising 571 patients with low-risk PCa; 1,074 with
intermediate-risk PCa and 906 with high-risk PCa. Three-dimensional conformal RT or IMRT were administered
[674]. The prostate dose ranged from 64.8 to 86.4 Gy; doses beyond 81 Gy were delivered during the last

64 PROSTATE CANCER - LIMITED UPDATE 2021


10 years of the study using image-guided IMRT. Complete androgen blockade was administered at the
discretion of the treating physician to 623 high-risk (69%), 456 intermediate-risk (42%) and 170 low-risk (30%)
PCa patients. The duration of ADT was 3 months for low-risk patients and 6 months for intermediate-risk and
high-risk patients, starting at 3 months before RT. The 10-year biochemical disease-free rate was significantly
improved by dose escalation: above 75.6 Gy in low-risk, and above 81 Gy for the intermediate- and high-
risk groups. It was also improved by adding 6 months of ADT in intermediate- and high-risk patients. In the
multivariate analysis, neither the dose > 81 Gy, nor adding ADT, influenced OS. Three RCTs have shown that
the benefits of ADT are independent of dose escalation, and that the use of ADT would not compensate for a
lower radiotherapy dose:
1. The GICOR study shows a better biochemical DFS in high-risk patients for 3D-CRT radiation dose > 72 Gy
when combined with long-term ADT [630].
2. DART01/05 GICOR shows improved biochemical control and OS in high-risk patients if 2 years of
adjuvant ADT is combined with high-dose RT [675].
3. EORTC trial 22991 shows that 6 months ADT improves biochemical and clinical DFS irrespective of the
dose (70, 74, 78 Gy) in intermediate-risk and low-volume high-risk localised PCa patients [676].

6.1.3.2 Proton beam therapy


In theory, proton beams are an attractive alternative to photon-beam RT for PCa, as they deposit almost all
their radiation dose at the end of the particle’s path in tissue (the Bragg peak), in contrast to photons which
deposit radiation along their path. There is also a very sharp fall-off for proton beams beyond their deposition
depth, meaning that critical normal tissues beyond this depth could be effectively spared. In contrast, photon
beams continue to deposit energy until they leave the body, including an exit dose.
One RCT on dose escalation (70.2 vs. 79.2 Gy) has incorporated protons for the boost doses
of either 19.8 or 28.8 Gy. This trial shows improved outcome with the higher dose but it cannot be used as
evidence for the superiority of proton therapy [617]. Thus, unequivocal information showing an advantage of
protons over IMRT photon therapy is still not available. Studies from the SEER database and from Harvard
describing toxicity and patient-reported outcomes do not point to an inherent superiority for protons [677, 678].
In terms of longer-term GI toxicity, proton therapy might even be inferior to IMRT [678].
A RCT comparing equivalent doses of proton-beam therapy with IMRT is underway. Meanwhile,
proton therapy must be regarded as an experimental alternative to photon-beam therapy.

6.1.3.3 Spacer during external beam radiation therapy


Biodegradable spacer insertion involves using a liquid gel or balloon to increase the distance between the
prostate and rectum and consequently reduce the amount of radiation reaching the rectum. Various materials
have been used with most evidence available for CE-marked hydrogel spacers [679]. A meta-analysis including
one RCT and six cohort studies using the hydrogel spacer demonstrated a 5–8% reduction in the rectal volume
receiving high-dose radiation, although heterogeneity between studies is found [680]. In the final analysis of
the RCT with a median follow-up of 37 months and with approximately two-thirds of patients evaluable, those
treated with spacer in situ had no deterioration from baseline bowel function whilst those treated without
spacer had a lower mean bowel summary score of 5.8 points which met the threshold for minimally important
difference of 4–6 points [681].
This meta-analysis highlights inconsistent reporting of procedural complications. In addition, with
more widespread clinical use safety reports describe uncommon, but severe and life changing, complications
including prostatic abscess, fistulae and sepsis [682]. Implantation is associated with a learning curve and
should only be undertaken by teams with experience of TRUS and transperineal procedures with robust audit
reporting in place [683]. Its role in the context of moderate or extreme hypofractionation is as yet unclear.

6.1.3.4 Brachytherapy
6.1.3.4.1 Low-dose rate (LDR) brachytherapy
Low-dose rate brachytherapy uses radioactive seeds permanently implanted into the prostate. There is a
consensus on the group of patients with the best outcomes after LDR monotherapy [684]: Stage cT1b-T2a
N0, M0; ISUP grade 1 with < 50% of biopsy cores involved with cancer or ISUP grade 2 with < 33% of biopsy
cores involved with cancer; an initial PSA level of < 10 ng/mL; an International Prostatic Symptom Score (IPSS)
< 12 and maximal flow rate > 15 mL/min on urinary flow tests [685]. In addition, with due attention to dose
distribution, patients having had a previous TURP can undergo brachytherapy without an increase in risk of
urinary toxicity. A minimal channel TURP is recommended and there should be at least a 3-month interval
between TURP and brachytherapy to allow for adequate healing [686-689].

The only available RCT comparing RP and brachytherapy as monotherapy was closed due to poor accrual
[690]. Outcome data are available from a number of large population cohorts with mature follow-up [691-698].

PROSTATE CANCER - LIMITED UPDATE 2021 65


The biochemical DFS for ISUP grade 1 patients after 5 and 10 years has been reported to range from 71% to
93% and 65% to 85%, respectively [691-698]. A significant correlation has been shown between the implanted
dose and biochemical control [699]. A D90 (dose covering 90% of the prostate volume) of > 140 Gy leads to a
significantly higher biochemical control rate (PSA < 1.0 ng/mL) after 4 years (92 vs. 68%). There is no benefit
in adding neoadjuvant or adjuvant ADT to LDR monotherapy [691]. Low-dose rate brachytherapy can be
combined with EBRT in good-, intermediate- and high-risk patients (see Section 6.2.3.2.3).

6.1.3.4.2 High-dose rate brachytherapy


High-dose rate (HDR) brachytherapy uses a radioactive source temporarily introduced into the prostate to
deliver radiation. The technical differences are outlined in Table 6.1.11. The use of the GEC (Groupe Européen
de Curiethérapie)/ESTRO Guidelines is strongly recommended [700]. High-dose rate brachytherapy can be
delivered in single or multiple fractions and is often combined with EBRT of at least 45 Gy [701]. A single RCT
of EBRT (55 Gy in 20 fractions) vs. EBRT (35.75 Gy in 13 fractions), followed by HDR brachytherapy (17 Gy in
two fractions over 24 hours) has been reported [702]. In 218 patients with organ-confined PCa the combination
of EBRT and HDR brachytherapy showed a significant improvement in the biochemical disease-free rate
(p = 0.04) at 5 and 10 year (75% and 46% compared to 61% and 39%). However, a very high, uncommon, rate
of early recurrences was observed in the EBRT arm alone, even after 2 years, possibly due to a dose lower
than the current standard used [702]. A systematic review of non-RCTs has suggested outcomes with EBRT
plus HDR brachytherapy are superior to brachytherapy alone, but this needs confirmation in a prospective RCT
[703].
Fractionated HDR brachytherapy as monotherapy can be offered to patients with low- and
intermediate-risk PCa, who should be informed that results are only available from limited series in very
experienced centres [704, 705]. Five-year PSA control rates over 90% are reported, with late grade 3+ GU
toxicity rates < 5% and no, or very minimal, grade 3+ GI toxicity rates [704, 705].

Table 6.1.11: Difference between LDR and HDR brachytherapy

Differences in prostate brachytherapy techniques


Low dose rate (LDR) • Permanent seeds implanted
• Uses Iodine-125 (I-125) (most common), Palladium-103 (Pd-103) or
Cesium-131 isotopes
• Radiation dose delivered over weeks and months
• Acute side effects resolve over months
• Radiation protection issues for patient and carers
High dose rate (HDR) • Temporary implantation
• Iridium-192 (IR-192) isotope introduced through implanted needles or
catheters
• Radiation dose delivered in minutes
• Acute side effects resolve over weeks
• No radiation protection issues for patient or carers

6.1.3.5 Acute side effects of external beam radiotherapy and brachytherapy


Gastrointestinal and urinary side effects are common during and after EBRT. In the EORTC 22991 trial,
approximately 50% of patients reported acute GU toxicity of grade 1, 20% of grade 2, and 2% grade 3. In
the same trial, approximately 30% of patients reported acute grade 1 GI toxicity, 10% grade 2, and less than
1% grade 3. Common toxicities included dysuria, urinary frequency, urinary retention, haematuria, diarrhoea,
rectal bleeding and proctitis [629]. In addition, general side effects such as fatigue are common. It should be
noted that the incidence of acute side effects is greater than that of late effects (see Section 8.2.2.1), implying
that most acute effects resolve. In a RCT of conventional dose EBRT vs. EBRT and LDR brachytherapy the
incidence of acute proctitis was reduced in the brachytherapy arm, but other acute toxicities were equivalent
[706]. Acute toxicity of HDR brachytherapy has not been documented in a RCT, but retrospective reports
confirm lower rates of GI toxicity compared with EBRT alone and grade 3 GU toxicity in 10%, or fewer,
patients, but a higher incidence of urinary retention [707]. Similar findings are reported using HFX; in a pooled
analysis of 864 patients treated using extreme HFX and stereotactic radiotherapy, declines in urinary and bowel
domains were noted at 3 months which returned to baseline, or better, by 6 months [708].

66 PROSTATE CANCER - LIMITED UPDATE 2021


6.1.4 Hormonal therapy
6.1.4.1 Introduction
6.1.4.1.1 Different types of hormonal therapy
Androgen deprivation can be achieved by either suppressing the secretion of testicular androgens or inhibiting
the action of circulating androgens at the level of their receptor. These two methods can be combined to
achieve what has been known as complete (or maximal or total) androgen blockade (CAB) using the old-
fashioned anti-androgens [709].

6.1.4.1.1.1 Testosterone-lowering therapy (castration)


6.1.4.1.1.1.1 Castration level
The castration level is < 50 ng/dL (1.7 nmol/L), which was defined more than 40 years ago when testosterone
testing was less sensitive. Current methods have shown that the mean value after surgical castration is
15 ng/dL [710]. Therefore, a more appropriate level should be defined as < 20 ng/dL (1 nmol/L). This definition
is important as better results are repeatedly observed with lower testosterone levels compared to 50 ng/dL
[711-713]. However, the castrate level considered by the regulatory authorities and in clinical trials addressing
castration in PCa is still the historical < 50 ng/dL (1.7 mmol/L).

6.1.4.1.1.1.2 Bilateral orchiectomy


Bilateral orchiectomy or subcapsular pulpectomy is still considered the primary treatment modality for ADT.
It is a simple, cheap and virtually complication-free surgical procedure. It is easily performed under local
anaesthesia and it is the quickest way to achieve a castration level which is usually reached within less than
twelve hours. It is irreversible and therefore does not allow for intermittent treatment [714].

6.1.4.1.1.1.3 Oestrogens
Treatment with oestrogens results in testosterone suppression and is not associated with bone loss [715].
Early studies tested oral diethylstilboestrol (DES) at several doses. Due to severe side effects, especially
thromboembolic complications, even at lower doses these drugs are not considered as standard first-line
treatment [716-718].

6.1.4.1.1.1.4 Luteinising-hormone-releasing hormone agonists


Long-acting LHRH agonists are currently the main forms of ADT. These synthetic analogues of LHRH are
delivered as depot injections on a 1-, 2-, 3-, 6-monthly, or yearly, basis. The first injection induces a transient
rise in luteinising hormone (LH) and follicle-stimulating hormone (FSH) leading to the ‘testosterone surge’
or ‘flare-up’ phenomenon which starts two to three days after administration and lasts for about one week.
This may lead to detrimental clinical effects (the clinical flare) such as increased bone pain, acute bladder
outlet obstruction, obstructive renal failure, spinal cord compression, and cardiovascular death due to
hypercoagulation status [719]. Patients at risk are usually those with high-volume symptomatic bony disease.
Concomitant therapy with an anti-androgen decreases the incidence of clinical flare but does not completely
remove the risk. Anti-androgen therapy is usually continued for 4 weeks but neither the timing nor the duration
of anti-androgen therapy are based on strong evidence. In addition, the long-term impact of preventing ‘flare-
up’ is unknown [720, 721].
Chronic exposure to LHRH agonists results in the down-regulation of LHRH-receptors, suppressing
LH and FSH secretion and therefore testosterone production. A castration level is usually obtained within
2 to 4 weeks [722]. Although there is no formal direct comparison between the various compounds, they are
considered to be equally effective [723]. No survival difference with orchiectomy has been reported despite the
lack of high-quality trials [724].
The different products have practical differences that need to be considered in everyday practice,
including the storage temperature, whether a drug is ready for immediate use or requires reconstitution, and
whether a drug is given by subcutaneous or intramuscular injection.

6.1.4.1.1.1.5 Luteinising-hormone-releasing hormone antagonists


Luteinising-hormone releasing hormone antagonists immediately bind to LHRH receptors, leading to a rapid
decrease in LH, FSH and testosterone levels without any flare. The practical shortcoming of these compounds
is the lack of a long-acting depot formulation with, so far, only monthly formulations being available. Degarelix
is a LHRH antagonist. The standard dosage is 240 mg in the first month followed by monthly injections of
80 mg. Most patients achieve a castrate level at day three [722]. A phase III RCT compared degarelix to
monthly leuprorelin following up patients for 12 months, suggesting a better PSA PFS for degarelix 240/80 mg
compared to monthly leuprorelin [725]. A systematic review did not show a major difference between agonists
and degarelix and highlighted the paucity of on-treatment data beyond 12 months as well as the lack of
survival data [726]. Its definitive superiority over the LHRH analogues remains to be proven.

PROSTATE CANCER - LIMITED UPDATE 2021 67


Relugolix is an oral gonadotropin-releasing hormone antagonist. It was compared to the LHRH agonist
leuprolid in a randomised phase III trial [727]. The primary endpoint was sustained testosterone suppression to
castrate levels through 48 weeks. There was a significant difference of 7.9 percentage points (95% CI: 4.1–11.8)
showing non-inferiority and superiority of relugolix. The incidence of major adverse cardiovascular events was
significantly lower with relugolix (prespecified safety analysis). Relugolix has been approved by the FDA [728].

6.1.4.1.1.1.6 Anti-androgens
These oral compounds are classified according to their chemical structure as:
• steroidal, e.g., cyproterone acetate (CPA), megestrol acetate and medroxyprogesterone acetate;
• non-steroidal or pure, e.g., nilutamide, flutamide and bicalutamide.

Both classes compete with androgens at the receptor level. This leads to an unchanged or slightly elevated
testosterone level. Conversely, steroidal anti-androgens have progestational properties leading to central
inhibition by crossing the blood-brain barrier.

6.1.4.1.1.1.6.1 Steroidal anti-androgens


These compounds are synthetic derivatives of hydroxyprogesterone. Their main pharmacological side effects
are secondary to castration (gynaecomastia is quite rare) whilst the non-pharmacological side effects are
cardiovascular toxicity (4–40% for CPA) and hepatotoxicity.

Cyproterone acetate was the first licensed anti-androgen but the least studied. Its most effective dose
as monotherapy is still unknown. Although CPA has a relatively long half-life (31–41 hours), it is usually
administered in two or three fractionated doses of 100 mg each. In one RCT CPA showed a poorer OS when
compared with LHRH analogues [729]. An underpowered RCT comparing CPA monotherapy with flutamide
in M1b PCa did not show any difference in DSS and OS at a median follow-up of 8.6 years [730]. Other CPA
monotherapy studies suffer from methodological limitations preventing firm conclusions.

6.1.4.1.1.1.6.2 Non-steroidal anti-androgens


Non-steroidal anti-androgen monotherapy with e.g., nilutamide, flutamide or bicalutamide does not suppress
testosterone secretion and it is claimed that libido, overall physical performance and bone mineral density
(BMD) are frequently preserved [731]. Non-androgen-related pharmacological side effects differ between
agents. Bicalutamide shows a more favourable safety and tolerability profile than flutamide and nilutamide
[732]. The dosage licensed for use in CAB is 50 mg/day, and 150 mg for monotherapy. The androgen
pharmacological side effects are mainly gynaecomastia (70%) and breast pain (68%). However, non-steroidal
anti-androgen monotherapy offers clear bone protection compared with LHRH analogues and probably LHRH
antagonists [731, 733]. All three agents share the potential for liver toxicity (occasionally fatal), requiring regular
monitoring of patients’ liver enzymes.

6.1.4.1.1.2 New androgen receptor pathway targetting agents (ARTA)


Once on ADT the development of castration-resistance (CRPC) is only a matter of time. It is considered to be
mediated through two main overlapping mechanisms: androgen-receptor (AR)-independent and AR-dependent
mechanisms (see Section 6.5 - Castrate-resistant PCa). In CRPC, the intracellular androgen level is increased
compared to androgen sensitive cells and an over-expression of the AR has been observed, suggesting an
adaptive mechanism [734]. This has led to the development of several new compounds targeting the androgen
axis. In mCRPC, abiraterone acetate plus prednisolone and enzalutamide have been approved. In addition
to ADT (sustained castration), abiraterone acetate plus prednisolone, apalutamide and enzalutamide have
been approved for the treatment of metastatic hormone sensitive PCa (mHSPC) by the FDA, abiraterone/
prednisolone and apalutamide also by the EMA (Note: EMA issued a positive opinion to extend the indication
for enzalutamide at the time of this publication, final approval is pending). For the updated approval status see
EMA and FDA websites [735-739]. Finally, apalutamide, darolutamide and enzalutamide have been approved
for non-metastatic CRPC (nmCRPC) at high risk of further metastases [740-744].

6.1.4.1.1.2.1 Abiraterone acetate


Abiraterone acetate is a CYP17 inhibitor (a combination of 17α-hydrolase and 17,20-lyase inhibition). By
blocking CYP17, abiraterone acetate significantly decreases the intracellular testosterone level by suppressing
its synthesis at the adrenal level and inside the cancer cells (intracrine mechanism). This compound must be
used together with prednisone/prednisolone to prevent drug-induced hyperaldosteronism [735].

6.1.4.1.1.2.2 Apalutamide, darolutamide, enzalutamide (alphabetical order)


These agents are novel non-steroidal anti-androgens with a higher affinity for the AR receptor than

68 PROSTATE CANCER - LIMITED UPDATE 2021


bicalutamide. While previous non-steroidal anti-androgens still allow transfer of ARs to the nucleus and would
act as partial agonists, all three agents also block AR transfer and therefore suppress any possible agonist-like
activity [739-741]. Darolutamide has structurally unique properties [740]. In particular, in preclinical studies, it
showed not to cross the blood-brain barrier [745, 746].

6.1.4.1.1.3 New compounds


6.1.4.1.1.3.1 PARP inhibitors
Poly (ADP-ribose) polymerase inhibitors (PARPi) block the enzyme poly ADP ribose polymerase (PARP) and
were developed aiming to selectively target cancer cells harbouring BRCA mutations and other mutations
inducing homologous recombination deficiency and high level of replication pressure with a sensitivity to PARPi
treatment. Due to the oncogenic loss of some DNA repair effectors and incomplete DNA repair repertoire, some
cancer cells are addicted to certain DNA repair pathways such as Poly (ADP-ribose) polymerase (PARP)-related
single-strand break repair pathway. The interaction between BRCA and PARP is a form of synthetic lethal effect
which means the simultaneously functional loss of two genes lead to cell death, while a defect in any single gene
only has a limited effect on cell viability [747]. The therapeutic indication for PCa is discussed in Section 6.5.8.1.

6.1.4.1.1.3.2 Immune checkpoint inhibitors


Immune checkpoints are key regulators of the immune system. Checkpoint proteins, such as B7-1/B7-2 on
antigen-presenting cells (APC) and CTLA-4 on T cells, help keep the immune responses in an equilibrium. The
binding of B7-1/B7-2 to CTLA-4 keeps the T cells in the inactive state while blocking the binding of B7-1/B7-2
to CTLA-4 whilst an immune checkpoint inhibitor (anti-CTLA-4 antibody) allows the T cells to be active and to
kill tumour cells.
Approved checkpoint inhibitors target the molecules CTLA4, programmed cell death protein 1
(PD-1), and programmed death-ligand 1 (PD-L1). Programmed death-ligand 1 is the transmembrane
programmed cell death 1 protein which interacts with PD-L1 (PD-1 ligand 1). Cancer-mediated upregulation of
PD-L1 on the cell surface may inhibit T cells. Antibodies that bind to either PD-1 or PD-L1 and therefore block
the interaction may allow the T cells to induce cell killing. Examples of PD-1 inhibitors are pembrolizumab and
nivolumab; of PD-L1 inhibitors, atezolizumab, avelumab and durvalumab and an example of CTLA4 inhibitors is
ipilimumab [748, 749]. Therapeutic use is discussed in Section 6.5 - Castration-resistant PCa.

6.1.4.1.1.3.3 Protein kinase B (AKT) inhibitors


Aberrant activation of the PI3K (phosphatidylinositol-4,5-bisphosphate 3-kinase)/AKT pathway, predominately
due to PTEN loss (phosphatase and tensin homologue deleted from chromosome 10), is common in PCa (40–
60% of mCRPC) and is associated with worse prognosis. The androgen receptor signaling and AKT pathway
are reciprocally cross-regulated, so that inhibition of one leads to upregulation of the other. AKT inhibitors are
small molecules which are designed to target and bind to all three isoforms of AKT, which is a key component
of the PI3K/AKT pathway and a key driver of cancer cell growth. Ipatasertib is an oral, highly specific, AKT
inhibitor which shows clinically significant activity when combined with abiraterone acetate in patients with loss
of the tumour suppressor protein PTEN (on IHC) within the tumour [750, 751]. The therapeutic indication for
PCa is discussed in Section 6.5.6.5.

6.1.5 Investigational therapies


6.1.5.1 Background
Besides RP, EBRT and brachytherapy, other modalities have emerged as potential therapeutic options in
patients with clinically localised PCa [752-755]. In this section, both whole gland and focal treatment will be
considered, looking particularly at high-intensity focused US (HIFU), cryotherapeutic ablation of the prostate
(cryotherapy) and focal photodynamic therapy, as sufficient data are available to form the basis of some
initial judgements. Other options such as radiofrequency ablation and electroporation, among others, are
considered to be in the early phases of evaluation [756]. In addition, a relatively newer development is focal
ablative therapy [756, 757] whereby lesion-targeted ablation is undertaken in a precise organ-sparing manner.
All these modalities have been developed as minimally invasive procedures with the aim of providing equivalent
oncological safety, reduced toxicity, and improved functional outcomes.

6.1.5.2 Cryotherapy
Cryotherapy uses freezing techniques to induce cell death by dehydration resulting in protein denaturation,
direct rupture of cellular membranes by ice crystals and vascular stasis and microthrombi, resulting in
stagnation of the microcirculation with consecutive ischaemic apoptosis [752-755]. Freezing of the prostate is
ensured by the placement of 17 gauge cryo-needles under TRUS guidance, placement of thermosensors at
the level of the external sphincter and rectal wall, and insertion of a urethral warmer. Two freeze-thaw cycles
are used under TRUS guidance resulting in a temperature of -40°C in the mid-gland and at the neurovascular

PROSTATE CANCER - LIMITED UPDATE 2021 69


bundle. Currently, third and fourth generation cryotherapy devices are mainly used. Since its inception,
cryotherapy has been used for whole-gland treatment in PCa either as a primary or salvage treatment option.
The main adverse effects of cryosurgery are ED (18%), urinary incontinence (2–20%), urethral
sloughing (0–38%), rectal pain and bleeding (3%) and recto-urethral fistula formation (0–6%) [758]. There is a
lack of prospective comparative data regarding oncological outcomes of whole-gland cryosurgery as a curative
treatment option for men with localised PCa, with most studies being non-comparative single-arm case series
with short follow-up [758].

6.1.5.3 High-intensity focused ultrasound


High-intensity focused ultrasound consists of focused US waves emitted from a transducer that cause
tissue damage by mechanical and thermal effects as well as by cavitation [759]. The goal of HIFU is to heat
malignant tissue above 65°C so that it is destroyed by coagulative necrosis. High-intensity focused US is
performed under general or spinal anaesthesia, with the patient lying in the lateral or supine position. High-
intensity focused US has previously been widely used for whole-gland therapy. The major adverse effects of
HIFU include acute urinary retention (10%), ED (23%), urethral stricture (8%), rectal pain or bleeding (11%),
recto-urethral fistula (0–5%) and urinary incontinence (10%) [758]. Disadvantages of HIFU include difficulty in
achieving complete ablation of the prostate, especially in glands larger than 40 mL, and in targeting cancers
in the anterior zone of the prostate. Similar to cryosurgery, the lack of any long-term prospective comparative
data on oncological outcomes prevents whole-gland HIFU from being considered as a reasonable alternative to
the established curative treatment options [758].

6.1.5.4 Focal therapy


During the past two decades, there has been a trend towards earlier diagnosis of PCa as a result of greater
public and professional awareness leading to the adoption of both formal and informal screening strategies.
The effect of this has been that men are identified at an earlier stage with smaller tumours that occupy only
5–10% of the prostate volume, with a greater propensity for unifocal or unilateral disease [760-762]. Most focal
therapies to date have been achieved with ablative technologies: cryotherapy, HIFU, photodynamic therapy,
electroporation, and focal RT by brachytherapy or CyberKnife® Robotic Radiosurgery System technology
(Accuray Inc., Sunnyvale, CA, USA). The main purpose of focal therapy is to ablate tumours selectively whilst
limiting toxicity by sparing the neurovascular bundles, sphincter and urethra [763-765].
A systematic review and network meta-analysis [758] on ablative therapy in men with localised
PCa performed a sub-group analysis of focal therapy vs. RP and EBRT. Nine case series reporting on focal
therapy were identified (5 studies reporting on focal cryosurgical ablation of the prostate [CSAP], three studies
on focal HIFU, and one study reported on both). For focal CSAP vs. RP or EBRT, no statistically significant
differences were found for BCR at 3 years. For focal HIFU vs. RP or EBRT there were neither comparable
data on oncological-, continence- nor potency outcomes at one year or more. More recently, Valerio et al.,
[757] performed a systematic review to summarise the evidence regarding the effectiveness of focal therapy
in localised PCa. Data from 3,230 patients across 37 studies were included, covering different energy sources
including HIFU, CSAP, photodynamic therapy, laser interstitial thermotherapy, focal brachytherapy, irreversible
electroporation and radiofrequency ablation. The overall quality of the evidence was low, due to the majority
of studies being single-centre, non-comparative and retrospective in design, heterogeneity of definitions and
approaches, follow-up strategies, outcomes, and duration of follow-up. Although the review suggests that focal
therapy has a favourable toxicity profile in the short-to-medium term, its oncological effectiveness remains
unproven due to lack of reliable comparative data against standard interventions such as RP and EBRT.

In order to update the evidence base, a systematic review incorporating a narrative synthesis was performed
by the Panel, including comparative studies assessing focal ablative therapy vs. radical treatment, AS or
alternative focal ablative therapy, published between 1st January 2000 to 12th June 2020 [766]. Only English
language papers were included in the review. Primary outcomes included oncological outcomes, adverse
events and functional outcomes. Only comparative studies recruiting at least 50 patients in every arm were
included. Relevant systematic reviews and ongoing prospective comparative studies with the same PICO
elements were included, and systematic reviews were quality assessed using AMSTAR criteria. The results
have been reported elsewhere [766]. In brief, out of 1,119 articles identified, 4 primary studies (1 RCT and 3
retrospective cohort studies) [767-771] recruiting 3,961 patients, and 10 systematic reviews were included
[758]. Only qualitative synthesis was possible due to clinical heterogeneity. Overall risk of bias (RoB) and
confounding were moderate to high. Comparative effectiveness data regarding focal therapy were inconclusive.
Data quality and applicability were poor due to clinical heterogeneity, RoB and confounding, lack of long-term
data, inappropriate outcome measures and poor external validity. The majority of systematic reviews had a low
or critically low confidence rating.
The only identified RCT, Azzouzi et al., [767] deserves discussion. The authors compared focal

70 PROSTATE CANCER - LIMITED UPDATE 2021


therapy using padeliporfin-based vascular-targeted photodynamic therapy (PDT) vs. AS in men with very low-
risk PCa. The study found, at a median follow-up of 24 months, that less patients progressed in the PDT arm
compared with the AS arm (adjusted HR: 0.34, 95% CI: 0.24–0.46), and needed less radical therapy (6% vs.
29%, p < 0.0001). In addition, more men in the PDT arm had a negative prostate biopsy at two years than
men in the AS arm (adjusted RR: 3.67, 95% CI: 2.53–5.33). Updated results were published in 2018 showing
that these benefits were maintained after four years [768]. Nevertheless, limitations of the study include
inappropriately comparing an intervention designed to destroy cancer tissue in men with low-risk PCa against
an intervention primarily aimed at avoiding unnecessary treatment in men with low-risk PCa, and an unusually
high observed rate of disease progression in the AS arm (58% in two years). Furthermore, more patients in the
AS arm chose to undergo radical therapy without a clinical indication which may have introduced confounding
bias. Finally, the AS arm did not undergo any confirmatory biopsy or any mpMRI scanning, which is not
representative of contemporary practice. Given the lack of robust comparative data on medium- to long-term
oncological outcomes for focal therapy against curative interventions (i.e. RP or EBRT), significant uncertainties
remain in regard to focal therapy as a proven alternative to either AS or radical therapy. Consequently, robust
prospective trials reporting standardised outcomes [772] are needed before recommendations in support of
focal therapy for routine clinical practice can be made [756, 772, 773]. At this time focal therapy should only be
performed within the context of a clinical trial setting or well-designed prospective cohort study.

6.1.6 General guidelines for the treatment of prostate cancer

Recommendations Strength rating


Inform patients that based on robust current data with up to 12 years of follow-up, no Strong
active treatment modality has shown superiority over any other active management options
or deferred active treatment in terms of overall- and PCa-specific survival for clinically
localised disease.
Offer a watchful waiting policy to asymptomatic patients with a life expectancy < 10 years Strong
(based on co-morbidities).
Inform patients that all active treatments have side effects. Strong
Surgical treatment
Inform patients that no surgical approach (open-, laparoscopic- or robotic radical Weak
prostatectomy) has clearly shown superiority in terms of functional or oncological results.
When a lymph node dissection (LND) is deemed necessary, perform an extended LND Strong
template for optimal staging.
Do not perform nerve-sparing surgery when there is a risk of ipsilateral extracapsular Weak
extension (based on cT stage, ISUP grade, nomogram, multiparametric magnetic resonance
imaging).
Do not offer neoadjuvant androgen deprivation therapy before surgery. Strong
Radiotherapeutic treatment
Offer intensity-modulated radiation therapy (IMRT) plus image-guided radiation therapy Strong
(IGRT) for definitive treatment of PCa by external-beam radiation therapy.
Offer moderate hypofractionation (HFX) with IMRT including IGRT to the prostate to patients Strong
with localised disease.
Ensure that moderate HFX adheres to radiotherapy protocols from trials with equivalent Strong
outcome and toxicity, i.e. 60 Gy/20 fractions in 4 weeks or 70 Gy/28 fractions in 6 weeks.
Offer low-dose rate (LDR) brachytherapy monotherapy to patients with good urinary Strong
function and low- or good prognosis intermediate-risk localised disease.
Offer LDR or high-dose rate brachytherapy boost combined with IMRT including IGRT to Strong
patients with good urinary function and intermediate-risk disease with adverse features or
high-risk disease.
Active therapeutic options outside surgery and radiotherapy
Only offer cryotherapy and high-intensity focused ultrasound within a clinical trial setting or Strong
well-designed prospective cohort study.
Only offer focal therapy within a clinical trial setting or well-designed prospective cohort Strong
study.

PROSTATE CANCER - LIMITED UPDATE 2021 71


6.2 Treatment by disease stages
6.2.1 Treatment of low-risk disease
6.2.1.1 Active surveillance
The main risk for men with low-risk disease is over treatment (see Sections 6.1.1.2 and 6.1.1.4); therefore, AS
should be considered for all such patients.

6.2.1.1.1 Active surveillance - inclusion criteria


Guidance regarding selection criteria for AS is limited by the lack of data from prospective RCTs. As a
consequence, the Panel undertook an international collaborative study involving healthcare practitioners
and patients to develop consensus statements for deferred treatment with curative intent for localised PCa,
covering all domains of AS (DETECTIVE Study) [252]. The criteria most often published include: ISUP grade 1,
clinical stage cT1c or cT2a, PSA < 10 ng/mL and PSA density < 0.15 ng/mL/cc [468, 774]. The latter threshold
remains controversial [774, 775]. These criteria were supported by the DETECTIVE consensus. There was no
agreement around the maximum number of cores that can be involved with cancer or the maximum percentage
core involvement although there was recognition that cT2c disease and extensive disease on MRI should
exclude men from AS.
The DETECTIVE Study concluded that men with favourable ISUP 2 cancer (PSA < 10 ng/mL, clinical
stage < cT2a and a low number of positive cores) should also be considered for deferred treatment [252]. In
this setting, re-biopsy within 6 to 12 months to exclude sampling error is mandatory [774, 776] even if this
could be modified in the future [777]. A systematic review and meta-analysis found three clinico-pathological
variables which were significantly associated with reclassification, which included PSA density, > 2 positive
cores, and African-American race [778].
The DETECTIVE consensus group were clear that those with ISUP 3 disease should not be
considered. In addition, a previous pathology consensus group suggested excluding men from AS when any
of the following features were present: predominant ductal carcinoma (including pure intraductal carcinoma),
sarcomatoid carcinoma, small cell carcinoma, EPE or LVI in needle biopsy [345] and perineal invasion [779],
and this view was supported.

Recently, a multidisciplinary consensus conference on germline testing attempted to develop a genetic


implementation framework for the management of PCa [139]. Based on consensus, BRCA2-gene testing
was recommended for AS discussions. However, the nature of such discussions and how a positive result
influences management were beyond the scope of the project. For now, if included in AS programmes, patients
with a BRCA2 mutation should be cautiously monitored; more robust data is needed to be able to make
definitive conclusions.

6.2.1.1.2 Tissue-based prognostic biomarker testing


Biomarkers, including Oncotype Dx®, Prolaris®, Decipher®, PORTOS and ProMark® are promising (see 5.2.8.3).
however, further data will be needed before such markers can be used in standard clinical practice [198].

6.2.1.1.3 Imaging for treatment selection


In men eligible for AS based upon systematic biopsy findings alone MRI can detect suspicious lesions
triggering reclassification at confirmatory biopsy [780, 781]. However, systematic biopsy retains substantial
added value at confirmatory biopsy.
A meta-analysis evaluated the proportion of men eligible for AS based on systematic TRUS-guided
biopsy in whom the cancer was upgraded by MRI-TBx (17%) and systematic biopsy (20%) at confirmatory
biopsy [782]. Ten percent of patients were upgraded by both biopsy methods, meaning MRI-TBx upgraded an
additional 7% (95% CI: 5–10%) of men, whilst systematic biopsy upgraded an additional 10% of men (95% CI:
8–14%). Even if the analysed series used different definitions for csPCa (and thus for cancer upgrading), MRI-
TBx and systematic biopsy appear to be complementary to each other, both missing a significant proportion of
cancer upgrading or reclassification. Therefore, combining the two biopsy techniques appears to be the best
way to select patients for AS at confirmatory biopsy.
Magnetic resonance imaging-positive men have approximately a three times higher chance
(RR: 2.77, 95% CI: 1.76–4.38) to cancer upgrading than MRI-negative men, having their first MRI [782]. Similar
results were obtained in an enrolled AS cohort (ISUP grade 1) which was compared to a surveillance cohort
from the pre-MRI era; the MRI-negative men had an HR of 0.61 (95% CI: 0.39–0.95, p = 0.03) for upgrading,
while the MRI-positive men had an HR of 1.96 (95% CI:1.36–2.82, p < 0.01) [783].
The Active Surveillance Magnetic Resonance Imaging Study (ASIST) randomised men on AS
scheduled for confirmatory biopsy to either 12-core systematic biopsy or to MRI with targeted biopsy (when
indicated) combined with systematic biopsy, up to 12 cores in total, avoiding oversampling in the MRI arm
[784]. The initial report showed little benefit from targeted biopsy. However, after 2 years of follow-up, use of

72 PROSTATE CANCER - LIMITED UPDATE 2021


MRI before confirmatory biopsy resulted in fewer failures of surveillance (19% vs. 35%, p = 0.017) and in fewer
patients progressing to ISUP grade > 2 cancer (9.9% vs. 23%, p = 0.048) [785].
At the DECTECTIVE consensus meeting it was agreed that men eligible for AS after combined
systematic- and MRI-targeted biopsy do not require a confirmatory biopsy [252].

6.2.1.1.4 Monitoring during active surveillance


Based on the DETECTIVE consensus study, the follow-up strategy should be based on serial DRE (at least
once yearly), PSA (at least once, every 6 months) and repeated biopsy [252]. Several authors have reported
data on sequential MRI evaluation, summarised in a review [786]. In MRI-positive men the overall upgrading
from ISUP 1 to ISUP > 2 PCa was 30% (81/269), following combined targeted and standard biopsies.
Upgrading occurred in 39% of these MRI-positive patients with MRI showing progression between entry
and follow-up MRI, while only 21% upgrading occurred in patients with MRI showing stable findings. These
data suggest that radiological progression is a predictor for upgrading. However, in three recent studies not
included in this review, an association between MRI progression and pathological upgrade was not observed
[787-789]. In a retrospective single-centre study of men on AS with ISUP 1 cancer and serial MRIs, the
upgrading at 1 year was 24% (29/122) [787]. On multivariable logistic regression, radiological progression
between serial mpMRI examinations was not predictive of upgrading. In a prospective single-centre study
of men on AS with ISUP 1 cancer and serial MRIs, the upgrading at 1 year was 32% (35/111) in the overall
population and 48% (30/63) in MRI-positive men. However, no difference in upgrading was observed between
MRI-positive men with radiological progression (41% [7/17], 95% CI: 18–67%) and MRI-positive men without
radiological progression (50% [23/46], 95% CI: 35–65%) [788]. These results were confirmed in a single-centre
retrospective analysis of 207 men with ISUP grade 1 cancer with a 3-year interval between diagnosis and
follow-up, with 42% (20/48) upgrading in men with radiological progression and 31% (41/132) in men with
stable disease. The authors concluded that increases in MRI score were not associated with reclassification
after adjusting for the 3-year MRI score (p = 0.9) [789].
Data are more limited on serial unchanged negative MRI findings. In a small side study including
75 men within the Prostate Cancer Research International Active Surveillance (PRIAS) study with an MRI at
baseline, 46 (61%) had a negative MRI. Fifteen percent (7/46) were reclassified to ISUP grade > 2 PCa at 12
months by systematic biopsy [790]. Other studies on serial negative MRI showed upgrading to ISUP grade > 2
PCa in 2% (1/56) [781], 5% (2/41) [787], 10% (5/48) [788], and 16% (8/51), respectively [789].
Data on the combination of serial MRI and PSA as a trigger for re-biopsy are even more limited.
Using MRI and PSA changes as the sole triggers for re-biopsy would have detected only 70% (14/20) of
progressions. Changes on MRI, prompting a non-protocol biopsy, accounted for 50% (10/20) of detected
upgrading and resulted in 10 (50%) additional biopsy procedures which failed to show pathological progression
[791]. Protocol-based re-biopsy, without MRI or PSA changes, however, detected pathological progression
in only 7% (6/87) of men. In another serial MRI study on AS, PSA velocity was significantly associated with
subsequent requirement for radical therapy in patients with no visible lesions (negative MRI); 0.98 (0.56–1.11)
ng/mL/year in progressed disease vs. 0.12 (0.16 to 0.51) ng/mL/year in non-progressed disease (p < 0.01)
[781]. Prostate-specific antigen doubling time was significant in patients with visible lesions (positive MRI);
3.2 (1.9–5.2) years in histopathological progressed disease vs. 5.7 (2.5–11.1) years in histopathological non-
progressed disease (p < 0.01). In patients with no visible lesions on their first MRI, a cut-off of 0.5 ng/mL/year
in PSA-velocity had a sensitivity of 89% (8/9 progressions identified) and a specificity of 75% for progression
to radical therapy [781]. Another study showed similar results on PSA-velocity and PSA doubling times in MRI-
negative men on AS [792]. For the PSA-velocity threshold value of > 0.5 ng/mL/year, sensitivity and specificity
were 92% and 86%, respectively. Prostate-specific antigen doubling time < 3 years in the first 2 years were
associated with tumour reclassification in 82% of cases (9/11) with a sensitivity and specificity for detection of
tumour progression of 69% and 97%, respectively.

The DETECTIVE consensus study concluded that repeat biopsy should be performed if there is a change in
mpMRI (i.e. increase in PI-RADS score, lesion volume or radiological T stage), or by DRE or PSA progression,
and if repeat biopsies are required they should be performed by MRI-guided targeted biopsies (including
in-bore, cognitive guidance or MRI fusion) combined with systematic biopsies [252]. The situation regarding
protocol-mandated, untriggered, biopsies or untriggered mpMRI biopsies remains less clear. The DETECTIVE
study failed to achieve consensus on these issues. Most contemporary long-term single-arm case series
on AS include protocol-mandated untriggered prostate biopsies at varying intervals, although comparative
effectiveness data remain lacking. Presently, it remains unclear if regular repeat mpMRI should be performed
in the absence of any triggers (i.e. protocol-mandated). Similarly, it remains unclear if protocol-mandated,
untriggered repeat prostate biopsies should be performed at regular intervals. As such, no recommendations
can be made at this time regarding these issues.

PROSTATE CANCER - LIMITED UPDATE 2021 73


6.2.1.1.5 Active Surveillance - when to change strategy
Men may remain on AS whilst they continue to consent, have a life expectancy of > 10 years and the disease
remains indolent. Patient anxiety about continued surveillance occurs in around 10% of patients on AS [793]
and was recognised as a valid reason for active treatment [252]. More common is the development of other
co-morbidities which may result in a decision to transfer to a WW strategy.
A PSA change alone (especially a PSA-DT < 3 years) is a less powerful indicator to change
management based on its weak link with grade progression [794, 795]. As a consequence, this should instead
trigger further investigation. There was clear agreement in the DETECTIVE consensus meeting that a change
in PSA should lead to repeat-MRI and repeat biopsy. It was also agreed that changes on follow-up MRI
needed a confirmatory biopsy before considering active treatment. However, there was no agreement on the
histopathology criteria (neither the extent of core involvement nor the number of cores involved) required to
trigger a change in management [252].

6.2.1.2 Alternatives to active surveillance for the treatment of low-risk disease


In terms of alternatives to AS in the management of patients with low-risk disease there is some data from
randomised studies. In the PIVOT trial (Section 6.1.1.3.1) which compared surgery vs. observation, only 42%
of patients had low-risk disease [478]. Sub-group analysis revealed that for low-risk disease there was no
statistically significant difference in all-cause mortality between surgery vs. observation (RR: 0.93, 95% CI:
0.78–1.11). In the ProtecT study (Section 6.1.1.4) which compared AM vs. surgery vs. EBRT, 56% of patients
had low-risk disease [467]. However, no sub-group analysis was performed on this group. The study found
no difference between the three arms in terms of OS and CSS, but AM had higher metastatic progression
compared with surgery or EBRT (6.0% vs. 2.6). There is no robust data comparing contemporary AS protocols
with either surgery or EBRT in patients with low-risk disease. On balance, although AS should be the default
management strategy in patients with low-risk disease and a life expectancy > 10 years, it would be reasonable
to consider surgery and EBRT as alternatives to AS in patients suitable for such treatments and who accept a
trade-off between toxicity and prevention of disease progression.

6.2.1.3 Summary of evidence and guidelines for the treatment of low-risk disease

Summary of evidence
Systematic biopsies have been scheduled in AS protocols, the number and frequency of biopsies varied,
there is no approved standard.
Although per-protocol MR scans are increasingly used in AS follow-up no conclusive evidence exist in terms
of their benefit/and whether biopsy may be ommitted based on the imaging findings.
Personalised risk-based approaches will ultimately replace protocol-based management of patients on AS.

Recommendations Strength rating


Active surveillance (AS)
Selection of patients
Offer AS to patients with a life expectancy > 10 years and low-risk disease. Strong
If a patient has had upfront multiparametric magnetic resonance imaging (mpMRI) followed Weak
by systematic and targeted biopsies there is no need for confirmatory biopsies.
Patients with intraductal and cribiform histology on biopsy should be excluded from AS. Strong
Perform a mpMRI before a confirmatory biopsy if no MRI has been performed before the Strong
initial biopsy.
Take both targeted biopsy (of any PI-RADS > 3 lesion) and systematic biopsy if a Strong
confirmatory biopsy is performed.
Follow-up strategy
Perform serum prostate-specific antigen (PSA) assessment every 6 months. Strong
Perform digital rectal examination (DRE) every 12 months. Strong
Counsel patients about the possibility of needing further treatment in the future. Strong
Active treatment
Offer surgery and radiotherapy as alternatives to AS to patients suitable for such treatments Weak
and who accept a trade-off between toxicity and prevention of disease progression.
Pelvic lymph node dissection (PLND)
Do not perform a PLND. Strong

74 PROSTATE CANCER - LIMITED UPDATE 2021


Radiotherapeutic treatment
Offer low-dose rate brachytherapy to patients with low-risk PCa, without a recent Strong
transurethral resection of the prostate and a good International Prostatic Symptom Score.
Use intensity-modulated radiation therapy plus image-guided radiation therapy with a total Strong
dose of 74-80 Gy or moderate hypofractionation (60 Gy/20 fx in 4 weeks or 70 Gy/28 fx in
6 weeks), without androgen deprivation therapy (ADT).
Other therapeutic options
Do not offer ADT monotherapy to asymptomatic men not able to receive any local Strong
treatment.
Only offer whole gland treatment (such as cryotherapy, high-intensity focused ultrasound, etc.) Strong
or focal treatment within a clinical trial setting or well-designed prospective cohort study.

6.2.2 Treatment of intermediate-risk disease


When managed with non-curative intent, intermediate-risk PCa is associated with 10-year and 15-year PCSM
rates of 13.0% and 19.6%, respectively [796].

6.2.2.1 Active Surveillance


In the ProtecT trial, up to 22% of the randomised patients in the AM arm had ISUP grade > 1 and 10% a PSA
> 10 ng/mL [467]. A Canadian consensus group proposes that low volume ISUP grade 2 (< 10% Gleason
pattern 4) may also be considered for AS. These recommendations have been endorsed by the American
Society of Clinical Oncology (ASCO) [797] and the recent DETECTIVE consensus meeting [252] for those
patients with a PSA < 10 ng/mL and low core positivity.

However, data is less consistent in other patient groups. It is clear that the presence of any grade 4 pattern is
associated with a 3-fold increased risk of metastases compared to ISUP grade 1, while a PSA up to 20 ng/mL
might be an acceptable threshold [776, 798, 799]. In addition, it is likely that mpMRI and targeted biopsies will
detect small focuses of Gleason 4 cancer that might have been missed with systematic biopsy. Therefore, care
must be taken when explaining this treatment strategy especially to patients with the longest life expectancy.

6.2.2.2 Surgery
Patients with intermediate-risk PCa should be informed about the results of two RCTs (SPCG-4 and PIVOT)
comparing RRP vs. WW in localised PCa. In the SPCG-4 study, death from any cause (RR: 0.71, 95% CI:
0.53–0.95), death from PCa (RR: 0.38, 95% CI: 0.23–0.62) and distant metastases (RR: 0.49, 95% CI:
0.32–0.74) were significantly reduced in intermediate-risk PCa at 18 years. In the PIVOT trial, according to a
pre-planned subgroup analysis among men with intermediate-risk tumours, RP significantly reduced all-cause
mortality (HR: 0.69, 95% CI: 0.49–0.98), but not death from PCa (0.50, 95% CI: 0.21–1.21) at 10 years. The risk
of having positive LNs in intermediate-risk PCa is between 3.7–20.1% [800]. An eLND should be performed
in intermediate-risk PCa if the estimated risk for pN+ exceeds 5% [381] or 7% if using the nomogram by
Gandaglia et al., which incorporates MRI-guided biopsies [521]. In all other cases eLND can be omitted, which
means accepting a low risk of missing positive nodes.

6.2.2.3 Radiation therapy


6.2.2.3.1 Recommended IMRT for intermediate-risk PCa
Patients suitable for ADT can be given combined IMRT with short-term ADT (4–6 months) [801-803]. For
patients unsuitable for ADT (e.g., due to co-morbidities) or unwilling to accept ADT (e.g. to preserve their sexual
health) the recommended treatment is IMRT (76–78 Gy) or a combination of IMRT and brachytherapy (see
Section 6.2.3.2.3).

6.2.2.3.2 Brachytherapy monotherapy


Low-dose rate brachytherapy can be offered to patients with ISUP grade 2 with < 33% of biopsy cores
involved with cancer and without a recent TURP and good IPSS Score. Fractionated HDR brachytherapy as
monotherapy can be offered to selected patients with intermediate-risk PCa although they should be informed
that results are only available from small series in very experienced centres. Five-year PSA control rates over
90% are reported, with late grade 3+ GU toxicity rates < 5% and no, or very minimal, grade 3+ GI toxicity rates
[704, 804]. There are no direct data to inform on the use of ADT in this setting. For the combination of EBRT
plus brachytherapy boost please see Section 6.2.3.2.3.

6.2.2.4 Other options for the primary treatment of intermediate-risk PCa (experimental therapies)
A prospective study on focal therapy using HIFU in patients with localised intermediate-risk disease was
published but the data was derived from an uncontrolled, single-arm case series [773]. There is a paucity of

PROSTATE CANCER - LIMITED UPDATE 2021 75


high-certainty data for either whole-gland or focal ablative therapy in the setting of intermediate-risk disease.
Consequently, neither whole-gland treatment nor focal treatment can be considered as standard therapy for
intermediate-risk patients and, if offered, it should only be in the setting of clinical trials [756].

Data regarding the use of ADT monotherapy for intermediate-risk disease have been inferred indirectly from
the EORTC 30891 trial, which was a RCT comparing deferred ADT vs. immediate ADT in 985 patients with
T0–4 N0–2 M0 disease [800]. The trial showed a small, but statistically significant, difference in OS in favour of
immediate ADT monotherapy but there was no significant difference in CSS, predominantly because the risk of
cancer-specific mortality was low in patients with PSA < 8 ng/mL. Consequently, the use of ADT monotherapy
for this group of patients is not considered as standard, even if they are not eligible for radical treatment.

6.2.2.5 Guidelines for the treatment of intermediate-risk disease

Recommendations Strength rating


Active surveillance (AS)
Offer AS to highly selected patients with ISUP grade group 2 disease (i.e. < 10% pattern 4, Weak
PSA < 10 ng/mL, < cT2a, low disease extent on imaging and biopsy) accepting the potential
increased risk of metastatic progression.
Radical prostatectomy (RP)
Offer RP to patients with intermediate-risk disease and a life expectancy of > 10 years. Strong
Offer nerve-sparing surgery to patients with a low risk of extracapsular disease. Strong
Pelvic lymph node dissection (ePLND)
Perform an ePLND in intermediate-risk disease (see Section 6.1.2.3.2). Strong
Radiotherapeutic treatment
Offer low-dose rate brachytherapy to intermediate-risk patients with ISUP grade 2 with Strong
< 33% of biopsy cores involved, without a recent transurethral resection of the prostate and
with a good International Prostatic Symptom Score.
For intensity-modulated radiotherapy (IMRT) plus image-guided radiotherapy (IGRT), Strong
use a total dose of 76–78 Gy or moderate hypofractionation (60 Gy/20 fx in 4 weeks or
70 Gy/28 fx in 6 weeks), in combination with short-term androgen deprivation therapy (ADT)
(4 to 6 months).
In patients not willing to undergo ADT, use a total dose of IMRT plus IGRT (76–78 Gy) Weak
or moderate hypofractionation (60 Gy/20 fx in 4 weeks or 70 Gy/28 fx in 6 weeks) or a
combination with brachytherapy.
Other therapeutic options
Only offer whole-gland ablative therapy (such as cryotherapy, high-intensity focused Strong
ultrasound, etc.) or focal ablative therapy for intermediate-risk disease within a clinical trial
setting or well-designed prospective cohort study.
Do not offer ADT monotherapy to intermediate-risk asymptomatic men not able to receive Weak
any local treatment.

6.2.3 Treatment of high-risk localised disease


Patients with high-risk PCa are at an increased risk of PSA failure, need for secondary therapy, metastatic
progression and death from PCa. Nevertheless, not all high-risk PCa patients have a uniformly poor prognosis
after RP [805]. When managed with non-curative intent, high-risk PCa is associated with 10-year and 15-year
PCSM rates of 28.8 and 35.5%, respectively [806]. There is no consensus regarding the optimal treatment of
men with high-risk PCa.

6.2.3.1 Radical prostatectomy


Provided that the tumour is not fixed to the pelvic wall or there is no invasion of the urethral sphincter, RP is a
reasonable option in selected patients with a low tumour volume. Extended PLND should be performed in all
high-risk PCa cases [381, 382]. Patients should be aware pre-operatively that surgery may be part of multi-
modal treatment.

6.2.3.1.1 ISUP grade 4–5


The incidence of organ-confined disease is 26–31% in men with an ISUP grade > 4 on systematic biopsy. A
high rate of downgrading exists between the biopsy ISUP grade and the ISUP grade of the resected specimen
[806]. Several retrospective case series have demonstrated CSS rates over 60% at 15 years after RP in the

76 PROSTATE CANCER - LIMITED UPDATE 2021


context of a multi-modal approach (adjuvant or salvage ADT and/or RT) in patients with a biopsy ISUP grade 5
[424, 494, 807, 808].

6.2.3.1.2 Prostate-specific antigen > 20 ng/mL


Reports in patients with a PSA > 20 ng/mL who underwent surgery as initial therapy within a multi-modal
approach demonstrated a CSS at 15 years of over 70% [424, 494, 501, 809-811].

6.2.3.1.3 R
 adical prostatectomy in cN0 patients who are found to have pathologically confirmed lymph node
invasion (pN1)
At 15 years follow-up cN0 patients who undergo RP but who were found to have pN1 were reported to have
an overall CSS and OS of 45% and 42%, respectively [812-818]. However, this is a very heterogeneous patient
group and further treatment must be individualised based on risk factors (see Sections 6.2.5.2 and 6.2.5.6).

6.2.3.2 External beam radiation therapy


6.2.3.2.1 Recommended external beam radiation therapy treatment policy for high-risk localised PCa
For high-risk localised PCa, a combined modality approach should be used consisting of IMRT plus long-term
ADT. The duration of ADT has to take into account PS, co-morbidities and the number of poor prognostic
factors. It is important to recognise that in several studies EBRT plus short-term ADT did not improve OS in
high-risk localised PCa and long-term ADT (at least 2 to 3 years) is currently recommended for these patients
[662, 663, 666].

6.2.3.2.2 Lymph node irradiation in cN0


There is no high level evidence for prophylactic whole-pelvic irradiation since RCTs have failed to show that
patients benefit from prophylactic irradiation (46–50 Gy) of the pelvic LNs in high-risk disease [819-821]. In the
RTOG 94-13 study there were no PFS differences between patients treated with whole-pelvic or prostate-only
RT but interactions between whole-pelvic RT and the duration of ADT were reported following the subgroup
analysis [666]. Furthermore, in most trials dealing with high-risk PCa irradiation of a whole pelvis field was
considered standard of care. The benefits of pelvic nodal irradiation using IMRT merit further investigation in
RCTs as conducted by the RTOG or the UK NCRI group. Performing an ePLND in order to decide whether
or not pelvic RT is required (in addition to combined prostate EBRT plus long-term ADT) remains purely
experimental in the absence of high level evidence.

6.2.3.2.3 Brachytherapy boost


In men with intermediate- or high-risk PCa, brachytherapy boost with supplemental EBRT and hormonal
treatment [822] may be considered. External beam RT (total dose of 78 Gy) has been compared with EBRT
(total dose 46 Gy) followed by LDR brachytherapy boost (prescribed dose 115 Gy) in intermediate-risk and
high-risk patients in a randomised trial with 12 months of ADT in both arms [823]. The LDR boost resulted in
5- and 7-year PSA PFS increase (89% and 86%, respectively, compared to 84% and 75%). This improvement
was achieved at a cost of increased late grade 3+ GU toxicity (18% compared to 8%) [824]. Toxicity resulted
mainly in the development of urethral strictures and incontinence and great care should be taken during
treatment planning.

See Section 6.1.3.4.2 for details on a RCT directly comparing EBRT alone and EBRT with HDR brachytherapy
boost. The TROG 03.04 RADAR trial randomised men with locally-advanced PCa to 6 or 18 months ADT
with an additional upfront radiation dose escalation stratification with dosing options of 66, 70, or 74 Gy
EBRT, or 46 Gy EBRT plus HDR brachytherapy boost. After a minimum follow-up of 10 years HDR boost
signficantly reduced distant progression, the study primary endpoint (subhazard ratio 0.68 [95% CI: 0.57–0.80];
p < 0.0001), when compared to EBRT alone and independent of duration of ADT [822]. Although radiation dose
escalation using brachytherapy boost provides much higher biological doses, the TROG 03.04 RADAR RCT
and systematic reviews show ADT use independently predicts better outcomes regardless of radiation dose
intensification [822, 825, 826]. Omitting ADT may result in inferior OS and based on current evidence ADT use
and duration should be in line with that used when delivering EBRT alone.

6.2.3.3 Options other than surgery and radiotherapy for the primary treatment of localised PCa
Currently there is a lack of evidence supporting any other treatment option apart from RP and radical RT in
localised high-risk PCa.

PROSTATE CANCER - LIMITED UPDATE 2021 77


6.2.3.4 Guidelines for radical treatment of high-risk localised disease

Recommendations Strength rating


Radical Prostatectomy (RP)
Offer RP to selected patients with high-risk localised PCa as part of potential multi-modal Strong
therapy.
Extended pelvic lymph node dissection (ePLND)
Perform an ePLND in high-risk PCa. Strong
Do not perform a frozen section of nodes during RP to decide whether to proceed with, or Strong
abandon, the procedure.
Radiotherapeutic treatment
In patients with high-risk localised disease, use intensity-modulated radiation therapy Strong
(IMRT) plus image-guided radiation therapy (IGRT) with 76–78 Gy in combination with
long-term androgen deprivation therapy (ADT) (2 to 3 years).
In patients with high-risk localised disease, use IMRT and IGRT with brachytherapy boost Weak
(either high-dose rate or low-dose rate), in combination with long-term ADT (2 to 3 years).
Therapeutic options outside surgery and radiotherapy
Do not offer either whole gland or focal therapy to patients with high-risk localised disease. Strong
Only offer ADT monotherapy to those patients unwilling or unable to receive any form of Strong
local treatment if they have a prostate-specific antigen (PSA)-doubling time < 12 months,
and either a PSA > 50 ng/mL or a poorly-differentiated tumour.

6.2.4 Treatment of locally advanced PCa


In the absence of high level evidence a recent systematic review could not define the most optimal treatment
option [827]. Randomised controlled trials are only available for EBRT. A local treatment combined with a
systemic treatment provides the best outcome, provided the patient is ready and fit enough to receive both.

6.2.4.1 Surgery
Surgery for locally advanced disease as part of a multi-modal therapy has been reported [806, 828, 829].
However, the comparative oncological effectiveness of RP as part of a multi-modal treatment strategy vs. upfront
EBRT with ADT for locally advanced PCa remains unknown, although a prospective phase III RCT (SPCG-15)
comparing RP (with or without adjuvant or salvage EBRT) against primary EBRT and ADT among patients with
locally advanced (T3) disease is currently recruiting [830]. Data from retrospective case series demonstrated over
60% CSS at 15 years and over 75% OS at 10 years [804, 806, 828, 829, 831-834]. For cT3b–T4 disease, PCa
cohort studies showed 10-year CSS of over 87% and OS of 65% [835-837]. The indication for RP in all previously
described stages assumes the absence of clinically detectable nodal involvement (cN0). In case of suspected
positive LNs during RP (initially considered cN0) the procedure should not be abandoned since RP may have
a survival benefit in these patients. Intra-operative frozen section analysis is not justified in this case [550]. An
ePLND is considered standard if a RP is planned.

6.2.4.2 Radiotherapy for locally advanced PCa


In locally advanced disease RCTs have clearly established that the additional use of long-term ADT combined
with RT produces better OS than ADT or RT alone (see Section 6.1.3.1.4 and Tables 6.1.9 and 6.1.10) [827].

6.2.4.3 Treatment of cN1 M0 PCa


Lymph node metastasised PCa is where options for local therapy and systemic therapies overlap.
Approximately 5% to 10% of newly diagnosed PCa patients have synchronous suspected pelvic nodal
metastases on conventional imaging (CT/bone scan) without bone or visceral mestatases (cN1 M0 stage).
Meta-analyses have shown that PSMA-PET/CT prior to primary treatment in advanced PCa detected disease
outside the prostate in 32% of cases despite prior negative conventional imaging using bone scan and pelvic
CT/MRI [397]. A RCT assessing PSMA-PET/CT as staging tool in high-risk PCa confirmed these findings and
showed a 32% increase in accuracy compared with conventional imaging for the detection of pelvic nodal
metastases [417]. Notably, more sensitive imaging also causes a stage shift with more cases classified as cN1,
but with, on average, lower nodal disease burden.

The management of cN1 PCa is mainly based on long-term ADT. The benefit of adding local treatment has
been assessed in various retrospective studies, summarised in one systematic review [839] including 5 studies
only dealing with cN1M0 patients [840-844]. The findings suggested an advantage in both OS and CSS after
local treatment (RT or RP) combined with ADT as compared to ADT alone. The main limitations of this analysis

78 PROSTATE CANCER - LIMITED UPDATE 2021


were the lack of randomisation, of comparisons between RP and RT, as well as the value of the extent of PLND
and of RT fields. Only limited evidence exists supporting RP for cN+ patients. Moschini et al., compared the
outcomes of 50 patients with cN+ with those of 252 patients with pN1, but cN0 at preoperative staging. cN+ was
not a significant predictor of CSS [838].
Based on the consistent benefit seen in retrospective studies including cN1 patients local therapy is
recommended in patients with cN1 disease at diagnosis in addition to long-term ADT (see Table 6.2.4.1).

The intensification of systemic treatment (abiraterone acetate, docetaxel, zoledronic acid) has been assessed
in unplanned sub-group analyses from the STAMPEDE multi-arm RCT by stratifying for cN+ and M+ status
[40, 843]. The analyses were balanced for nodal involvement and for planned RT use in STAMPEDE at
randomisation and at analysis. Abiraterone acetate was associated with a non-significant OS improvement
(HR: 0.75, 95% CI: 0.48–1.18) in non-metastatic patients (N0/N+M0), but OS data were still immature with a
low number of events. Furthermore, this was an underpowered subgroup analysis and hypothesis generating
at best. Moreover, subgroup analyses were performed according to the metastatic/non-metastatic status and
to the nodal status (any M) without specific data for the N+M0 population (n = 369; 20% of the overall cohort).
The same would apply for the docetaxel arm in the STAMPEDE trial for which no specific subgroup analysis
of newly diagnosed N+M0 PCa (n = 171, 14% of the overall cohort) was performed. However, the addition of
docetaxel, zoledronic acid, or their combination, did not provide any OS benefit when stratifying by M0 and N+
status.

Table 6.2.4.1: Selected studies assessing local treatment in (any cT) cN1 M0 prostate cancer patients

Study n Design Study period/ Treatment Effect on survival


follow-up arms
Bryant, et al. n = 648 Retrospective 2000-2015 ADT ± EBRT Significant benefit for
2018 [845] (National combined treatment only
Veterans Affairs) 61 mo. if PSA levels less than the
median (26 ng/mL)
All-cause mortality HR: 0.50
CSS, HR: 0.38
Sarkar, et al. n = 741 Retrospective 2000-2015 ADT ± local Significant benefit for RP
2019 [846] (National treatment All cause mortality HR 0.36
Veterans Affairs) 51 mo. (surgery or RT) CSS, HR: 0.32

No statistical difference for


RP vs. RT (p > 0.1)
All-cause mortality HR: 047
CSS, HR: 0.88
Lin, et al. n = 983 Retrospective 2004-2006 ADT ± EBRT Significant benefit for
2015 [841] before (NCDB) combined treatment
propensity 48 mo. 5-yr OS: 73% vs. 52%
score HR: 0.5
matching
Tward, et al. n = 1,100 Retrospective 1988-2006 EBRT Significant benefit for EBRT
2013 [840] (SEER) (n = 397) vs. 5-yr CSS (not OS?) 78% vs.
64 mo. no EBRT 71%
(n=703) HR: 0.66
No information 5-yr. OS: 68% vs. 56%,
on ADT) HR: 0.70
Rusthoven, n = 796 Retrospective 1995-2005 EBRT vs. no Significant benefit for EBRT
et al. 2014 [844] (SEER) EBRT (no 10-yr OS: 45% vs. 29%
61 mo. information on HR: 0.58
ADT)
Seisen, et al. n = 1,987 Retrospective 2003-2011 ADT ± local Significant benefit for
2018 [842] (NCDB) treatment combined treatment
50 mo. (surgery or RT) 5-yr OS: 78.8% vs. 49.2%
HR: 0.31
No difference between RP
and RT

PROSTATE CANCER - LIMITED UPDATE 2021 79


James, et al. n = 177 Unplanned sub- 2005-2014 ADT ± EBRT Significant benefit for
2016 [843] group analysis combined treatment
RCT 17 mo. 5-yr OS: 93% vs. 71%
2-yr FFS: 81% vs 53%
FFS, HR: 0.48
ADT = androgen deprivation therapy; EBRT = external beam radiotherapy in standard fractionation;
FFS = failure-free survival; HR = hazard ratio; mo = months; n = number of patients; OS = overall survival;
RT = radiotherapy; wk = week; yr = year.

6.2.4.3.1 Guidelines for the management of cN1 M0 prostate cancer

Recommendations Strength rating


Offer patients with cN1 disease a local treatment (either radical prostatectomy or intensity Weak
modulated radiotherapy plus image-guided radiotherapy) plus long-term ADT.

6.2.4.4 Options other than surgery and radiotherapy for primary treatment
6.2.4.4.1 Investigational therapies
Currently cryotherapy, HIFU or focal therapies have no place in the management of locally-advanced PCa.

6.2.4.4.2 Androgen deprivation therapy monotherapy


The deferred use of ADT as single treatment modality has been answered by the EORTC 30891 trial [800].
Nine hundred and eighty-five patients with T0-4 N0-2 M0 PCa received ADT alone, either immediately or after
symptomatic progression or occurrence of serious complications. After a median follow-up of 12.8 years, the
OS favoured immediate treatment (HR: 1.21, 95% CI: 1.05–1.39). Surprisingly, no different disease-free or
symptom-free survival was observed, raising the question of survival benefit. In locally-advanced T3-T4 M0
disease unsuitable for surgery or RT, immediate ADT may only benefit patients with a PSA > 50 ng/mL and a
PSA-DT < 12 months or those that are symptomatic [800, 847]. The median time to start deferred treatment
was 7 years. In the deferred treatment arm 25.6% died without needing treatment.

6.2.4.5 Guidelines for radical treatment of locally-advanced disease

Recommendations Strength rating


Radical Prostatectomy (RP)
Offer RP to selected patients with locally-advanced PCa as part of multi-modal therapy. Strong
Extended pelvic lymph node dissection (ePLND)
Perform an ePLND prior to RP in locally-advanced PCa. Strong
Radiotherapeutic treatments
In patients with locally-advanced disease, offer intensity-modulated radiation therapy (IMRT) Strong
plus image-guide radiation therapy in combination with long-term androgen deprivation
therapy (ADT).
Offer long-term ADT for at least 2 years. Weak
Therapeutic options outside surgery and radiotherapy
Do not offer whole gland treatment or focal treatment to patients with locally-advanced Strong
PCa.
Only offer ADT monotherapy to those patients unwilling or unable to receive any form of Strong
local treatment if they have a prostate-specific antigen (PSA)-doubling time < 12 months,
and either a PSA > 50 ng/mL, a poorly-differentiated tumour or troublesome local
disease-related symptoms.
Offer patients with cN1 disease a local treatment (either RP or IMRT plus IGRT) plus Strong
long-term ADT.

6.2.5 Adjuvant treatment after radical prostatectomy


6.2.5.1 Introduction
Adjuvant treatment is by definition additional to the primary or initial therapy with the aim of decreasing the risk
of relapse. A post-operative detectable PSA is an indication of persistent prostate cells (see Section 6.2.6). All
information listed below refers to patients with a post-operative undetectable PSA.

80 PROSTATE CANCER - LIMITED UPDATE 2021


6.2.5.2 Risk factors for relapse
Patients with ISUP grade > 2 in combination with extraprostatic extension (pT3a) and particularly those with
SV invasion (pT3b) and/or positive surgical margins are at high risk of progression, which can be as high as
50% after 5 years [850]. Irrespective of the pT stage, the number of removed nodes [851-858], tumour volume
within the LNs and capsular perforation of the nodal metastases are predictors of early recurrence after RP
for pN1 disease [859]. A LN density (defined as ‘the percentage of positive LNs in relation to the total number
of analysed/removed LNs’) of over 20% was found to be associated with poor prognosis [860]. The number
of involved nodes seems to be a major factor for predicting relapse [853, 854, 861]; the threshold considered
is less than 3 positive nodes from an ePLND [512, 853, 861]. However, prospective data are needed before
defining a definitive threshold value.

6.2.5.2.1 Biomarker-based risk stratification after radical prostatectomy


The Decipher® gene signature consists of a 22-gene panel representing multiple biological pathways and was
developed to predict systemic progression after definitive treatment. A meta-analysis of five studies analysed
the performance of the Decipher® Genomic Classifier (GC) test on men post-RP. The authors showed in
multivariable analysis that Decipher® GC remained a statistically significant predictor of metastasis (HR: 1.30,
95% CI: 1.14–1.47, p < 0.001) per 0.1 unit increase in score and concluded that it can independently improve
prognostication of patients post-RP within nearly all clinicopathologic, demographic, and treatment subgroups
[862].
Further studies are needed to establish how to best incorporate Decipher® GC in clinical decision-
making.

6.2.5.3 Immediate (adjuvant) post-operative external irradiation after RP (cN0 or pN0)


Four prospective RCTs have assessed the role of immediate post-operative RT (adjuvant RT [ART]),
demonstrating an advantage (endpoint, development of BCR) in high-risk patients (e.g., pT2/pT3 with positive
surgical margins and GS 8–10) post-RP (Table 6.2.5.1). In the ARO 96-02 trial, 80% of the pT3/R1/GS 8–10
patients randomised to observation developed BCR within 10 years. It must be emphasised that PSA was
undetectable at inclusion only in the ARO 96-02 trial which presents a major limitation interpreting these
findings as patients with a detectable PSA would now be considered for salvage therapy rather than ART [863].

Table 6.2.5.1: O
 verview of all four randomised trials for adjuvant surgical bed radiation therapy after RP*
(without ADT)

Reference n Inclusion Randomisation Definition of Median Biochemical Overall


criteria BCR PSA (ng/ FU (mo) Progression- survival
mL) free survival
SWOG 8794 431 pT3 cN0 ± 60-64 Gy vs. > 0.4 152 10 yr.: 53% vs. 10 yr.:
2009 [863] involved SM observation 30% 74% vs.
(p < 0.05) 66%
Median
time:
15.2 vs.
13.3 yr.,
p = 0.023
EORTC 1,005 pT3 ± 60 Gy vs. > 0.2 127 10 yr.: 60.6% 81% vs.
22911 2012 involved SM observation vs. 41% 77% n.s.
[864] pN0 pT2 (p < 0.001)
involved SM
pN0
ARO 96-02 388 pT3 (± 60 Gy vs. > 0.05 + 112 10 yr.: 56% vs. 10 yr.:
2014 [865] involved SM) observation confirmation 35% 82% vs.
pN0 PSA (p = 0.0001) 86% n.s.
post-RP
undetectable

PROSTATE CANCER - LIMITED UPDATE 2021 81


FinnProstate 250 pT2,R1/ 66.6 Gy vs. > 0.4 (in 2 112 vs. 10 yr.: 82% vs. 10 yr.:
Group 2019 pT3a observation successive 103 61% 92% vs.
[866] (+SRT) measurements) (patients p < 0.001 87% n.s.
alive)
*See Section 6.3.5.1 for delayed (salvage) post-radical prostatectomy external irradiation.
BCR = biochemical recurrence; FU = follow-up; mo = months; n = number of patients; n.s. = not significant;
OS = overall survival; PSA = prostate-specific antigen; RP = radical prostatectomy; SM = surgical margin.

6.2.5.4 Comparison of adjuvant- and salvage radiotherapy


Two retrospective matched studies (510 and 149 patients receiving ART) failed to show an advantage for
metastasis-free survival [867, 868]. However, both studies were underpowered for high-risk patients (pT3b/R1/
ISUP grade 4–5 PCa).
In contrast to these studies, a propensity score-matched retrospective analysis of two cohorts of
366 pT3 and/or R1 patients found that compared to SRT at a PSA between 0.1 and 0.5 ng/mL, ART given at
an undetectable PSA (< 0.1 ng/mL) improved all three endpoints; BCR, metastasis-free survival, and OS [869].

Both approaches (ART and early SRT) together with the efficacy of adjuvant ADT are compared in three
prospective RCTs: the Medical Research Council (MRC) Radiotherapy and Androgen Deprivation In
Combination After Local Surgery (RADICALS) [870], the Trans-Tasman Oncology Group (TROG) Radiotherapy
Adjuvant Versus Early Salvage (RAVES) [871], and Groupe d’Etude des Tumeurs Uro-Génitales (GETUG-AFU
17) [872]. In addition, a preplanned meta-analysis of all three trials has been published (Table 6.2.5.2) [873].
Two trials closed early after randomising 333/470 patients (RAVES) and 424/718 (GETUG-AFU-17) patients.
RADICALS-RT included 1,396 patients with the option of subsequent inclusion in RADICALS-HT; 154/649
(24%) of patients starting in the adjuvant RT group also received neo-adjuvant or adjuvant hormone therapy;
90 patients for 6 months/45 for 2 years/19 patients outside RADICALS-HT. From the SRT group, 61/228 (27%)
received neo-adjuvant or adjuvant hormone therapy for 6 months (n = 33) and 2 years (n = 13). Fifteen of these
patients were treated outside the trial [870]. All men in the GETUG-AFU-17 trial (n = 424) received 6 months of
hormone treatment. All together, 684 out of 2,153 patients received additional ADT for at least 6 months across
both trials [873]. Radiotherapy to the pelvic lymphatics was allowed in the GETUG-AFU and in the RADICALS-
RT Trials. To be included in these trails, patients must have had a post-RP PSA < 0.2 ng/mL (RADICALS-RT)
or < 0.1 ng/mL (RAVES, GETUG-AFU17) (see Table 6.2.5.2). Biochemical progression was defined as a PSA
> 0.4 ng/mL after RT.
The primary endpoint for RAVES and GETUG-AFU 17 was biochemical PFS and for RADICALS-RT
metastasis-free survival. So far only PFS data has been reported, and not metastasis-free survival or OS data.
With a median follow-up between 4.9 years and 6.25 years there was no statistical significant difference for
biochemical PFS (BPFS) for both treatments in all three trials (see Table 6.2.5.2) indicating that in the majority
of patients adjuvant irradiation should be avoided. Additionally, there was a significant lower rate of grade
> 2 GU late side effects and grade 3–4 urethral strictures in favour of early SRT; which may also be caused by
the low number of patients with PSA-progression and subsequent need for early SRT at the time of analysis
(40% of patients).

It is important to note that the indication for ART changed over the last ten years with the introduction of ultra-
sensitive PSA-tests, favouring early SRT. Therefore many patients, randomised in these three trials (accruing
2006–2008) are not likely to benefit from ART as there is a low risk of biochemical progression (~20–30%) in,
for example, pT3R0- or pT2R1-tumours. The median pre-SRT-PSA in all 3 trials was 0.24 ng/mL which is much
lower than the conventional cut-off level of PSA < 0.5 ng/mL used to base ´early´ SRT on. Therefore, patients
with ‘low-risk factors’ of biochemical progression after RP should be closely followed up with ultra-senstive
assays and SRT should start as soon as PSA starts to rise, which has to be confirmed by a second PSA
measurement. The proportion of ´very high risk´ patients (with at least two out of three features: pT3/R1/GS
8-10) in all three trials was low (between 10–20%) and therefore even the meta-analysis may be underpowered
to show an outcome in favour of SRT [873]. In addition, the side-effect profile may have been impacted with a
larger proportion of ART patients receiving treatment with older 3D-treatment planning techniques as compared
to SRT patients (GETUG-AFU 17: ART, 69% 3D vs. 46% SRT) and patients treated more recently were more
likely to undergo IMRT techniques with a proven lower rate of late side effects [610].

For these reasons 10-year results and results of metastasis-free survival endpoints should be awaited before
drawing final conclusions. Due to the small number of patients with a combination of high-risk features
included in these three trials (between 10–20%) ART remains a recommended treatment option in highly
selected patients with at least two out of three high-risk features (‘very high-risk patients’) such as pT3/R1/
ISUP > 3 [854, 874].

82 PROSTATE CANCER - LIMITED UPDATE 2021


Table 6.2.5.2 O
 verview of all four randomised trials and one meta-analysis for patients treated with
adjuvant versus early salvage RT after radical prostatectomy

Reference n Inclusion
Randomisation Definition Median BPFS OS or Side effects
criteria of BCR PSA FU (yr) MFS
(ng/mL)
RAVES 333 pT3a/pT3b 64 Gy ART > 0.4 post 6.1 5 yr.: n.r. LT grade
TROG 08.03/ target any T - SM+ PSA: (< 0.1) vs. RT 86% vs. > GU:
ANZUP was PSA post- 64 Gy early SRT 87% 70% vs.
2020 [871] 470 RP: < 0.1 at PSA > 0.2 (p > 0.05) 54%
early med. pre-SRT: (p = 0.002)
closed n.r.
RADICALS- 1,396 pT3a/ 52.5 Gy (20 Fx) > 0.4 or 2 at 4.9 5 yr.: n.r. SR urinary
RT pT3b/pT4 or 66 Gy (33 Fx) any time 85% vs. incontinence
2020 [870] PSA > 10 ART 88% 1yr: 4.8 vs.
pre-RP early SRT (p = 0.56) 4 (p = 0.023)
any T, SM+ identical urethral
Gleason at PSA > 0.1 stricture
7-10 med.pre-SRT: grade 3/4
PSA post 0.2 2 yr:
RP: < 0.2 6% vs. 4%
(p = 0.02)
GETUG-AFU 424 pT3a/pT3b/ 66 Gy (ART) vs. > 0.4 6.25 5 yr: n.r. LT grade > 2
17 target pT4a and 66 Gy early SRT 92% vs. GU 27% vs.
2020 [872] was SM+ at PSA 0.1 90% 7%
718 PSA both groups: (p = 0.42) (p < 0.001)
early post-RP: 6 mo. LHRH-A ED: 28% vs.
closed < 0.1 med. pre-SRT 8%
0.24 (p < 0.001)
ARTISTIC- 2,153 see above see above see above 4.5 5 yr.: n.r. n.r.
Meta-analysis 89% vs.
2020 [873] 88%
p = 0.7
ART = adjuvant radiotherapy; BCR = biochemical recurrence; BPFS = biochemical progression-free survival;
ED = erectile dysfunction; FU = follow-up; Fx = fraction; GU = genitourinary; LT = late toxicity; mo = months;
med = median; MFS = metastasis-free survival; n.r. = not reported; OS = overall survival; RP = radical
prostatectomy; RT = radiotherapy; SR = self reported; SRT = salvage radiotherapy; + = positive; yr = year.

6.2.5.5 Adjuvant androgen ablation in men with N0 disease


Adjuvant androgen ablation with bicalutamide 150 mg daily did not improve PFS in localised disease while it
did for locally-advanced disease after RT. However this never translated to an OS benefit [875]. A systematic
review showed a possible benefit for PFS but not OS for adjuvant androgen ablation [489].

The TAX3501 trial comparing the role of leuprolide (18 months) with and without docetaxel (6 cycles) ended
prematurely due to poor accrual. A recent phase III RCT comparing adjuvant docetaxel against surveillance
after RP for locally-advanced PCa showed that adjuvant docetaxel did not confer any oncological benefit [876].
Consequently, adjuvant chemotherapy after RP should only be considered in a clinical trial [877].

6.2.5.6 Adjuvant treatment in pN1 disease


6.2.5.6.1 Adjuvant androgen ablation alone
The combination of RP and early adjuvant HT in pN+ PCa has been shown to achieve a 10-year CSS rate of
80% and has been shown to significantly improve CSS and OS in prospective RCTs [848, 849]. However, these
trials included mostly patients with high-volume nodal disease and multiple adverse tumour characteristics and
these findings may not apply to men with less extensive nodal metastases.

6.2.5.6.2 Adjuvant radiotherapy combined with ADT in pN1 disease


In a retrospective multi-centre cohort study, maximal local control with RT to the prostatic fossa appeared to
be beneficial in PCa patients with pN1 after RP, treated ‘adjuvantly‘ with continuous ADT (within 6 months after
surgery irrespective of PSA). The beneficial impact of adjuvant RT on survival in patients with pN1 PCa was

PROSTATE CANCER - LIMITED UPDATE 2021 83


highly influenced by tumour characteristics. Men with low-volume nodal disease (< 3 LNs), ISUP grade 2–5 and
pT3–4 or R1, as well as men with 3 to 4 positive nodes were more likely to benefit from RT after surgery, while
the other subgroups did not [878].
These results were confirmed by a US National Cancer Database analysis based on 5,498 patients
[879]. Another US National Cancer Database study including 8,074 pN1 patients reports improved OS after
ADT + EBRT (including pelvic LNs) vs. observation and vs. ADT alone in all men with single or multiple adverse
pathological features. Men without any adverse pathological features did not benefit from immediate adjuvant
therapy [880].

In a series of 2,596 pN1 patients receiving ADT (n = 1,663) or ADT plus RT (n = 906), combined treatment was
associated with improved OS, with a HR of 1.5 for ADT alone [881]. In a SEER retrospective population-based
analysis, adding RT to RP showed a non-significant trend for improved OS but not PCa-specific survival, but
data on the extent of additional RT is lacking in this study [844]. Radiotherapy should be given to the pelvic
lymphatics and the prostatic fossa [878, 882-884]. For pN1 patients no data are available regarding adjuvant
EBRT without ADT.

6.2.5.6.3 Observation of pN1 patients after radical prostatectomy and extended lymph node dissection
Several retrospective studies and a systematic review addressed the management of patients with pN1 PCa at
RP [861, 878, 884-886]. A subset of patients with limited nodal disease (1–2 positive LNs) showed favourable
oncological outcomes and did not require additional treatment.
An analysis of 209 pN1 patients with one or two positive LNs at RP showed that 37% remained
metastasis-free without need of salvage treatment at a median follow-up of 60.2 months [885]. Touijer et al.
reported their results of 369 LN-positive patients (40 with and 329 without adjuvant treatment) and showed that
higher pathologic grade group and three or more positive LNs were significantly associated with an increased
risk of BCR on multivariable analysis [861]. Biochemical-free survival rates in pN1 patients without adjuvant
treatment ranged from 43% at 4 years to 28% at 10 years [886]. Reported CSS rates were 78% at 5 years
and 72% at 10 years. The majority of these patients were managed with initial observation after surgery, had
favourable disease characteristics, and 63% had only one positive node [886].

6.2.5.7 Guidelines for adjuvant treatment in pN0 and pN1 disease after radical prostatectomy

Recommendations Strength rating


Do not prescribe adjuvant androgen deprivation therapy (ADT) in pN0 patients. Strong
Only offer adjuvant intensity-modulated radiation therapy (IMRT) plus image-guided Strong
radiation therapy (IGRT) to high-risk patients (pN0) with at least two out of three high-risk
features (ISUP grade group 4–5, pT3 ± positive margins).
Discuss three management options with patients with pN1 disease after an extended lymph Weak
node dissection, based on nodal involvement characteristics:
1. Offer adjuvant ADT;
2. Offer adjuvant ADT with additional IMRT plus IGRT;
 ffer observation (expectant management) to a patient after eLND and < 2 nodes and a
3. O
PSA < 0.1 ng/mL.

6.2.5.8 Guidelines for non-curative or palliative treatments in prostate cancer

Recommendations Strength rating


Watchful waiting (WW) for localised prostate cancer
Offer WW to asymptomatic patients not eligible for local curative treatment and those with a Strong
short life expectancy.
Watchful waiting for locally-advanced prostate cancer
Offer a deferred treatment policy using androgen deprivation monotherapy to M0 Weak
asymptomatic patients with a prostate-specific antigen (PSA) doubling time > 12 months, a
PSA < 50 ng/mL and well-differentiated tumour, who are unwilling or unable to receive any
form of local treatment.

84 PROSTATE CANCER - LIMITED UPDATE 2021


6.2.6 Persistent PSA after radical prostatectomy
Between 5 and 20% of men continue to have detectable or persistent PSA after RP (when defined in the
majority of studies as detectable post-RP PSA of > 0.1 ng/mL within 4 to 8 weeks of surgery) [887, 888]. It may
result from persistent local disease, pre-existing metastases or residual benign prostate tissue.

6.2.6.1 Natural history of persistently elevated PSA after RP


Several studies have shown that persistent PSA after RP is associated with more advanced disease (such as
positive surgical margins (PSM), pathologic stage > T3a, positive nodal status and/or pathologic ISUP grade
> 3) and poor prognosis (See table 6.2.6.1). Initially defined as > 0.1 ng/mL, improvements in the sensitivity of
PSA assays now allow for the detection of PSA at much lower levels.
Moreira et al., demonstrated that failure to achieve a PSA of less than 0.03 ng/mL within 6 months
of surgery was associated with an increased risk of BCR and overall mortality [889, 890]. However, since the
majority of the published literature is based on the > 0.1 ng/mL PSA cut-off there is significantly more long-
term data for this definition.

Predictors of PSA persistence were higher BMI, higher pre-operative PSA and ISUP grade > 3 [890]. In patients
with PSA persistence, one and 5-year BCR-free survival were 68% and 36%, compared to 95% and 72%,
respectively, in men without PSA persistence [889]. Ten-year OS in patients with and without PSA persistence
was 63% and 80%, respectively. In line with these data, Ploussard et al. reported that approximately 74% of
patients with persistent PSA develop BCR [887]. Spratt et al., confirmed that a persistently detectable PSA
after RP represents one of the worst prognostic factors associated with oncological outcome [891]. Of 150
patients with a persistent PSA, 95% received RT before detectable metastasis. In a multivariable analysis
the presence of a persistently detectable PSA post-RP was associated with a 4-fold increase in the risk of
developing metastasis. This was confirmed by recent data from Preisser et al. who showed that persistent PSA
is prognostic of an increased risk of metastasis and death [892]. At 15 years after RP, metastasis-free survival
rates, OS and CSS rates were 53.0 vs. 93.2% (p < 0.001), 64.7 vs. 81.2% (p < 0.001) and 75.5 vs. 96.2%
(p < 0.001) for persistent vs. undetectable PSA, respectively. The median follow-up was 61.8 months for
patients with undetectable PSA vs. 46.4 months for patients with persistent PSA. In multivariable Cox
regression models, persistent PSA represented an independent predictor for metastasis (HR: 3.59, p < 0.001),
death (HR: 1.86, p < 0.001) and cancer-specific death (HR: 3.15, p < 0.001).
However, not all patients with persistent PSA after RP experience disease recurrence. Xiang et al.,
showed a 50% 5-year BCR-free survival in men who had a persistent PSA level > 0.1 but < 0.2 ng/mL at
6-8 weeks after RP [893].

Rogers et al., assessed the clinical outcome of 160 men with a persistently detectable PSA level after RP [894].
No patient received adjuvant therapy before documented metastasis. In their study, 38% of patients had no
evidence of metastases for > 7 years while 32% of the patients were reported to develop metastases within 3
years. Noteworthy is that a significant proportion of patients had low-risk disease. In multivariable analysis the
PSA slope after RP (as calculated using PSA levels 3 to 12 months after surgery) and pathological ISUP grade
were significantly associated with the development of distant metastases.

Table 6.2.6.1: Studies on the natural history of patients with persistent PSA after RP

Authors Study n Definition PSA Treatment Outcome Other details/comments


population persistence
Ploussard 496 men PSA > 0.1 ng/mL 74.4% with
et al., pN0 with at 6 wk. BCR
J Urol persistent 5% with
2013 PSA 14 metastasis
[887] centres
1998 - 2011

PROSTATE CANCER - LIMITED UPDATE 2021 85


Moreira 901 men 230 (8 pN1) PSA No RT info Increased Relative to men with
et al., Shared persistence risk for BCR undetectable PSA levels,
BJUI Equal definition of a after surgery those with a PSA nadir of
2010 Access PSA nadir > 0.03 0.03 (HR: 3.88, p < 0.001),
[890] Regional ng/mL 0.04 (HR: 4.87, p < 0.001),
Cancer 0.05-0.09 (HR: 12.69, p <
Hospital 0.001), 0.1-0.19 (HR: 13.17,
(SEARCH) p < 0.001), and 0.2 ng/mL
database. (HR: 13.23, p < 0.001) were at
2001-2008 increased risk of BCR while
men with a nadir of 0.01 (HR:
1.36, p = 0.400) and 0.02 (HR:
1.64, p = 0.180) were not.
Moreira 1,156 men 291 PSA > 0.03 ng/mL No RT info Increased Median FU 48 mo. In patients
et al., Shared (10 within 6 mo. BCR and with persistent PSA 1 and
J Urol Equal pN1) overall 5-yr. BFS was 68% and 36%,
2009 Access mortality significantly lower than 95%
[889] Regional and 72%, respectively, in men
Cancer without persistent PSA.
Hospital 10-year OS in patients
(SEARCH) with vs. without persistent
database. PSA was 63% vs. 80%. In
After 1997 men with persistent PSA
independent predictors of
BCR were higher PSA
nadir (HR: 2.19, p < 0.001),
positive surgical margins
(HR: 1.75, p = 0.022) and high
pathological ISUP grade (4-5
vs. 1, HR: 2.40, p = 0.026).
Independent predictors of
OM were a higher PSA nadir
(HR: 1.46, p 0.013) and
seminal vesicle invasion
(HR: 3.15, p = 0.047)
Rogers 224 men 160 >
PSA 0.1 ng/mL No Metastasis- Mean FU 5.3 yr. 75 men
et al., Single- men at 3 mo. treatment free survival (47%) developed distant
Cancer centre (58 before at 3, 5 and metastases after RP (median
2004 (Johns pN1) onset of 10 yr. was time to metastases 5.0 yr.;
[894] Hopkins) metastasis 68%, 49%, range, 0.5-13 yr.).
1989 - 2002 and 22%,
respectively. The slope of PSA changes
approximately 3-12 mo.
after RP at a cut-off value
> 0.05 ng/mL was found
to be predictive of distant
metastasis-free survival
(HR: 2.9, p < 0.01).
BCR = biochemical recurrence; FU = follow-up; HR = hazard ratio; mo = months; n = number of patients;
OM = overall mortality; PSA = prostate-specific antigen; RT = radiotherapy.

6.2.6.2 Imaging in patients with persistently elevated PSA after RP


Standard imaging with bone scan and MRI has a low pick-up rate in men with a PSA below 2 ng/mL. However,
PSMA PET/CT has been shown to identify residual cancer with positivity rates of 33%, 46%, 57%, 82%, and
97%, in men with post-RP PSA ranges of 0–0.19, 0.2–0.49, 0.5–0.99, 1–1.99, and > 2 ng/mL, respectively
[397, 895-899] which can guide SRT planning [900]. Based on these post-RP PSA ranges, Schmidt-Hegemann
et al. studied 129 patients who had either persistent PSA (52%) or BCR (48%) after RP, showing that men with
a persistent PSA had significantly more pelvic nodal involvement on PSMA PET/CT than those developing
a detectable PSA [901]. In a multi-centre retrospective study including 191 patients, 68Ga-PSM localised
biochemical persistence after RP in more than two-thirds of patients with high-risk PCa features. The obturator

86 PROSTATE CANCER - LIMITED UPDATE 2021


and presacral/mesorectal nodes were identified as high risk for residual disease [902]. Another retrospective
study included 150 patients with persistent PSA after RARP who were re-staged with both 68Ga-PSMA and
18F-DCFPyL PSMA. The authors found that in the presence of persistent PSA the majority of patients already

had metastatic pelvic LNs or distant metastases which would support a role of PSMA PET/CT imaging in
guiding (salvage) treatment strategies [903]. At present there is uncertainty regarding the best treatment if
PSMA PET/CT shows metastatic disease.

6.2.6.3 Impact of post-operative RT and/or ADT in patients with persistent PSA


The benefit of SRT in patients with persistent PSA remains unclear due to a lack of RCTs, however, it would
appear that men with a persistent PSA do less well than men with BCR undergoing RT.

Wiegel et al., [888] showed that following salvage RT to the prostate bed, patients with a detectable PSA after
RP had significantly worse oncological outcomes when compared with those who achieved an undetectable
PSA. Ten-year metastasis-free survival was 67% vs. 83% and OS was 68% vs. 84%, respectively. Recent
data from Preisser et al., [892] also compared oncological outcomes of patients with persistent PSA who
received SRT vs. those who did not. In the subgroup of patients with persistent PSA, after 1:1 propensity score
matching between patients with SRT vs. no RT, OS rates at 10 years after RP were 86.6 vs. 72.6% in the entire
cohort (p < 0.01), 86.3 vs. 60.0% in patients with positive surgical margin (p = 0.02), 77.8 vs. 49.0% in pT3b
disease (p < 0.001), 79.3 vs. 55.8% in ISUP grade 1 disease (p < 0.01) and 87.4 vs. 50.5% in pN1 disease
(p < 0.01), respectively. Moreover, CSS rates at 10 years after RP were 93.7 vs. 81.6% in the entire cohort
(p < 0.01), 90.8 vs. 69.7% in patients with positive surgical margin (p = 0.04), 82.7 vs. 55.3% in pT3b disease
(p < 0.01), 85.4 vs. 69.7% in ISUP grade 1 disease (p < 0.01) and 96.2 vs. 55.8% in pN1 disease (p < 0.01), for
SRT vs. no RT, respectively. In multivariable models, after 1:1 propensity score matching, SRT was associated
with lower risk for death (HR: 0.42, p = 0.02) and lower cancer-specific death (HR: 0.29, p = 0.03). These
survival outcomes in patients with persistent PSA who underwent SRT suggest they benefit but outcomes are
worse than for men experiencing BCR.

It is clear from a number of studies that poor outcomes are driven by the level of pre-RT PSA, the presence of
ISUP grade > 4 in the RP histology and pT3b disease [888, 904-908]. Fossati et al., suggested that only men
with a persistent PSA after RP and ISUP grade < 3 benefit significantly [909], although this is not supported by
Preisser et al. [892]. The current data does not allow making any clear treatment decisions.

Addition of ADT may improve PFS [904]. Choo et al. studied the addition of 2-year ADT to immediate RT to
the prostate bed in patients with pathologic T3 disease (pT3) and/or positive surgical margins after RP [904].
Twenty-nine of the 78 included patients had persistently detectable post-operative PSA. The relapse-free rate
was 85% at 5 years and 68% at 7 years, which was superior to the 5-year progression-free estimates of 74%
and 61% in the post-operative RT arms of the EORTC and the SWOG studies, respectively, which included
patients with undetectable PSA after RP [863, 864]. Patients with persistently detectable post-operative PSA
comprised approximately 50% and 12%, respectively, of the study cohorts in the EORTC and the SWOG
studies.
In the ARO 96-02, a prospective RCT, 74 patients with PSA persistence (20%) received immediate
SRT only (66 Gy per protocol [arm C]). The 10-year clinical relapse-free survival was 63% [888]. The GETUG-22
trial comparing RT with RT plus short-term ADT for post-RP PSA persistence (0.2–2.0 ng/mL) reported good
tolerability of the combined treatment. The oncological endpoints are yet to be published [910].

Ploussard and colleagues recently performed a systematic review of oncologic outcomes and effectiveness
of salvage therapies in men with persistent PSA after RP. Their findings confirmed a strong correlation of
PSA persistence with poor oncologic outcomes [911]. The authors also report that SRT was associated with
improved survival outcomes, although the available evidence is of low quality [911].

6.2.6.4 Conclusion
The available data suggest that patients with PSA persistence after RP may benefit from early aggressive multi-
modality treatment, however, the lack of prospective RCTs makes firm recommendations difficult.

PROSTATE CANCER - LIMITED UPDATE 2021 87


6.2.6.5 Recommendations for the management of persistent PSA after radical prostatectomy

Recommendations Strength rating


Offer a prostate-specific membrane antigen positron emission tomography (PSMA PET) Weak
scan to men with a persistent prostate-specific antigen > 0.2 ng/mL if the results will
influence subsequent treatment decisions.
Treat men with no evidence of metastatic disease with salvage radiotherapy and additional Weak
hormonal therapy.

6.3 Management of PSA-only recurrence after treatment with curative intent


Follow-up will be addressed in Chapter 7 and is not discussed here.

6.3.1 Background
Between 27% and 53% of all patients undergoing RP or RT develop a rising PSA (PSA recurrence). Whilst
a rising PSA level universally precedes metastatic progression, physicians must inform the patient that the
natural history of PSA-only recurrence may be prolonged and that a measurable PSA may not necessarily
lead to clinically apparent metastatic disease. Physicians treating patients with PSA-only recurrence face a
difficult set of decisions in attempting to delay the onset of metastatic disease and death while avoiding over-
treating patients whose disease may never affect their OS or QoL. It should be emphasised that the treatment
recommendations for these patients should be given after discussion in a multidisciplinary team.

6.3.2 Definitions of clinically relevant PSA relapse


The PSA level that defines treatment failure depends on the primary treatment. Patients with rising PSA
after RP or primary RT have different risks of subsequent symptomatic metastatic disease based on various
parameters, including the PSA level. Therefore, physicians should carefully interpret BCR endpoints when
comparing treatments.
After RP, the threshold that best predicts further metastases is a PSA > 0.4 ng/mL and rising [912-
914]. However, with access to ultra-sensitive PSA testing, a rising PSA much below this level will be a cause
for concern for patients. After primary RT, with or without short-term hormonal manipulation, the RTOG-ASTRO
Phoenix Consensus Conference definition of PSA failure (with an accuracy of > 80% for clinical failure) is ‘any
PSA increase > 2 ng/mL higher than the PSA nadir value, regardless of the serum concentration of the nadir’
[915].
After HIFU or cryotherapy no endpoints have been validated against clinical progression or
survival; therefore, it is not possible to give a firm recommendation of an acceptable PSA threshold after these
alternative local treatments [916].

6.3.3 Natural history of biochemical recurrence


Once a PSA relapse has been diagnosed, it is important to determine whether the recurrence has developed
at local or distant sites. A recent systematic review and meta-analysis investigated the impact of BCR on
hard endpoints and concluded that patients experiencing BCR are at an increased risk of developing distant
metastases, PCa-specific and overall mortality [916]. However, the effect size of BCR as a risk factor for
mortality is highly variable. After primary RP its impact ranges from HR 1.03 (95% CI: 1.004–1.06) to HR 2.32
(95% CI: 1.45–3.71) [917, 918]. After primary RT, OS rates are approximately 20% lower at 8 to 10 years
follow-up even in men with minimal co-morbidity [919, 920]. Still, the variability in reported effect sizes of
BCR remains high and suggests that only certain patient subgroups with BCR might be at an increased risk of
mortality.
The risk of subsequent metastases, PCa-specific- and overall mortality may be predicted by the
initial clinical and pathologic factors (e.g., T-category, PSA, ISUP grade) and PSA kinetics (PSA-DT and interval
to PSA failure), which was further investigated by this systematic review [916].

For patients with BCR after RP, the following outcomes were found to be associated with significant prognostic
factors:
• distant metastatic recurrence: positive surgical margins, high RP specimen pathological ISUP grade, high
pT category, short PSA-DT, high pre-SRT PSA;
• prostate-cancer-specific mortality: high RP specimen pathological ISUP grade, short interval to
biochemical failure as defined by investigators, short PSA-DT;
• overall mortality: high RP specimen pathological ISUP grade, short interval to biochemical failure, high
PSA-DT.

88 PROSTATE CANCER - LIMITED UPDATE 2021


For patients with BCR after RT, the corresponding outcomes are:
• distant metastatic recurrence: high biopsy ISUP grade, high cT category, short interval to biochemical
failure;
• prostate-cancer-specific mortality: short interval to biochemical failure;
• overall mortality: high age, high biopsy ISUP grade, short interval to biochemical failure, high initial
(pre-treatment) PSA.

Based on this meta-analysis, proposal is to stratify patients into ‘EAU Low-Risk BCR’ (PSA-DT > 1 year AND
pathological ISUP grade < 4 for RP; interval to biochemical failure > 18 months AND biopsy ISUP grade < 4 for
RT) or ‘EAU High-Risk BCR’ (PSA-DT < 1 year OR pathological ISUP grade 4–5 for RP, interval to biochemical
failure < 18 months OR biopsy ISUP grade 4–5 for RT), since not all patients with BCR will have similar
outcomes. The stratification into ‘EAU Low-Risk’ or ‘EAU High-Risk’ BCR has recently been validated in a
European cohort [921].

6.3.4 The role of imaging in PSA-only recurrence


Imaging is only of value if it leads to a treatment change which results in an improved outcome. In practice,
however, there are very limited data available regarding the outcomes consequent on imaging at relapse.

6.3.4.1 Assessment of metastases


6.3.4.1.1 Bone scan and abdominopelvic CT
Because BCR after RP or RT precedes clinical metastases by 7 to 8 years on average [855, 922], the diagnostic
yield of common imaging techniques (bone scan and abdominopelvic CT) is low in asymptomatic patients
[923]. In men with PSA-only relapse after RP the probability of a positive bone scan is < 5%, when the PSA
level is < 7 ng/mL [924, 925]. Only 11–14% of patients with BCR after RP have a positive CT [924]. In a series
of 132 men with BCR after RP the mean PSA level and PSA velocity associated with a positive CT were
27.4 ng/mL and 1.8 ng/mL/month, respectively [926].

6.3.4.1.2 Choline PET/CT


In two different meta-analyses the combined sensitivities and specificities of choline PET/CT for all sites of
recurrence in patients with BCR were 86–89% and 89–93%, respectively [927, 928].
Choline PET/CT may detect multiple bone metastases in patients showing a single metastasis on
bone scan [929] and may be positive for bone metastases in up to 15% of patients with BCR after RP and
negative bone scan [930]. The specificity of choline PET/CT is also higher than bone scan with fewer false-
positive and indeterminate findings [401]. Detection of LN metastases using choline PET/CT remains limited
by the relatively poor sensitivity of the technique (see Section 5.3.2.3). Choline PET/CT sensitivity is strongly
dependent on the PSA level and kinetics [409, 931, 932]. In patients with BCR after RP, PET/CT detection
rates are only 5–24% when the PSA level is < 1 ng/mL but rises to 67–100% when the PSA level is > 5 ng/mL.
Despite its limitations, choline PET/CT may change medical management in 18–48% of patients with BCR after
primary treatment [933-935].
Choline PET/CT should only be recommended in patients fit enough for curative loco-regional
salvage treatment. The sensitivity of choline PET is known to be strongly influenced by PSA level and kinetics
and drops to sub-optimal values in patients with a low PSA [932]; after RP a possible PSA cut-off level for
choline PET/CT analysis seems to be between 1 and 2 ng/mL [932].

6.3.4.1.3 Fluoride PET and PET/CT


18F-NaF PET/CT has a higher sensitivity than bone scan in detecting bone metastases [936]. However,
18F-NaF PET is limited by a relative lack of specificity and by the fact that it does not assess soft-tissue

metastases [937].

6.3.4.1.4 Fluciclovine PET/CT


18F-Fluciclovine PET/CT has a slightly higher sensitivity than choline PET/CT in detecting the site of relapse in
BCR [938]. In a recent multi-centre trial evaluating 596 patients with BCR in a mixed population (33.3% after
RP, 59.5% after RT ± RP, 7.1% other) fluciclovine PET/CT showed an overall detection rate of 67.7%, with a
sensitivity of 62.7% (95% CI: 56–69%); lesions could be visualised either at local level (38.7%) or in LNs and
bones (9%) [939]. As for choline PET/CT, fluciclovine PET/CT sensitivity is dependent on the PSA level, with a
sensitivity likely inferior to 50% at PSA < 1 ng/mL.
A retrospective study investigated 18F-fluciclovine PET/CT and pelvic MRI for PCa at PSA
relapse after initial treatment. In the examined cohort of 129 patients 18F-Fluciclovine PET/CT and pelvic
MRI demonstrated a high degree of concordance (~94%). The sensitivity, specificity, PPV and NPV for
18F-fluciclovine PET/CT and MRI were 96.6%, 94.3%, 93.4%, and 97%, and 91.5%, 95.7%, 94.7%, and 93%,

PROSTATE CANCER - LIMITED UPDATE 2021 89


respectively, demonstrating similar diagnostic performance between the two modalities [940].
18F-fluciclovine has been approved in the U.S. and Europe and it is therefore the only PCa-specific

radiotracer widely commercially available.

6.3.4.1.5 Prostate-specific membrane antigen based PET/CT


Prostate-specific membrane antigen PET/CT has shown good potential in patients with BCR, although most
studies are limited by their retrospective design. Reported predictors of 68Ga-PSMA PET in the recurrence
setting were recently updated based on a high-volume series (see Table 6.3.1) [397]. Positivity rates in the
prostate bed were significantly different in patients after RP (22%) and RT (52%). High sensitivity (75%) and
specificity (99%) were observed on per-lesion analysis.

Table 6.3.1: PSMA-positivity separated by PSA level category [397]

PSA (ng/mL) 68Ga-PMSA PET positivity


< 0.2 33% (CI: 16–51)
02–0.49 45% (CI: 39–52)
0.5–0.99 59% (CI: 50–68)
1.0–1.99 75% (CI: 66–84)
2.0+ 95% (CI: 92–97)
PSA = prostate-specific antigen; 68Ga-PSMA PET = Gallium-68 prostate-specific membrane antigen positron
emission tomography.

Prostate-specific membrane antigen PET/CT seems substantially more sensitive than choline PET/CT,
especially for PSA levels < 1 ng/mL [941, 942]. In a study of 314 patients with BCR after treatment and a
median PSA level of 0.83 ng/mL, 68Ga-PSMA PET/CT was positive in 197 patients (67%). Of the 88 patients
with negative choline PET/CT, 59 (67%) were positive on 68Ga-PSMA PET/CT [943]. In another prospective
multi-centre trial including 635 patients with BCR after RP (41%), RT (27%), or both (32%), PPV for 68Ga-PSMA
PET/CT was 0.84 (95% CI: 0.75–0.90) by histopathologic validation (primary endpoint, n = 87) and 0.92
(95% CI: 0.75–0.90) by a composite reference standard. Detection rates significantly increased with PSA, from
38% for PSA < 0.5 ng/mL (n = 136) to 57% for PSA of 0.5–1.0 ng/mL (n = 79), 84% for PSA of 1.0–2.0 ng/mL
(n = 89), 86% for PSA of 2.0–5.0 ng/mL (n = 158), and 97% for PSA > 5.0 ng/mL (n = 173, p < 0.001) [944].
It is worth noting that the term ‘PSMA PET’ refers to several different radiopharmaceuticals; the
majority of published studies used 68Ga-PSMA-11 [397, 895, 897, 945, 946], but other authors are reporting
data with 18F-labelled PSMA [898, 899]. At present there are no conclusive data about comparison of such
tracers [947].

6.3.4.1.6 Whole-body and axial MRI


Little is known regarding the accuracy of whole-body or axial MRI in patients with BCR after RP or RT [948].
Therefore, the role of these techniques in detecting occult bone or LN metastases in the case of BCR requires
further assessment.

6.3.4.2 Assessment of local recurrences


6.3.4.2.1 Local recurrence after radical prostatectomy
Because the sensitivity of anastomotic biopsies is low, especially for PSA levels < 1 ng/mL [923], salvage RT
is usually decided on the basis of BCR without histological proof of local recurrence. The dose delivered to the
prostatic fossa tends to be uniform since it has not been demonstrated that a focal dose escalation at the site
of recurrence improves the outcome. Therefore, most patients undergo salvage RT without local imaging.

Magnetic resonance imaging can detect local recurrences in the prostatic bed but its sensitivity in patients
with a PSA level < 0.5 ng/mL remains controversial [949, 950]. Choline PET/CT is less sensitive than MRI
when the PSA level is < 1 ng/mL [951]. In a retrospective study of 53 patients with BCR after RP (median PSA
level 1.5 ng/mL) who underwent 18F-choline whole body hybrid PET/MRI, MRI identified more local relapses
(17 vs. 14, p = 0.453) while PET outperformed whole-body MRI for detection of regional (16 vs. 9, p = 0.016)
and distant (12 vs. 6, p = 0.031) metastases [952].
The detection rates of 68Ga-PSMA PET/CT in patients with BCR after RP are 11.3–50% for PSA
levels < 0.2 ng/mL, 20–72% for PSA levels of 0.20–0.49 ng/mL and 25–87.5% for PSA levels of 0.5–1 ng/mL
[953]. Prostate-specific membrane antigen PET/CT sudies showed that a substantial part of recurrences after
RP were located outside the prostatic fossa even at low PSA levels. In a retrospective study of 125 patients
with a median PSA level of 0.40 ng/mL (interquartile range 0.28–0.63), PSMA-expressing disease was found in

90 PROSTATE CANCER - LIMITED UPDATE 2021


66 patients (53%), of whom 25 patients (38%) had PSMA-avid lesions outside the pelvis, 33 (50%) had lesions
confined to the pelvic LNs and prostate bed, and 8 (12%) had recurrence only in the prostate fossa [954]. In
another retrospective study of 119 men with a mean PSA level of 0.34 ng/mL, 68Ga-PSMA PET/CT was positive
in 41 patients (34.4%); only 3 were positive in the prostatic fossa, while 21 had uptake in the skeleton, 4 in
retroperitoneal LNs and 21 in pelvic LNs [895]. Combining 68Ga-PSMA PET and MRI may improve the detection
of local recurrences, as compared to 68Ga-PSMA PET/CT [955].

6.3.4.2.2 Local recurrence after radiation therapy


In patients with BCR after RT biopsy status is a major predictor of outcome, provided the biopsies are obtained
18–24 months after initial treatment. Given the morbidity of local salvage options it is necessary to obtain
histological proof of the local recurrence before treating the patient [923].
Transrectal US is not reliable in identifying local recurrence after RT. In contrast, MRI has yielded
excellent results and can be used for biopsy targeting and guiding local salvage treatment [923, 956-959],
even if it slightly underestimates the volume of the local recurrence [960]. Detection of recurrent cancer is
also feasible with choline PET/CT [961], but choline PET/CT has not yet been compared to mpMRI. Prostate-
specific membrane antigen PET/CT can also play a role in the detection of local recurrences after RT [397].

6.3.4.3 Summary of evidence on imaging in case of biochemical recurrence


In patients with BCR imaging can detect both local recurences and distant metastases, however, the sensitivity
of detection depends on the PSA level. After RP, PSMA PET/CT seems to be the imaging modality with the
highest sensitivity at low PSA levels (< 0.5 ng/mL) and may help distinguishing patients with recurrences
confined to the prostatic fossa from those with distant metastases which may impact the design and use of
post-RP salvage RT. After RT, MRI has shown excellent results at detecting local recurrences and guiding
prostate biopsy. Given the substantial morbidity of post-RT salvage local treatments, distant metastases
must be ruled out in patients with local recurrences and who are fit for these salvage therapies. Choline-,
fluciclovine- or PSMA-PET/CT can be used to detect metastases in these patients but for this indication PSMA
PET/CT seems the most sensitive technique.

6.3.4.4 Guidelines for imaging in patients with biochemical recurrence

Recommendations Strength rating


Prostate-specific antigen (PSA) recurrence after radical prostatectomy
Perform prostate-specific membrane antigen (PSMA) positron emission tomography (PET) Weak
computed tomography (CT) if the PSA level is > 0.2 ng/mL and if the results will influence
subsequent treatment decisions.
In case PSMA PET/CT is not available, and the PSA level is > 1 ng/mL, perform fluciclovine Weak
PET/CT or choline PET/CT imaging if the results will influence subsequent treatment
decisions.
PSA recurrence after radiotherapy
Perform prostate magnetic resonance imaging to localise abnormal areas and guide Weak
biopsies in patients fit for local salvage therapy.
Perform PSMA PET/CT (if available) or fluciclovine PET/CT or choline PET/CT in patients fit Strong
for curative salvage treatment.

6.3.5 Treatment of PSA-only recurrences


The timing and treatment modality for PSA-only recurrences after RP or RT remain a matter of controversy
based on the limited evidence.

6.3.5.1 Treatment of PSA-only recurrences after radical prostatectomy


6.3.5.1.1 Salvage radiotherapy for PSA-only recurrence after radical prostatectomy (cTxcN0M0, without PET/ CT)
Early SRT provides the possibility of cure for patients with an increasing PSA after RP. Boorjian et al., reported
a 75% reduced risk of systemic progression with SRT when comparing 856 SRT patients with 1,801 non-SRT
patients [962]. The RAVES and RADICAL trials assessing SRT in post-RP patients with PSA levels exceeding
0.1–0.2 ng/mL showed 5-year freedom from BCR and BCR-free survival rates of 88% [870, 963].
The PSA level at BCR was shown to be prognostic [962]. More than 60% of patients who
are treated before the PSA level rises to > 0.5 ng/mL will achieve an undetectable PSA level [964-
967], corresponding to a ~80% chance of being progression-free 5 years later [968]. A retrospective
analysis of 635 patients who were followed after RP and experienced BCR and/or local recurrence and
either received no salvage treatment (n = 397) or salvage RT alone (n = 160) within 2 years of BCR

PROSTATE CANCER - LIMITED UPDATE 2021 91


showed that salvage RT was associated with a 3-fold increase in PCa-specific survival relative to those
who received no salvage treatment (p < 0.001). Salvage RT has been shown to be effective mainly
in patients with a short PSA-DT [969]. According to a Panel systematic review which also proposed
a scoring system [916], men with a PSA-DT > 1 year and a pathological GS < 8 have a significantly
lower risk of clinical progression and can be classified as ‘EAU Low Risk’; men with a PSA-DT < 1 year
or a pathological pGS 8–10 are at increased risk of clinical progression and should be classified as
‘EAU High Risk’. These definitions have been externally validated and may be helpful for individualised
treatment decisions [921]. Despite the indication for salvage RT, a ‘wait and see‘ strategy remains an
option in patients with a PSA-DT of more than 12 months and other favourable factors such a time to BCR
> 3 years, < pT3a, ISUP grade < 2/3 [916, 970]. For an overview see Table 6.3.2.

Although biochemical progression is now widely accepted as a surrogate marker of PCa recurrence; metastatic
disease, disease-specific and OS are more meaningful endpoints to support clinical decision-making. A
systematic review and meta-analysis on the impact of BCR after RP reports SRT to be favourable for OS and
PCa-specific mortality. In particular SRT should be initiated in patients with rapid PSA kinetics after RP and
with a PSA cut-off of 0.4 ng/mL [916]. A recent international multi-institutional analysis of pooled data from
RCTs has suggested that metastasis-free survival is the most valid surrogate endpoint with respect to impact
on OS [971, 972]. Table 6.3.3 summarises results of recent studies on clinical endpoints after SRT.

Table 6.3.2: S
 elected studies of post-prostatectomy salvage radiotherapy, stratified by pre-salvage
radiotherapy PSA level* (cTxcN0M0, without PET/CT)

Reference n Median FU pre-SRT RT dose ADT bNED/PFS 5-yr. results


(mo) PSA (ng/mL) (year)
median
Bartkowiak, 464 71 0.31 66.6 Gy 54% (5.9) 73% vs. 56%; PSA
et al. 2018 < 0.2 vs. > 0.2 ng/mL
[973] p < 0.0001
Soto, et al. 441 36 < 1 (58%) 68 Gy 63/55% (3) 44/40% ADT/no ADT
2012 [974] 24% ADT ADT/no ADT p < 0.16
Stish, et al. 1,106 107 0.6 68 Gy 50% (5) 44% vs. 58%; PSA
2016 [964] 16% ADT 36% (10) < 0.5 vs. > 0.5 ng/mL
p < 0.001
Tendulkar, 2,460 60 0.5 66 Gy 56% (5) SRT; PSA < 0.2 ng/mL
et al. 2016 16% ADT 71% 0.21-0.5 ng/mL
[975] 63% 0.51-1.0 ng/mL
54% 1.01-2.0 ng/mL
43% > 2 ng/mL 37%
p < 0.001
*Androgen deprivation therapy can influence the outcome ‘biochemically no evidence of disease (bNED)’ or
‘progression-free survival’. To facilitate comparisons, 5-year bNED/PFS read-outs from Kaplan-Meier plots are
included.
ADT = androgen deprivation therapy; bNED = biochemically no evidence of disease; FU = follow up;
mo = months; n = number of patients; PFS = progression-free survival; PSA = prostate-specific antigen;
SRT = salvage radiotherapy; yr = year.

Table 6.3.3: R
 ecent studies reporting clinical endpoints after SRT (cTxcN0M0, without PET/CT)
(the majority of included patients did not receive ADT)

Reference n Median FU Regimen Outcome


(mo)
Bartkowiak, 464 71 66.6 (59.4-72) Gy 5.9 yr. OS
et al. 2018 no ADT post-SRT PSA < 0.1 ng/mL 98%
[973] post-SRT PSA > 0.1 ng/mL 92%
p = 0.005

92 PROSTATE CANCER - LIMITED UPDATE 2021


Jackson, 448 64 68.4 Gy no ADT 5 yr. DM
et al. 2014 post-SRT PSA < 0.1 ng/mL 5%
[976] post-SRT PSA > 0.1 ng/mL 29%
p < 0.0001
5 yr. DSM
post-SRT PSA < 0.1 ng/mL 2%
post-SRT PSA > 0.1 ng/mL 7%
p < 0.0001
OS
post-SRT PSA < 0.1 ng/mL 97%
post-SRT PSA > 0.1 ng/mL 90%
p < 0.0001
Stish, 1,106 107 68 (64.8-70.2) Gy 5 and 8.9 yr. DM
et al. 2016 39% 2D treatment SRT: PSA < 0.5 ng/mL 7% and 12%
[964] planning SRT: PSA > 0.5 ng/mL 14% and 23%
incl. 16% ADT p < 0.001
5 and 8.9 yr. DSM
SRT: PSA < 0.5 ng/mL < 1% and 6%
SRT: PSA > 0.5 ng/mL 5% and 10%
p = 0.02 5 and 8.9 yr. OS
SRT: PSA < 0.5 ng/mL 94% and 86%
SRT: PSA > 0.5 ng/mL 91% and 78%
p = 0.14
Tendulkar, 2,460 60 66 (64.8-68.4) Gy 10 yr. DM
et al. 2016 incl. 16% ADT SRT: PSA 0.01-0.2 ng/mL 9%
[975] SRT: PSA 0.21-0.50 ng/mL 15%
SRT: PSA 0.51-1.0 ng/mL 19%
SRT: PSA 1.01-2.0 ng/mL 20%
SRT: PSA > 2 ng/mL 37%, p < 0.001
ADT = androgen deprivation therapy; DM = distant metastasis; DSM = disease specific mortality;
FU = follow up; mo. = month; n = number of patients; OS = overall survival; PSA = prostate specific antigen;
SRT = salvage radiotherapy.

6.3.5.1.2 Salvage radiotherapy combined with androgen deprivation therapy (cTxcN0, without PET/CT)
Data from RTOG 9601 [977] suggest both CSS and OS benefit when adding 2 years of bicalutamide (150 mg o.d.)
to SRT. According to GETUG-AFU 16 also 6-months treatment with a LHRH-analogue can significantly improve
10-year BCR, biochemical PFS and, modestly, metastasis-free survival. However, SRT combined with either
goserelin or placebo showed similar DSS and OS rates [978]. Table 6.3.4 provides an overview of these two RCTs.

These RCTs support adding ADT to SRT. However when interpreting these data it has to be kept in mind that
RTOG 9601 used outdated radiation dosages (< 66 Gy) and technique. The question with respect to the patient
risk profile, whether to offer combination treatment or not, and the optimal combination (LHRH or bicalutamide)
remains, as yet, unsolved. The EAU risk classification may offer guidance in this respect [916, 921].

One of these RCTs reports improved OS (RTOG 96-01) and the other improved metastasis-free survival but
due to methodological discrepancies also related to follow-up and risk patterns, it is, as yet, not evident which
patients should receive ADT, which type of ADT and for how long. Men at high risk of further progression
(e.g., with a PSA > 0.7 ng/mL and GS > 8) may benefit from SRT combined with two years of ADT; for those at
lower risk (e.g., PSA < 0.7 ng/mL and GS 8) SRT combined with 6 months of ADT may be sufficient. Men with
a low-risk profile (PSA < 0.5 ng/mL and GS < 8) may receive SRT alone. In a sub-analysis of men with a PSA
of 0.61 to 1.5 (n = 253) there was an OS benefit associated with anti-androgen assignment (HR: 0.61, 95%
CI: 0.39–0.94). In those receiving early SRT (PSA 0.6 ng/mL, n = 389), there was no improvement in OS
(HR: 1.16, 95% CI: 0.79–1.70), an increased other-cause mortablity hazard (subdistribution HR: 1.94, 95%
CI: 1.17–3.20, p = 0.01) and increased odds of late grades 3–5 cardiac and neurologic toxic effects (OR: 3.57,
95% CI: 1.09–15.97, p = 0.05). These results suggest that pre-SRT PSA level may be a prognostic biomarker
for outcomes of anti-androgen treatment with SRT. In patients receiving late SRT (PSA > 0.6 ng/mL, hormone
therapy was associated with improved outcomes. In men receiving early SRT (PSA < 0.6 ng/mL), long-term
anti-androgen treatment was not associated with improved OS [979].
A recent review addressing the benefit from combining HT with SRT suggested risk stratification of
patients based on the pre-SRT PSA (< 0.5, 0.6–1, > 1 ng/mL), margin status and ISUP grade as a framework

PROSTATE CANCER - LIMITED UPDATE 2021 93


to individualise treatment [980]. In a retrospective multi-centre-study including 525 patients, only in patients
with more aggressive disease characteristics (pT3b/4 and ISUP grade > 4 or pT3b/4 and PSA at early SRT
> 0.4 ng/mL) the administration of concomitant ADT was associated with a reduction in distant metastasis
[981]. Similarly, in a retrospective analysis of 1,125 patients, stage > pT3b, GS > 8 and a PSA level at SRT
> 5 ng/mL were identified as risk factors for clinical recurrence. A significant effect of long-term ADT was
observed in patients with > 2 adverse features. For patients with a single risk factor, short-term HT was
sufficient whilst patients without risk factors showed no significant benefit from concomitant ADT [982].

Table 6.3.4: R
 andomised controlled trials comparing salvage radiotherapy combined with androgen
deprivation therapy vs. salvage radiotherapy alone

Reference n Risk groups Median Regimen Outcome


FU (mo)
GETUG-AFU 16 369 RT + ISUP grade 112 66 Gy + 6 mo. GnRH 10 yr. PFS 64%
2019 [978] ADT 374 RT < 2/3 89%, analogue p < 0.0001
6 mo. 66 Gy 10 yr. PFS 49%
ISUP grade 10 yr. MFS 75%
> 4 11% p = 0.034
cN0 10 yr. MFS 69%
RTOG 9601 384 RT + pT2 R1, pT3 156 64.8 Gy + bicalutamide 12 yr. DM 14%
2017 [977] ADT 376 RT cN0 24 mo. 64.8 Gy + p = 0.005
placebo 12 yr. DM 23%
12 yr. OS 76%
p = 0.04
12 yr. OS 71%
12 yr. DSM 5.8%
p < 0.001
12 yr. DSM 13.4%
ADT = androgen deprivation therapy; DM = distant metastasis; DSM = disease specific mortality; PFS = 
progression free survival; FU = follow-up; GnRH = gonadotropin-releasing hormone; MFS = metastasis-free
survival; OS = overall survival; PFS = progression-free survival; mo = months; n = number of patients;
RT = radiotherapy; yr = year.

6.3.5.1.2.1 Target volume, dose, toxicity


There have been various attempts to define common outlines for ‘clinical target volumes‘ of PCa [983-986]
and for organs at risk of normal tissue complications [987]. However, given the variations of techniques and
dose-constraints, a satisfactory consensus has not yet been achieved. A benefit in biochemical PFS but not
metastasis-free survival has been reported in patients receiving whole pelvis SRT (± ADT) but the advantages
must be weighed against possible side effects [988].

The optimal SRT dose has not been well defined. It should be at least 66 Gy to the prostatic fossa (± the base
of the SVs, depending on the pathological stage after RP) [965, 989]. In a systematic review, the pre-SRT
PSA level and SRT dose both correlated with BCR, showing that relapse-free survival decreased by 2.4%
per 0.1 ng/mL PSA and improved by 2.6% per Gy, suggesting that a treatment dose above 70 Gy should be
administered at the lowest possible PSA level [990]. The combination of pT stage, margin status and ISUP
grade and the PSA at SRT seems to define the risk of biochemical progression, metastasis and overall mortality
[867, 991, 992]. In a study on 894 node-negative PCa patients, doses ranging from 64 to > 74 Gy were
assigned to twelve risk groups defined by their pre-SRT PSA classes < 0.1, 0.1–0.2, 0.2–0.4, and > 0.4 ng/mL
and ISUP grade, < 1 vs. 2/3 vs. > 4 [993]. The updated Stephenson nomograms incorporate the SRT and ADT
doses as predictive factors for biochemical failure and distant metastasis [975].

Salvage RT is also associated with toxicity. In one report on 464 SRT patients receiving median 66.6
(max. 72) Gy, acute grade 2 toxicity was recorded in 4.7% for both the GI and GU tract. Two men had late
grade 3 reactions of the GI tract. Severe GU tract toxicity was not observed. Late grade 2 complications
occurred in 4.7% (GI tract) and 4.1% (GU tract), respectively, and 4.5% of the patients developed moderate
urethral stricture [973].
In a RCT on dose escalation for SRT involving 350 patients, acute grade 2 and 3 GU toxicity was
observed in 13.0% and 0.6%, respectively, with 64 Gy and in 16.6% and 1.7%, respectively, with 70 Gy.

94 PROSTATE CANCER - LIMITED UPDATE 2021


Gastrointestinal tract toxicity of grades 2 and 3 occurred in 16.0% and 0.6%, respectively, with 64 Gy, and in
15.4% and 2.3%, respectively, with 70 Gy. Late effects have yet to be reported [994, 995].

With dose escalation over 72 Gy and/or up to a median of 76 Gy, the rate of severe side effects, especially
genitourinary symptoms, clearly increases, even with newer planning and treatment techniques [996, 997]. In
particular, when compared with 3D-CRT, IMRT was associated with a reduction in grade 2 GI toxicity from 10.2
to 1.9% (p = 0.02) but had no differential effect on the relatively high level of GU toxicity (5-year, 3D-CRT 15.8%
vs. IMRT 16.8%) [996]. After a median salvage IMRT dose of 76 Gy, the 5-year risk of grade 2–3 toxicity rose
to 22% for GU and 8% for GI symptoms, respectively [997]. Doses of at least 66 Gy and up to 72 Gy can be
recommended [973, 994].

6.3.5.1.2.2 Salvage RT with or without ADT (cTx CN0/1) (with PET/CT)


In a prospective multi-centre study of 323 patients with BCR, PSMA PET/CT changed the management intent
in 62% of patients as compared to conventional staging. This was due to a significant reduction in the number
of men in whom the site of disease recurrence was unknown (77% vs. 19%, p < 0.001) and a significant
increase in the number of men with metastatic disease (11% vs. 57%) [415]. A recent prospective study in
a subgroup of 119 BCR patients with low PSA (< 0.5 ng/mL) reported a change in the intended treatment in
30.2% of patients [895]; however, no data exist on the impact on final outcome. Another prospective study in
272 patients with early biochemical recurrent PCa after RP showed that 68Ga-PSMA-ligand PET/CT may tailor
further therapy decisions (e.g., local vs. systemic treatment) at low PSA values (0.2–1 ng/mL) [897].
A single-centre study retrospectively assessed 164 men from a prospective database who underwent
imaging with PSMA PET/CT for a rising PSA after RP with PSA levels < 0.5 ng/mL. In men with a negative PSMA
PET/CT who received salvage RT, 85% (23 out of 27) demonstrated a treatment response compared to a further
PSA increase in 65% of those not treated (22 out of 34). In the 36/99 men with disease confined to the prostate
fossa on PSMA, 83% (29 out of 36) responded to salvage RT [945]. Thus, PSMA PET/CT might stratify men
into a group with high response (negative findings or recurrence confined to the prostate) and poor response
(positive nodes or distant disease) to salvage RT. As there are no prospective phase III data (in particular not for
PCa-specific survival or OS) these results have to be confirmed before a recommendation can be provided.

6.3.5.1.2.3 Metastasis-directed therapy for rcN+ (with PET/CT)


Radiolabelled PSMA PET/CT is increasingly used as a diagnostic tool to assess metastatic disease burden
in patients with BCR following prior definitive therapy. A review including 30 studies and 4,476 patients
showed overall estimates of positivity in a restaging setting of 38% in pelvic LNs and 13% in extra-pelvic
LN metastases [397]. The percentage positivity of PSMA PET/CT was proven to increase with higher PSA
values, from 33% (95% CI: 16–51) for a PSA of < 0.2 ng/mL, to 45% (39–52), 59% (50–68), 75% (66–84), and
95% (92–97) for PSA subgroup values of 0.2–0.49, 0.5–0.99, 1.00–1.99, and > 2.00 ng/mL, respectively [397].
Results of this review demonstrated high sensitivity and specificity of 68Ga-PSMA PET in advanced PCa with a
per-lesion-analised sensitivity and specificity of 75% and 99%, respectively.

In patients relapsing after a local treatment (including cN+ and higly selected M1 patients), a metastases-
targeting therapy has been proposed, with the aim to delay systemic treatment. Metastasis-directed (MDT)
therapy in PET/CT detected nodal oligo-recurrent PCa after RP was assessed in a large retrospective multi-
institutional study (263 patients received MDT and 1,816 patients standard of care as control group [matched
3:1]). Metastasis-targeting therapy consisted of salvage LN resection (n = 166) and stereotactic ablation RT
(SABR) (n = 97). After a median follow-up of 70 months, the MDT-group showed significantly better CSS (5-year
survival 98.6% vs. 95.7%, p < 0.01, respectively) [998].

Another retrospective study compared SABR with elective nodal irradiation (ENRT) in PET/CT-detected nodal
oligo-recurrent PCa (n = 506 patients, 365 of which with N1 pelvic recurrence). With a median follow-up of 36
months, ENRT (n = 197) was associated with a significant reduction of nodal recurrences compared with SABR
(n = 309) of 2% vs. 18%, respectively, but at the cost of higher side effects of ENRT [999]. These results have
to be confirmed in prospective trials before any recommendations can be made. In these situations SABR
should be used in highly selected patients only. For MDT in M1-patients see Section 6.4.8.

6.3.5.1.3 Salvage lymph node dissection


The surgical management of (recurrent) nodal metastases in the pelvis has been the topic of several
retrospective analyses [1000-1002] and a systematic review [1003]. The reported 5-year BCR-free survival rates
ranged from 6% to 31%. Five-year OS was approximately 84% [1003]. Biochemical recurrence rates were
found to be dependent on PSA at surgery and location and number of positive nodes [1004]. Addition of RT to
the lymphatic template after salvage LN dissection may improve the BCR rate [1005]. In a recent multi-centre

PROSTATE CANCER - LIMITED UPDATE 2021 95


retrospective study long-term outcomes of salvage LN dissection were reported to be worse than previously
described in studies with shorter follow-up [1006]. Biochemical recurrence-free survival at 10 years was 11%.
Patients with a PSA response after salvage LN dissection and patients receiving ADT within 6 months from
salvage LN dissection had a lower risk of death from PCa [1006]. High level evidence for the oncological value
of salvage LN dissection is still lacking [1003].

6.3.5.1.4 Comparison of adjuvant- and salvage radiotherapy


Section 6.2.5.4 - ART, is referred to for more details. Main findings are that after RP the vast majority of patients
do not need ART which is supported by the results of 3 phase III RCTs comparing adjuvant RT and early
salvage RT with a median follow-up of 5 years [870-874].
However, longer term (10-year) results and results of metastasis-free survival endpoints are needed
before final conclusions can be drawn. Due to the small number of patients with a combination of high-risk
features included in these 3 trails (only approximately 20%) ART remains a recommended treatment option in
highly selected patients with at least two out of three high-risk features, such as pT3/R1/ISUP > 3 [854, 874].

6.3.5.2 Management of PSA failures after radiation therapy


Therapeutic options in these patients are ADT or salvage local procedures. A recent systematic review and
meta-analysis included all studies comparing the efficacy and toxicity of salvage RP, salvage HIFU, salvage
cryotherapy, SBRT, salvage LDR brachytherapy, and salvage HDR brachytherapy in the management of locally
recurrent PCa after primary radical EBRT [1007]. The outcomes were BCR-free survival at 2 and 5 years.
No significant differences with regards to recurrence-free survival between these modalities
was found [1007]. Five-year recurrence-free survival ranged from 50% after cryotherapy to 60% after HDR
brachytherapy and SBRT [1007]. The authors reported that severe GU toxicity exceeded 21% for HIFU and
RP, whereas it ranged from 4.2% to 8.1% with re-irradiation. Differences in severe GI toxicity also appeared to
favour re-irradiation, particularly HDR brachytherapy [1007]. Due to the methodological limitations of this review
(the majority of the included studies were uncontrolled single-arm case series and there was considerable
heterogeneity in the definitions of core outcomes) the available evidence for these treatment options is of low
quality and strong recommendations regarding the choice of any of these techniques cannot be made. The
following is an overview of the most important findings for each of these techniques.

6.3.5.2.1 Salvage radical prostatectomy


Salvage RP after RT is associated with a higher likelihood of adverse events compared to primary surgery
because of the risk of fibrosis and poor wound healing due to radiation [1008].

6.3.5.2.1.1 Oncological outcomes


In a systematic review of the literature, Chade, et al., showed that SRP provided 5- and 10-year BCR-free
survival estimates ranging from 47–82% and from 28-53%, respectively. The 10-year CSS and OS rates ranged
from 70–83% and from 54–89%, respectively. The pre-SRP PSA value and prostate biopsy ISUP grade were
the strongest predictors of the presence of organ-confined disease, progression, and CSS [1009]. In a recent
multi-centre analysis including 414 patients, 5-year BCR-free survival, CSS and OS were 56.7%, 97.7% and
92.1%, respectively [1010]. Pathological T stage > T3b (OR: 2.348) and GS (up to OR 7.183 for GS > 8) were
independent predictors for BCR.

Table 6.3.5: O
 ncological results of selected salvage radical prostatectomy case series

Reference n Median Pathologic PSM Lymph-node BCR-free CSS Time


FU Organ- (%) involvement probability (%) probability
(mo) confined (%) (%) (%)
Chade, et al. 2011 404 55 55 25 16 37 83 10 yr.
[1011]
Mandel, et al. 55 36 50 27 22 49 89 5 yr.
2016 [1012]
Ogaya-Pinies, 96 14 50 17 8 85* - 14 mo.
et al. 2018 [1013]
Marra, et al. 2021 414 36 46 30 16 57 98 5 yr.
[1010]
*Percentage of patients without BCR.
BCR = biochemical recurrence; CSS = cancer-specific survival; FU = follow-up; mo = months; n = number of
patients; PSM = positive surgical margin.

96 PROSTATE CANCER - LIMITED UPDATE 2021


6.3.5.2.1.2 Morbidity
Compared to primary open RP, SRP is associated with a higher risk of later anastomotic stricture (47 vs. 5.8%),
urinary retention (25.3% vs. 3.5%), urinary fistula (4.1% vs. 0.06%), abscess (3.2% vs. 0.7%) and rectal injury
(9.2 vs. 0.6%) [1014]. In more recent series, these complications appear to be less common [1008, 1009, 1012].
Functional outcomes are also worse compared to primary surgery, with urinary incontinence ranging
from 21% to 90% and ED in nearly all patients [1009, 1012].

Table 6.3.6: P
 eri-operative morbidity in selected salvage radical prostatectomy case series

Reference n Rectal injury (%) Anastomotic Clavien 3-5 (%) Blood loss, mL,
stricture (%) mean, range
Ward, et al. 2005 [1015] 138 5 22 - -
Sanderson, et al. 2006 [1016] 51 2 41 6 -
Gotto, et al. 2010 [1014] 98 9 41 25 -
Gontero, et al. 2019 [1008] 395 1.6 25 3.6 10.1
n = number of patients.

6.3.5.2.1.3 Summary of salvage radical prostatectomy


In general, SRP should be considered only in patients with low co-morbidity, a life expectancy of at least 10
years, a pre-SRP PSA < 10 ng/mL and initial biopsy ISUP grade < 2/3, no LN involvement or evidence of
distant metastatic disease pre-SRP, and those whose initial clinical staging was T1 or T2 [1009].
A meta-analysis and systematic review of local salvage therapies after RT for PCa has suggested
that re-irradiation with SBRT, HDR brachytherapy or LDR brachytherapy appears to result in less severe GU
toxicity than RP, and re-irradiation with HDR brachytherapy in less severe GI toxicity than RP [1007].
Salvage RP might be associated with more severe GU and GI toxicity compared to re-irradiation [1007].

6.3.5.2.2 Salvage cryoablation of the prostate


6.3.5.2.2.1 Oncological outcomes
Salvage cryoablation of the prostate (SCAP) has been proposed as an alternative to salvage RP, as it has a
potentially lower risk of morbidity and equal efficacy.
In a recent systematic review a total of 32 studies assessed SCAP, recruiting a total of 5,513
patients. The overwhelming majority of patients (93%) received whole-gland SCAP. The adjusted pooled
analysis for 2-year BCR-free survival for SCAP was 67.49% (95% CI: 61.68–72.81%), and for 5-year BCR-free
survival was 50.25% (95% CI: 44.10–56.40%). However, the certainty of the evidence was low. Table 6.3.7
summarises the results of a selection of the largest series on SCAP to date in relation to oncological outcomes
(BCR only) [1007].

Table 6.3.7: O
 ncological results of selected salvage cryoablation of the prostate case series, including at
least 250 patients

Reference n Median Time point BCR-free probability Definition of failure


FU (mo) of outcome
measurement
(yr)
Ginsburg, et al. 2017 [1017] 898 19.0 5 71.3% Phoenix criteria
Spiess, et al. 2010 [1018] 450 40.8 3.4 39.6% PSA > 0.5 ng/mL
Li, et al. 2015 [1019] 486 18.2 5 63.8% Phoenix criteria
Kovac, et al. 2016 [1020] 486 18.2 5 75.5% Phoenix criteria
(nadir PSA < 0.4 ng/mL);
22.1%
(nadir PSA > 0.4 ng/mL)
Ahmad, et al. 2013 [1021] 283 23.9 3 67.0% Phoenix criteria
(nadir PSA < 1 ng/mL);
14.0%
(nadir PSA > 1 ng/mL)
Pisters, et al. 2008 [1022] 279 21.6 5 58.9% (ASTRO); ASTRO and Phoenix
54.5% (Phoenix) criteria
ASTRO = American Society for Therapeutic Radiology and Oncology; BCR = biochemical recurrence;
FU = follow-up; mo. = months; n = number of patients; PSA = prostate-specific antigen; yr. = year.

PROSTATE CANCER - LIMITED UPDATE 2021 97


6.3.5.2.2.2 Morbidity
The main adverse effects and complications relating to SCAP include urinary incontinence, urinary retention
due to bladder outflow obstruction, recto-urethral fistula, and erectile dysfunction. The recent SR and
meta-analysis [1007] showed an adjusted pooled analysis for severe GU toxicity for SCAP of 15.44% (95%
CI: 10.15–21.54%). As before, the certainty of the evidence was low. Table 6.3.8 summarises the results of a
selection of the largest series on SCAP to date in relation to GU outcomes.

Table 6.3.8: P
 eri-operative morbidity, erectile function and urinary incontinence in selected salvage
cryoablation case series, including at least 100 patients

Study n Time point Incontinence Obstruction/ Rectourethral ED (%)


of outcome (%) Retention (%) fistula (%)
measurement
(mo)
Li, et al. 2015 [1019] 486 12 33.3 21.7 4.7 71.3
Ahmad, et al. 2013 [1021] 283 12 12.0 8.1 1.8 83.0
Pisters, et al. 2008 [1022] 279 12 4.4 3.2 1.2 NA
Cespedes, et al. 1997 [1023] 143 Median 27.0 28.0 14.0 NA NA
Chin, et al. 2001 [1024] 118 Median 18.6 6.7 NA 3.3 NA
ED = erectile dysfunction; mo = months; n = number of patients.

6.3.5.2.2.3 Summary of salvage cryoablation of the prostate


In general, the evidence base relating to the use of SCAP is poor, with significant uncertainties relating to long-
term oncological outcomes, and SCAP appears to be associated with significant morbidity. Consequently,
SCAP should only be performed in selected patients in experienced centres as part of a clinical trial or well-
designed prospective cohort study.

6.3.5.2.3 Salvage re-irradiation


6.3.5.2.3.1 Salvage brachytherapy for radiotherapy failure
Carefully selected patients with a good PS, primary localised PCa, good urinary function and histologically
proven local recurrence are candidates for salvage brachytherapy using either HDR- or LDR.
In a recent systematic review a total of 16 studies (4 prospective) and 32 studies (2 prospective)
assessed salvage HDR and LDR brachytherapy, respectively [1007] with the majority (> 85%) receiving whole-
gland brachytherapy rather than focal treatment. The adjusted pooled analysis for 2-year BCR-free survival for
HDR was 77% (95% CI: 70–83%) and for LDR was 81% (95% CI:74–86%). The 5-year BCR-free survival for
HDR was 60% (95% CI: 52–67%) and for LDR was 56% (95% CI: 48–63%). As noted above, brachytherapy
techniques are associated with lower rates of severe GU toxicity when compared to RP or HIFU, at 8% for
HDR (95% CI: 5.1–11%) and 8.1% for LDR (95% CI: 4.3–13%). Rates of severe GI toxicity are reported to
be very low at 0% for HDR (95% CI: 0–0.2%) and 1.5% for LDR (95% CI: 0.2–3.4%). High-dose-rate or LDR
brachytherapy are effective treatment options with an acceptable toxicity profile. However, the published series
are small and likely under-report toxicity. Consequently this treatment should be offered in experienced centres
ideally within randomised clinical trials or prospective registry studies.

Table 6.3.9: T
 reatment-related toxicity and BCR-free probablity in selected salvage brachytherapy
studies including at least 100 patients.

Author Study design n and BT type Median FU Treatment toxicity BCR-free probability
(mo)
Lopez, et al. multi-centre 75 HDR 52 23.5% late G3+ GU 5 yr 71%
2019 [1025] retrospective 44 LDR (95% CI: 65.9-75.9%)
Crook, et al. multi-centre 100 LDR 54 14% late G3 n.r.
2019 [1026] prospective combined GI/GU
Smith, et al. single-centre 108 LDR 76 15.7%/2.8% late G3 5 yr. 63.1%
2020 [1027] retrospective GU/GI 10 yr. 52%
Lyczek, et al. single-centre 115 HDR n.r. 12.2%/0.9% 60% at 40 mo.
2009 [1028] retrospective late G3+ GU/GI
BT = brachytherapy; GI = gastrointestinal; GU = genito-urinary; HDR = high-dose rate; LDR = low-dose rate;
mo = months; n = number of patients; n.r. = not reported; yr = year.

98 PROSTATE CANCER - LIMITED UPDATE 2021


6.3.5.2.3.2 Salvage stereotactic ablative body radiotherapy for radiotherapy failure
6.3.5.2.3.2.1 Oncological outcomes and morbidity
Stereotactic ablative body radiotherapy (CyberKnife or Linac-based treatment) is a potentially viable new
option to treat local recurrence after RT. Carefully selected patients with good IPSS-score, without obstruction,
good PS and histologically proven localised local recurrence are potential candidates for SABR. In a recent
meta-analysis and systematic review five mostly retrospective studies including 206 patients were treated with
CyberKnife or linac-based-treatment showing 2-year RFS estimates (61.6%, 95% CI: 52.6–69.9%) [1007]. In
a retrospective multi-centre study (n = 100) the median pre-salvage PSA was 4.3 ng/mL with 34% of patients
having received ADT for twelve months (median). All recurrences were biopsy proven. Patients were treated
with the CyberKnife with a single dose of 6 Gy in six daily fractions (total dose 36 Gy). With a median follow-up
of 30 months the estimated 3-year second BCR-free survival was 55% [1029].
In a smaller retrospective series including 50 men with histologically proven local recurrence with
a median pre-salvage PSA of 3.9 ng/mL only 15% had received additional ADT. The estimated 5-year second
BCR-free survival was 60% (median follow-up of 44 months) which is an outcome comparable to series
treating patients with RP, HIFU or brachytherapy [1030]. Table 6.3.10 summarises the results of the two larger
SABR series addressing oncological outcomes and morbidity.

Table 6.3.10: T
 reatment-related toxicity and BCR-free survival in selected SABR studies including at least
50 patients

Author Study design n and Median Fractionation ADT Treatment BCR-free


RT-type FU (mo) (SD/TD) toxicity survival
Fuller, et al. single-centre 50 44 SD 6.8 Gy 7/50 5 yr: 8% late 5 yr. 60%
2020 [1030] retrospective Cyber Knife TD 34 Gy G3+ GU
Pasquier, multi-centre 100 30 SD 6 Gy 34/100 3 yr. grade 2+ 3 yr. 55%
et al. 2020 retrospective Cyber Knife TD 36 Gy median GU 20.8%
[1029] 12 mo. GI 1%
BCR = biochemical recurrence-free; FU = follow-up; mo = months; n = number of patients; RT-type = type of
radiotherapy; SD = single dose; TD = total dose; yr = year.

6.3.5.2.3.2.2 Morbidity
In a retrospective single-centre study with 50 consecutive patients chronic significant toxicity was only seen
for the GU domain with 5-year grade 2+ and grade 3+ GU rates of 17% and 8%, respectively. No GI toxicity
> grade 1 was seen. Of note, of the fifteen patients who were sexually potent pre-salvage SBRT, twelve
subsequently lost potency [1030]. In a retrospective French (GETUG) multi-centre series (n = 100) the 3-year
late grade 2+ GU and GI toxicity was 20.8% (95% CI: 13–29%) and 1% (95% CI: 0.1–5.1%), respectively
[1029].

6.3.5.2.3.2.3 Summary of salvage stereotactic ablative body radiotherapy


Despite the encouraging results so far the number of patients treated with SABR is relatively limited. In view of
the rates of higher grade 2+ GU side effects, SABR should only be offered to selected patients, in experienced
centres as part of a clinical trial or well-designed prospective study.

6.3.5.2.4 Salvage high-intensity focused ultrasound


6.3.5.2.4.1 Oncological outcomes
Salvage HIFU has emerged as an alternative thermal ablation option for radiation-recurrent PCa. Being
relatively newer than SCAP the data for salvage HIFU are even more limited. A recent systematic review and
meta-analysis also included salvage HIFU [1007]. A total of 20 studies assessed salvage HIFU, recruiting
1,783 patients. The overwhelming majority of patients (86%) received whole-gland salvage HIFU. The adjusted
pooled analysis for 2-year BCR-free survival for salvage HIFU was 54.14% (95% CI: 47.77–60.38%) and for
5-year BCR-free survival 52.72% (95% CI: 42.66–62.56%). However, the certainty of the evidence was low.
Table 6.3.11 summarises the results of a selection of the largest series on salvage HIFU to date in relation to
oncological outcomes (BCR only).

PROSTATE CANCER - LIMITED UPDATE 2021 99


Table 6.3.11: O
 ncological results of selected salvage cryoablation of the prostate case series, including
at least 250 patients

Study n Median Time point BCR-free Definition of failure


FU (mo) of outcome probability
measurement
(yr)
Crouzet, et al. 418 39.6 5 49.0% Phoenix criteria
2017 [1031]
Murat, et al. 167 Mean 3 25.0% (high-risk); Phoenix criteria or positive
2009 [1032] 18.1 53.0% (low-risk)* biopsy or initiation of post-HIFU
salvage therapy
Kanthabalan, et al. 150 35.0 3 48.0% Phoenix criteria
2017 [1033]
Jones, et al. 100 12.0 1 50.0% Nadir PSA > 0.5 ng/mL or
2018 [1034] positive biopsy
*Results stratified by pre-EBRT D’Amico risk groups
BCR = biochemical recurrence; CSS = cancer-specific survival; FU = follow-up; mo = months; n = number of
patients; yr = year.

6.3.5.2.4.2 Morbidity
The main adverse effects and complications relating to salvage HIFU include urinary incontinence, urinary
retention due to bladder outflow obstruction, rectourethral fistula and erectile dysfunction. The recent
systematic review and meta-analysis showed an adjusted pooled analysis for severe GU toxicity for salvage
HIFU of 22.66% (95% CI: 16.98–28.85%) [1007]. The certainty of the evidence was low. Table 6.3.12
summarises the results of a selection of the largest series on salvage HIFU to date in relation to GU outcomes.

Table 6.3.12: P
 eri-operative morbidity, erectile function and urinary incontinence in selected salvage
HIFU case series, including at least 100 patientss

Study n Time point Incontinence* Obstruction/ Rectourethral ED (%)


of outcome (%) retention (%) fistula (%)
measurement
(yr)
Crouzet, et al. 418 Median 39.6 42.3 18.0 2.3 n.r.
2017 [1031]
Murat, et al. 167 Median 18.1 49.5 7.8 3.0 n.r.
2009 [1032]
Kanthabalan, et al. 150 24 12.5 8.0 2.0 41.7
2017 [1033]
Jones, et al. 100 12 42.0 49.0 5.0 74.0
2018 [1034]
*Incontinence was heterogeneously defined; figures represent at least 1 pad usage.
ED = erectile dysfunction; mo = months; n.r. = not reported; n = number of patients.

6.3.5.2.4.3 Summary of salvage high-intensity focused ultrasound


There is a lack of high-certainty data which prohibits any recommendations regarding the indications for
salvage HIFU in routine clinical practice. There is also a risk of significant morbidity associated with its use in
the salvage setting. Consequently, salvage HIFU should only be performed in selected patients in experienced
centres as part of a clinical trial or well-designed prospective cohort study.

6.3.6 Hormonal therapy for relapsing patients


The Panel conducted a systematic review including studies published from 2000 onwards [1035]. Conflicting
results were found on the clinical effectiveness of HT after previous curative therapy of the primary tumour.
Some studies reported a favourable effect of HT, including the only RCT addressing the research question
of this review (86% vs. 79% advantage in OS in the early HT group) [1036]. Other studies did not find any
differences between early vs. delayed, or no, HT. One study found an unfavourable effect of HT [1037]. This
may be the result of selecting clinically unfavourable cases for (early) HT and more intensive diagnostic work-
up and follow-up in these patients.

100 PROSTATE CANCER - LIMITED UPDATE 2021


The studied population is highly heterogeneous regarding their tumour biology and therefore clinical
course. Predictive factors for poor outcomes were; CRPC, distant metastases, CSS, OS, short PSA-DT, high
ISUP grade, high PSA, increased age and co-morbidities. In some studies, such as the Boorjian, et al. study
[970], high-risk patients, mainly defined by a high ISUP grade and a short PSA-DT (most often less than 6
months) seem to benefit most from (early) HT, especially in men with a long life expectancy.

No data were found on the effectiveness of different types of HT, although it is unlikely that this will have a
significant impact on survival outcomes in this setting. Non-steroidal anti-androgens have been claimed to
be inferior compared to castration, but this difference was not seen in M0 patients [969]. One of the included
RCTs suggested that intermittent HT is not inferior to continuous HT in terms of OS and CSS [1038]. A small
advantage was found in some QoL domains but not overall QoL outcomes. An important limitation of this
RCT is the lack of any stratifying criteria such as PSA-DT or initial risk factors. Based on the lack of definitive
efficacy and the undoubtedly associated significant side effects, patients with recurrence after primary curative
therapy should not receive standard HT since only a minority of them will progress to metastases or PCa-
related death. The objective of HT should be to improve OS, postpone distant metastases, and improve QoL.
Biochemical response to only HT holds no clinical benefit for a patient. For older patients and those with
co-morbidities the side effects of HT may even decrease life expectancy; in particular cardiovascular risk
factors need to be considered [1039, 1040]. Early HT should be reserved for those at the highest risk of disease
progression defined mainly by a short PSA-DT at relapse (< 6–12 months) or a high initial ISUP grade (> 2/3)
and a long life expectancy.

6.3.7 Observation
In unselected relapsing patients the median actuarial time to the development of metastasis will be 8 years and
the median time from metastasis to death will be a further 5 years [855]. For patients with EAU Low-Risk BCR
features (see Section 6.3.3), unfit patients with a life expectancy of less than 10 years or patients unwilling to
undergo salvage treatment, active follow-up may represent a viable option.

6.3.8 Guidelines for second-line therapy after treatment with curative intent

Local salvage treatment Strength rating


Recommendations for biochemical recurrence (BCR) after radical prostatectomy
Offer monitoring, including prostate-specific antigen (PSA), to EAU Low-Risk BCR patients. Weak
Offer early salvage intensity-modulated radiotherapy plus image-guided radiotherapy to Strong
men with two consecutive PSA rises.
A negative PET/CT scan should not delay salvage radiotherapy (SRT), if otherwise indicated. Strong
Do not wait for a PSA threshold before starting treatment. Once the decision for SRT has Strong
been made, SRT (at least 66 Gy) should be given as soon as possible.
Offer hormonal therapy in addition to SRT to men with BCR. Weak
Recommendations for BCR after radiotherapy
Offer monitoring, including PSA to EAU Low-Risk BCR patients. Weak
Only offer salvage radical prostatectomy (RP), brachytherapy, high-intensity focused Strong
ultrasound, or cryosurgical ablation to highly selected patients with biopsy proven
local recurrence within a clinical trial setting or well-designed prospective cohort study
undertaken in experienced centres.
Salvage RP should only be performed in experienced centres. Weak
Recommendations for systemic salvage treatment
Do not offer androgen deprivation therapy to M0 patients with a PSA-doubling time > 12 Strong
months.

6.4 Treatment: Metastatic prostate cancer


6.4.1 Introduction
All prospective data available rely on the definition of M1 disease based on CT scan and bone scan. The
influence on treatment and outcome of newer, more sensitive, imaging has not been assessed yet.

6.4.2 Prognostic factors


Median survival of patients with newly diagnosed metastases is approximately 42 months [1041] with ADT alone,
however, it is highly variable since the M1 population is heterogeneous. Several prognostic factors for survival
have been suggested including the number and location of bone metastases, presence of visceral metastases,
ISUP grade, PS status and initial PSA alkaline phosphatase, but only few have been validated [1042-1045].

PROSTATE CANCER - LIMITED UPDATE 2021 101


‘Volume‘ of disease as a potential predictor was introduced by CHAARTED (Chemo-hormonal
Therapy versus Androgen Ablation Randomized Trial for Extensive Disease in Prostate Cancer) [1045-1047]
and has been shown to be predictive in a powered subgroup analysis for benefit of addition of prostate
radiotherapy ADT [1048].
Based on a large SWOG 9346 cohort, the PSA level after 7 months of ADT was used to create 3
prognostic groups (see Table 6.4.1) [1049]. A PSA < 0.2 ng/mL at 7 months has been confirmed as a prognostic
marker for men receiving ADT for metastatic disease in the CHAARTED study independent of the addition of
docetaxel [1050].

Table 6.4.1 Definition of high- and low-volume and risk in CHAARTED [1045-1047] and LATITIDE [744]

High Low
CHAARTED > 4 Bone metastasis including > 1 outside vertebral column or pelvis Not high
(volume) OR
Visceral metastasis
LATITUDE > 2 high-risk features of: Not high
(risk) • > 3 Bone metastasis
• Visceral metastasis
• > ISUP grade 4

Table 6.4.2: Prognostic factors based on the SWOG 9346 study [1049]

PSA after 7 months of castration Median survival on ADT monotherapy


< 0.2 ng/mL 75 months
0.2 < 4 ng/mL 44 months
> 4 ng/mL 13 months

6.4.3 First-line hormonal treatment


Primary ADT has been the standard of care for over 50 years [709]. There is no high level evidence in favour
of a specific type of ADT, neither for orchiectomy or for an LHRH analogue or antagonist, with the exception
of patients with impending spinal cord compression for whom either a bilateral orchidectomy or LHRH
antagonists are the preferred options.

6.4.3.1 Non-steroidal anti-androgen monotherapy


Based on a Cochrane review comparing non-steroidal anti-androgen (NSAA) monotherapy to castration (either
medical or surgical), NSAA was considered to be less effective in terms of OS, clinical progression, treatment
failure and treatment discontinuation due to adverse events [1051]. The evidence quality of the studies included
in this review was rated as moderate.

6.4.3.2 Intermittent versus continuous androgen deprivation therapy


Three independent reviews [1052-1054] and two meta-analyses [1055, 1056] looked at the clinical efficacy
of intermittent androgen deprivation (IAD) therapy. All of these reviews included 8 RCTs of which only 3 were
conducted in patients with exclusively M1 disease. The 5 remaining trials included different patient groups,
mainly locally-advanced and metastatic patients relapsing.

So far, the SWOG 9346 is the largest trial addressing IAD in M1b patients [1057]. Out of 3,040 screened
patients, only 1,535 patients met the inclusion criteria. This highlights that, at best, only 50% of M1b patients
can be expected to be candidates for IAD, i.e. the best PSA responders. This was a non-inferiority trial leading
to inconclusive results: the actual upper limit was above the pre-specified 90% upper limit of 1.2 (HR: 1.1,
CI: 0.99–1.23), the pre-specified non-inferiority limit was not achieved, and the results did not show a
significant inferiority for any treatment arm. However, based on this study inferior survival with IAD cannot be
completely ruled out.
Other trials did not show any survival difference with an overall HR for OS of 1.02 (0.94–1.11) [1052].
These reviews and the meta-analyses came to the conclusion that a difference in OS or CSS between IAD
and continuous ADT is unlikely. A recent review of the available phase III trials highlighted the limitations of
most trials and suggested a cautious interpretation of the non-inferiority results [1058]. None of the trials that
addressed IAD vs. continuous ADT in M1 patients showed a survival benefit but there was a constant trend
towards improved OS and PFS with continuous ADT. However, most of these trials were non-inferiority trials. In

102 PROSTATE CANCER - LIMITED UPDATE 2021


some cohorts the negative impact on sexual function was less pronounced with IAD. There is a trend favouring
IAD in terms of QoL, especially regarding treatment-related side effects such as hot flushes [1059, 1060].

6.4.3.3 Immediate versus deferred androgen deprivation therapy


In symptomatic patients immediate treatment is mandatory, however, controversy still exists for asymptomatic
metastatic patients due to the lack of high quality studies. A first Cochrane review extracted four RCTs: the
VACURG I and II trials, the MRC trial, and the ECOG 7887 study [1050, 1051]. These studies were conducted
in the pre-PSA era and included patients with advanced metastatic or non-metastatic PCa who received
immediate vs. deferred ADT [1061]. No improvement in PCa CSS was observed, although immediate ADT
significantly reduced disease progression. The Cochrane analysis was updated in 2019 and concluded that
early ADT probably extends time to death of any cause and time to death from PCa [1062]. Since the analysis
included only a very limited number of M1 patients who were not evaluated separately, the benefit of immediate
ADT in this setting remains unclear.

6.4.4 Combination therapies


All of the following combination therapies have been studied with continuous ADT, not intermittent ADT.

6.4.4.1 ‘Complete’ androgen blockade


The largest RCT in 1,286 M1b patients found no difference between surgical castration with or without
flutamide [1063]. However, results with other anti-androgens or castration modalities have differed and
systematic reviews have shown that CAB using a NSAA appears to provide a small survival advantage
(< 5%) vs. monotherapy (surgical castration or LHRH agonists) [1064, 1065] beyond 5 years of survival [1066]
but this minimal advantage in a small subset of patients must be balanced against the increased side effects
associated with long-term use of NSAAs.

6.4.4.2 Androgen deprivation combined with other agents


6.4.4.2.1 Androgen deprivation therapy combined with chemotherapy
Three large RCTs were conducted [801, 1045, 1067]. All trials compared ADT alone as the standard of care with
ADT combined with immediate docetaxel (75 mg/sqm, every 3 weeks within 3 months of ADT initiation). The
primary objective in all three studies was to assess OS. The key findings are summarised in Table 6.4.3.

Table 6.4.3: Key findings - Hormonal treatment combined with chemotherapy

STAMPEDE [801, 1068] GETUG [1067] CHAARTED [1045,1046]


ADT ADT + Docetaxel + P ADT ADT + Docetaxel ADT ADT + Docetaxel
n 1,184 592 193 192 393 397
Newly 58% 59% 75% 67% 73% 73%
diagnosed M+
Key inclusion Patients scheduled for long- Metastatic disease Metastatic disease
criteria term ADT Karnofsky score > 70% ECOG PS 0, 1 or 2
- newly diagnosed M1 or N+
situations
- locally advanced (at least two
of cT3 cT4, ISUP grade > 4,
PSA > 40 ng/mL)
- relapsing locally treated
disease with a PSA > 4 ng/mL
and a PSA-DT < 6 mo.

or PSA > 20 ng/mL,


or nodal
or metastatic relapse
Primary OS OS OS
objective
Median follow 43; 78.2 (update M1) 50 54 (update)
up (mo)
HR (95% CI) 0.78 (0.66-0.93) 1.01 (0.75-1.36) 0.72 (0.59-0.89)

PROSTATE CANCER - LIMITED UPDATE 2021 103


M1 only
n 1,086 - -
HR (95% CI) 0.81 (0.69-0.95) - -
ADT = androgen deprivation therapy; ECOG = Eastern Cooperative Oncology Group; FU = follow-up;
HR = hazard ratio; ISUP = International Society for Urological Pathology; mo = month; n = number of patients;
OS = overall survival; P = prednisone; PSA-DT = prostate-specific antigen-doubling time.

In the GETUG 15 trial, all patients had newly diagnosed M1 PCa, either de novo or after a primary treatment
[1067]. They were stratified based on previous treatment and Glass risk factors [1042]. In the CHAARTED trial
the same inclusion criteria applied and patients were stratified according to disease volume; high volume being
defined as either presence of visceral metastases or four, or more, bone metastases, with at least one outside
the spine and pelvis [1045].

STAMPEDE is a multi-arm multi-stage trial in which the reference arm (ADT monotherapy) included 1,184
patients. One of the experimental arms was docetaxel combined with ADT (n = 593), another was docetaxel
combined with zoledronic acid (n = 593). Patients were included with either M1 or N1, or having two of the
following 3 criteria: T3/4, PSA > 40 ng/mL or ISUP grade 4–5. Also relapsed patients after local treatment
were included if they met one of the following criteria: PSA > 4 ng/mL with a PSA-DT < 6 months or a PSA
> 20 ng/mL, N1 or M1. No stratification was used regarding metastatic disease volume (high/low volume) [801].
In all 3 trials toxicity was mainly haematological with around 12–15% grade 3–4 neutropenia, and 6–12% grade
3–4 febrile neutropenia. The use of granulocyte colony-stimulating factor receptor (GCSF) was shown to be
beneficial in reducing febrile neutropenia. Primary or secondary prophylaxis with GCSF should be based on
available guidelines [1069, 1070].

Based on these data, upfront docetaxel combined with ADT should be considered as a standard in men
presenting with metastases at first presentation provided they are fit enough to receive the drug [1070].
Docetaxel is used at the standard dose of 75 mg/sqm combined with steroids as pre-medication. Continuous
oral corticosteroid therapy is not mandatory.
In subgroup analyses from GETUG-AFU 15 and CHAARTED the beneficial effect of the addition of
docetaxel to ADT is most evident in men with de novo metastatic high-volume disease [1046, 1047], while it
was in the same range whatever the volume in the post-hoc analysis from STAMPEDE [1068]. The effects were
less apparent in men who had prior local treatment although the numbers were small and the event rates lower.
A systematic review and meta-analysis which included these 3 trials showed that the addition
of docetaxel to standard of care improved survival [1070]. The HR of 0.77 (95% CI: 0.68–0.87, p < 0.0001)
translates into an absolute improvement in 4-year survival of 9% (95% CI: 5–14). Docetaxel in addition to
standard of care also improves failure-free survival, with a HR of 0.64 (0.58–0.70, p < 0.0001) translating into a
reduction in absolute 4-year failure rates of 16% (95% CI: 12–19).

6.4.4.2.2 Combination with the new hormonal treatments (abiraterone, apalutamide, enzalutamide)
In two large RCTs (STAMPEDE, LATITUDE) the addition of abiraterone acetate (1000 mg daily) plus prednisone
(5 mg daily) to ADT in men with mHSPC was studied [40, 744, 1071]. The primary objective of both trials was
an improvement in OS. Both trials showed a significant OS benefit but in LATITUDE in high-risk metastatic
patients only with a HR of 0.62 (0.51–0.76) [744]. The HR in STAMPEDE was very similar with 0.63 (0.52–0.76)
in the total patient population (metastatic and non-metastatic) and a HR of 0.61 in the subgroup of metastatic
patients [40]. The inclusion criteria in the two trials differed but both trials were positive for OS. While only
high-risk patients were included in the LATITUDE trial a post-hoc analysis from STAMPEDE showed the same
benefit whatever the risk or the volume stratification [1072].

All secondary objectives such as PFS, time to radiographic progression, time to pain, or time to chemotherapy
were positive and in favour of the combination. The key findings are summarised in Table 6.4.4. No difference in
treatment-related deaths was observed with the combination of ADT plus abiraterone acetate and prednisone
compared to ADT monotherapy (HR: 1.37 [0.82–2.29]). However, twice as many patients discontinued
treatment due to toxicity in the combination arms in STAMPEDE (20%) compared to LATITUDE (12%). Based
on these data upfront abiraterone acetate plus prednisone combined with ADT should be considered as a
standard in men presenting with metastases at first presentation, provided they are fit enough to receive the
drug (see Table 6.4.4) [1071].

In three large RCTs (ENZAMET, ARCHES and TITAN) the addition of AR antagonists to ADT in men with
mHSPC was tested [742, 743, 1073]. In ARCHES the primary endpoint was radiographic progression-free
survival (rPFS). Radiographic PFS was significantly improved for the combination of enzalutamide and ADT

104 PROSTATE CANCER - LIMITED UPDATE 2021


with a HR of 0.39 (0.3–0.5). Approximately 36% of the patients had low-volume disease; around 25% had prior
local therapy and 18% of the patients had received prior docetaxel. In ENZAMET the primary endpoint was
OS. The addition of enzalutamide to ADT improved OS with a HR of 0.67 (0.52–0.86). Approximately half of the
patients had concomitant docetaxel; about 40% had prior local therapy and about half of the patients had low-
volume disease [743]. In the TITAN trial, ADT plus apalutamide was used and rPFS and OS were co-primary
endpoints. Radiographic PFS was significantly improved by the addition of apalutamide with a HR of 0.48
(0.39–0.6); OS at 24 months was improved for the combination with a HR of 0.67 (0.51–0.89). In this trial 16%
of patients had prior local therapy, 37% had low-volume disease and 11% received prior docetaxel [742].
In summary, the addition of AR antagonists significantly improves clinical outcomes with no
convincing evidence of differences between subgroups. The majority of patients treated had de novo
metastatic disease and the evidence is most compelling in this situation. In the trials with the AR antagonists,
a proportion of patients had metachronous disease (see Table 6.4.5); therefore a combination should also be
considered for men progressing after radical local therapy. Lastly, whether the addition of an AR antagonist
plus docetaxel adds further OS benefit is currently not observed. Longer follow-up data are needed before a
definitive conclusion is possible. At the moment, since toxicity clearly increases, AR antagonists plus docetaxel
should not be given outside of clinical trials.

Table 6.4.4: Results from the STAMPEDE arm G and LATITUDE studies

STAMPEDE [40] LATITUDE [744]


ADT ADT + AA + P ADT + placebo ADT + AA + P
n 957 960 597 602
Newly diagnosed N+ 20% 19% 0 0
Newly diagnosed M+ 50% 48% 100% 100%
Key inclusion criteria Patients scheduled for long-term ADT Newly diagnosed M1 disease and 2 out
- newly diagnosed M1 or N+ situations of the 3 risk factors: ISUP grade > 4,
- locally advanced (at least two of cT3 cT4, > 3 bone lesions, measurable visceral
ISUP grade > 4, PSA > 40 ng/mL) metastasis
- relapsing locally treated disease with a
PSA > 4 ng/mL and a PSA-DT < 6 mo.

or PSA > 20 ng/mL


or nodal
or metastatic relapse
Primary objective OS OS
Radiographic PFS
Median follow up (mo) 40 30.4
3-yr. OS 83% (ADT + AA + P) 66% (ADT + AA + P)
76% (ADT) 49% (ADT + placebo)
HR (95% CI) 0.63 (0.52 - 0.76) 0.62 (0.51-0.76)
M1 only
n 1,002 1,199
3-yr. OS NA 66% (ADT + AA + P)
49% (ADT + placebo)
HR (95% CI) 0.61 (0.49-0.75) 0.62 (0.51-0.76)
HR FFS (biological, radiological, clinical or Radiographic PFS:
death): 0.29 (0.25-0.34) 0.49 (0.39-0.53)
AA = abiraterone acetate; ADT = androgen deprivation therapy; CI = confidence interval;
ECOG = Eastern Cooperative Oncology Group; FFS = failure-free survival; HR = hazard ratio; mo = month;
n = number of patients; NA = not available; OS = overall survival; P = prednisone; PFS = progression-free
survival; PSA = prostate-specific antigen; yr. = year.

PROSTATE CANCER - LIMITED UPDATE 2021 105


Table 6.4.5 Results from the ENZAMET and TITAN studies

ENZAMET [1073] TITAN [742]


ADT+ older ADT + enzalutamide ADT + placebo ADT +
antagonist +/-docetaxel apalutamide
+/-docetaxel (SOC)
n 562 563 527 525
Newly diagnosed M+ 72.1% 72.5% 83.7% 78.3%
(ENZAMET incl.
unknown)
Low volume 47% 48% 36% 38%
Primary objective OS OS
Radiographic PFS
Median follow up (mo) 34 30.4
3-yr. OS 3-yr survival: 2-yr survival:
80% (ADT + enzalutamide) 84% (ADT + apalutamide)
72% (SOC) 74% (ADT + placebo)
HR (95% CI) for OS 0.67 (0.52-0.86) 0.67 (0.51-0.89)
ADT = androgen deprivation therapy; CI = confidence interval; HR = hazard ratio; mo = month; n = number of
patients; OS = overall survival; SOC = standard of care; PFS = progression-free survival; yr = year.

6.4.5 Treatment selection and patient selection


There are no head-to-head data comparing 6 cycles of docetaxel and the long-term use of abiraterone acetate
plus prednisone in newly diagnosed mHSPC. However, for a period, patients in STAMPEDE were randomised
to either the addition of abiraterone or docetaxel to standard of care. Data from the two experimental arms has
been extracted although this was not pre-specified in the protocol and therefore the data were not powered for
this comparison. The survival advantage for both drugs appeared similar [1074]. A recent meta-analysis also
found no significant OS benefit for either drug [1075]. Limitations of network meta-analyses include variable
patient populations with different treatment benefits and follow-up periods. In the STOPCAP systematic review
and meta-analysis, abiraterone acetate plus prednisone was found to have the highest probability of being
the most effective treatment [1076]. Both modalities have different and agent-specific side effects and require
strict monitoring of side effects during treatment. Therefore, the choice will most likely be driven by patient
preference, the specific side effects, fitness for docetaxel, availability and cost.

There have been several network meta-analyses of the published data concluding that combination therapy
is more efficient than ADT alone, but none of the combination therapies has been clearly proven to be
superior over another [1077, 1078]. Life expectancy has to be taken into account when deciding on offering a
combination therapy vs. ADT alone. Radiographic PFS is significantly prolonged in all trials for the combination
therapies, e.g., from 14.8 months to 33 months in the LATITUDE trial, therefore suggesting that men with a life
expectancy below 15 months are not likely to profit clinically from receiving a combination therapy.

6.4.6 Deferred treatment for metastatic PCa (stage M1)


The only candidates with metastasised disease who may possibly be considered for deferred treatment are
asymptomatic patients with a strong wish to avoid treatment-related side effects. However, since the median
survival is only 42 months the time without treatment (before symptoms) is short in most cases. The risk of
developing symptoms, and even dying from PCa, without receiving any benefit from hormone treatment has been
highlighted [796, 805]. Patients with deferred treatment for advanced PCa must be amenable to close follow-up.

6.4.7 Treatment of the primary tumour in newly diagnosed metastatic disease


The first reported trial evaluating prostate RT in men with metastatic castration-sensitive disease was the
HORRAD trial. Four hundred and thirty-two patients were randomised to ADT alone or ADT plus IMRT with
IGRT to the prostate. Overall survival was not significantly different (HR: 0.9 [0.7-1.14]), median time to PSA
progression was significantly improved in the RT arm (HR: 0.78 [0.63–0.97]) [1079]. The STAMPEDE trial
evaluated 2,061 men with mCSPC who were randomised to ADT alone vs. ADT plus RT to the prostate. This
trial confirmed that RT to the primary tumour did not improve OS in unselected patients [1048]. However,
following the results from CHAARTED and prior to analysing the data, the original screening investigations were
retrieved and patients categorised as low- or high volume. In the low-volume subgroup (n = 819) there was a
significant OS benefit by the addition of prostate RT and it must be highlighted that this benefit was obtained
without an increased dose. The doses and template used in STAMPEDE should be considered (55 Gy in 20 daily

106 PROSTATE CANCER - LIMITED UPDATE 2021


fractions over 4 weeks or 36 Gy in 6-weekly fractions of 6 Gy or a biological equivalent total dose of 72 Gy).
Therefore RT of the prostate only in patients with low-volume metastatic disease should be considered. Of
note, only 18% of these patients had additional docetaxel and no patients had additional abiraterone acetate
plus prednisone, so no clear recommendation can be made about triple combinations. In addition, it is not
clear if these data can be extrapolated to RP as local treatment as results of ongoing trials are awaited.
In a recent systematic review and meta-analysis including the above two RCTs, the authors found
that, overall, there was no evidence that the addition of prostate RT to ADT improved survival in unselected
patients (HR: 0.92, 95% CI: 0.81–1.04, p = 0.195) [1080]. However, there was a clear difference in the effect of
metastatic burden on survival with an absolute improvement of 7% in 3-year survival in men who had four or
fewer bone metastases.

6.4.8 Metastasis-directed therapy in M1-patients


In patients relapsing after a local treatment, a metastases-targeting therapy has been proposed, with the aim
to delay systemic treatment. There are two randomised phase II trials testing metastasis-directed therapy
(MDT) using surgery ± SABR vs. surveillance [1081] or SABR vs. surveillance in men with oligo-recurrent PCa
[1082]. Oligo-recurrence was defined as < 3 lesions on choline-PET/CT only [1081] or conventional imaging
with MRI/CT and/or bone scan [1082]. The sample size was small with 62 and 54 patients, respectively, and
a substantial proportion of them had nodal disease only [1081]. Androgen deprivation therapy-free survival
was the primary endpoint in one study which was longer with MDT than with surveillance [1081]. The primary
endpoint in the ORIOLE trial was progression after 6 months which was significantly lower with SBRT than with
surveillance (19% vs. 61%, p = 0.005) [1082]. Currently there is no data to suggest an improvement in OS.
Two comprehensive reviews highlighted MDT (SABR) as a promising therapeutic approach that must still be
considered as experimental until the results of the ongoing RCT are available [1083, 1084].

6.4.9 Guidelines for the first-line treatment of metastatic disease

Recommendations Strength rating


Offer immediate systemic treatment with androgen deprivation therapy (ADT) to palliate Strong
symptoms and reduce the risk for potentially serious sequelae of advanced disease (spinal
cord compression, pathological fractures, ureteral obstruction) to M1 symptomatic patients.
Offer luteinising hormone-releasing hormone (LHRH) antagonists, especially to patients with Weak
an impending spinal cord compression or bladder outlet obstruction.
Offer surgery and/or local radiotherapy to any patient with M1 disease and evidence of Strong
impending complications such as spinal cord compression or pathological fracture.
Offer immediate systemic treatment to M1 patients asymptomatic from their tumour. Weak
Discuss deferred ADT with well-informed M1 patients asymptomatic from their tumour since Weak
it lowers the treatment-related side effects, provided the patient is closely monitored.
Offer short-term administration of an older generation androgen receptor (AR) antagonist to Weak
M1 patients starting LHRH agonist to reduce the risk of the ‘flare-up’ phenomenon.
Do not offer AR antagonist monotherapy to patients with M1 disease. Strong
Discuss combination therapy including ADT plus systemic therapy with all M1 patients. Strong
Do not offer ADT monotherapy to patients whose first presentation is M1 disease if they Strong
have no contraindications for combination therapy and have a sufficient life expectancy to
benefit from combination therapy and are willing to accept the increased risk of side effects.
Offer ADT combined with chemotherapy (docetaxel) to patients whose first presentation is Strong
M1 disease and who are fit for docetaxel.
Offer ADT combined with abiraterone acetate plus prednisone or apalutamide or Strong
enzalutamide to patients whose first presentation is M1 disease and who are fit enough for
the regimen.
Offer ADT combined with prostate radiotherapy (using the doses from the STAMPEDE Strong
study) to patients whose first presentation is M1 disease and who have low volume of
disease by CHAARTED criteria.
Do not offer ADT combined with any local treatment (radiotherapy/surgery) to patients with Strong
high volume (CHAARTED criteria) M1 disease outside of clinical trials (except for symptom
control).
Do not offer ADT combined with surgery to M1 patients outside of clinical trials. Strong
Only offer metastasis-directed therapy to M1 patients within a clinical trial setting or well- Strong
designed prospective cohort study.

PROSTATE CANCER - LIMITED UPDATE 2021 107


6.5 Treatment: Castration-resistant PCa (CRPC)
6.5.1 Definition of CRPC
Castrate serum testosterone < 50 ng/dL or 1.7 nmol/L plus either:
a. Biochemical progression: Three consecutive rises in PSA at least one week apart resulting in two 50%
increases over the nadir, and a PSA > 2 ng/mL
or
b. Radiological progression: The appearance of new lesions: either two or more new bone lesions on
bone scan or a soft tissue lesion using RECIST (Response Evaluation Criteria in Solid Tumours) [1085].
Symptomatic progression alone must be questioned and subject to further investigation. It is not
sufficient to diagnose CRPC.

6.5.2 Management of mCRPC - general aspects


Selection of treatment for mCRPC is multifactorial and in general dependent on:
• previous treatment for mHSPC and for non-mHSPC;
• previous treatment for mCRPC;
• quality of response and pace of progression on previous treatment;
• known cross resistance between androgen receptor targeted agents (ARTA);
• co-medication and known drug interactions (see approved summary of product characteristics);
• known genetic alterations;
• known histological variants and DNA repair deficiency (consider platinum or targeted therapy like poly-
ADP ribose);
• local approval status of drugs and reimbursement situation;
• available clinical trials;
• The patient and his co-morbidities.

6.5.2.1 Molecular diagnostics


All metastatic patients should be offered somatic genomic testing for homologous repair and MMR defects,
preferably on metastatic carcinoma tissue but testing on primary tumour may also be performed. Alternatively,
but still less common, genetic testing on circulating tumour DNA (ctDNA) is an option and has been used
in some trials. One test, the FoundationOne® Liquid CDx has been FDA approved [1086]. Defective MMR
assessment can be performed by immunohistochemistry for MMR proteins (MSH2, MSH6, MLH1 and PMS2)
and/or by next-generation sequencing assays [1087]. Germline testing for BRCA1/2, ATM and MMR is
recommended for high risk and particularly for metastatic PCa if clinically indicated.
Molecular diagnostics should be performed by a certified (accredited) institution using a standard
NGS (Next Generation Sequencing) multiplication procedure (minimum depth of coverage of 200 X). The genes
and respective exons should be listed; not only DNA for mutations but RNA needs to be examined for fusions
and protein expression to obtain all clinically relevant information. A critical asset is the decision support
helping to rate the mutations according to their clinical relevance [1088, 1089].
Level 1 evidence for the use of PARP-inhibitors has been reported [1090-1092].
Microsatellite instability (MSI)-high (or mismatch repair deficiency [MMR]) is rare in PCa, but for those patients,
pembrolizumab has been approved by the FDA and could be a valuable additional treatment option [1093, 1094].
Germline molecular testing is discussed in Section 5.1.3 - Genetic testing for inherited PCa.
Recommendations for germline testing are provided in Section 5.1.4.

6.5.3 Treatment decisions and sequence of available options


Approved agents for the treatment of mCRPC in Europe are docetaxel, abiraterone/prednisolone,
enzalutamide, cabazitaxel, olaparib and radium-223. In general, sequencing of ARTA like abiraterone and
enzalutamide is not recommended particularly if the time of response to ADT and to the first ARTA was short
(< 12 months) and high-risk features of rapid progression are present [1095, 1096].
The use of chemotherapy with docetaxel and subsequent cabazitaxel in the treatment sequence
is recommended and should be applied early enough when the patient is still fit for chemotherapy. This is
supported by high level evidence [1095].

6.5.4 Non-metastatic CRPC


Frequent PSA testing in men treated with ADT has resulted in earlier detection of biochemical progression. Of
these men approximately one-third will develop bone metastases detectable on bone scan within two years [1097].
In men with CRPC and no detectable clinical metastases using bone scan and CT-scan, baseline
PSA level, PSA velocity and PSA-DT have been associated with time to first bone metastasis, bone metastasis-
free survival and OS [1097, 1098]. These factors may be used when deciding which patients should be
evaluated for metastatic disease. A consensus statement by the PCa Radiographic Assessments for Detection

108 PROSTATE CANCER - LIMITED UPDATE 2021


of Advanced Recurrence (RADAR) group suggested a bone scan and a CT scan when the PSA reached 2
ng/mL and if this was negative, it should be repeated when the PSA reached 5 ng/mL, and again after every
doubling of the PSA based on PSA testing every three months in asymptomatic men [1099]. Symptomatic
patients should undergo relevant investigation regardless of PSA level. With more sensitive imaging techniques
like PSMA PET/CT or whole-body MRI, more patients are diagnosed with early mCRPC [1100]. It remains
unclear if the use of PSMA PET/CT in this setting improves outcome.

Three large phase III RCTs, PROSPER [1101], SPARTAN [1102] and ARAMIS [1103], evaluated metastasis-free
survival as the primary endpoint in patients with nmCRPC (M0 CRPC) treated with enzalutamide (PROSPER)
vs. placebo or apalutamide (SPARTAN) vs. placebo or darolutamide vs. placebo (ARAMIS), respectively. The
M0 status was established by CT and bone scans. Only patients at high risk for the development of metastasis
with a short PSA-DT of < 10 months were included. Patient characteristics in both trials revealed that about
two-thirds of participants had a PSA-DT of < 6 months. All trials showed a significant metastasis-free
survival benefit (PROSPER: median metastasis-free survival was 36.6 months in the enzalutamide group vs.
14.7 months in the placebo group [HR for metastasis or death: 0.29, 95% CI: 0.24–0.35, p < 0.001]; SPARTAN:
median metastasis-free survival was 40.5 months in the apalutamide group vs. 16.2 months in the placebo
group [HR for metastasis or death, 0.28, 95% CI: 0.23–0.35, p < 0.001]; ARAMIS: median metastasis-free
survival was 40.4 months in the darolutamide group vs. 18.4 months in the placebo group (HR: 0.41, 95%
CI: 0.34–0.50; 2-sided p < 0.0001]). All 3 trials showed a survival benefit after a follow-up of more than
30 months. In view of the long-term treatment with these AR targeting agents in asymptomatic patients,
potential adverse events need to be taken into consideration and the patient informed accordingly.

6.5.5 Metastatic CRPC


The remainder of this section focuses on the management of men with proven metastatic CRPC (mCRPC) on
conventional imaging.

6.5.5.1 Conventional androgen deprivation in CRPC


Eventually men with PCa will show evidence of disease progression despite castration. Two trials have
shown only a marginal survival benefit for patients remaining on LHRH analogues during second- and third-
line therapies [1104, 1105]. However, in the absence of prospective data, the modest potential benefits of
continuing castration outweigh the minimal risk of treatment. In addition, all subsequent treatments have been
studied in men with ongoing androgen suppression, therefore, it should be continued in these patients.

Table 6.5.1: Randomised phase III controlled trials - first-line treatment of mCRPC

Author Intervention Comparison Selection criteria Main outcomes


DOCETAXEL
SWOG 99-16 docetaxel/EMP, mitoxantrone, OS: 17.52 vs. 15.6 mo.
2004 [1106] every 3 weeks, every 3 weeks, (p = 0.02, HR: 0.80;
60 mg/m2, EMP 12 mg/m2 95% CI: 0.67-0.97)
3 x 280 mg/day prednisone 5 mg PFS: 6.3 vs. 3.2 mo.
BID (p < 0.001)
TAX 327 2008 docetaxel, every mitoxantrone, OS: 19.2 for 3 weekly
[1107, 1108] 3 weeks, 75 mg/m2 every 3 weeks, vs. 17.8 mo. 4-weekly
prednisone 5 mg 12 mg/m2, and 16.3 in the control
BID Prednisone 5 mg group.
or BID (p = 0.004, HR: 0.79,
docetaxel, weekly, 95% CI: 0.67-0.93)
30 mg/m2
prednisone 5 mg
BID
ABIRATERONE
COU-AA-302 abiraterone + placebo + - No previous docetaxel. OS: 34.7 vs. 30.3 mo.
2013 [1109-1111] prednisone prednisone - ECOG 0-1. (HR: 0.81, p = 0.0033).
- PSA or radiographic FU: 49.2 mo. rPFS:
progression. 16.5 vs. 8.3 mo.
- No or mild symptoms. (p < 0.0001)
- No visceral metastases.

PROSTATE CANCER - LIMITED UPDATE 2021 109


ENZALUTAMIDE
PREVAIL enzalutamide placebo - No previous docetaxel. OS: 32.4 vs. 30.2 mo.
2014 [1112] - ECOG 0-1. (p < 0.001). FU: 22 mo.
- PSA or radiographic (p < 0.001 HR: 0.71,
progression. 95% CI: 0.60-0.84)
- No or mild symptoms. rPFS: 20.0 mo. vs.
- 10% had visceral mets. 5.4 mo. HR: 0.186
(95% CI: 0.15-0.23)
p < 0.0001)
SIPULEUCEL-T
SIPULEUCEL-T sipuleucel-T [1113] placebo [1113] - Some with previous OS: 25.8 vs. 21.7 mo.
2010 [1113] docetaxel. (p = 0.03 HR: 0.78,
- ECOG 0-1. 95% CI: 0.61-0.98).
- Asymptomatic or FU: 34.1 mo. PFS:
minimally symptomatic. 3.7 vs. 3.6 mo. (no
difference)
2006 [1114] sipuleucel-T [1114] placebo [1114] - ECOG 0-1. OS: 25.9 vs. 21.4 mo.
- No visceral met. (p = 0.1). FU: 36 mo.
- No corticosteroids. PFS: 11.7 vs. 10.0 wk.
IPATASERTIB
IPAtential150 ipatasertib abiraterone + Previously untreated for rPFS in PTEN loss
2020 [1115] (400 mg/d) + prednisolone mCRPC, asymptomatic/ (IHC) population:
abiraterone plus ipatasertib mildly symptomatic, 18.5 vs. 16.5 mo.
(1000 mg/d) + or abiraterone + PTEN loss by IHC. (p = 0.0335, HR: 0.77
prednisone (5 mg prednisolone + 95% CI: 0.61-0.98)
bid) placebo.
BID = twice a day; CI = confidence interval; ECOG = Eastern Cooperative Oncology Group; EMP = 
estramustine; FU = follow-up; HR = hazard ratio; mo = month; PFS = progression-free survival; (r)PFS = 
(radiographic) progression-free survival; OS = overall survival; ICH = immunohistochemistry.

6.5.6 First-line treatment of metastatic CRPC


6.5.6.1 Abiraterone
Abiraterone was evaluated in 1,088 chemo-naïve, asymptomatic or mildly symptomatic mCRPC patients in the
phase III COU-AA-302 trial. Patients were randomised to abiraterone acetate or placebo, both combined with
prednisone [1109]. Patients with visceral metastases were excluded. The main stratification factors were ECOG
PS 0 or 1 and asymptomatic or mildly symptomatic disease. Overall survival and rPFS were the co-primary
endpoints. After a median follow-up of 22.2 months there was significant improvement of rPFS (median 16.5
vs. 8.2 months, HR: 0.52, p < 0.001) and the trial was unblinded. At the final analysis with a median follow-up
of 49.2 months, the OS endpoint was significantly positive (34.7 vs. 30.3 months, HR: 0.81, 95% CI: 0.70–0.93,
p = 0.0033) [1111]. Adverse events related to mineralocorticoid excess and liver function abnormalities were
more frequent with abiraterone, but mostly grade 1–2. Sub-set analysis of this trial showed the drug to be
equally effective in an elderly population (> 75 years) [1116].

6.5.6.2 Enzalutamide
A randomised phase III trial (PREVAIL) included a similar patient population and compared enzalutamide
and placebo [1112]. Men with visceral metastases were eligible but the numbers included were small.
Corticosteroids were allowed but not mandatory. PREVAIL was conducted in a chemo-naïve mCRPC
population of 1,717 men and showed a significant improvement in both co-primary endpoints, rPFS
(HR: 0.186, CI: 0.15–0.23, p < 0.0001), and OS (HR: 0.706, CI: 0.6–0.84, p < 0.001). A > 50% decrease in
PSA was seen in 78% of patients. The most common clinically relevant adverse events were fatigue and
hypertension. Enzalutamide was equally effective and well tolerated in men > 75 years [1117] as well as in
those with or without visceral metastases [1118]. However, for men with liver metastases, there seems to be no
discernible benefit [1118, 1119].
Enzalutamide has also been compared with bicalutamide 50 mg/day in a randomised double blind
phase II study (TERRAIN) [1120] showing a significant improvement in PFS (15.7 months vs. 5.8 months,
HR: 0.44, p < 0.0001) in favour of enzalutamide. With extended follow-up and final analysis the benefit in OS
and rPFS were confirmed [1121].

110 PROSTATE CANCER - LIMITED UPDATE 2021


6.5.6.3 Docetaxel
A statistically significant improvement in median survival of 2.0–2.9 months has been shown with docetaxel-
based chemotherapy compared to mitoxantrone plus prednisone [1108, 1122]. The standard first-line
chemotherapy is docetaxel 75 mg/m2, 3-weekly doses combined with prednisone 5 mg twice a day (BID), up
to 10 cycles. Prednisone can be omitted if there are contraindications or no major symptoms. The following
independent prognostic factors; visceral metastases, pain, anaemia (Hb < 13 g/dL), bone scan progression,
and prior estramustine may help stratify the response to docetaxel. Patients can be categorised into three
risk groups: low risk (0 or 1 factor), intermediate (2 factors) and high risk (3 or 4 factors), and show three
significantly different median OS estimates of 25.7, 18.7 and 12.8 months, respectively [1123].
Age by itself is not a contraindication to docetaxel [1124] but attention must be paid to careful
monitoring and co-morbidities as discussed in Section 5.4 - Estimating life expectancy and health status
[1125]. In men with mCRPC who are thought to be unable to tolerate the standard dose and schedule,
docetaxel 50 mg/m2 every two weeks seems to be well tolerated with less grade 3–4 adverse events and a
prolonged time to treatment failure [1126].

6.5.6.4 Sipuleucel-T
In 2010 a phase III trial of sipuleucel-T showed a survival benefit in 512 asymptomatic or minimally
symptomatic mCRPC patients [1113]. After a median follow-up of 34 months, the median survival was
25.8 months in the sipuleucel-T group compared to 21.7 months in the placebo group, with a HR of 0.78
(p = 0.03). No PSA decline was observed and PFS was similar in both arms. The overall tolerance was very
good, with more cytokine-related adverse events grade 1–2 in the sipuleucel-T group, but the same grade 3–4
adverse events in both arms. Sipuleucel-T is not available in Europe (and has had its licence withdrawn).

6.5.6.5 Ipatasertib
The AKT inhibitor ipatasertib in combination with abiraterone plus prednisone was studied in asymptomatic
or mildly symptomatic patients with PTEN loss by IHC and previously untreated for mCRPC. The randomised
phase III trial (IPAtential) showed a significant benefit for the first endpoint rPFS in the PTEN loss (IHC)
population (18.5 vs. 16.5 mo; p = 0,0335, HR: 0.77, 95% CI: 0.61–0,98). The OS results are still pending. Side
effects of the AKT inhibitor ipatasertib include rash and diarrhoea [750]. This combination is still investigational
[1115].

Table 6.5.2: Randomised controlled phase III - second-line/third-line trials in mCRPC

Author Intervention Comparison Selection criteria Main outcomes


ABIRATERONE
COU-AA-301 abiraterone + placebo + Previous docetaxel. OS: 15.8 vs. 11.2 mo.
2012 [1127] prednisone HR prednisone ECOG 0-2. PSA (p < 0.0001, HR: 0.74, 95%
or radiographic CI: 0.64-0.86; p < 0.0001).
progression. FU: 20.2 mo.
rPFS: no change
COU-AA-301 OS: 14.8 vs. 10.9 mo.
2011 [1128] (p < 0.001 HR: 0.65; 95%
CI: 0.54-0.77).
FU: 12.8 mo.
rPFS: 5.6 vs. 3.6 mo.
Radium-223
ALSYMPCA radium-223 placebo Previous or no OS: 14.9 vs. 11.3 mo.
2013 [1129] previous docetaxel. (p = 0.002, HR: 0.61; 95%
ECOG 0-2. Two or CI: 0.46-0.81).
more symptomatic All secondary endpoints show a
bone metastases. No benefit over best standard of care.
visceral metastases.

PROSTATE CANCER - LIMITED UPDATE 2021 111


CABAZITAXEL
TROPIC cabazitaxel + mitoxantrone Previous docetaxel. OS: 318/378 vs. 346/377 events
2013 [1130] prednisone + prednisone ECOG 0-2. (OR: 2.11; 95% CI: 1.33-3.33).
FU: 25.5 months OS > 2 yr 27% vs.
16% PFS: -
TROPIC OS: 15.1 vs. 12.7 mo.
2010 [1131] (p < 0.0001, HR: 0.70;
95% CI: 0.59-0.83). FU: 12.8 mo.
PFS: 2.8 vs. 1.4 mo.
(p < 0.0001, HR: 0.74,
95% CI: 0.64-0.86)
CARD Cabazitaxel ARTA: Previous docetaxel. Med OS 13.6 vs. 11.0 mo.
2019 [1095] (25 mg/m2 Abiraterone+ Progression < 12 mo. (p = 0.008, HR: 0.64, 95%
Q3W) prednisone on prior alternative CI: 0.46-0.89).
+ prednisone OR ARTA (either before or rPFS 8.0 vs. 3.7 mo.
+ G-CSF Enzalutamide after docetaxel) (p < 0.001, HR: 0.54,
95% CI: 0.40-0.73).
FU: 9.2 mo.
ENZALUTAMIDE
AFFIRM enzalutamide placebo Previous docetaxel. OS: 18.4 vs. 13.6 mo.
2012 [1132] ECOG 0-2. (p < 0.001, HR: 0.63; 95%
CI: 0.53-0.75).
FU: 14.4 mo.
rPFS: 8.3 vs. 2.9 mo.
(HR: 0.40; 95% CI: 0.35-0.47,
p < 0.0001).
PARP inhibitor
PROfound Olaparib Abiraterone + Previous ARTA, rPFS: 7.39 vs. 3.55 mo.
2020 [1091, prednisolone alterations in HRR (p < 0.0001, HR: 0.34; 95%
1092, 1133] or mutated genes CI: 0.25-0.47), conf. ORR
enzalutamide; 33.3% vs. 2.3% (OR 20.86, 95%
cross-over CI: 4.18-379.18).
allowed at OS: 19.1 mo vs. 14.7 mo
progression (in pts with BRCA1/2, ATM
alterations)
(p = 0.0175; HR 0.69, 95%
CI: 0.5-0.97).
*Only studies reporting survival outcomes as primary endpoints have been included.
ARTA = androgen receptor targeting agents; CI = confidence interval; ECOG = Eastern Cooperative Oncology
Group; FU = follow-up; HR = hazard ratio; mo = months OS = overall survival; OR = odds ratio; ORR = objective
response rate; PSA = prostate-specific antigen; rPFS = radiographic progression-free survival; yr = year;
HRR= homologous recombination repair.

6.5.7 Second-line treatment for mCRPC and sequence


All patients who receive treatment for mCRPC will eventually progress. All treatment options in this setting are
presented in Table 6.5.2. High level evidence exists for second-line treatments after first-line treatment with
docetaxel and for third-line therapy.

6.5.7.1 Cabazitaxel
Cabazitaxel is a novel taxane with activity in docetaxel-resistant cancers. It was studied in a large prospective,
randomised, phase III trial (TROPIC) comparing cabazitaxel plus prednisone vs. mitoxantrone plus prednisone
in 755 patients with mCRPC, who had progressed after or during docetaxel-based chemotherapy [1131].
Patients received a maximum of ten cycles of cabazitaxel (25 mg/m2) or mitoxantrone (12 mg/m2) plus
prednisone (10 mg/day). Overall survival was the primary endpoint which was significantly longer with
cabazitaxel (median: 15.1 vs. 12.7 months, p < 0.0001). There was also a significant improvement in PFS
(median: 2.8 vs. 1.4 months, p < 0.0001), objective RECIST response (14.4% vs. 4.4%, p < 0.005), and PSA
response rate (39.2% vs. 17.8%, p < 0.0002). Treatment-associated WHO grade 3–4 adverse events developed
significantly more often in the cabazitaxel arm, particularly haematological (68.2% vs. 47.3%, p < 0.0002)
but also non-haematological (57.4 vs. 39.8%, p < 0.0002) toxicity [1134]. In two post-marketing randomised

112 PROSTATE CANCER - LIMITED UPDATE 2021


phase III trials, cabazitaxel was shown not to be superior to docetaxel in the first-line setting; in the second-line
setting in terms of OS, 20 mg/m2 cabazitaxel was not inferior to 25 mg/m2, but less toxic. Therefore, the lower
dose should be preferred [1135, 1136]. Cabazitaxel should preferably be given with prophylactic granulocyte
colony-stimulating factor (G-CSF) and should be administered by physicians with expertise in handling
neutropenia and sepsis [1137].

6.5.7.2 Abiraterone acetate after prior docetaxel


Positive results of the large phase III trial (COU-AA-301) were reported after a median follow-up of 12.8 months
[1128] and confirmed by the final analysis [1127]. A total of 1,195 patients with mCRPC were randomised 2:1 to
abiraterone acetate plus prednisone or placebo plus prednisone. All patients had progressive disease based on
the Prostate Cancer Clinical Trials Working Group 2 (PCWG2) criteria after docetaxel therapy (with a maximum
of two previous chemotherapeutic regimens). The primary endpoint was OS, with a planned HR of 0.8 in favour
of abiraterone. After a median follow-up of 20.2 months, the median survival in the abiraterone group was 15.8
months compared to 11.2 months in the placebo arm (HR: 0.74, p < 0.0001). The benefit was observed in all
subgroups and all the secondary objectives were in favour of abiraterone (PSA, radiologic tissue response, time
to PSA or objective progression). The incidence of the most common grade 3–4 adverse events did not differ
significantly between arms, but mineralocorticoid-related side effects were more frequent in the abiraterone
group, mainly grade 1–2 (fluid retention, oedema and hypokalaemia).

6.5.7.3 Enzalutamide after docetaxel


The planned interim analysis of the AFFIRM study was published in 2012 [1132]. This trial randomised
1,199 patients with mCRPC in a 2:1 fashion to enzalutamide or placebo. The patients had progressed after
docetaxel treatment, according to the PCWG2 criteria. Corticosteroids were not mandatory, but could be
prescribed, and were received by about 30% of the patients. The primary endpoint was OS, with an expected
HR benefit of 0.76 in favour of enzalutamide. After a median follow-up of 14.4 months, the median survival in
the enzalutamide group was 18.4 months compared to 13.6 months in the placebo arm (HR: 0.63, p < 0.001).
This led to the recommendation to halt and unblind the study. The benefit was observed irrespective of age,
baseline pain intensity, and type of progression. In the final analysis with longer follow-up the OS results were
confirmed despite crossover and extensive post-progression therapies [1121]. Enzalutamide was active also in
patients with visceral metastases.
All the secondary objectives were in favour of enzalutamide (PSA, soft tissue response, QoL, time
to PSA or objective progression). No difference in terms of side effects was observed in the two groups, with a
lower incidence of grade 3–4 adverse events in the enzalutamide arm. There was a 0.6% incidence of seizures
in the enzalutamide group compared to none in the placebo arm.

6.5.7.4 Radium-223
The only bone-specific drug that is associated with a survival benefit is the α-emitter radium-223. In a large
phase III trial (ALSYMPCA) 921 patients with symptomatic mCRPC, who failed or were unfit for docetaxel, were
randomised to six injections of 50 kBq/kg radium-223 or placebo, plus standard of care. The primary endpoint
was OS. Radium-223 significantly improved median OS by 3.6 months (HR: 0.70, p < 0.001) and was also
associated with prolonged time to first skeletal event, improvement in pain scores and improvement in QoL
[1129]. The associated toxicity was mild and, apart from slightly more haematologic toxicity and diarrhoea with
radium-223, it did not differ significantly from that in the placebo arm [1129]. Radium-223 was effective and
safe whether or not patients were docetaxel pre-treated [1138]. Due to safety concerns, use of radium-223
was recently restricted to after docetaxel and at least one AR targeted agent [1139]. In particular, the use of
radium-223 in combination with abiraterone acetate plus prednisolone showed significant safety risks related
to fractures and more deaths. This was most striking in patients without the concurrent use of anti-resorptive
agents [1140].

6.5.8 Treatment after docetaxel and one line of hormonal treatment for mCRPC
For men progressing quickly on AR targeted therapy (< 12 months) it is now clear that cabazitaxel is the
treatment supported by the best data. The CARD trial, an open label randomised phase III trial, evaluated
cabazitaxel after docetaxel and one line of ARTA (either abiraterone plus prednisolone or enzalutamide) [1095].
It included patients progressing in less than 12 months on previous abiraterone or enzalutamide for mCRPC.
Cabazitaxel more than doubled rPFS vs. another ARTA and reduced the risk of death by 36% vs. ARTA. The
rPFS with cabazitaxel remained superior regardless of the ARTA sequence and if docetaxel was given before,
or after, the first ARTA.
The choice of further treatment after docetaxel and one line of hormonal treatment for mCRPC
is open for patients who have a > 12 months response to first-line abiraterone or enzalutamide for mCRPC
[1141]. Either radium-223 or second-line chemotherapy (cabazitaxel) are reasonable options. In general,

PROSTATE CANCER - LIMITED UPDATE 2021 113


subsequent treatments in unselected patients are expected to have less benefit than with earlier use [1142,
1143] and there is evidence of cross-resistance between enzalutamide and abiraterone [1144, 1145].

Poly (ADP-ribose) polymerase inhibitors have shown high rates of response in men with somatic homologous
recombination repair (HRR) deficiency in initial studies. Men previously treated with both docetaxel and at
least one ARTA and whose tumours demonstrated homozygous deletions or deleterious mutations in DNA-
repair genes showed an 88% response rate to olaparib [1146] and in another confirmatory trial a confirmed
composite response of 54.3% (95% CI: 39·0–69.1) in the 400 mg cohort and in 18 of 46 (39.1%; 25.1–54.6)
evaluable patients in the 300 mg cohort [1147].

6.5.8.1 PARP inhibitors for mCRPC


So far, two PARP inhibitors, olaparib and rucaparib, are licenced by the FDA (EMA only approved olaparib) and
several other PARP inhibitors are under investigation (e.g., talazoparib, niraparib).

A randomised phase III trial (PROfound) compared the PARP inhibitor olaparib to an alternative ARTA in
mCRPC with alterations in > 1 of any qualifying gene with a role in HRR and progression on an ARTA. Most
patients were heavily pre-treated with 1–2 chemotherapies and up to 2 ARTAs [1091, 1092]. Radiographic
PFS by blinded independent central review in the BRCA1/2 or ATM mutated population (Cohort A) was the
first endpoint and significantly favoured olaparib (HR: 0.49, 95% CI: 0.38–0.63). The final results for OS
demonstrated a significant improvement among men with BRCA1/2 or ATM mutations (Cohort A) (p = 0.0175;
HR: 0.69, 95% CI: 0.50– 0.97). This was not significant in men with any (other) HRR alteration (Cohort B) (HR:
0.96, 95% CI: 0.63–1.49). Of note, patients in the physician’s choice of enzalutamide/abiraterone-arm who
progressed, 66% (n = 86/131) crossed over to olaparib. When looking specifically at the Cohort B patients,
there was no advantage to olaparib in the rPFS by blinded independent central review (HR: 0.88, 95%
CI: 0.58–1.36) or in OS (HR: 0.73, 95% CI: 0.45–1.23), however, there was a benefit to olaparib for rPFS by
investigator assessment (HR: 0.60, 95% CI: > 0.39–0.93) [1092, 1133].
The most common adverse events were anaemia (46.1% vs. 15.4%), nausea (41.4% vs. 19.2%),
decreased appetite (30.1% vs. 17.7%) and fatigue (26.2% vs. 20.8%) for olaparib vs. enzalutamide/
abiraterone. Among patients receiving olaparib 16.4% discontinued treatment secondary to an adverse event,
compared to 8.5% of patients receiving enzalutamide/abiraterone. Interestingly, 4.3% of patients receiving
olaparib had a pulmonary embolism, compared to 0.8% among those receiving enzalutamide/abiraterone,
none of which were fatal. There were no reports of myelodysplastic syndrome or acute myeloid leukaemia. This
is the first trial to show a benefit for genetic testing and precision medicine in mCRPC.

The olaparib approval by the FDA is for patients with deleterious or suspected deleterious germline or
somatic homologous recombination repair (HRR) gene-mutated mCRPC, who have progressed following prior
treatment with enzalutamide or abiraterone. The EMA approved olaparib for patients with BRCA1 and BRCA2
alterations [1148]. The recommended olaparib dose is 600 mg daily (300 mg taken orally twice daily), with or
without food.

Rucaparib has been approved for patients with deleterious BRCA mutations (germline and/or somatic) who
have been treated with ARTA and a taxane-based chemotherapy [1149]. Approval was not based on OS
data but on the results of the single-arm TRITON2 trial (NCT02952534). The confirmed ORR per independent
radiology review in 62 patients with deleterious BRCA mutations was 43.5% (95% CI: 31–57) [1150].

6.5.8.2 Sequencing treatment


6.5.8.2.1 ARTA -> ARTA (chemotherapy-naïve patients)
The use of sequential ARTAs in mCRPC showed limited benefit in retrospective series as well as in one
prospective trial [1151-1158]. In particular in patients who had a short response to the first ARTA for mCRPC
(< 12 months), this sequence should be avoided because of known cross resistance and the availability of
chemotherapy and PARP inhibitors (if a relevant mutation is present).
In highly selected patients treated for more than 24 weeks with abiraterone plus prednisolone, the
sequence with enzalutamide showed some activity with a median rPFS of 8.1 months (95% CI: 6.1–8.3) and
an unconfirmed PSA response rate of 27% [1159]. In case the patient is unfit for chemotherapy and a PARP
inhibitor best supportive care should be considered in case no other appropriate treatment option is available
(clinical trial: immunotherapy if MSI-high). An ARTA-ARTA sequence should never be the preferred option but
might be considered in such patients if the PS still allows for active treatment and the potential side effects
seem manageable.
First prospective cross-over data on an ARTA-ARTA sequence [1151] and a systematic review
and meta-analysis suggest that for the endpoints PFS and PSA PFS, but not for OS, abiraterone followed by
enzalutamide is the preferred choice [1160].

114 PROSTATE CANCER - LIMITED UPDATE 2021


6.5.8.2.2 ARTA -> PARP inhibitor/olaparib
This sequence in patients with deleterious or suspected deleterious germline or somatic homologous
recombination repair (HRR) gene-mutated mCRPC is supported by data from the randomised phase III PROfound
trial [1092]. A subgoup of patients in this trial was pre-treated with one or two ARTAs and no chemotherapy. The
ARTA – docetaxel - PARP inhibitor vs. ARTA – PARP inhibitor - docetaxel sequences are still under investigation.

6.5.8.2.3 Docetaxel for mHSPC -> docetaxel rechallenge


There is limited evidence for second- or third-line use of docetaxel after treatment with docetaxel for mHSPC.
Docetaxel seems to be less active than ARTA at progression to mCRPC following docetaxel for mHSPC [1161].

6.5.8.2.4 ARTA -> docetaxel or docetaxel -> ARTA followed by PARP inhibitor
Both olaparib and rucaparib are active in biomarker-selected mCRPC patients after ARTA and docetaxel in
either sequence [1092, 1149].

6.5.8.2.5 ARTA before or after docetaxel


There is level 1 evidence for both sequences (see Table 6.5.2).

6.5.8.2.6 ARTA –> docetaxel -> cabazitaxel or docetaxel –> ARTA -> cabazitaxel
Both third-line treatment sequences are supported by level 1 evidence. Of note, there is high level evidence
favouring cabazitaxel vs. a second ARTA after docetaxel and one ARTA. CARD is the first prospective
randomised phase III trial adressing this question (see Table 6.5.2) [1095].

6.5.9 Prostate-specific membrane antigen (PSMA) therapy


6.5.9.1 Background
During the 90s several radiopharmaceuticals including phosphorous-32, strontium-89, yttrium-90,
samarium-153, and rhenium-186 [1162] were developed for the treatment of bone pain secondary to
metastasis from PCa. They were effective at palliation; relieving pain and improving QoL, especially in the
setting of diffuse bone metastasis. However, they never gained widespread adoption. The first radioisotope to
demonstrate a survival benefit was radium-223 (see Section 6.5.7.4).

6.5.9.2 PSMA-based therapy


The increasing use of PSMA PET as a diagnostic tracer and the realisation that this allowed identification
of a greater number of metastatic deposits led to attempts to treat cancer by replacing the imaging isotope
with a therapeutic isotope which accumulates where the tumour is demonstrated (theranostics) [1164].
Therefore, after identification of the target usually with diagnostic 68Gallium-labelled PSMA, therapeutic
radiopharmaceuticals labelled with beta (lutetium-177 or ytrium-90) or a (actinium-225) emitting isotopes could
be used to treat metastatic PCa. At present, all of these agents should be regarded as investigational.

The PSMA therapeutic radiopharmaceutical supported with the most robust data is 177Lu-PSMA-617. The
first patient was treated in 2014 and early clinical studies evaluating the safety and efficacy of Lu-PSMA
therapy have demonstrated promising results, despite the fact that a significant proportion of men had
already progressed on multiple therapies [1165]. Nonetheless, most of the literature is based on single-centre
experience and RCTs are lacking [1166]. Recently, data from uncontrolled prospective phase II trials have been
published reporting high response rates with low toxic effects [1167, 1168]. Positive signals are coming from
a randomised phase II trial comparing Lu-PSMA with cabazitaxel in ARTA and docetaxel pre-treated patients.
The primary endpoint of PSA reduction > 50% was achieved in highly selected patients (PSMA- and FDG PET/
CT criteria) was superior with Lu-PSMA [1169]. More robust data are expected from ongoing trials.

6.5.10 Immunotherapy for mCRPC


The immune checkpoint inhibitor pembrolizumab was approved by the FDA for all MMR–deficient cancers or in
those with instable microsatellite status (MSI-high) [1093]. This also applies to PCa but is a very rare finding in
this tumour entity [1094]. In all other PCa patients pembrolizumab monotherapy is still experimental. It shows
limited anti-tumour activity with an acceptable safety profile, again in a small subset of patients. A phase II trial
enrolled 258 patients treated with pembrolizumab [1163]. The objective response rate was around 4%, but
those responses were durable. Combination immunotherapy is under investigation.

6.5.11 Monitoring of treatment


Baseline examinations should include a medical history, clinical examination as well as baseline blood
tests (PSA, total testosterone level, full blood count, renal function, baseline liver function tests, alkaline
phosphatase), bone scan and CT of chest, abdomen and pelvis [1170, 1171]. The use of choline or PSMA

PROSTATE CANCER - LIMITED UPDATE 2021 115


PET/CT scans for progressing CRPC is unclear and most likely not as beneficial as for patients with BCR or
hormone-naïve disease. Flares, PSMA upregulation and discordant results compared with PSA response or
progression on ARTA have been described [1172]. Prostate-specific antigen alone is not reliable enough [1173]
for monitoring disease activity in advanced CRPC since visceral metastases may develop in men without
rising PSA [1174]. Instead, the PCWG2 recommends a combination of bone scintigraphy and CT scans, PSA
measurements and clinical benefit in assessing men with CRPC [1122]. A majority of experts at the 2015
Advanced Prostate Cancer Consensus Conference (APCCC) suggested regular review and repeating blood
profile every two to three months with bone scintigraphy and CT scans at least every six months, even in the
absence of a clinical indication [1170]. This reflects that the agents with a proven OS benefit all have potential
toxicity and considerable cost and patients with no objective benefit should have their treatment modified. The
APCCC participants stressed that such treatments should not be stopped for PSA progression alone. Instead,
at least two of the three criteria (PSA progression, radiographic progression and clinical deterioration) should
be fulfilled to stop treatment. For trial purposes, the updated PCWG3 put more weight on the importance of
documenting progression in existing lesions and introduced the concept of no longer ‘clinically benefiting‘ to
underscore the distinction between first evidence of progression and the clinical need to terminate or change
treatment [1175]. These recommendations also seem valid for clinical practice outside trials.

6.5.12 When to change treatment


The timing of mCRPC treatment change remains a matter of debate in mCRPC although it is clearly advisable
to start or change treatment immediately in men with symptomatic progressing metastatic disease. Preferably,
any treatment change should precede development of de novo symptoms or worsening of existing symptoms.
Although, the number of effective treatments is increasing, head-to-head comparisons are still rare, as are
prospective data assessing the sequencing of available agents. Therefore it is not clear how to select the
most appropriate ‘second-line‘ treatment, in particular in patients without HRR alterations or other biomarkers.
A positive example, however, is the CARD trial which clearly established cabazitaxel as the better third-line
treatment in docetaxel pre-treated patients after one ARTA compared to the use of a second ARTA [1095].
The ECOG PS has been used to stratify patients. Generally men with a PS of 0–1 are likely to
tolerate treatments and those with a PS of > 2 are less likely to benefit. However, it is important that treatment
decisions are individualised, in particular when symptoms related to disease progression are impacting on PS.
In such cases, a trial of active life-prolonging agents to establish if a given treatment will improve the PS may
be appropriate. Sequencing of treatment is discussed in a summery paper published following the St. Gallen
Advanced Prostate Cancer Consensus Conference 2019 [1176].

6.5.13 Symptomatic management in metastatic CRPC


Castration-resistant PCa is usually a debilitating disease often affecting the elderly male. A multidisciplinary
approach is required with input from urologists, medical oncologists, radiation oncologists, nurses,
psychologists and social workers [1176, 1177]. Critical issues of palliation must be addressed when
considering additional systemic treatment, including management of pain, constipation, anorexia, nausea,
fatigue and depression.

6.5.13.1 Common complications due to bone metastases


Most patients with CRPC have painful bone metastases. External beam radiotherapy is highly effective, even
as a single fraction [1178, 1179]. A single infusion of a third generation bisphosphonate could be considered
when RT is not available [1180]. Common complications due to bone metastases include vertebral collapse
or deformity, pathological fractures and spinal cord compression. Cementation can be an effective treatment
for painful spinal fracture whatever its origin, clearly improving both pain and QoL [1181]. It is important to
offer standard palliative surgery, which can be effective for managing osteoblastic metastases [1182, 1183].
Impending spinal cord compression is an emergency. It must be recognised early and patients should be
educated to recognise the warning signs. Once suspected, high-dose corticosteroids must be given and MRI
performed as soon as possible. A systematic neurosurgery or orthopaedic surgeon consultation should be
planned to discuss a possible decompression, followed by EBRT [1184]. Otherwise, EBRT with, or without,
systemic therapy, is the treatment of choice.

6.5.13.2 Preventing skeletal-related events


6.5.13.2.1 Bisphosphonates
Zoledronic acid has been evaluated in mCRPC to reduce skeletal-related events (SRE). This study was
conducted when no active anti-cancer treatments, but for docetaxel, were available. Six hundred and forty
three patients who had CRPC with bone metastases were randomised to receive zoledronic acid, 4 or 8 mg
every three weeks for 15 consecutive months, or placebo [1185]. The 8 mg dose was poorly tolerated and
reduced to 4 mg but did not show a significant benefit. However, at 15 and 24 months of follow-up, patients

116 PROSTATE CANCER - LIMITED UPDATE 2021


treated with 4 mg zoledronic acid had fewer SREs compared to the placebo group (44 vs. 33%, p = 0.021)
and in particular fewer pathological fractures (13.1 vs. 22.1%, p = 0.015). Furthermore, the time to first SRE
was longer in the zoledronic acid group. No survival benefit has been seen in any prospective trial with
bisphosphonates.

6.5.13.2.2 RANK ligand inhibitors


Denosumab is a fully human monoclonal antibody directed against RANKL (receptor activator of nuclear factor
kappa-B ligand), a key mediator of osteoclast formation, function, and survival. In M0 CRPC, denosumab
has been associated with increased bone-metastasis-free survival compared to placebo (median benefit:
4.2 months, HR: 0.85, p = 0.028) [1178]. This benefit did not translate into a survival difference (43.9 compared
to 44.8 months, respectively) and neither the FDA or the EMA have approved denosumab for this indication
[1186].
The efficacy and safety of denosumab (n = 950) compared with zoledronic acid (n = 951) in
patients with mCRPC was assessed in a phase III trial. Denosumab was superior to zoledronic acid in delaying
or preventing SREs as shown by time to first on-study SRE (pathological fracture, radiation or surgery to
bone, or spinal cord compression) of 20.7 vs. 17.1 months, respectively (HR: 0.82, p = 0.008). Both urinary
N-telopeptide and bone-specific alkaline phosphatase were significantly suppressed in the denosumab arm
compared with the zoledronic acid arm (p < 0.0001 for both). However, these findings were not associated with
any survival benefit and in a recent post-hoc re-evaluation of endpoints, denosumab showed identical results
when comparing SREs and symptomatic skeletal events [1187].

The potential toxicity (e.g., osteonecrosis of the jaw, hypocalcaemia) of these drugs must always be kept in
mind (5–8.2% in M0 CRPC and mCRPC, respectively) [1187-1189]. Patients should have a dental examination
before starting therapy as the risk of jaw necrosis is increased by several risk factors including a history
of trauma, dental surgery or dental infection [1190]. Also, the risk for osteonecrosis of the jaw increased
numerically with the duration of use in a pivotal trial [1191] (one year vs. two years with denosumab), but
this was not statistically significant when compared to zoledronic acid [1186]. According to the EMA,
hypocalcaemia is a concern in patients treated with denosumab and zoledronic acid. Hypocalcaemia must be
corrected by adequate intake of calcium and vitamin D before initiating therapy [1192]. Hypocalcaemia should
be identified and prevented during treatment with bone protective agents (risk of severe hypocalcaemia is
8% and 5% for denosumab and zoledronic acid, respectively) [1189]. Serum calcium should be measured in
patients starting therapy and monitored during treatment, especially during the first weeks and in patients with
risk factors for hypocalcaemia or on other medication affecting serum calcium. Daily calcium (> 500 mg) and
vitamin D (> 400 IU equivalent) are recommended in all patients, unless in case of hypercalcaemia [1189, 1193,
1194].

6.5.14 Summary of evidence and guidelines for life-prolonging treatments of castrate-resistant


disease

Summary of evidence LE
First-line treatment for mCRPC will be influenced by which treatments were used when metastatic 4
cancer was first discovered.
No clear-cut recommendation can be made for the most effective drug for first-line CRPC treatment 3
(i.e., hormone therapy, chemotherapy or radium-223) as no validated predictive factors exist.

Recommendations Strength rating


Ensure that testosterone levels are confirmed to be < 50 ng/dL before diagnosing castrate- Strong
resistant PCa (CRPC).
Counsel, manage and treat patients with metastatic CRPC (mCRPC) in a multidisciplinary Strong
team.
Treat patients with mCRPC with life-prolonging agents. Strong
Offer mCRPC patients somatic and/or germline molecular testing as well as testing for Strong
mismatch repair deficiencies or microsatellite instability.

PROSTATE CANCER - LIMITED UPDATE 2021 117


6.5.15 Guidelines for systematic treatments of castrate-resistant disease

Recommendations Strength rating


Base the choice of treatment on the performance status, symptoms, co-morbidities, Strong
location and extent of disease, genomic profile, patient preference, and on the previous
treatment for hormone-sensitive metastatic PCa (mHSPC) (alphabetical order: abiraterone,
cabazitaxel, docetaxel, enzalutamide, olaparib, radium-223, sipuleucel-T).
Offer patients with mCRPC who are candidates for cytotoxic therapy and are chemotherapy Strong
naïve docetaxel with 75 mg/m2 every 3 weeks.
Offer patients with mCRPC and progression following docetaxel chemotherapy further Strong
life-prolonging treatment options, which include abiraterone, cabazitaxel, enzalutamide,
radium-223 and olaparib in case of DNA homologous recombination repair (HRR) alterations.
Base further treatment decisions of mCRPC on performance status, previous treatments, Strong
symptoms, co-morbidities, genomic profile, extent of disease and patient preference.
Offer abiraterone or enzalutamide to patients previously treated with one or two lines of Strong
chemotherapy.
Avoid sequencing of androgen receptor targeted agents. Weak
Offer chemotherapy to patients previously treated with abiraterone or enzalutamide. Strong
Offer cabazitaxel to patients previously treated with docetaxel. Strong
Offer cabazitaxel to patients previously treated with docetaxel and progressing within 12 Strong
months of treatment with abiraterone or enzalutamide.
Novel agents
Offer poly(ADP-ribose) polymerase (PARP) inhibitors to pre-treated mCRPC patients with Strong
relevant DNA repair gene mutations.

6.5.16 Guidelines for supportive care of castrate-resistant disease


These recommendations are in addition to appropriate systemic therapy.

Recommendations Strength rating


Offer bone protective agents to patients with mCRPC and skeletal metastases to prevent Strong
osseous complications.
Monitor serum calcium and offer calcium and vitamin D supplementation when prescribing Strong
either denosumab or bisphosphonates.
Treat painful bone metastases early on with palliative measures such as intensity-modulated Strong
radiation therapy plus image-guided radiation therapy and adequate use of analgesics.
In patients with spinal cord compression start immediate high-dose corticosteroids and Strong
assess for spinal surgery followed by irradiation. Offer radiation therapy alone if surgery is
not appropriate.

6.5.17 Guideline for non-metastatic castrate-resistant disease

Recommendation Strength rating


Offer apalutamide, darolutamide or enzalutamide to patients with M0 CRPC and a high risk Strong
of developing metastasis (PSA-DT < 10 months) to prolong time to metastases and overall
survival.

6.6 Summary of guidelines for the treatment of prostate cancer

Table 6.6.1: EAU risk groups for biochemical recurrence of localised and locally-advanced prostate

Definition
Low-risk Intermediate-risk High-risk
PSA < 10 ng/mL PSA 10-20 ng/mL PSA > 20 ng/mL any PSA
and GS < 7 (ISUP grade 1) or GS 7 (ISUP grade 2/3) or GS > 7 (ISUP grade 4/5) any GS (any ISUP grade)
and cT1-2a or cT2b or cT2c cT3-4 or cN+
Localised Locally advanced
GS = Gleason score; ISUP = International Society for Urological Pathology; PSA = prostate-specific antigen.

118 PROSTATE CANCER - LIMITED UPDATE 2021


6.6.1 General guidelines recommendations for treatment of prostate cancer

Recommendations Strength rating


Inform patients that based on robust current data with up to 12 years of follow-up, no Strong
active treatment modality has shown superiority over any other active management options
or deferred active treatment in terms of overall- and PCa-specific survival for clinically
localised disease.
Offer a watchful waiting policy to asymptomatic patients with a life expectancy < 10 years Strong
(based on co-morbidities).
Inform patients that all active treatments have side effects. Strong
Surgical treatment
Inform patients that no surgical approach (open-, laparoscopic- or robotic radical Weak
prostatectomy) has clearly shown superiority in terms of functional or oncological results.
When a lymph node dissection (LND) is deemed necessary, perform an extended LND Strong
template for optimal staging.
Do not perform nerve-sparing surgery when there is a risk of ipsilateral extracapsular extension Weak
(based on cT stage, ISUP grade, nomogram, multiparametric magnetic resonance imaging).
Do not offer neoadjuvant androgen deprivation therapy before surgery. Strong
Radiotherapeutic treatment
Offer intensity-modulated radiation therapy (IMRT) plus image-guided radiation therapy Strong
(IGRT) for definitive treatment of PCa by external-beam radiation therapy.
Offer moderate hypofractionation (HFX) with IMRT including IGRT to the prostate, to Strong
patients with localised disease.
Ensure that moderate HFX adheres to radiotherapy protocols from trials with equivalent Strong
outcome and toxicity, i.e. 60 Gy/20 fractions in 4 weeks or 70 Gy/28 fractions in 6 weeks.
Offer low-dose rate (LDR) brachytherapy monotherapy to patients with good urinary Strong
function and low- or good prognosis intermediate-risk localised disease.
Offer LDR or high-dose rate brachytherapy boost combined with IMRT including IGRT to
patients with good urinary function and intermediate-risk disease with adverse features or
high-risk disease.
Active therapeutic options outside surgery and radiotherapy
Only offer cryotherapy and high-intensity focused ultrasound within a clinical trial setting or Strong
well-designed prospective cohort study.
Only offer focal therapy within a clinical trial setting or well-designed prospective cohort study. Strong

6.6.2 Guidelines recommendations for the various disease stages

Recommendations Strength rating


Low-risk disease
Active Selection of patients
surveillance (AS) Offer AS to patients with life expectancy > 10 years and low-risk Strong
disease.
If a patient has had upfront multiparametric magnetic resonance Weak
imaging (mpMRI) followed by systematic and targeted biopsies there
is no need for confirmatory biopsies.
Patients with intraductal and cribiform histology on biopsy should be Strong
excluded from AS.
Perform a mpMRI before a confirmatory biopsy if no mpMRI has been Strong
performed before the initial biopsy.
Take both targeted biopsy (of any PI-RADS > 3 lesion) and systematic Strong
biopsy if a confirmatory biopsy is performed.
Follow-up strategy
Perform serum prostate-specific antigen (PSA) assessment every 6 Strong
months.
Perform digital rectal examination (DRE) every 12 months. Strong
Counsel patients about the possibility of needing further treatment in Strong
the future.

PROSTATE CANCER - LIMITED UPDATE 2021 119


Active treatment Offer surgery and radiotherapy (RT) as alternatives to AS to patients Weak
suitable for such treatments and who accept a trade-off between
toxicity and prevention of disease progression.
Pelvic lymph node Do not perform a PLND (estimated risk for pN+ < 5%). Strong
dissection (PLND)
Radiotherapeutic Offer low-dose rate (LDR) brachytherapy to patients with low-risk PCa, Strong
treatment without a recent transurethral resection of the prostate (TURP) and
with a good International Prostatic Symptom Score (IPSS).
Use intensity-modulated radiation therapy (IMRT) plus image-guided Strong
radiation therapy (IGRT) with a total dose of 74-80 Gy or moderate
hypofractionation (60 Gy/20 fx in 4 weeks, or 70 Gy/28 fx in 6 weeks),
without androgen deprivation therapy (ADT).
Other therapeutic Do not offer ADT monotherapy to asymptomatic men not able to Strong
options receive any local treatment.
Only offer whole gland treatment (such as cryotherapy, high-intensity Strong
focused ultrasound [HIFU], etc.) or focal treatment within a clinical trial
setting or well-designed prospective cohort study.
Intermediate-risk disease
Active surveillance Offer AS to highly selected patients with ISUP grade group 2 disease Weak
(i.e. < 10% pattern 4, PSA < 10 ng/mL, < cT2a, low disease extent
on imaging and biopsy) accepting the potential increased risk of
metastatic progression.
Radical Offer RP to patients with intermediate-risk disease and a life Strong
Prostatectomy expectancy > 10 years.
(RP) Offer nerve-sparing surgery to patients with a low risk of extracapsular Strong
disease.
Extended pelvic Perform an ePLND in intermediate-risk disease if the estimated risk for Strong
lymph node positive lymph nodes exceeds 5%.
dissection (ePLND)
Radiotherapeutic Offer LDR brachytherapy to intermediate-risk patients with ISUP grade Strong
treatment 2 with < 33% of biopsy cores involved, without a recent TURP and
with a good IPSS Score.
For IMRT plus IGRT, use a total dose of 76-78 Gy or moderate Strong
hypofractionation (60 Gy/20 fx in 4 weeks or 70 Gy/28 fx in 6 weeks),
in combination with short-term ADT (4 to 6 months).
In patients not willing to undergo ADT, use a total dose of IMRT plus Weak
IGRT (76-78 Gy) or moderate hypofractionation (60 Gy/20 fx in 4 weeks
or 70 Gy/28 fx in 6 weeks) or a combination with brachytherapy.
Other therapeutic Only offer whole-gland ablative therapy (such as cryotherapy, HIFU, Strong
options etc.) or focal ablative therapy for intermediate-risk disease within a
clinical trial setting or well-designed prospective cohort study.
Do not offer ADT monotherapy to intermediate-risk asymptomatic men Weak
not able to receive any local treatment.
High-risk localised disease
Radical Offer RP to selected patients with high-risk localised PCa, as part of Strong
prostatectomy potential multi-modal therapy.
Extended pelvic Perform an ePLND in high-risk PCa. Strong
lymph node Do not perform a frozen section of nodes during RP to decide whether Strong
dissection to proceed with, or abandon, the procedure.
Radiotherapeutic In patients with high-risk localised disease, use IMRT plus IGRT with Strong
treatments 76-78 Gy in combination with long-term ADT (2 to 3 years).
In patients with high-risk localised disease, use IMRT and IGRT with Weak
brachytherapy boost (either HDR or LDR), in combination with long-
term ADT (2 to 3 years).
Therapeutic Do not offer either whole gland nor focal therapy to patients with high- Strong
options outside risk localised disease.
surgery and Only offer ADT monotherapy to those patients unwilling or unable to Strong
radiotherapy receive any form of local treatment if they have a PSA-doubling time
< 12 months, and either a PSA > 50 ng/mL or a poorly-differentiated
tumour.

120 PROSTATE CANCER - LIMITED UPDATE 2021


Locally-advanced disease
Radical Offer RP to selected patients with locally-advanced PCa as part of Strong
prostatectomy multi-modal therapy.
Extended pelvic Perform an ePLND prior to RP in locally-advanced PCa. Strong
lymph node
dissection
Radiotherapeutic In patients with locally-advanced disease, offer IMRT plus IGRT in Strong
treatments combination with long-term ADT.
Offer long-term ADT for at least two years. Weak
Therapeutic Do not offer whole gland treatment or focal treatment to patients with Strong
options outside locally-advanced PCa.
surgery and Only offer ADT monotherapy to those patients unwilling or unable to Strong
radiotherapy receive any form of local treatment if they have a PSA-doubling time
< 12 months, and either a PSA > 50 ng/mL, a poorly-differentiated
tumour or troublesome local disease-related symptoms.
Offer patients with cN1 disease a local treatment (either RP or IMRT Weak
plus IGRT) plus long-term ADT.
Adjuvant treatment after radical prostatectomy
pN0 & pN1 Do not prescribe adjuvant ADT in pN0 patients. Strong
disease Only offer adjuvant intensity-modulated radiation therapy (IMRT) plus Strong
image-guided radiation therapy (IGRT) to high-risk patients (pN0) with
at least two out of three high-risk features (ISUP grade group 4–5,
pT3 ± positive margins).
Discuss three management options with patients with pN1 disease Weak
after an ePLND, based on nodal involvement characteristics:
1. Offer adjuvant ADT;
2. Offer adjuvant ADT with additional IMRT plus IGRT;
3. Offer observation (expectant management) to a patient after
eLND and < 2 nodes and a PSA < 0.1 ng/mL.
Non-curative or palliative treatments in a first-line setting
Localised disease
Watchful waiting Offer WW to asymptomatic patients not eligible for local curative Strong
treatment and those with a short life expectancy.
Locally-advanced disease
Watchful waiting Offer a deferred treatment policy using ADT monotherapy to M0 Weak
asymptomatic patients with a PSA-DT > 12 months, a PSA < 50 ng/mL
and well-differentiated tumour who are unwilling or unable to receive
any form of local treatment.
Persistent PSA after radical prostatectomy
Offer a prostate-specific membrane antigen positron-emission Weak
tomography (PSMA PET) scan to men with a persistent PSA
> 0.2 ng/mL if the results will influence subsequent treatment
decisions
Treat men with no evidence of metastatic disease with salvage RT and Weak
additional hormonal therapy.

PROSTATE CANCER - LIMITED UPDATE 2021 121


6.6.3 Guidelines for metastatic disease, second-line and palliative treatments

Recommendations Strength rating


Metastatic disease in a first-line setting
M1 patients Offer immediate systemic treatment with ADT to palliate symptoms Strong
and reduce the risk for potentially serious sequelae of advanced
disease (spinal cord compression, pathological fractures, ureteral
obstruction) to M1 symptomatic patients.
Offer luteinising hormone-releasing hormone (LHRH) antagonists, Weak
especially to patients with an impending spinal cord compression or
bladder outlet obstruction.
Offer surgery and/or local radiotherapy to any patient with M1 disease Strong
and evidence of impending complications such as spinal cord
compression or pathological fracture.
Offer immediate systemic treatment to M1 patients asymptomatic Weak
from their tumour.
Discuss deferred ADT with well-informed M1 patients asymptomatic Weak
from their tumour since it lowers the treatment-related side effects,
provided the patient is closely monitored.
Offer short-term administration of an older generation androgen Weak
receptor (AR) antagonist to M1 patients starting LHRH agonist to
reduce the risk of the ‘flare-up’ phenomenon.
Do not offer AR antagonists monotherapy to patients with M1 disease. Strong
Discuss combination therapy including ADT plus systemic therapy Strong
with all M1 patients.
Do not offer ADT monotherapy to patients whose first presentation is Strong
M1 disease if they have no contraindications for combination therapy
and have a sufficient life expectancy to benefit from combination
therapy and are willing to accept the increased risk of side effects.
Offer ADT combined with chemotherapy (docetaxel) to patients whose Strong
first presentation is M1 disease and who are fit for docetaxel.
Offer ADT combined with abiraterone acetate plus prednisone or Strong
apalutamide or enzalutamide to patients whose first presentation is
M1 disease and who are fit for the regimen.
Offer ADT combined with prostate radiotherapy (using the doses Strong
from the STAMPEDE study) to patients whose first presentation is M1
disease and who have low volume of disease by CHAARTED criteria.
Do not offer ADT combined with any local treatment (radiotherapy/ Strong
surgery) to patients with high-volume M1 disease (CHAARTED criteria)
outside of clinical trials (except for symptom control).
Do not offer ADT combined with surgery to M1 patients outside of Strong
clinical trials.
Only offer metastasis-directed therapy to M1 patients within a clinical Strong
trial setting or well-designed prospective cohort study.
Biochemical recurrence after treatment with curative intent
Biochemical Offer monitoring, including PSA, to EAU Low-Risk BCR patients. Weak
recurrence Offer early salvage IMRT plus IGRT to men with two consecutive PSA Strong
after radical rises.
prostatectomy A negative PET/CT scan should not delay salvage radiotherapy (SRT), Strong
(RP) if otherwise indicated.
Do not wait for a PSA threshold before starting treatment. Once the Strong
decision for SRT has been made, SRT (at least 66 Gy) should be given
as soon as possible.
Offer hormonal therapy in addition to SRT to men with biochemical Weak
recurrence (BCR).

122 PROSTATE CANCER - LIMITED UPDATE 2021


Biochemical Offer monitoring, including PSA, to EAU Low-Risk BCR patients. Weak
recurrence after Only offer salvage RP, brachytherapy, HIFU, or cryosurgical ablation Strong
RT to highly selected patients with biopsy proven local recurrence within
a clinical trial setting or well-designed prospective cohort study
undertaken in experienced centres.
Systemic salvage Do not offer ADT to M0 patients with a PSA-DT > 12 months. Strong
treatment
Life-prolonging treatments of castration-resistant disease
Ensure that testosterone levels are confirmed to be < 50 ng/dL, before Strong
diagnosing castration-resistant PCa (CRPC).
Counsel, manage and treat patients with metastatic CRPC (mCRPC) Strong
in a multidisciplinary team.
Treat patients with mCRPC with life-prolonging agents. Strong
Offer mCRPC patients somatic and/or germline molecular testing Strong
as well as testing for mismatch repair deficiencies or microsatellite
instability.
Systemic treatments of castrate-resistant disease
Base the choice of treatment on the performance status (PS),
symptoms, co-morbidities, location and extent of disease, genomic
profile, patient preference, and on the previous treatment for hormone-
sensitive metastatic PCa (mHSPC) (alphabetical order: abiraterone,
cabazitaxel, docetaxel, enzalutamide, olaparib, radium-223,
sipuleucel-T).
Offer patients with mCRPC who are candidates for cytotoxic therapy Strong
and are chemotherapy naïve docetaxel with 75 mg/m2 every 3 weeks.
Offer patients with mCRPC and progression following docetaxel Strong
chemotherapy further life-prolonging treatment options, which include
abiraterone, cabazitaxel, enzalutamide, radium-223 and olaparib in
case of DNA homologous recombination repair (HRR).
Base further treatment decisions of mCRPC on pre-treatment PS Strong
status, previous treatments, symptoms, co-morbidities, genomic
profile, extent of disease and patient preference.
Offer abiraterone or enzalutamide to patients previously treated with Strong
one or two lines of chemotherapy.
Avoid sequencing of androgen receptor targeted agents. Weak
Offer chemotherapy to patients previously treated with abiraterone or Strong
enzalutamide.
Offer cabazitaxel to patients previously treated with docetaxel. Strong
Offer cabazitaxel to patients previously treated with docetaxel and Strong
progressing within 12 months of treatment with abiraterone or
enzalutamide.
Novel agents
Offer poly(ADP-ribose) polymerase (PARP) inhibitors to pre-treated Strong
mCRPC patients with relevant DNA repair gene mutations.
Supportive care of castration-resistant disease
Offer bone protective agents to patients with mCRPC and skeletal Strong
metastases to prevent osseous complications.
Monitor serum calcium and offer calcium and vitamin D Strong
supplementation when prescribing either denosumab or
bisphosphonates.
Treat painful bone metastases early on with palliative measures such Strong
as IMRT plus IGRT and adequate use of analgesics.
In patients with spinal cord compression start immediate high-dose Strong
corticosteroids and assess for spinal surgery followed by irradiation.
Offer radiation therapy alone if surgery is not appropriate.
Non-metastatic castrate-resistant disease
Offer apalutamide, darolutamide or enzalutamide to patients with M0 Strong
CRPC and a high risk of developing metastasis (PSA-DT < 10 months)
to prolong time to metastases and overall survival.

PROSTATE CANCER - LIMITED UPDATE 2021 123


7. FOLLOW-UP
The rationale for following up patients is to assess immediate- and long-term oncological results, ensure
treatment compliance and allow initiation of further therapy, when appropriate. In addition, follow-up allows
monitoring of side effects or complications of therapy, functional outcomes and an opportunity to provide
psychological support to PCa survivors, all of which is covered in Chapter 8.

7.1 Follow-up: After local treatment


7.1.1 Definition
Local treatment is defined as RP or RT, either by IMRT plus IGRT or LDR- or HDR-brachytherapy, or any
combination of these, including neoadjuvant and adjuvant therapy. Unestablished alternative treatments
such as HIFU, cryosurgery and focal therapy options do not have a well-defined, validated, PSA cut-off to
define BCR but follow the general principles as presented in this section. In general, a confirmed rising PSA is
considered a sign of disease recurrence.

7.1.2 Why follow-up?


The first post-treatment clinic visit focuses on detecting treatment-related complications and assist patients
in coping with their new situation apart from providing information on the pathological analysis. Men with PCa
are at increased risk of depression and attention for mental health status is required [1195, 1196]. Tumour or
patient characteristics may prompt changing the follow-up schedule.

7.1.3 How to follow-up?


The procedures indicated at follow-up visits vary according to the clinical situation. A disease-specific history
is mandatory at every follow-up visit and includes psychological aspects, signs of disease progression, and
treatment-related complications. Evaluation of treatment-related complications in the post-treatment period
is highlighted in Sections 6.1.2.4, 6.1.2.4.3, 6.3.9.2, 6.3.10.2.2, 6.3.11.2 and 8.2. The examinations used for
cancer-related follow-up after curative surgery or RT are discussed below.

7.1.3.1 Prostate-specific antigen monitoring


Measurement of PSA is the cornerstone of follow-up after local treatment. While PSA thresholds depend on the
local treatment used, PSA recurrence almost always precedes clinical recurrence [914, 1197]. The key question
is to establish when a PSA rise is clinically significant since not all PSA increases have the same clinical value
(see Section 6.3) [916].

7.1.3.1.1 Active surveillance follow-up


Patients included in an AS programme should be monitored according to the recommendations presented in
Section 6.2.2.

7.1.3.1.2 Prostate-specific antigen monitoring after radical prostatectomy


Following RP, the PSA level is expected to be undetectable (< 0.1 ng/mL). Prostate-specific antigen is generally
determined every 3 months in the first year, every 6 months until 3 years and yearly thereafter but evidence
for a specific interval is low [467] and mainly based on the observation that early recurrences are more likely
to be associated with more rapid progression [916, 1198, 1199]. Prostate-specific antigen level is
expected to be undetectable 2 months after a successful RP [1200]. As mentioned in Section 6.3.2 the
PSA-threshold at relapse best predicting metastases after RP is > 0.4 ng/mL. Persistently measurable PSA in
patients treated with RP is discussed in Section 6.2.6 - Persistent PSA.

Ultrasensitive PSA assays remain controversial for routine follow-up after RP. Men with an ultrasensitive PSA
nadir < 0.01 ng/mL have a high (96%) likelihood of remaining relapse free within 2 years [1201]. In addition,
post-RP ultrasensitive PSA levels > 0.01 ng/mL in combination with clinical characteristics such as ISUP grade
and surgical margin status may predict PSA progression and can be useful to establish follow-up intervals
[1199]. Lastly, PSA and associated PSA-DT [1202] calculated prior to 0.2 ng/mL may help identify suitable
candidates for early intervention [1203]. Prostate-specific antigen monitoring after salvage RT to the prostatic
fossa is done at similar intervals and an early PSA rise predicts more rapid progression [1198].

7.1.3.1.3 Prostate-specific antigen monitoring after radiotherapy


Following RT, PSA drops more slowly as compared to post RP. A nadir < 0.5 ng/mL is associated with a
favourable outcome after RT although the optimal cut-off value remains controversial [1204]. The interval
before reaching the nadir can be up to 3 years, or more. At the 2006 RTOG-ASTRO Consensus Conference
the Phoenix definition of radiation failure was proposed to establish a better correlation between definition

124 PROSTATE CANCER - LIMITED UPDATE 2021


and clinical outcome (mainly metastases), namely, an increase of 2 ng/mL above the post-treatment PSA nadir
[915]. This definition also applies to patients who received HT [915].

7.1.3.1.4 Digital rectal examination


Local recurrence after curative treatment is possible without a concomitant rise in PSA level [1205]. However,
this has only been proven in patients with unfavourable undifferentiated tumours. Prostate-specific antigen
measurement and DRE comprise the most useful combination for first-line examination in follow-up after RT
but the role of DRE was questioned since it failed to detect any local recurrence in the absence of a rising PSA
in a series of 899 patients [1206]. In a series of 1,118 prostatectomy patients, no local histologically proven
recurrence was found by DRE alone and PSA measurement may be the only test needed after RP [1207, 1208].

7.1.3.1.5 Transrectal ultrasound, bone scintigraphy, CT, MRI and PET/CT


Imaging techniques have no place in routine follow-up of localised PCa as long as the PSA is not rising.
Imaging is only justified in patients for whom the findings will affect treatment decisions, either in case of BCR
or in patients with symptoms (see Section 6.3.4 for a more detailed discussion).

7.1.4 How long to follow-up?


Most patients who fail treatment for PCa do so within 7 years after local therapy [482]. Patients should be
followed up more closely during the initial post-treatment period when risk of failure is highest. Prostate-
specific antigen measurement, disease-specific history and DRE (if considered) are recommended every 6
months until 3 years and then annually. Whether follow-up should be stopped if PSA remains undetectable
(after RP) or stable (after RT) remains an unanswered question.

7.1.5 Summary of evidence and guidelines for follow-up after treatment with curative intent

Summary of evidence LE
A rising PSA must be differentiated from a clinically meaningful relapse. 3
The PSA threshold that best predicts further metastases after RP is > 0.4 ng/mL and > NADIR + 2
after IMRT plus IGRT (± ADT].
Palpable nodules combined with increasing serum PSA suggest at least local recurrence. 2a

Recommendations Strength rating


Routinely follow up asymptomatic patients by obtaining at least a disease-specific history Strong
and serum prostate-specific antigen (PSA) measurement. These should be performed at 3,
6 and 12 months after treatment, then every 6 months until 3 years, and then annually.
At recurrence, only perform imaging if the result will affect treatment planning. Strong

7.2 Follow-up: During first line hormonal treatment (androgen sensitive period)
7.2.1 Introduction
Androgen deprivation therapy is used in various situations: combined with radiotherapy for localised or
locally-advanced disease, as monotherapy for a relapse after a local treatment, or in the presence of
metastatic disease often in combination with other treatments. All these situations are based on the benefits
of testosterone suppression either by drugs (LHRH agonists or antagonists) or orchidectomy. Inevitably, the
disease will become castrate-resistant, although ADT will be maintained.
This paragraph addresses the general principles of follow-up of patients on ADT alone. As treatment
of CRPC and follow-up are closely linked, Section 6.5.7 includes further information on other drug treatments.
Furthermore the specific follow-up needed for every single drug is outside the scope of this text.
Regular clinical follow-up is mandatory and cannot be replaced by imaging or laboratory tests
alone. Complementary investigations must be restricted to those that are clinically helpful to avoid unnecessary
examinations and costs.

7.2.2 Purpose of follow-up


The main objectives of follow-up in patients receiving ADT are to ensure treatment compliance, to monitor
treatment response, to detect side effects early, and to guide treatment at the time of CRPC. After the initiation
of ADT, it is recommended that patients are evaluated every 3 to 6 months. This must be individualised and
each patient should be advised to contact his physician in the event of troublesome symptoms.

PROSTATE CANCER - LIMITED UPDATE 2021 125


7.2.3 General follow-up of men on ADT
Patients under ADT require regular follow-up, including monitoring of serum testosterone, creatinine, liver
function and metabolic parameters at 3 to 6 month intervals. Men on ADT can experience toxicity independent
of their disease stage.

7.2.3.1 Testosterone monitoring


Testosterone monitoring should be considered standard clinical practice in men on ADT. Many men receiving
medical castration will achieve a castrate testosterone level (< 20 ng/dL), and most a testosterone level
< 50 ng/dL. However, approximately 13–38% of patients fail to achieve these levels and up to 24% of men may
experience temporary testosterone surges (testosterone > 50 ng/dL) during long-term treatment [1200] referred
to as ‘acute on-chronic effect’ or ‘breakthrough response’ [1209]. Breakthrough rates for the < 20 ng/dL
threshold were found to be more frequent (41.3%) and an association with worse clinical outcomes was
suggested [1209].
The timing of measurements is not clearly defined. A 3 to 6-month testosterone level assessment
has been suggested to ensure castration is achieved (especially during medical castration) and maintained.
In case the castrate testosterone level is not reached, switching to another agonist or antagonist or to an
orchiectomy should be considered. In patients with a confirmed rising PSA and/or clinical progression,
serum testosterone must be evaluated in all cases to confirm a castration-resistant state. Ideally, suboptimal
testosterone castrate levels should be confirmed with mass spectrometry or an immunoassay [1210, 1211].

7.2.3.2 Liver function monitoring


Liver function tests will detect treatment toxicity (especially applicable for NSAA), but rarely indicate disease
progression. Men on combined ADT should have their transaminase levels checked at least twice a year in view
of potential liver toxicity but a more frequent check is needed with some drugs (like abiraterone acetate). Alkaline
phosphatase may increase secondary to bone metastases and androgen-induced osteoporosis, therefore it may
be helpful to determine bone-specific isoenzymes as none are directly influenced by ADT [1212].

7.2.3.3 Serum creatinine and haemoglobine


Estimated glomerular filtration rate monitoring is good clinical practice as an increase may be linked to ureteral
obstruction or bladder retention. A decline in Hb is a known side effect of ADT. A significant decline after 3
months of ADT is independently associated with shorter progression-free and OS rates and might explain
significant fatigue although other causes should be considered [1213]. Anaemia is often multi-factorial and
other possible aetiologies should be excluded.

7.2.3.4 Monitoring of metabolic complications


The most severe complications of androgen suppression are metabolic syndrome, cardiovascular morbidity,
mental health problems, and bone resorption (see Section 8.2.4.5).

All patients should be screened for diabetes by checking fasting glucose and HbA1c (at baseline and routinely)
in addition to checking blood lipid levels. Men with impaired glucose tolerance and/or diabetes should be
referred for an endocrine consultation. Prior to starting ADT a cardiology consultation should be considered in
men with a history of cardiovascular disease and in men older than 65 years. Men on ADT are at increased risk
of cardiovascular problems and hypertension and regular checks are required [1214].

7.2.3.5 Monitoring bone problems


Androgen deprivation therapy increases the risk of osteoporosis. Adminstration of ADT for more than a year, as
compared to less than one year, showed a higher risk of osteoperosis (HR: 1.77 and 1.38, respectively) [1215].
Several scores (e.g., Fracture Risk Assessment Tool [FRAX score], Osteoporosis Self-Assessment Tool [OST],
Osteoporosis Risk Assessment Instrument [ORAI], Osteoporosis Index of Risk [OSIRIS], Osteoporosis Risk
Estimation [SCORE]) can help identify men at risk of osteoporotic complications but validation of these scores
in the ADT setting is required (see Section 8.3.2.2) [1216-1218].
Routine bone monitoring for osteoporosis should be performed using DEXA scan [1219-1221].
Presence of osteoporosis should prompt the use of bone protective agents. The criteria for initiation of bone
protective agents are mentioned in Section 8.3.2.2. If no bone protective agents are given, a DEXA scan should
be done regularly, at least every 2 years [1222]
A review summarising the incidence of bone fractures showed an almost doubling of the risk of
fractures when using ADT depending on patients’ age and duration and type of ADT with the highest incidence in
older men and men on additional novel ARTA medication across the entire spectrum of disease [1223]. In case of
an osteoporotic fracture a bone protective agent is mandatory. Vitamin D and calcium levels should be regularly
monitored when patients receive ADT and patients should be supplemented if needed. (see Section 6.5.15).

126 PROSTATE CANCER - LIMITED UPDATE 2021


7.2.3.6 Monitoring lifestyle and cognition
Lifestyle (e.g., diet, exercise, smoking status, etc.) affects QoL and potentially outcome [1212, 1213]. During
follow-up men should be counselled on the beneficial effects of exercise to avoid ADT-related toxicity [1224].
Androgen deprivation therapy may affect mental and cognitive health and men on ADT are three times more
likely to report depression [1225]. Attention to mental health should therefore be an integral part of the follow-
up scheme.

7.2.4 Methods of follow-up in men on ADT without metastases


7.2.4.1 Prostate-specific antigen monitoring
Prostate-specific antigen is a key marker for following the course of androgen-sensitive non-metastasised PCa.
Imaging should be considered when PSA is rising > 2 ng/mL or in case of symptoms suggestive of metastasis.

7.2.4.2 Imaging
In general, asymptomatic patients with a stable PSA level do not require further imaging, although care
needs to be taken in patients with aggressive variants when PSA levels may not reflect tumour progression
[1226]. New bone pain requires at least targeted imaging and potentially a bone scan. When PSA progression
suggests CRPC status and treatment modification is considered, imaging, by means of a bone and CT scan, is
recommended for restaging. Detection of metastases greatly depends on imaging (see Section 6.3.4).

7.2.5 Methods of follow-up in men under ADT for metastatic hormone-sensitive PCa
In metastatic patients it is of the utmost importance to counsel about early signs of spinal cord compression,
urinary tract complications (ureteral obstruction, bladder outlet obstruction) or bone lesions that are at an
increased fracture risk. The intervals for follow-up in M1 patients should be guided by patients’ complaints
and can vary. Since most men will receive another anti-cancer therapy combined with ADT such as ARTA,
chemotherapy or local radiotherapy, follow-up frequency should also be dependent on the treatment modality.

7.2.5.1 PSA monitoring


In men on ADT alone, a PSA decline to < 4 ng/mL suggests a likely prolonged response and follow-up visits
may be scheduled every 3 to 6 months provided the patient is asymptomatic or clinically improving. Depending
on symptoms and risk assessment, more frequent visits may be indicated. Treatment response may be
evaluated based on a change in serum PSA level [1049, 1050] and bone- and CT scan although there is no
consensus about how frequently these should be performed [1176]. A rise in PSA level usually precedes the
onset of clinical symptoms by several months. A rising PSA should prompt assessment of testosterone level,
which is mandatory to define CRPC status, as well as restaging using imaging. However it is now recognised
that a stable PSA during ADT is not enough to characterise a non-progressive situation [1227].

7.2.5.2 Imaging as a marker of response in metastatic PCa


Treatment response in soft-tissue metastases can be assessed by morphological imaging methods using the
Response Evaluation Criteria in Solid Tumours (RECIST) criteria. However, these criteria cannot be used in
bone where response assessment is difficult [1228, 1229].
When bone scan is used to follow bone metastases, a quantitative estimation of tracer uptake at
bone scan can be obtained through automated methods such as the Bone Scan Index [1230]. Nonetheless,
bone scan is limited by the so-called ‘flare’ phenomenon which is defined by the development of new images
induced by treatment on a first follow-up scan which, after longer observation, actually represent a favourable
response. Flare is observed within 8 to 12 weeks of treatment initiation and can lead to a false-positive
diagnosis of disease progression. Computed tomography cannot be used to monitor sclerotic bone lesions
because bone sclerosis can occur under effective treatment and reflects bone healing. Magnetic resonance
imaging can directly assess the bone marrow and demonstrate progression based on morphologic criteria or
changes in apparent diffusion coefficient. A standardisation for reporting is available [1231]. The ability of PET/
CT to assess response has been evaluated in a few studies. Until further data are available, MRI and PET/CT
should not be used outside trials for treatment monitoring in metastatic patients [1232].
Men with metastasised PCa on ADT should also in the absence of PSA rise be followed up with
regular imaging since twenty-five percent of men on ADT in the ECOG3805 CHAARTED trial developed
progression on imaging without a PSA rise [1227].

PROSTATE CANCER - LIMITED UPDATE 2021 127


7.2.6 Guidelines for follow-up during hormonal treatment

Recommendations Strength rating


The follow-up strategy must be individualised based on stage of disease, prior symptoms, Strong
prognostic factors and the treatment given.
In patients with stage M0 disease, schedule follow-up at least every 6 months. As a Strong
minimum requirement, include a disease-specific history, serum prostate-specific antigen
(PSA) determination, as well as liver and renal function in the diagnostic work-up.
In M1 patients, schedule follow-up at least every 3 to 6 months. Strong
In patients on long-term androgen deprivation therapy (ADT), measure initial bone mineral Strong
density to assess fracture risk.
During follow-up of patients receiving ADT, check PSA and testosterone levels and monitor Strong
patients for symptoms associated with metabolic syndrome as a side effect of ADT.
As a minimum requirement, include a disease-specific history, haemoglobin, serum Strong
creatinine, alkaline phosphatase, lipid profiles and HbA1c level measurements.
Counsel patients (especially with M1b status) about the clinical signs suggestive of spinal Strong
cord compression.
When disease progression is suspected, restaging is needed and the subsequent follow-up Strong
adapted/individualised.
In M1 patients perform regular imaging (CT and bone scan) even without PSA progression. Weak
In patients with suspected progression, assess the testosterone level. By definition, Strong
castration-resistant PCa requires a testosterone level < 50 ng/dL (< 1.7 nM/L).

8. QUALITY OF LIFE OUTCOMES IN PROSTATE


CANCER
This chapter is presented in two parts. The first (8.2) will summarise long-term consequences (> 12 months)
of therapies for PCa. Based on two systematic reviews, the second (8.3) provides evidence-based
recommendations for supporting patients when selecting primary treatment options for localised disease and
also supportive interventions aimed at improving disease-specific QoL across all stages of disease.

8.1 Introduction
Quality of life and personalised care go hand in hand. Treating PCa can affect an individual both physically and
mentally, as well as close relations and work or vocation. These multifaceted issues all have a bearing on an
individual‘s perception of QoL [1233]. Approaching care from a holistic point of view requires the intervention
of a multi-disciplinary team including urologists, medical oncologists, radiation oncologists, oncology nurses,
behavioural practitioners and many others including fellow patients. Attention to the psychosocial concerns of
people with PCa is integral to quality clinical care, and this can include the needs of carers and partners [1234].
Prostate cancer care should not be reduced to focusing on the organ in isolation: side effects or late adverse
effects of treatment can manifest systemically and have a major influence on the patient’s QoL. Taking QoL into
consideration relies on understanding the patient’s values and preferences so that optimal treatment proposals
can be formulated and discussed.

8.2 Adverse effects of PCa therapies


8.2.1 Surgery
The absence of standardisation in reporting surgical complications for RP and the introduction of different
techniques has resulted in a wide variation in the types of complications reported, as well as variation in the
overall incidence of complications [1235-1238]. The most common post-operative issue is ED but other related
issues to consider include dry ejaculation, which occurs with removal of the prostate, change in the quality
of orgasm and occasional pain on orgasm. Men also complain of loss of penile length (3.73%, 19/510 men)
[1239]. The second most commonly occurring complication is long-term incontinence [1235-1238] but voiding
difficulties may also occur associated with bladder neck contracture (e.g., 1.1% after RALP) [1240].
For those undergoing minimally invasive procedures port site hernia has been reported in 0.66%
after inserting 12 mm bladeless trocar [1241] and can occur more rarely with 8 mm and 5 mm trocars [1241].
A key consideration is whether long-term consequences of surgery are reduced by using newer techniques

128 PROSTATE CANCER - LIMITED UPDATE 2021


such as RALP. Systematic reviews have documented complication rates after RALP [500, 591-594], and can
be compared with contemporaneous reports after RRP [595]. From these reports, the mean continence rates at
12 months were 89–100% for patients treated with RALP and 80–97% for patients treated with RRP. A
prospective controlled non-randomised trial of patients undergoing RP in 14 centres using RALP or RRP
demonstrated that at 12 months after RALP, 21.3% were incontinent, as were 20.2% after RRP. The unadjusted
OR was 1.08 (95% CI: 0.87–1.34). Erectile dysfunction was observed in 70.4% after RALP and 74.7% after
RRP. The unadjusted OR was 0.81 (95% CI: 0.66–0.98) [596, 1242]. Further follow-up demonstrates similar
functional outcomes with both techniques at 24 months [1242, 1243]. A single-centre randomised phase III
study comparing RALP and RRP (n = 326) also demonstrates similar functional outcomes with both techniques
at 24 months [495].

8.2.2 Radiotherapy
8.2.2.1 Side-effects of external beam radiotherapy
Analysis of the toxicity outcomes of the Prostate Testing for Cancer and Treatment (ProtecT) trial [1244] shows
that patients treated with EBRT and 6 months of ADT report bowel toxicity including persistent diarrhoea,
bowel urgency and/or incontinence and rectal bleeding (described in detail in Section 8.3.1.1 below).
Participants in the ProtecT study were treated with 3D-CRT and more recent studies using IMRT demonstrate
less bowel toxicity than noted previously with 3D-CRT [1245].
A systematic review and meta-analysis of observational studies comparing patients exposed or
unexposed to radiotherapy in the course of treatment for PCa demonstrates an increased risk of developing
second cancers for bladder (OR: 1.39), colorectal (OR: 1.68) and rectum (OR: 1.62) with similar risks over lag
times of 5 and 10 years. Absolute excess risks over 10 years are small (1–4%) but should be discussed with
younger patients in particular [1246].

8.2.2.2 Side effects from brachytherapy


Some patients experience significant urinary complications following implantation such as urinary retention
(1.5-22%), with post-implantation TURP reported as being required in up to 8.7% of cases, and incontinence
(0–19%) [1247]. Chronic urinary morbidity is more common with combined EBRT and BT and can occur in
up to 20% of patients, depending on the severity of the symptoms before brachytherapy. Urethral strictures
account for at least 50% of urinary complications and can be resolved with dilation in the majority [822, 824].
Prevention of morbidity depends on careful patient selection, and expert assessment of IPSS score, backed up
by urodynamic studies.

8.2.3 Local primary whole-gland treatments other than surgery or radiotherapy


8.2.3.1 Cryosurgery
In Ramsay et al.’s systematic review and meta-analysis there was evidence that the rate of urinary incontinence
at one year was lower for cryotherapy than for RP, but the size of the difference decreased with longer follow-
up [758]. There was no significant difference between cryotherapy vs. EBRT in terms of urinary incontinence
at one year (< 1%); cryotherapy had a similar ED rate (range 0–40%) to RP at one year. There were insufficient
data to compare cryotherapy vs. EBRT in terms of ED.

8.2.3.2 High-intensity focused ultrasound


In terms of toxicity there are insufficient data on urinary incontinence, ED or bowel dysfunction to draw any
conclusions, although at one year HIFU had lower incontinence rates than RP (OR: 0.06, 95% CI: 0.01–0.48) [758].

8.2.4 Hormonal therapy


A summary of impacts on psychological factors due to the use of ADT such as sexual function, mood,
depression, cognitive function and impact on partners can be found in two clinical reviews [1248, 1249].
A small RCT evaluated the QoL at one-year follow-up in patients with non-localised PCa, between
various ADT regimens, or no treatment. Patients treated by ADT reported a significant decline in spatial
reasoning, spatial abilities and working memory as well as increased depression, tension, anxiety, fatigue and
irritability during treatment [1250]. Conversely, a prospective observational study with follow-up to 3 years
failed to demonstrate an association with cognitive decline in men on ADT when compared to men with PCa
not treated with ADT and healthy controls [1251]. A prospective observational study of non-metastatic PCa
found that immediate ADT was associated with a lower overall QoL compared to deferred treatment [1252].
Another retrospective non-randomised study suggested that men receiving LHRH agonists reported more
worry and physical discomfort and poorer overall health, and were less likely to believe themselves free of
cancer than patients undergoing orchiectomy. The stage at diagnosis had no effect on health outcomes [1253].
Using a specific non-validated questionnaire, bicalutamide monotherapy showed a significant
advantage over castration in the domains of physical capacity and sexual interest (not sexual function) at

PROSTATE CANCER - LIMITED UPDATE 2021 129


12 months [1254]. A post-hoc analysis, including only patients with sexual interest suggested that bicalutamide
was associated with better sexual preservation, including maintained sexual interest, feeling sexually attractive
[1255], preserved libido and erectile function [1256]. Intermittent androgen deprivation has been discussed
elsewhere (see Section 6.4.3.2).

8.2.4.1 Sexual function


Cessation of sexual activity is very common in people undergoing ADT, affecting up to 93% [1257]. Androgen
deprivation therapy reduces both libido and the ability to gain and maintain erections. The management of
acquired ED is mostly non-specific [1258].

8.2.4.2 Hot flushes


Hot flushes are a common side-effect of ADT (prevalence estimated between 44–80% of men on ADT)
[1257]. They appear 3 months after starting ADT, usually persist long-term and have a significant impact on
QoL. Oestrogen-receptor modulators or low-dose oestrogen therapies, e.g. DES, 0.5–1 mg/day, reduce the
frequency and severity of hot flushes. Both treatments carry a risk of cardiovascular complications [1259].

Serotonin re-uptake inhibitors (e.g., venlafaxine or sertraline) also appear to be effective in men but
less than hormone therapies based on a prospective RCT comparing venlafaxine, 75 mg daily, with
medroxyprogesterone, 20 mg daily, or cyproterone acetate, 100 mg daily [1260]. After 6 months of LHRH
(n = 919), 311 men had significant hot flushes and were randomised to one of the treatments. Based on median
daily hot-flush score, venlafaxine was inferior -47.2% (interquartile range -74.3 to -2.5) compared to -94.5%
(-100.0 to -74.5) in the cyproterone group, and -83.7% (-98.9 to -64.3) in the medroxyprogesterone group. With
a placebo effect influencing up to 30% of patients [1261], the efficacy of clonidine, veralipride, gabapentine
[1262] and acupuncture [1263] need to be compared in prospective RCTs.

8.2.4.3 Non-metastatic bone fractures


Due to increased bone turnover and decreased bone mineral density (BMD) in a time-dependent manner, ADT
use is linked to an increased risk of fracture (up to 45% RR with long-term ADT) [1264]. Severe fractures in men
are associated with a significant risk of death [1265]. A precise evaluation of BMD should be performed by dual
emission X-ray absorptiometry (DEXA), ideally before or shortly after starting long-term ADT. An initial low BMD
(T-score < -2.5 or < -1, with other risk factors) indicates a high risk of subsequent non-metastatic fracture and
causes should be investigated. Other risk factors include increasing age, body mass index of 19 or less, history
of previous fracture or parent with fractured hip, current smoking, use of glucocorticoids, rheumatoid arthritis,
alcohol consumption > 2 units per day, history of falls and a number of other chronic medical conditions [1266].
Fracture risk alorithms which combine BMD and clinical risk factors such as FRAX score can be used to guide
treatment decisions but uncertainty exists regarding the optimal intervention threshold, therefore no specific
risk algorithm can be recommended for men on ADT for PCa. Obesity (increase in body fat mass by up to
10%) and sarcopenia (decrease in lean tissue mass by up to 3%) as well as weight loss are common and occur
during the first year of ADT [1267]. These changes increase the fracture risk [1268].

Bicalutamide monotherapy may have less impact on BMD but is limited by its suboptimal efficacy [1269, 1270]
(see Section 6.1.4.1.1.5.2.3). The intermittent LHRH-agonist modality might be associated with less bone
impact [1271].

8.2.4.4 Metabolic effects


Lipid alterations are common and may occur as early as the first 3 months of treatment [1267]. Androgen
deprivation therapy also decreases insulin sensitivity and increases fasting plasma insulin levels, which is a
marker of insulin resistance. In diabetic patients, metformin appears to be an attractive option for protection
against metabolic effects based on retrospective analysis [1272], but there is insufficient data to recommend its
use in non-diabetic patients.

Metabolic syndrome is an association of independent cardiovascular disease risk factors, often associated with
insulin resistance. The definition requires at least three of the following criteria [1273]:
• waist circumference > 102 cm;
• serum triglyceride > 1.7 mmol/L;
• blood pressure > 130/80 mmHg or use of medication for hypertension;
• high-density lipoprotein (HDL) cholesterol < 1 mmol/L;
• glycaemia > 5.6 mmol/L or the use of medication for hyperglycaemia.

The prevalence of a metabolic-like syndrome is higher during ADT compared with men not receiving ADT [1274].

130 PROSTATE CANCER - LIMITED UPDATE 2021


Skeletal muscle mass heavily influences basal metabolic rate and is in turn heavily influenced by endocrine
pathways [1275]. Androgen deprivation therapy-induced hypogonadism results in negative effects on skeletal
muscle health. A prospective longitudinal study involving 252 men on ADT for a median of 20.4 months
reported lean body mass decreases progressively over 3 years; 1.0% at one year, 2.1% at 2 years, and 2.4% at
3 years which appears more pronounced in men at > 70 years of age [1276].

8.2.4.5 Cardiovascular morbidity


Cardiovascular mortality is a common cause of death in PCa patients [1040, 1277, 1278]. Several
studies showed that ADT after only 6 months was associated with an increased risk of diabetes mellitus,
cardiovascular disease, and myocardial infarction [1279]. The RTOG 92-02 [1280] and 94-08 [1281] trials
confirmed an increased cardiovascular risk, unrelated to the duration of ADT and not accompanied by an
overall increased cardiovascular mortality. No increase in cardiovascular mortality has been reported in a
systematic meta-analysis of trials RTOG 8531, 8610, 9202, EORTC 30891 or EORTC 22863 [1282]. However,
serious concerns about the conclusions of this meta-analysis have been raised due to poor consideration
of bias in the included studies [1283, 1284]. Meta-analysis of observational data reports consistent links
between ADT and the risk of cardiovascular disease patients treated for PCa e.g. the associations between
GnRH agonists and nonfatal or fatal myocardial infarction or stroke RR: 1.57 (95% CI: 1.26–1.94) and RR: 1.51
(95% CI: 1.24–1.84), respectively [1285]. An increase in cardiovascular mortality has been reported in patients
suffering from previous congestive heart failure or myocardial infarction in a retrospective database analysis
[1286] or presenting with a metabolic syndrome [1287]. It has been suggested that LHRH antagonists might
be associated with less cardiovascular morbidity compared to agonists [1288]. In a phase III RCT the use of
relugolix, an oral LHRH antagonist, was associated with a reduced risk of major adverse cardiovascular events
when compared to leuprolide, an injectable LHRH agonists, at 2.9% vs. 6.2%, respectively, over a follow-up
time of of 48 weeks (HR 0.46, 95% CI: 0.24–0.88 [727].
These concerns resulted in an FDA warning and consensus paper from the American Heart, Cancer
Society and Urological Associations [1039]. Preventive advice includes non-specific measures such as loss of
weight, increased exercise, minimising alcohol intake, improved nutrition and smoking cessation [69, 1289].

8.2.4.6 Fatigue
Fatigue often develops as a side-effect of ADT. Regular exercise appears to be the best protective measure.
Anaemia may be a cause of fatigue [1257, 1290]. Anaemia requires an aetiological diagnosis (medullar invasion,
renal insufficiency, iron deficiency, chronic bleeding) and individualised treatment. Iron supplementation (using
injectable formulations only) must be systematic if deficiency is observed. Regular blood transfusions may be
required in patients with severe anaemia. Erythropoiesis-stimulating agents might be considered in dedicated
cases, taking into account the possible increased risk of thrombovascular events [1291].

8.2.4.7 Neurological side effects


Castration seems also to be associated with an increased risk of stroke [1292], and is suspect to be associated
with an increased risk for depression and cognitive decline such as Alzheimer disease [1293].

8.3 Overall quality of life in men with PCa


Living longer with PCa does not necessarily equate to living well [1233, 1234]. There is clear evidence of unmet
needs and ongoing support requirements for some individuals after diagnosis and treatment for PCa [1294].
Cancer impacts on the wider family and cognitive behavioural therapy can help reduce depression, anxiety
and stress in caregivers [1295]. Radical treatment for PCa can negatively impact long-term QoL (e.g., sexual,
urinary and bowel dysfunction) as can ADT used in short- or long-term treatment, e.g., sexual problems,
fatigue, psychological morbidity, adverse metabolic sequelae and increased cardiovascular and bone fracture
risk [1248, 1296]. Direct symptoms from advanced or metastatic cancer, e.g., pain, hypercalcaemia, spinal
cord compression and pathological fractures, also adversely affect health [1297, 1298]. Patient’s QoL including
domains such as sexual function, urinary function and bowel function is worse after treatment for PCa
compared to non-cancer controls [1299, 1300].

The concept of ‘quality of life’ is subjective and can mean different things to different people, but there are
some generally common features across virtually all patients. Drawing from these common features, specific
tools or ‘patient-reported outcome measures’ (PROMs) have been developed and validated for men with PCa.
These questionnaires assess common issues after PCa diagnosis and treatment and generate scores which
reflect the impact on perceptions of HRQoL. During the process of undertaking two dedicated systematic
reviews around cancer-specific QoL outcomes in patients with PCa as the foundation for our guideline
recommendations, the following validated PROMs were found in our searches (see Table 8.3.1).

PROSTATE CANCER - LIMITED UPDATE 2021 131


Table 8.3.1: PROMs assessing cancer specific quality of life

Questionnaire Domains/items
Functional Assessment of Cancer Therapy-General Physical well-being, social/family well-being,
(FACT-G) [1301] emotional well-being, and functional well-being.
Functional Assessment of Cancer Therapy-Prostate 12 cancer site specific items to assess for
(FACT-P) [1302] prostate-related symptoms. Can be combined with
FACT-G or reported separately.
European Organisation for Research and Treatment of Five functional scales (physical, role, cognitive,
Cancer QLQ-C30 (EORTC QLQ-C30) [1303] emotional, and social); three symptom scales
(fatigue, pain, and nausea and vomiting); global
health status/QoL scale; and a number of single
items assessing additional symptoms commonly
reported by cancer patients (dyspnoea, loss of
appetite, insomnia, constipation and diarrhoea)
and perceived financial impact of the disease.
European Organisation for Research and Treatment of Urinary, bowel and treatment-related symptoms,
Cancer QLQ-PR 25 (EORTC QLQ-PR 25) [1304] as well as sexual activity and sexual function.
Expanded prostate cancer index composite (EPIC) [1305] Urinary, bowel, sexual, and hormonal symptoms.
Expanded prostate cancer index composite short form Urinary, sexual, bowel, and hormonal domains.
26 (EPIC 26) [1306]
UCLA Prostate Cancer Index (UCLA PCI) [1307] Urinary, bowel, and sexual domains.
Prostate Cancer Quality of Life Instrument (PCQoL) Urinary, sexual, and bowel domains, supplemented
[1308] by a scale assessing anxiety.
Prostate Cancer Outcome Study Instrument [1309] Urinary, bowel, and sexual domains.

8.3.1 Long-term (> 12 months) quality of life outcomes in men with localised disease
8.3.1.1 Men undergoing local treatments
The results of the Prostate Testing for Cancer and Treatment (ProtecT) trial (n = 1,643 men) reported no
difference in EORTC QLQ-C30 assessed global QoL, up to 5 years of follow-up in men aged 50–69 years
with T1-T2 disease randomised for treatment with AM, RP or RT with 6 months of ADT [1244]. However, EPIC
urinary summary scores (at 6 years) were worse in men treated with RP compared to AM or RT (88.7 vs. 89.0
vs. 91.4, respectively) as were urinary incontinence (80.9 vs. 85.8 vs. 89.4, respectively) and sexual summary,
function and bother scores (32.3 vs. 40.6 vs. 41.3 for sexual summary, 23.7 vs. 32.5 vs. 32.7 for sexual function
and 51.4 vs. 57.9 vs. 60.1 for sexual bother, respectively) at 6 years of follow-up. Minimal clinically important
differences for the 50 item EPIC questionnaire are not available. For men receiving RT with 6 months of ADT,
EPIC bowel scores were poorer compared to AM and RP in all domains: function (90.8 vs. 92.3 vs. 92.3,
respectively), bother (91.7 vs. 94.2 vs. 93.7, respectively) and summary (91.2 vs. 93.2 vs. 93.0, respectively) at
6 years of follow-up in the ProtecT trial.

The findings regarding RP and RT are supported by other observational studies [1238, 1310]. The Prostate
Cancer Outcomes Study (PCOS) [1238] studied a cohort of 1,655 men, of whom 1,164 had undergone RP
and 491 RT. The study reported that at 5 years of follow-up, men who underwent RP had a higher prevalence
of urinary incontinence and ED, while men treated with RT had a higher prevalence of bowel dysfunction.
However, despite these differences detected at 5 years, there were no significant differences in the adjusted
odds of urinary incontinence, bowel dysfunction or ED between RP and RT at 15 years. More recently,
investigators reported that although EBRT was associated with a negative effect in bowel function, the
difference in bowel domain score was below the threshold for clinical significance 12 months after treatment
[1245]. As 81% of patients in the EBRT arm of the study received IMRT, these data suggest that the risk of
side effects is reduced with IMRT compared to older 3D-CRT techniques. This is supported by a contemporary
5-year prospective, population-based cohort study where PROs were compared in men with favourable- and
unfavourable-risk localised disease [1310]. In the 1,386 men with favourable risk, comparison between AS
and nerve-sparing prostatectomy, EBRT or LDR brachytherapy demonstrates that surgery is associated with
worse urinary incontinence at 5 years and sexual dysfunction at 3 years when compared to AS. External beam
RT is associated with changes not clinically different from AS, and LDR brachytherapy is associated with
worse irritative urinary-, bowel- and sexual symptoms at one year. In 619 men with unfavourable risk disease,
comparison between non-nerve sparing RP and EBRT with ADT demonstrates that surgery is associated with
worse urinary incontinence and sexual function through 5 years.

132 PROSTATE CANCER - LIMITED UPDATE 2021


With respect to brachytherapy cancer-specific QoL outcomes, one small RCT (n = 200) evaluated bilateral
nerve-sparing RP and brachytherapy in men with localised disease (up to T2a), which reported worsening of
physical functioning as well as irritative urinary symptomatology in 20% of brachytherapy patients at one year
of follow-up. However, there were no significant differences in EORTC QLQ-C30/PR-25 scores at 5 years of
follow-up when compared to pre-treatment values [1311]. It should be noted of this trial, within group tests
only were reported. In a subsequent study by the same group comparing bilateral nerve-sparing RARP and
brachytherapy (n = 165), improved continence was noted with brachytherapy in the first 6 months but lower
potency rates up to 2 years [1312]. These data and a synthesis of 18 randomised and non-randomised studies
in a systematic review involving 13,604 patients are the foundation of the following recommendations [1313].

8.3.1.2 Guidelines for quality of life in men undergoing local treatments

Recommendations Strength rating


Advise eligible patients for active surveillance that global quality of life is equivalent for up to Strong
5 years compared to radical prostatectomy or external beam radiotherapy.
Discuss the negative impact of surgery on urinary and sexual function, as well as the Strong
negative impact of radiotherapy on bowel function with patients.
Advise patients treated with brachytherapy of the negative impact on irritative urinary Weak
symptomatology at one year but not after 5 years.

8.3.2 Improving quality of life in men who have been diagnosed with PCa
8.3.2.1 Men undergoing local treatments
In men with localised disease, nurse-led multi-disciplinary rehabilitation (addressing sexual functioning, cancer
worry, relationship issues depression, managing bowel and urinary function problems) provided positive short-
term effects (4 months) on sexual function (effect size 0.45) and long-term (12 months) positive effects on
sexual limitation (effect size 0.5) and cancer worry (effect size 0.51) [1314].
In men with post-surgical urinary incontinence, conservative management options include pelvic
floor muscle training with or without biofeedback, electrical stimulation, extra-corporeal magnetic innervation
(ExMI), compression devices (penile clamps), lifestyle changes, or a combination of methods. Uncertainty
around the effectiveness and value of these conservative interventions remains [1315]. Surgical interventions
including sling and artificial urinary sphincter significantly decrease the number of pads used per day and
increase the QoL compared with before intervention. The overall cure rate is around 60% and results in
improvement in incontinence by about 25% [1316].
The use of PDE5 inhibitors in penile rehabilitation has been subject to some debate. A single
centre, double blind RCT of 100 men undergoing nerve-sparing surgery reported no benefit of nightly sildenafil
(50 mg) compared to on-demand use [1317]. However, a multi-centre double blind RCT (n = 423) in men aged
< 68 years, with normal pre-treatment erectile function undergoing either open, conventional or robot-assisted
laparoscopic nerve-sparing RP, tadalafil (5 mg) once per day improved participants EPIC sexual domain-scores
(least squares mean difference +9.6, 95% CI: 3.1–16.0) when compared to 20 mg ‘on demand’ or placebo at
9 months of follow-up [622]. Therefore, based on discordant results, no clear recommendation is possible, even
if a trend exists for early use of PDE5 inhibitors after RP for penile rehabilitation [1318]. A detailed discussion
can be found in the EAU Sexual and Reproductive Health Guidelines [1319].

8.3.2.2 Men undergoing systemic treatments


Similar to men treated with a radical approach (see above), in men with T1-T3 disease undergoing RT and
ADT, a combined nurse-led psychological support and physiotherapist-led multi-disciplinary rehabilitation has
reported improvements in QoL. Specifically this intervention involved action planning around patients’ needs
related to lifestyle changes, weight control, toilet habits, sexuality, and psychological problems. This was
complemented with pelvic floor muscle therapy. Improvements in urinary (adjusted mean 4.5, 95% CI: 0.6–8.4),
irritative (adjusted mean 5.8, 95% CI: 1.4–10.3) and hormonal (adjusted mean 4.8, 95% CI: 0.8–8.8) EPIC
domains were found up to 22 weeks of follow-up [1320].
Providing supervised aerobic and resistance exercise training of a moderate intensity improves
EORTC QLQ-C30 role (adjusted mean 15.8, 95% CI: 6.6–24.9) and cognitive domain outcomes (adjusted
mean 11.4, 95% CI: 3.3–19.6) as well as symptom scales for fatigue (adjusted mean 11.0, 95% CI: 20.2–1.7),
nausea (adjusted mean 4.0, 95% CI: 7.4–0.25), and dyspnoea (adjusted mean 12.4, 95% CI: 22.5–2.3) up to 3
months in men treated with ADT [1321]. Such interventions have also reported clinically relevant improvements
in FACT-P (mean difference 8.9, 95% CI: 3.7–14.2) in men on long-term ADT [1322, 1323]. These findings are
supported by a systematic review which reported improvements up to 12 weeks in cancer-specific QoL in a
meta-analysis of high quality trials (SMD 0.33, 95%, CI: 0.08–0.58) [1290].

PROSTATE CANCER - LIMITED UPDATE 2021 133


In case dietary intake is not adequate, vitamin D and calcium supplementation should be offered, as
there is evidence that vitamin D and calcium have modest effects on bone in men on ADT [1324]. Online tools
are available to calculate daily calcium intake for individual patients. For vitamin D deficiency a dose of at least
800 IU/day colecalciferol can be recommended. Use of a 25(OH) assay may be helpful to measure vitamin D
levels [1325, 1326].
Anti-resorptive therapy is recommended for men on ADT for > 6 months with either a BMD T score
of < -2.5 or with an additional risk factor for osteoporosis or annual bone loss confirmed to exceed 5%, or
in cases of severe fracture. Referral to a bone specialist should be considered in complex cases with severe
fracture and/or multiple risk factors. Alendronate, risedronate, zoledronate and denosumab have all been
shown to prevent bone loss in men with locally-advanced PCa on ADT [1222, 1327-1329]. Patients should
be warned about the < 5% risk of osteonecrosis of the jaw and/or atypical femoral fractures associated with
these drugs. Bisphosphonates increase BMD in the hip and spine by up to 7% in one year [1328, 1330]. The
optimal regimen for zoledronic acid remains unclear: quarterly [1331] or yearly [1332] injections. The question
is relevant as the risk of jaw necrosis is both dose- and time-related [1333]. A quarterly regimen could be
considered for a BMD < 2.5 as a yearly injection is unlikely to provide sufficient protection [1334]. Care should
be taken when discontinuing treatment as rebound increased bone resorption can occur.

In M0 patients, denosumab has been shown to increase the lumbar BMD by 5.6% compared to a 1% decrease
in the placebo arm after 2 years, using a 60 mg subcutaneous regimen every 6 months [1335]. This was
associated with a significant decrease in vertebral fracture risk (1.5% vs. 3.9%, p = 0.006). The benefits were
similar whatever the age (< or > 70 years), the duration or type of ADT, the initial BMD, the patient’s weight or
the initial BMI. This benefit was not associated with any significant toxicity, e.g. jaw osteonecrosis or delayed
healing in vertebral fractures. In M0 patients, with the use of a higher dosage (120 mg every 4 weeks), a delay
in bone metastases of 4.2 months has been shown [1187] without any impact on OS, but with an increase in
side effects. Therefore, this later regimen cannot be recommended.

8.3.2.3 Decision regret


Several treatments with curative intent for localised PCa are available all with comparable 10-year OS [467].
They vary in terms of the incidence of major side effects, including urinary symptoms, bowel symptoms and
compromised sexual functioning [1244, 1245, 1336]. For this reason, patients’ treatment preferences, in which
they weigh expected benefits against likely side effects, are a central consideration in shared decision-making
and in making informed treatment decisions [1337-1339].
It remains challenging, however, to evaluate whether the decision-making process can be viewed
as successful; that is, whether the choice of treatment best reflects the patient’s preferences and expectations
[1340, 1341]. According to Decision Justification Theory (DJT), it is the more specific information on which
treatment experiences lead to regret that decision regret needs to be better understood and to minimise it
in future patients [1342]. Maguire et al. found that about 25% of men with PCa undergoing either single or
combined modality treatments report experiencing worse side effects than expected [1343]. Schroeck et al.
found urinary incontinence most strongly correlated with regret after prostatectomy [1344].
Unmet expectations are comparable among the treatment groups, except for fatigue. Fatigue is
less frequently reported as worse than expected by patients who received BT when compared to patients who
received RP or EBRT. This could be explained by the less invasive treatment course of BT in comparison to
EBRT with or without ADT and RP [1345]. Unmet expectations were more frequently reported by patients with
positive surgical margins following surgery; having had a passive role in the decision-making process; and
who had higher scores on the decisional conflict scale (i.e. more uncertainty about the treatment decision).
Interestingly, positive surgical margins are not directly associated with an increased risk of PC-related mortality
[969]. Active participation and support in the process of forming a preference increases the chance of choosing
a treatment that is in line with patients’ expectations [1339, 1346, 1347].
While it may seem desirable to tailor the patients’ role in decision-making to their initial preference,
and particularly to a preference for deferring to the advice of the clinician, this does not result in less decisional
conflict or regret. Increasing patients knowledge regardless of initial preference may actually be preferable
[1344].

8.3.2.4 Decision aids in prostate cancer


Shared decision-making can increase patients’ comfort when confronted with management decisions but has
been shown to improve health outcome [1348] and more training seems needed for health care professionals
guiding patients [1349]. Patient education decreased PSA testing [1350] and increased adherence to active
surveillance protocols [1351, 1352]. Autonomous active decision-making by patients was associated with
less regret after prostatectomy regardless of the method chosen and decision aids reduce decisional conflict
[1353]. Still guidance is needed to optimise patients’ understanding of the options [1354]. Patients prioritised

134 PROSTATE CANCER - LIMITED UPDATE 2021


effectiveness and pain control over mode of administration and risk of fatigue when confronted with treatment
choice in metastasised PCa [1355]. When implementing decision aids clinical validity and utility should
be carefully evaluated and distinguished [1356]. A decision aid should educate as well as promote shared
decision-making to optimise efficacy [1357] and pay attention to communicative aspects [1358].

8.3.2.5 Guidelines for quality of life in men undergoing systemic treatments

Recommendations Strength rating


Offer men on androgen deprivation therapy (ADT), 12 weeks of supervised (by trained Strong
exercise specialists) combined aerobic and resistance exercise.
Advise men on ADT to maintain a healthy weight and diet, to stop smoking, reduce alcohol Strong
to < 2 units daily and have yearly screening for diabetes and hypercholesterolemia. Ensure
that calcium and vitamin D meet recommended levels.
Offer men with T1-T3 disease specialist nurse-led, multi-disciplinary rehabilitation based Strong
on the patients’ personal goals addressing incontinence, sexuality, depression and fear of
recurrence, social support and positive lifestyle changes after any radical treatment.
Offer men starting on long-term ADT dual emission X-ray absorptiometry (DEXA) scanning Strong
to assess bone mineral density.
Offer anti-resorptive therapy to men on long term ADT with either a BMD T-score of < -2.5 Strong
or with an additional clinical risk factor for fracture or annual bone loss on ADT is confirmed
to exceed 5%.

9. REFERENCES
1. Mottet, N., et al. EAU-EANM-ESTRO-ESUR-SIOG Guidelines on Prostate Cancer – 2020 update.
Part 1: Screening, Diagnosis, and Local Treatment with Curative Intent. Eur Urol, 2021. 79: 342.
https://pubmed.ncbi.nlm.nih.gov/33172724/
2. Cornford, P., et al. EAU-EANM-ESTRO-ESUR-SIOG Guidelines on Prostate Cancer. Part II 2020
update: Treatment of Relapsing, Metastatic, and Castration-Resistant Prostate Cancer. Eur Urol,
2021. 79: 263.
https://pubmed.ncbi.nlm.nih.gov/33039206/
3. Guyatt, G.H., et al. What is “quality of evidence” and why is it important to clinicians? BMJ, 2008.
336: 995.
https://pubmed.ncbi.nlm.nih.gov/18456631
4. Guyatt, G.H., et al. GRADE: an emerging consensus on rating quality of evidence and strength of
recommendations. BMJ, 2008. 336: 924.
https://pubmed.ncbi.nlm.nih.gov/18436948
5. Phillips B, et al. Oxford Centre for Evidence-based Medicine Levels of Evidence. 1998. Updated by
Jeremy Howick March 2009 [access data March 2021].
https://www.cebm.net/2009/06/oxford-centre-evidence-based-medicine-levels-evidence-march-2009/
6. Guyatt, G.H., et al. Going from evidence to recommendations. BMJ, 2008. 336: 1049.
https://pubmed.ncbi.nlm.nih.gov/18467413
7. Willemse, P.-P.M., et al. Systematic review of deferred treatment with curative intent for
localised prostate cancer to explore heterogeneity of definitions, thresholds and criteria and
clinical effectiveness. PROSPERO International prospective register of systematic reviews, 2018.
CRD42018071780.
https://www.crd.york.ac.uk/prospero/display_record.php?RecordID=71780
8. Van den Broeck, T., et al. The impact of surgeon and hospital volume on oncological and
non-oncological outcomes in patients undergoing radical prostatectomy for non-metastatic
prostate cancer. PROSPERO International prospective register of systematic reviews, 2020.
CRD42020186466
https://www.crd.york.ac.uk/prospero/display_record.php?ID=CRD42020186466
9. Moris, L., et al. Oncological outcomes of patients who underwent nerve sparing surgery compared
to non-nerve-sparing radical prostatectomy for non-metastatic prostate cancer. PROSPERO
International prospective register of systematic reviews, 2020. CRD42020186493
https://www.crd.york.ac.uk/prospero/display_record.php?ID=CRD42020186493

PROSTATE CANCER - LIMITED UPDATE 2021 135


10. Lardas, M., et al. Patient- and tumor-related prognostic factors for post-operative incontinence after
radical prostatectomy for non-metastatic prostate cancer: A systematic review and meta-analysis.
PROSPERO International prospective register of systematic reviews, 2020 CRD42020186524
https://www.crd.york.ac.uk/prospero/display_record.php?RecordID=186524
11. Ferlay, J., et al. Cancer incidence and mortality worldwide: sources, methods and major patterns in
GLOBOCAN 2012. Int J Cancer, 2015. 136: E359.
https://pubmed.ncbi.nlm.nih.gov/25220842
12. Haas, G.P., et al. The worldwide epidemiology of prostate cancer: perspectives from autopsy
studies. Can J Urol, 2008. 15: 3866.
https://pubmed.ncbi.nlm.nih.gov/18304396
13. Bell, K.J., et al. Prevalence of incidental prostate cancer: A systematic review of autopsy studies. Int
J Cancer, 2015. 137: 1749.
https://pubmed.ncbi.nlm.nih.gov/25821151
14. Fleshner, K., et al. The effect of the USPSTF PSA screening recommendation on prostate cancer
incidence patterns in the USA. Nat Rev Urol, 2017. 14: 26.
https://pubmed.ncbi.nlm.nih.gov/27995937
15. Hemminki, K. Familial risk and familial survival in prostate cancer. World J Urol, 2012. 30: 143.
https://pubmed.ncbi.nlm.nih.gov/22116601
16. Jansson, K.F., et al. Concordance of tumor differentiation among brothers with prostate cancer. Eur
Urol, 2012. 62: 656.
https://pubmed.ncbi.nlm.nih.gov/22386193
17. Randazzo, M., et al. A positive family history as a risk factor for prostate cancer in a population-
based study with organised prostate-specific antigen screening: results of the Swiss European
Randomised Study of Screening for Prostate Cancer (ERSPC, Aarau). BJU Int, 2016. 117: 576.
https://pubmed.ncbi.nlm.nih.gov/26332304
18. Beebe-Dimmer, J.L., et al. Risk of Prostate Cancer Associated With Familial and Hereditary Cancer
Syndromes. J Clin Oncol, 2020. 38: 1807.
https://pubmed.ncbi.nlm.nih.gov/32208047
19. Bratt, O., et al. Family History and Probability of Prostate Cancer, Differentiated by Risk Category: A
Nationwide Population-Based Study. J Natl Cancer Inst, 2016. 108.
https://pubmed.ncbi.nlm.nih.gov/27400876
20. Amin Al Olama, A., et al. Multiple novel prostate cancer susceptibility signals identified by fine-
mapping of known risk loci among Europeans. Hum Mol Genet, 2015. 24: 5589.
https://pubmed.ncbi.nlm.nih.gov/26025378
21. Eeles, R.A., et al. Identification of 23 new prostate cancer susceptibility loci using the iCOGS
custom genotyping array. Nat Genet, 2013. 45: 385.
https://pubmed.ncbi.nlm.nih.gov/23535732
22. Schumacher, F.R., et al. Association analyses of more than 140,000 men identify 63 new prostate
cancer susceptibility loci. Nat Genet, 2018. 50: 928.
https://pubmed.ncbi.nlm.nih.gov/29892016
23. Giri, V.N., et al. Germline genetic testing for inherited prostate cancer in practice: Implications for
genetic testing, precision therapy, and cascade testing. Prostate, 2019. 79: 333.
https://pubmed.ncbi.nlm.nih.gov/30450585
24. Nicolosi, P., et al. Prevalence of Germline Variants in Prostate Cancer and Implications for Current
Genetic Testing Guidelines. JAMA Oncol, 2019. 5: 523.
https://pubmed.ncbi.nlm.nih.gov/30730552
25. Pritchard, C.C., et al. Inherited DNA-Repair Gene Mutations in Men with Metastatic Prostate Cancer.
N Engl J Med, 2016. 375: 443.
https://pubmed.ncbi.nlm.nih.gov/27433846
26. Ewing, C.M., et al. Germline mutations in HOXB13 and prostate-cancer risk. N Engl J Med, 2012.
366: 141.
https://pubmed.ncbi.nlm.nih.gov/22236224
27. Lynch, H.T., et al. Screening for familial and hereditary prostate cancer. Int J Cancer, 2016. 138: 2579.
https://pubmed.ncbi.nlm.nih.gov/26638190
28. Nyberg, T., et al. Prostate Cancer Risks for Male BRCA1 and BRCA2 Mutation Carriers: A
Prospective Cohort Study. Eur Urol, 2020. 77: 24.
https://pubmed.ncbi.nlm.nih.gov/31495749
29. Castro, E., et al. Germline BRCA mutations are associated with higher risk of nodal involvement,
distant metastasis, and poor survival outcomes in prostate cancer. J Clin Oncol, 2013. 31: 1748.
https://pubmed.ncbi.nlm.nih.gov/23569316

136 PROSTATE CANCER - LIMITED UPDATE 2021


30. Castro, E., et al. Effect of BRCA Mutations on Metastatic Relapse and Cause-specific Survival After
Radical Treatment for Localised Prostate Cancer. Eur Urol, 2015. 68: 186.
https://pubmed.ncbi.nlm.nih.gov/25454609
31. Na, R., et al. Germline Mutations in ATM and BRCA1/2 Distinguish Risk for Lethal and Indolent
Prostate Cancer and are Associated with Early Age at Death. Eur Urol, 2017. 71: 740.
https://pubmed.ncbi.nlm.nih.gov/27989354
32. Page, E.C., et al. Interim Results from the IMPACT Study: Evidence for Prostate-specific Antigen
Screening in BRCA2 Mutation Carriers. Eur Urol, 2019. 76: 831.
https://pubmed.ncbi.nlm.nih.gov/31537406
33. Mano, R., et al. Malignant Abnormalities in Male BRCA Mutation Carriers: Results From a
Prospectively Screened Cohort. JAMA Oncol, 2018. 4: 872.
https://pubmed.ncbi.nlm.nih.gov/29710070
34. Leitzmann, M.F., et al. Risk factors for the onset of prostatic cancer: age, location, and behavioral
correlates. Clin Epidemiol, 2012. 4: 1.
https://pubmed.ncbi.nlm.nih.gov/22291478
35. Breslow, N., et al. Latent carcinoma of prostate at autopsy in seven areas. The International Agency
for Research on Cancer, Lyons, France. Int J Cancer, 1977. 20: 680.
https://pubmed.ncbi.nlm.nih.gov/924691
36. Esposito, K., et al. Effect of metabolic syndrome and its components on prostate cancer risk: meta-
analysis. J Endocrinol Invest, 2013. 36: 132.
https://pubmed.ncbi.nlm.nih.gov/23481613
37. Blanc-Lapierre, A., et al. Metabolic syndrome and prostate cancer risk in a population-based case-
control study in Montreal, Canada. BMC Public Health, 2015. 15: 913.
https://pubmed.ncbi.nlm.nih.gov/26385727
38. Preston, M.A., et al. Metformin use and prostate cancer risk. Eur Urol, 2014. 66: 1012.
https://pubmed.ncbi.nlm.nih.gov/24857538
39. Freedland, S.J., et al. Statin use and risk of prostate cancer and high-grade prostate cancer: results
from the REDUCE study. Prostate Cancer Prostatic Dis, 2013. 16: 254.
https://pubmed.ncbi.nlm.nih.gov/23567655
40. James, N.D., et al. Abiraterone for Prostate Cancer Not Previously Treated with Hormone Therapy.
N Engl J Med, 2017. 377: 338.
https://pubmed.ncbi.nlm.nih.gov/28578639
41. Vidal, A.C., et al. Obesity increases the risk for high-grade prostate cancer: results from the
REDUCE study. Cancer Epidemiol Biomarkers Prev, 2014. 23: 2936.
https://pubmed.ncbi.nlm.nih.gov/25261967
42. Davies, N.M., et al. The effects of height and BMI on prostate cancer incidence and mortality:
a Mendelian randomization study in 20,848 cases and 20,214 controls from the PRACTICAL
consortium. Cancer Causes Control, 2015. 26: 1603.
https://pubmed.ncbi.nlm.nih.gov/26387087
43. Dickerman, B.A., et al. Alcohol intake, drinking patterns, and prostate cancer risk and mortality: a
30-year prospective cohort study of Finnish twins. Cancer Causes Control, 2016. 27: 1049.
https://pubmed.ncbi.nlm.nih.gov/27351919
44. Zhao, J., et al. Is alcohol consumption a risk factor for prostate cancer? A systematic review and
meta-analysis. BMC Cancer, 2016. 16: 845.
https://pubmed.ncbi.nlm.nih.gov/27842506
45. Key, T.J. Nutrition, hormones and prostate cancer risk: results from the European prospective
investigation into cancer and nutrition. Recent Results Cancer Res, 2014. 202: 39.
https://pubmed.ncbi.nlm.nih.gov/24531775
46. Alexander, D.D., et al. Meta-Analysis of Long-Chain Omega-3 Polyunsaturated Fatty Acids
(LComega-3PUFA) and Prostate Cancer. Nutr Cancer, 2015. 67: 543.
https://pubmed.ncbi.nlm.nih.gov/25826711
47. Lippi, G., et al. Fried food and prostate cancer risk: systematic review and meta-analysis. Int J Food
Sci Nutr, 2015. 66: 587.
https://pubmed.ncbi.nlm.nih.gov/26114920
48. Chen, P., et al. Lycopene and Risk of Prostate Cancer: A Systematic Review and Meta-Analysis.
Medicine (Baltimore), 2015. 94: e1260.
https://pubmed.ncbi.nlm.nih.gov/26287411
49. Rowles, J.L., 3rd, et al. Processed and raw tomato consumption and risk of prostate cancer: a
systematic review and dose-response meta-analysis. Prostate Cancer Prostatic Dis, 2018. 21: 319.
https://pubmed.ncbi.nlm.nih.gov/29317772

PROSTATE CANCER - LIMITED UPDATE 2021 137


50. Ilic, D., et al. Lycopene for the prevention and treatment of benign prostatic hyperplasia and
prostate cancer: a systematic review. Maturitas, 2012. 72: 269.
https://pubmed.ncbi.nlm.nih.gov/22633187
51. Bylsma, L.C., et al. A review and meta-analysis of prospective studies of red and processed meat,
meat cooking methods, heme iron, heterocyclic amines and prostate cancer. Nutr J, 2015. 14: 125.
https://pubmed.ncbi.nlm.nih.gov/26689289
52. Zhang, M., et al. Is phytoestrogen intake associated with decreased risk of prostate cancer? A
systematic review of epidemiological studies based on 17,546 cases. Andrology, 2016. 4: 745.
https://pubmed.ncbi.nlm.nih.gov/27260185
53. Applegate, C.C., et al. Soy Consumption and the Risk of Prostate Cancer: An Updated Systematic
Review and Meta-Analysis. Nutrients, 2018. 10.
https://pubmed.ncbi.nlm.nih.gov/29300347
54. Kristal, A.R., et al. Plasma vitamin D and prostate cancer risk: results from the Selenium and Vitamin
E Cancer Prevention Trial. Cancer Epidemiol Biomarkers Prev, 2014. 23: 1494.
https://pubmed.ncbi.nlm.nih.gov/24732629
55. Nyame, Y.A., et al. Associations Between Serum Vitamin D and Adverse Pathology in Men
Undergoing Radical Prostatectomy. J Clin Oncol, 2016. 34: 1345.
https://pubmed.ncbi.nlm.nih.gov/26903577
56. Cui, Z., et al. Serum selenium levels and prostate cancer risk: A MOOSE-compliant meta-analysis.
Medicine (Baltimore), 2017. 96: e5944.
https://pubmed.ncbi.nlm.nih.gov/28151881
57. Allen, N.E., et al. Selenium and Prostate Cancer: Analysis of Individual Participant Data From Fifteen
Prospective Studies. J Natl Cancer Inst, 2016. 108.
https://pubmed.ncbi.nlm.nih.gov/27385803
58. Lippman, S.M., et al. Effect of selenium and vitamin E on risk of prostate cancer and other cancers:
the Selenium and Vitamin E Cancer Prevention Trial (SELECT). JAMA, 2009. 301: 39.
https://pubmed.ncbi.nlm.nih.gov/19066370
59. Kramer, B.S., et al. Use of 5-alpha-reductase inhibitors for prostate cancer chemoprevention:
American Society of Clinical Oncology/American Urological Association 2008 Clinical Practice
Guideline. J Clin Oncol, 2009. 27: 1502.
https://pubmed.ncbi.nlm.nih.gov/19252137
60. Andriole, G.L., et al. Effect of dutasteride on the risk of prostate cancer. N Engl J Med, 2010.
362: 1192.
https://pubmed.ncbi.nlm.nih.gov/20357281
61. Thompson, I.M., et al. The influence of finasteride on the development of prostate cancer. N Engl
J Med, 2003. 349: 215.
https://pubmed.ncbi.nlm.nih.gov/12824459
62. Haider, A., et al. Incidence of prostate cancer in hypogonadal men receiving testosterone therapy:
observations from 5-year median followup of 3 registries. J Urol, 2015. 193: 80.
https://pubmed.ncbi.nlm.nih.gov/24980615
63. Watts, E.L., et al. Low Free Testosterone and Prostate Cancer Risk: A Collaborative Analysis of 20
Prospective Studies. Eur Urol, 2018. 74: 585.
https://pubmed.ncbi.nlm.nih.gov/30077399
64. Burns, J.A., et al. Inflammatory Bowel Disease and the Risk of Prostate Cancer. Eur Urol, 2019.
75: 846.
https://pubmed.ncbi.nlm.nih.gov/30528221
65. Zhou, C.K., et al. Male Pattern Baldness in Relation to Prostate Cancer-Specific Mortality: A
Prospective Analysis in the NHANES I Epidemiologic Follow-up Study. Am J Epidemiol, 2016.
183: 210.
https://pubmed.ncbi.nlm.nih.gov/26764224
66. Lian, W.Q., et al. Gonorrhea and Prostate Cancer Incidence: An Updated Meta-Analysis of 21
Epidemiologic Studies. Med Sci Monit, 2015. 21: 1902.
https://pubmed.ncbi.nlm.nih.gov/26126881
67. Rao, D., et al. Does night-shift work increase the risk of prostate cancer? a systematic review and
meta-analysis. Onco Targets Ther, 2015. 8: 2817.
https://pubmed.ncbi.nlm.nih.gov/26491356
68. Islami, F., et al. A systematic review and meta-analysis of tobacco use and prostate cancer mortality
and incidence in prospective cohort studies. Eur Urol, 2014. 66: 1054.
https://pubmed.ncbi.nlm.nih.gov/25946735

138 PROSTATE CANCER - LIMITED UPDATE 2021


69. Brookman-May, S.D., et al. Latest Evidence on the Impact of Smoking, Sports, and Sexual Activity
as Modifiable Lifestyle Risk Factors for Prostate Cancer Incidence, Recurrence, and Progression: A
Systematic Review of the Literature by the European Association of Urology Section of Oncological
Urology (ESOU). Eur Urol Focus, 2019. 5: 756.
https://pubmed.ncbi.nlm.nih.gov/29576530
70. Ju-Kun, S., et al. Association Between Cd Exposure and Risk of Prostate Cancer: A PRISMA-
Compliant Systematic Review and Meta-Analysis. Medicine (Baltimore), 2016. 95: e2708.
https://pubmed.ncbi.nlm.nih.gov/26871808
71. Russo, G.I., et al. Human papillomavirus and risk of prostate cancer: a systematic review and meta-
analysis. Aging Male, 2018: 1.
https://pubmed.ncbi.nlm.nih.gov/29571270
72. Multigner, L., et al. Chlordecone exposure and risk of prostate cancer. J Clin Oncol, 2010. 28: 3457.
https://pubmed.ncbi.nlm.nih.gov/20566993
73. Bhindi, B., et al. The Association Between Vasectomy and Prostate Cancer: A Systematic Review
and Meta-analysis. JAMA Intern Med, 2017. 177: 1273.
https://pubmed.ncbi.nlm.nih.gov/28715534
74. Cremers, R.G., et al. Self-reported acne is not associated with prostate cancer. Urol Oncol, 2014.
32: 941.
https://pubmed.ncbi.nlm.nih.gov/25011577
75. Huang, T.B., et al. Aspirin use and the risk of prostate cancer: a meta-analysis of 24 epidemiologic
studies. Int Urol Nephrol, 2014. 46: 1715.
https://pubmed.ncbi.nlm.nih.gov/24687637
76. Bhindi, B., et al. The impact of the use of aspirin and other nonsteroidal anti-inflammatory drugs on
the risk of prostate cancer detection on biopsy. Urology, 2014. 84: 1073.
https://pubmed.ncbi.nlm.nih.gov/25443907
77. Lin, S.W., et al. Prospective study of ultraviolet radiation exposure and risk of cancer in the United
States. Int J Cancer, 2012. 131: E1015.
https://pubmed.ncbi.nlm.nih.gov/22539073
78. Pabalan, N., et al. Association of male circumcision with risk of prostate cancer: a meta-analysis.
Prostate Cancer Prostatic Dis, 2015. 18: 352.
https://pubmed.ncbi.nlm.nih.gov/26215783
79. Rider, J.R., et al. Ejaculation Frequency and Risk of Prostate Cancer: Updated Results with an
Additional Decade of Follow-up. Eur Urol, 2016. 70: 974.
https://pubmed.ncbi.nlm.nih.gov/27033442
80. Brierley, J.D., et al., TNM classification of malignant tumors. UICC International Union Against
Cancer. 8th edn. 2017.
https://www.uicc.org/resources/tnm/publications-resources
81. Cooperberg, M.R., et al. The University of California, San Francisco Cancer of the Prostate Risk
Assessment score: a straightforward and reliable preoperative predictor of disease recurrence after
radical prostatectomy. J Urol, 2005. 173: 1938.
https://pubmed.ncbi.nlm.nih.gov/15879786
82. Ploussard, G., et al. Decreased accuracy of the prostate cancer EAU risk group classification in the
era of imaging-guided diagnostic pathway: proposal for a new classification based on MRI-targeted
biopsies and early oncologic outcomes after surgery. World J Urol, 2020. 38: 2493.
https://pubmed.ncbi.nlm.nih.gov/31838560
83. Epstein, J.I., et al. The 2005 International Society of Urological Pathology (ISUP) Consensus
Conference on Gleason Grading of Prostatic Carcinoma. Am J Surg Pathol, 2005. 29: 1228.
https://pubmed.ncbi.nlm.nih.gov/16096414
84. Epstein, J.I., et al. The 2014 International Society of Urological Pathology (ISUP) Consensus
Conference on Gleason Grading of Prostatic Carcinoma: Definition of Grading Patterns and
Proposal for a New Grading System. Am J Surg Pathol, 2016. 40: 244.
https://pubmed.ncbi.nlm.nih.gov/26492179
85. van Leenders, G., et al. The 2019 International Society of Urological Pathology (ISUP) Consensus
Conference on Grading of Prostatic Carcinoma. Am J Surg Pathol, 2020. 44: e87.
https://pubmed.ncbi.nlm.nih.gov/32459716
86. Epstein, J.I., et al. A Contemporary Prostate Cancer Grading System: A Validated Alternative to the
Gleason Score. Eur Urol, 2016. 69: 428.
https://pubmed.ncbi.nlm.nih.gov/26166626

PROSTATE CANCER - LIMITED UPDATE 2021 139


87. Kane, C.J., et al. Variability in Outcomes for Patients with Intermediate-risk Prostate Cancer
(Gleason Score 7, International Society of Urological Pathology Gleason Group 2-3) and
Implications for Risk Stratification: A Systematic Review. Eur Urol Focus, 2017. 3: 487.
https://pubmed.ncbi.nlm.nih.gov/28753804
88. Zelic, R., et al. Predicting Prostate Cancer Death with Different Pretreatment Risk Stratification
Tools: A Head-to-head Comparison in a Nationwide Cohort Study. Eur Urol, 2020. 77: 180.
https://pubmed.ncbi.nlm.nih.gov/31606332
89. Dess, R.T., et al. Development and Validation of a Clinical Prognostic Stage Group System for
Nonmetastatic Prostate Cancer Using Disease-Specific Mortality Results From the International
Staging Collaboration for Cancer of the Prostate. JAMA Oncol, 2020. 6: 1912.
https://pubmed.ncbi.nlm.nih.gov/33090219
90. Zumsteg, Z.S., et al. Unification of favourable intermediate-, unfavourable intermediate-, and very
high-risk stratification criteria for prostate cancer. BJU Int, 2017. 120: E87.
https://pubmed.ncbi.nlm.nih.gov/28464446
91. IARC France All Cancers (excluding non-melanoma skin cancer) Estimated Incidence, Mortality and
Prevalence Worldwide in 2012.
https://gco.iarc.fr/today/data/pdf/fact-sheets/cancers/cancer-fact-sheets-29.pdf
92. Etzioni, R., et al. Limitations of basing screening policies on screening trials: The US Preventive
Services Task Force and Prostate Cancer Screening. Med Care, 2013. 51: 295.
https://pubmed.ncbi.nlm.nih.gov/23269114
93. Loeb, S. Guideline of guidelines: prostate cancer screening. BJU Int, 2014. 114: 323.
https://pubmed.ncbi.nlm.nih.gov/24981126
94. Moyer, V.A. Screening for prostate cancer: U.S. Preventive Services Task Force recommendation
statement. Ann Intern Med, 2012. 157: 120.
https://pubmed.ncbi.nlm.nih.gov/22801674
95. Carter, H.B., et al. Early detection of prostate cancer: AUA Guideline. J Urol, 2013. 190: 419.
https://pubmed.ncbi.nlm.nih.gov/23659877
96. Drazer, M.W., et al. National Prostate Cancer Screening Rates After the 2012 US Preventive Services
Task Force Recommendation Discouraging Prostate-Specific Antigen-Based Screening. J Clin
Oncol, 2015. 33: 2416.
https://pubmed.ncbi.nlm.nih.gov/26056181
97. Hu, J.C., et al. Increase in Prostate Cancer Distant Metastases at Diagnosis in the United States.
JAMA Oncol, 2017. 3: 705.
https://pubmed.ncbi.nlm.nih.gov/28033446
98. Jemal, A., et al. Prostate Cancer Incidence and PSA Testing Patterns in Relation to USPSTF
Screening Recommendations. Jama, 2015. 314: 2054.
https://pubmed.ncbi.nlm.nih.gov/26575061
99. Gaylis, F.D., et al. Change in prostate cancer presentation coinciding with USPSTF screening
recommendations at a community-based urology practice. Urol Oncol, 2017. 35: 663.e1.
https://pubmed.ncbi.nlm.nih.gov/28736250
100. Shah, N., et al. Prostate Biopsy Characteristics: A Comparison Between the Pre- and Post-2012
United States Preventive Services Task Force (USPSTF) Prostate Cancer Screening Guidelines. Rev
Urol, 2018. 20: 77.
https://pubmed.ncbi.nlm.nih.gov/30288144
101. Siegel, R.L., et al. Cancer statistics, 2019. CA Cancer J Clin, 2019. 69: 7.
https://pubmed.ncbi.nlm.nih.gov/30620402
102. Kelly, S.P., et al. Past, Current, and Future Incidence Rates and Burden of Metastatic Prostate
Cancer in the United States. Eur Urol Focus, 2018. 4: 121.
https://pubmed.ncbi.nlm.nih.gov/29162421
103. Fenton, J.J., et al. Prostate-Specific Antigen-Based Screening for Prostate Cancer: Evidence Report
and Systematic Review for the US Preventive Services Task Force. Jama, 2018. 319: 1914.
https://pubmed.ncbi.nlm.nih.gov/29801018
104. Ilic, D., et al. Prostate cancer screening with prostate-specific antigen (PSA) test: a systematic
review and meta-analysis. BMJ, 2018. 362: k3519.
https://pubmed.ncbi.nlm.nih.gov/30185521
105. Grossman, D.C., et al. Screening for Prostate Cancer: US Preventive Services Task Force
Recommendation Statement. JAMA, 2018. 319: 1901.
https://pubmed.ncbi.nlm.nih.gov/2680553

140 PROSTATE CANCER - LIMITED UPDATE 2021


106. Bibbins-Domingo, K., et al. The US Preventive Services Task Force 2017 Draft Recommendation
Statement on Screening for Prostate Cancer: An Invitation to Review and Comment. JAMA, 2017.
317: 1949.
https://pubmed.ncbi.nlm.nih.gov/28397958
107. U.S. Preventive Services Task Force. Final Recommendation Statement. Prostate Cancer: Screening
2018 [accessed March 2021].
https://www.uspreventiveservicestaskforce.org/uspstf/recommendation/prostate-cancer-screening
108. Arnsrud Godtman, R., et al. Opportunistic testing versus organized prostate-specific antigen
screening: outcome after 18 years in the Goteborg randomized population-based prostate cancer
screening trial. Eur Urol, 2015. 68: 354.
https://pubmed.ncbi.nlm.nih.gov/25556937
109. Ilic, D., et al. Screening for prostate cancer. Cochrane Database Syst Rev, 2013: Cd004720.
https://pubmed.ncbi.nlm.nih.gov/23440794
110. Hayes, J.H., et al. Screening for prostate cancer with the prostate-specific antigen test: a review of
current evidence. JAMA, 2014. 311: 1143.
https://pubmed.ncbi.nlm.nih.gov/24643604
111. Borghesi, M., et al. Complications After Systematic, Random, and Image-guided Prostate Biopsy.
Eur Urol, 2017. 71: 353.
https://pubmed.ncbi.nlm.nih.gov/27543165
112. Booth, N., et al. Health-related quality of life in the Finnish trial of screening for prostate cancer. Eur
Urol, 2014. 65: 39.
https://pubmed.ncbi.nlm.nih.gov/23265387
113. Vasarainen, H., et al. Effects of prostate cancer screening on health-related quality of life: results of
the Finnish arm of the European randomized screening trial (ERSPC). Acta Oncol, 2013. 52: 1615.
https://pubmed.ncbi.nlm.nih.gov/23786174
114. Heijnsdijk, E.A., et al. Quality-of-life effects of prostate-specific antigen screening. N Engl J Med,
2012. 367: 595.
https://pubmed.ncbi.nlm.nih.gov/22894572
115. Martin, R.M., et al. Effect of a Low-Intensity PSA-Based Screening Intervention on Prostate Cancer
Mortality: The CAP Randomized Clinical Trial. JAMA, 2018. 319: 883.
https://pubmed.ncbi.nlm.nih.gov/29509864
116. Hugosson, J., et al. A 16-yr Follow-up of the European Randomized study of Screening for Prostate
Cancer. Eur Urol, 2019. 76: 43.
https://pubmed.ncbi.nlm.nih.gov/30824296
117. The benefits and harms of breast cancer screening: an independent review. Lancet, 2012. 380: 1778.
https://pubmed.ncbi.nlm.nih.gov/23117178
118. Pinsky, P.F., et al. Extended follow-up for prostate cancer incidence and mortality among
participants in the Prostate, Lung, Colorectal and Ovarian randomized cancer screening trial. BJU
Int, 2019. 123: 854.
https://pubmed.ncbi.nlm.nih.gov/30288918
119. Brandt, A., et al. Age-specific risk of incident prostate cancer and risk of death from prostate cancer
defined by the number of affected family members. Eur Urol, 2010. 58: 275.
https://pubmed.ncbi.nlm.nih.gov/20171779
120. Carlsson, S., et al. Screening for Prostate Cancer Starting at Age 50-54 Years. A Population-based
Cohort Study. Eur Urol, 2017. 71: 46.
https://pubmed.ncbi.nlm.nih.gov/27084245
121. Albright, F., et al. Prostate cancer risk prediction based on complete prostate cancer family history.
Prostate, 2015. 75: 390.
https://pubmed.ncbi.nlm.nih.gov/25408531
122. Kamangar, F., et al. Patterns of cancer incidence, mortality, and prevalence across five continents:
defining priorities to reduce cancer disparities in different geographic regions of the world. J Clin
Oncol, 2006. 24: 2137.
https://pubmed.ncbi.nlm.nih.gov/16682732
123. Chornokur, G., et al. Disparities at presentation, diagnosis, treatment, and survival in African
American men, affected by prostate cancer. Prostate, 2011. 71: 985.
https://pubmed.ncbi.nlm.nih.gov/21541975
124. Karami, S., et al. Earlier age at diagnosis: another dimension in cancer disparity? Cancer Detect
Prev, 2007. 31: 29.
https://pubmed.ncbi.nlm.nih.gov/17303347

PROSTATE CANCER - LIMITED UPDATE 2021 141


125. Sanchez-Ortiz, R.F., et al. African-American men with nonpalpable prostate cancer exhibit greater
tumor volume than matched white men. Cancer, 2006. 107: 75.
https://pubmed.ncbi.nlm.nih.gov/16736511
126. Bancroft, E.K., et al. Targeted Prostate Cancer Screening in BRCA1 and BRCA2 Mutation Carriers:
Results from the Initial Screening Round of the IMPACT Study. Eur Urol, 2014. 66: 489.
https://pubmed.ncbi.nlm.nih.gov/24484606
127. Gulati, R., et al. Screening Men at Increased Risk for Prostate Cancer Diagnosis: Model Estimates of
Benefits and Harms. Cancer Epidemiol Biomarkers Prev, 2017. 26: 222.
https://pubmed.ncbi.nlm.nih.gov/27742670
128. Vickers, A.J., et al. Strategy for detection of prostate cancer based on relation between prostate
specific antigen at age 40-55 and long term risk of metastasis: case-control study. BMJ, 2013.
346: f2023.
https://pubmed.ncbi.nlm.nih.gov/23596126
129. Carlsson, S., et al. Influence of blood prostate specific antigen levels at age 60 on benefits and
harms of prostate cancer screening: population based cohort study. Bmj, 2014. 348: g2296.
https://pubmed.ncbi.nlm.nih.gov/24682399
130. Naji, L., et al. Digital Rectal Examination for Prostate Cancer Screening in Primary Care: A
Systematic Review and Meta-Analysis. Ann Fam Med, 2018. 16: 149.
https://pubmed.ncbi.nlm.nih.gov/29531107
131. Loeb, S., et al. Pathological characteristics of prostate cancer detected through prostate specific
antigen based screening. J Urol, 2006. 175: 902.
https://pubmed.ncbi.nlm.nih.gov/16469576
132. Gelfond, J., et al. Intermediate-Term Risk of Prostate Cancer is Directly Related to Baseline Prostate
Specific Antigen: Implications for Reducing the Burden of Prostate Specific Antigen Screening.
J Urol, 2015. 194: 46.
https://pubmed.ncbi.nlm.nih.gov/25686543
133. Roobol, M.J., et al. Is additional testing necessary in men with prostate-specific antigen levels of
1.0 ng/mL or less in a population-based screening setting? (ERSPC, section Rotterdam). Urology,
2005. 65: 343.
https://pubmed.ncbi.nlm.nih.gov/15708050
134. Boyle, H.J., et al. Updated recommendations of the International Society of Geriatric Oncology on
prostate cancer management in older patients. Eur J Cancer, 2019. 116: 116.
https://pubmed.ncbi.nlm.nih.gov/31195356
135. Kretschmer, A., et al. Biomarkers in prostate cancer - Current clinical utility and future perspectives.
Crit Rev Oncol Hematol, 2017. 120: 180.
https://pubmed.ncbi.nlm.nih.gov/29198331
136. Roobol, M.J., et al. Improving the Rotterdam European Randomized Study of Screening for Prostate
Cancer Risk Calculator for Initial Prostate Biopsy by Incorporating the 2014 International Society of
Urological Pathology Gleason Grading and Cribriform growth. Eur Urol, 2017. 72: 45.
https://pubmed.ncbi.nlm.nih.gov/28162815
137. Louie, K.S., et al. Do prostate cancer risk models improve the predictive accuracy of PSA
screening? A meta-analysis. Ann Oncol, 2015. 26: 848.
https://pubmed.ncbi.nlm.nih.gov/25403590
138. Mortezavi, A., et al. Head-to-head Comparison of Conventional, and Image- and Biomarker-based
Prostate Cancer Risk Calculators. Eur Urol Focus, 2020.
https://pubmed.ncbi.nlm.nih.gov/32451315
139. Giri, V.N., et al. Implementation of Germline Testing for Prostate Cancer: Philadelphia Prostate
Cancer Consensus Conference 2019. J Clin Oncol, 2020. 38: 2798.
https://pubmed.ncbi.nlm.nih.gov/32516092
140. John, E.M., et al. Prevalence of pathogenic BRCA1 mutation carriers in 5 US racial/ethnic groups.
Jama, 2007. 298: 2869.
https://pubmed.ncbi.nlm.nih.gov/18159056
141. Edwards, S.M., et al. Two percent of men with early-onset prostate cancer harbor germline
mutations in the BRCA2 gene. Am J Hum Genet, 2003. 72: 1.
https://pubmed.ncbi.nlm.nih.gov/12474142
142. van Asperen, C.J., et al. Cancer risks in BRCA2 families: estimates for sites other than breast and
ovary. J Med Genet, 2005. 42: 711.
https://pubmed.ncbi.nlm.nih.gov/16141007

142 PROSTATE CANCER - LIMITED UPDATE 2021


143. Agalliu, I., et al. Rare germline mutations in the BRCA2 gene are associated with early-onset
prostate cancer. Br J Cancer, 2007. 97: 826.
https://pubmed.ncbi.nlm.nih.gov/17700570
144. Leongamornlert, D., et al. Frequent germline deleterious mutations in DNA repair genes in familial
prostate cancer cases are associated with advanced disease. Br J Cancer, 2014. 110: 1663.
https://pubmed.ncbi.nlm.nih.gov/24556621
145. Wang, Y., et al. CHEK2 mutation and risk of prostate cancer: a systematic review and meta-analysis.
Int J Clin Exp Med, 2015. 8: 15708.
https://pubmed.ncbi.nlm.nih.gov/26629066
146. Zhen, J.T., et al. Genetic testing for hereditary prostate cancer: Current status and limitations.
Cancer, 2018. 124: 3105.
https://pubmed.ncbi.nlm.nih.gov/29669169
147. Leongamornlert, D., et al. Germline BRCA1 mutations increase prostate cancer risk. Br J Cancer,
2012. 106: 1697.
https://pubmed.ncbi.nlm.nih.gov/22516946
148. Thompson, D., et al. Cancer Incidence in BRCA1 mutation carriers. J Natl Cancer Inst, 2002. 94: 1358.
https://pubmed.ncbi.nlm.nih.gov/12237281
149. Karlsson, R., et al. A population-based assessment of germline HOXB13 G84E mutation and
prostate cancer risk. Eur Urol, 2014. 65: 169.
https://pubmed.ncbi.nlm.nih.gov/22841674
150. Storebjerg, T.M., et al. Prevalence of the HOXB13 G84E mutation in Danish men undergoing radical
prostatectomy and its correlations with prostate cancer risk and aggressiveness. BJU Int, 2016.
118: 646.
https://pubmed.ncbi.nlm.nih.gov/26779768
151. Ryan, S., et al. Risk of prostate cancer in Lynch syndrome: a systematic review and meta-analysis.
Cancer Epidemiol Biomarkers Prev, 2014. 23: 437.
https://pubmed.ncbi.nlm.nih.gov/24425144
152. Rosty, C., et al. High prevalence of mismatch repair deficiency in prostate cancers diagnosed in
mismatch repair gene mutation carriers from the colon cancer family registry. Fam Cancer, 2014.
13: 573.
https://pubmed.ncbi.nlm.nih.gov/25117503
153. Richie, J.P., et al. Effect of patient age on early detection of prostate cancer with serum prostate-
specific antigen and digital rectal examination. Urology, 1993. 42: 365.
https://pubmed.ncbi.nlm.nih.gov/7692657
154. Carvalhal, G.F., et al. Digital rectal examination for detecting prostate cancer at prostate specific
antigen levels of 4 ng./ml. or less. J Urol, 1999. 161: 835.
https://pubmed.ncbi.nlm.nih.gov/10022696
155. Okotie, O.T., et al. Characteristics of prostate cancer detected by digital rectal examination only.
Urology, 2007. 70: 1117.
https://pubmed.ncbi.nlm.nih.gov/18158030
156. Gosselaar, C., et al. The role of the digital rectal examination in subsequent screening visits in the
European randomized study of screening for prostate cancer (ERSPC), Rotterdam. Eur Urol, 2008.
54: 581.
https://pubmed.ncbi.nlm.nih.gov/18423977
157. Stamey, T.A., et al. Prostate-specific antigen as a serum marker for adenocarcinoma of the prostate.
N Engl J Med, 1987. 317: 909.
https://pubmed.ncbi.nlm.nih.gov/2442609
158. Catalona, W.J., et al. Comparison of digital rectal examination and serum prostate specific antigen
in the early detection of prostate cancer: results of a multicenter clinical trial of 6,630 men. J Urol,
1994. 151: 1283.
https://pubmed.ncbi.nlm.nih.gov/7512659
159. Semjonow, A., et al. Discordance of assay methods creates pitfalls for the interpretation of prostate-
specific antigen values. Prostate Suppl, 1996. 7: 3.
https://pubmed.ncbi.nlm.nih.gov/8950358
160. Thompson, I.M., et al. Prevalence of prostate cancer among men with a prostate-specific antigen
level < or =4.0 ng per milliliter. N Engl J Med, 2004. 350: 2239.
https://pubmed.ncbi.nlm.nih.gov/15163773
161. Dong, F., et al. Validation of pretreatment nomograms for predicting indolent prostate cancer:
efficacy in contemporary urological practice. J Urol, 2008. 180: 150.
https://pubmed.ncbi.nlm.nih.gov/18485398

PROSTATE CANCER - LIMITED UPDATE 2021 143


162. Carter, H.B., et al. Longitudinal evaluation of prostate-specific antigen levels in men with and
without prostate disease. JAMA, 1992. 267: 2215.
https://pubmed.ncbi.nlm.nih.gov/1372942
163. Schmid, H.P., et al. Observations on the doubling time of prostate cancer. The use of serial prostate-
specific antigen in patients with untreated disease as a measure of increasing cancer volume.
Cancer, 1993. 71: 2031.
https://pubmed.ncbi.nlm.nih.gov/7680277
164. Arlen, P.M., et al. Prostate Specific Antigen Working Group guidelines on prostate specific antigen
doubling time. J Urol, 2008. 179: 2181.
https://pubmed.ncbi.nlm.nih.gov/18423743
165. Heidenreich, A. Identification of high-risk prostate cancer: role of prostate-specific antigen, PSA
doubling time, and PSA velocity. Eur Urol, 2008. 54: 976.
https://pubmed.ncbi.nlm.nih.gov/18640768
166. Ramirez, M.L., et al. Current applications for prostate-specific antigen doubling time. Eur Urol, 2008.
54: 291.
https://pubmed.ncbi.nlm.nih.gov/18439749
167. O’Brien, M.F., et al. Pretreatment prostate-specific antigen (PSA) velocity and doubling time are
associated with outcome but neither improves prediction of outcome beyond pretreatment PSA
alone in patients treated with radical prostatectomy. J Clin Oncol, 2009. 27: 3591.
https://pubmed.ncbi.nlm.nih.gov/19506163
168. Vickers, A.J., et al. Systematic review of pretreatment prostate-specific antigen velocity and
doubling time as predictors for prostate cancer. J Clin Oncol, 2009. 27: 398.
https://pubmed.ncbi.nlm.nih.gov/19064972
169. Choo, R., et al. Wide variation of prostate-specific antigen doubling time of untreated, clinically
localized, low-to-intermediate grade, prostate carcinoma. BJU Int, 2004. 94: 295.
https://pubmed.ncbi.nlm.nih.gov/15291854
170. Cannon, G.M., Jr., et al. Prostate-specific antigen doubling time in the identification of patients at
risk for progression after treatment and biochemical recurrence for prostate cancer. Urology, 2003.
62 Suppl 1: 2.
https://pubmed.ncbi.nlm.nih.gov/14747037
171. Daskivich, T.J., et al. Prostate specific antigen doubling time calculation: not as easy as 1, 2, 4.
J Urol, 2006. 176: 1927.
https://pubmed.ncbi.nlm.nih.gov/17070213
172. Loberg, R.D., et al. Prostate-specific antigen doubling time and survival in patients with advanced
metastatic prostate cancer. Urology, 2003. 62 Suppl 1: 128.
https://pubmed.ncbi.nlm.nih.gov/14747050
173. Stephan, C., et al. The influence of prostate volume on the ratio of free to total prostate specific
antigen in serum of patients with prostate carcinoma and benign prostate hyperplasia. Cancer,
1997. 79: 104.
https://pubmed.ncbi.nlm.nih.gov/8988733
174. Catalona, W.J., et al. Use of the percentage of free prostate-specific antigen to enhance
differentiation of prostate cancer from benign prostatic disease: a prospective multicenter clinical
trial. JAMA, 1998. 279: 1542.
https://pubmed.ncbi.nlm.nih.gov/9605898
175. Huang, Y., et al. Value of free/total prostate-specific antigen (f/t PSA) ratios for prostate cancer
detection in patients with total serum prostate-specific antigen between 4 and 10 ng/mL: A meta-
analysis. Medicine (Baltimore), 2018. 97: e0249.
https://pubmed.ncbi.nlm.nih.gov/29595681
176. Bryant, R.J., et al. Predicting high-grade cancer at ten-core prostate biopsy using four kallikrein
markers measured in blood in the ProtecT study. J Natl Cancer Inst, 2015. 107.
https://pubmed.ncbi.nlm.nih.gov/25863334
177. Loeb, S., et al. The Prostate Health Index: a new test for the detection of prostate cancer. Ther Adv
Urol, 2014. 6: 74.
https://pubmed.ncbi.nlm.nih.gov/24688603
178. de la Calle, C., et al. Multicenter Evaluation of the Prostate Health Index to Detect Aggressive
Prostate Cancer in Biopsy Naive Men. J Urol, 2015. 194: 65.
https://pubmed.ncbi.nlm.nih.gov/25636659
179. Catalona, W.J., et al. A multicenter study of [-2]pro-prostate specific antigen combined with prostate
specific antigen and free prostate specific antigen for prostate cancer detection in the 2.0 to
10.0 ng/ml prostate specific antigen range. J Urol, 2011. 185: 1650.
https://pubmed.ncbi.nlm.nih.gov/21419439

144 PROSTATE CANCER - LIMITED UPDATE 2021


180. Nordstrom, T., et al. Comparison Between the Four-kallikrein Panel and Prostate Health Index for
Predicting Prostate Cancer. Eur Urol, 2015. 68: 139.
https://pubmed.ncbi.nlm.nih.gov/25151013
181. Klein, E.A., et al. The Single-parameter, Structure-based IsoPSA Assay Demonstrates Improved
Diagnostic Accuracy for Detection of Any Prostate Cancer and High-grade Prostate Cancer
Compared to a Concentration-based Assay of Total Prostate-specific Antigen: A Preliminary Report.
Eur Urol, 2017. 72: 942.
https://pubmed.ncbi.nlm.nih.gov/28396176
182. Stovsky, M., et al. Clinical Validation of IsoPSA™, a Single Parameter, Structure Based Assay for
Improved Detection of High Grade Prostate Cancer. J Urol, 2019. 201: 1115.
https://pubmed.ncbi.nlm.nih.gov/30810464
183. Deras, I.L., et al. PCA3: a molecular urine assay for predicting prostate biopsy outcome. J Urol,
2008. 179: 1587.
https://pubmed.ncbi.nlm.nih.gov/18295257
184. Hessels, D., et al. DD3(PCA3)-based molecular urine analysis for the diagnosis of prostate cancer.
Eur Urol, 2003. 44: 8.
https://pubmed.ncbi.nlm.nih.gov/12814669
185. Nakanishi, H., et al. PCA3 molecular urine assay correlates with prostate cancer tumor volume:
implication in selecting candidates for active surveillance. J Urol, 2008. 179: 1804.
https://pubmed.ncbi.nlm.nih.gov/18353398
186. Hessels, D., et al. Predictive value of PCA3 in urinary sediments in determining clinico-pathological
characteristics of prostate cancer. Prostate, 2010. 70: 10.
https://pubmed.ncbi.nlm.nih.gov/19708043
187. Auprich, M., et al. Contemporary role of prostate cancer antigen 3 in the management of prostate
cancer. Eur Urol, 2011. 60: 1045.
https://pubmed.ncbi.nlm.nih.gov/21871709
188. Nicholson, A., et al. The clinical effectiveness and cost-effectiveness of the PROGENSA(R) prostate
cancer antigen 3 assay and the Prostate Health Index in the diagnosis of prostate cancer: a
systematic review and economic evaluation. Health Technol Assess, 2015. 19: 1.
https://pubmed.ncbi.nlm.nih.gov/26507078
189. Wei, J.T., et al. Can urinary PCA3 supplement PSA in the early detection of prostate cancer? J Clin
Oncol, 2014. 32: 4066.
https://pubmed.ncbi.nlm.nih.gov/25385735
190. Van Neste, L., et al. Detection of High-grade Prostate Cancer Using a Urinary Molecular Biomarker-
Based Risk Score. Eur Urol, 2016. 70: 740.
https://pubmed.ncbi.nlm.nih.gov/27108162
191. Tomlins, S.A., et al. Recurrent fusion of TMPRSS2 and ETS transcription factor genes in prostate
cancer. Science, 2005. 310: 644.
https://pubmed.ncbi.nlm.nih.gov/16254181
192. Tomlins, S.A., et al. Urine TMPRSS2:ERG Plus PCA3 for Individualized Prostate Cancer Risk
Assessment. Eur Urol, 2016. 70: 45.
https://pubmed.ncbi.nlm.nih.gov/25985884
193. Donovan, M.J., et al. A molecular signature of PCA3 and ERG exosomal RNA from non-DRE urine is
predictive of initial prostate biopsy result. Prostate Cancer Prostatic Dis, 2015. 18: 370.
https://pubmed.ncbi.nlm.nih.gov/26345389
194. McKiernan, J., et al. A Novel Urine Exosome Gene Expression Assay to Predict High-grade Prostate
Cancer at Initial Biopsy. JAMA Oncol, 2016. 2: 882.
https://pubmed.ncbi.nlm.nih.gov/27032035
195. Seisen, T., et al. Accuracy of the prostate health index versus the urinary prostate cancer antigen 3
score to predict overall and significant prostate cancer at initial biopsy. Prostate, 2015. 75: 103.
https://pubmed.ncbi.nlm.nih.gov/25327361
196. Russo, G.I., et al. A Systematic Review and Meta-analysis of the Diagnostic Accuracy of Prostate
Health Index and 4-Kallikrein Panel Score in Predicting Overall and High-grade Prostate Cancer. Clin
Genitourin Cancer, 2017. 15: 429.
https://pubmed.ncbi.nlm.nih.gov/28111174
197. Vedder, M.M., et al. The added value of percentage of free to total prostate-specific antigen, PCA3,
and a kallikrein panel to the ERSPC risk calculator for prostate cancer in prescreened men. Eur Urol,
2014. 66: 1109.
https://pubmed.ncbi.nlm.nih.gov/25168616

PROSTATE CANCER - LIMITED UPDATE 2021 145


198. Lamy, P.J., et al. Prognostic Biomarkers Used for Localised Prostate Cancer Management: A
Systematic Review. Eur Urol Focus, 2018. 4: 790.
https://pubmed.ncbi.nlm.nih.gov/28753865
199. Partin, A.W., et al. Clinical validation of an epigenetic assay to predict negative histopathological
results in repeat prostate biopsies. J Urol, 2014. 192: 1081.
https://pubmed.ncbi.nlm.nih.gov/24747657
200. Moore, C.K., et al. Prognostic significance of high grade prostatic intraepithelial neoplasia and
atypical small acinar proliferation in the contemporary era. J Urol, 2005. 173: 70.
https://pubmed.ncbi.nlm.nih.gov/15592031
201. Smeenge, M., et al. Role of transrectal ultrasonography (TRUS) in focal therapy of prostate cancer:
report from a Consensus Panel. BJU Int, 2012. 110: 942.
https://pubmed.ncbi.nlm.nih.gov/22462566
202. Rouviere, O., et al. Use of prostate systematic and targeted biopsy on the basis of multiparametric
MRI in biopsy-naive patients (MRI-FIRST): a prospective, multicentre, paired diagnostic study.
Lancet Oncol, 2019. 20: 100.
https://pubmed.ncbi.nlm.nih.gov/30470502
203. Wysock, J.S., et al. HistoScanning(TM) to Detect and Characterize Prostate Cancer-a Review of
Existing Literature. Curr Urol Rep, 2017. 18: 97.
https://pubmed.ncbi.nlm.nih.gov/29064054
204. Correas, J.M., et al. Advanced ultrasound in the diagnosis of prostate cancer. World J Urol, 2021.
39: 661.
https://pubmed.ncbi.nlm.nih.gov/32306060
205. Lughezzani, G., et al. Comparison of the Diagnostic Accuracy of Micro-ultrasound and Magnetic
Resonance Imaging/Ultrasound Fusion Targeted Biopsies for the Diagnosis of Clinically Significant
Prostate Cancer. Eur Urol Oncol, 2019. 2: 329.
https://pubmed.ncbi.nlm.nih.gov/31200848
206. Cornud, F., et al. MRI-directed high-frequency (29MhZ) TRUS-guided biopsies: initial results of a
single-center study. Eur Radiol, 2020. 30: 4838.
https://pubmed.ncbi.nlm.nih.gov/32350662
207. Bratan, F., et al. Influence of imaging and histological factors on prostate cancer detection and
localisation on multiparametric MRI: a prospective study. Eur Radiol, 2013. 23: 2019.
https://pubmed.ncbi.nlm.nih.gov/23494494
208. Borofsky, S., et al. What Are We Missing? False-Negative Cancers at Multiparametric MR Imaging of
the Prostate. Radiology, 2018. 286: 186.
https://pubmed.ncbi.nlm.nih.gov/29053402
209. Johnson, D.C., et al. Detection of Individual Prostate Cancer Foci via Multiparametric Magnetic
Resonance Imaging. Eur Urol, 2019. 75: 712.
https://pubmed.ncbi.nlm.nih.gov/30509763
210. Drost, F.H., et al. Prostate MRI, with or without MRI-targeted biopsy, and systematic biopsy for
detecting prostate cancer. Cochrane Database Syst Rev, 2019. 4: CD012663.
https://pubmed.ncbi.nlm.nih.gov/31022301
211. Dickinson, L., et al. Magnetic resonance imaging for the detection, localisation, and characterisation
of prostate cancer: recommendations from a European consensus meeting. Eur Urol, 2011. 59: 477.
https://pubmed.ncbi.nlm.nih.gov/21195536
212. Weinreb, J.C., et al. PI-RADS Prostate Imaging – Reportin49g and Data System: 2015, Version 2.
Eur Urol, 2016. 69: 16.
https://pubmed.ncbi.nlm.nih.gov/26427566
213. Turkbey, B., et al. Prostate Imaging Reporting and Data System Version 2.1: 2019 Update of
Prostate Imaging Reporting and Data System Version 2. Eur Urol, 2019. 76: 340.
https://pubmed.ncbi.nlm.nih.gov/30898406
214. Barkovich, E.J., et al. A Systematic Review of the Existing Prostate Imaging Reporting and Data
System Version 2 (PI-RADSv2) Literature and Subset Meta-Analysis of PI-RADSv2 Categories
Stratified by Gleason Scores. AJR Am J Roentgenol, 2019. 212: 847.
https://pubmed.ncbi.nlm.nih.gov/30807218
215. Westphalen, A.C., et al. Variability of the Positive Predictive Value of PI-RADS for Prostate MRI
across 26 Centers: Experience of the Society of Abdominal Radiology Prostate Cancer Disease-
focused Panel. Radiology, 2020. 296: 76.
https://pubmed.ncbi.nlm.nih.gov/32315265

146 PROSTATE CANCER - LIMITED UPDATE 2021


216. Goldberg, H., et al. Comparison of Magnetic Resonance Imaging and Transrectal Ultrasound
Informed Prostate Biopsy for Prostate Cancer Diagnosis in Biopsy Naïve Men: A Systematic Review
and Meta-Analysis. J Urol, 2020. 203: 1085.
https://pubmed.ncbi.nlm.nih.gov/31609177
217. Kasivisvanathan, V., et al. MRI-Targeted or Standard Biopsy for Prostate-Cancer Diagnosis. N Engl
J Med, 2018. 378: 1767.
https://pubmed.ncbi.nlm.nih.gov/29552975
218. van der Leest, M., et al. Head-to-head Comparison of Transrectal Ultrasound-guided Prostate
Biopsy Versus Multiparametric Prostate Resonance Imaging with Subsequent Magnetic Resonance-
guided Biopsy in Biopsy-naive Men with Elevated Prostate-specific Antigen: A Large Prospective
Multicenter Clinical Study. Eur Urol, 2019. 75: 570.
https://pubmed.ncbi.nlm.nih.gov/30477981
219. Wegelin, O., et al. The FUTURE Trial: A Multicenter Randomised Controlled Trial on Target Biopsy
Techniques Based on Magnetic Resonance Imaging in the Diagnosis of Prostate Cancer in Patients
with Prior Negative Biopsies. Eur Urol, 2019. 75: 582.
https://pubmed.ncbi.nlm.nih.gov/30522912
220. Exterkate, L., et al. Is There Still a Need for Repeated Systematic Biopsies in Patients with Previous
Negative Biopsies in the Era of Magnetic Resonance Imaging-targeted Biopsies of the Prostate? Eur
Urol Oncol, 2020. 3: 216.
https://pubmed.ncbi.nlm.nih.gov/31239236
221. Stabile, A., et al. Factors Influencing Variability in the Performance of Multiparametric Magnetic
Resonance Imaging in Detecting Clinically Significant Prostate Cancer: A Systematic Literature
Review. Eur Urol Oncol, 2020. 3: 145.
https://pubmed.ncbi.nlm.nih.gov/32192942
222. Farrell, C., et al. Prostate Multiparametric Magnetic Resonance Imaging Program Implementation
and Impact: Initial Clinical Experience in a Community Based Health System. Urology Practice,
2018. 5: 165.
https://www.sciencedirect.com/science/article/abs/pii/S2352077917300729
223. Meng, X., et al. The Institutional Learning Curve of Magnetic Resonance Imaging-Ultrasound Fusion
Targeted Prostate Biopsy: Temporal Improvements in Cancer Detection in 4 Years. J Urol, 2018.
200: 1022.
https://pubmed.ncbi.nlm.nih.gov/29886090
224. Raeside, M., et al. Prostate MRI evolution in clinical practice: Audit of tumour detection and staging
versus prostatectomy with staged introduction of multiparametric MRI and Prostate Imaging
Reporting and Data System v2 reporting. J Med Imaging Radiat Oncol, 2019. 63: 487.
https://pubmed.ncbi.nlm.nih.gov/30951248
225. Shaish, H., et al. Impact of a Structured Reporting Template on Adherence to Prostate Imaging
Reporting and Data System Version 2 and on the Diagnostic Performance of Prostate MRI for
Clinically Significant Prostate Cancer. J Am Coll Radiol, 2018. 15: 749.
https://pubmed.ncbi.nlm.nih.gov/29506919
226. Niaf, E., et al. Prostate focal peripheral zone lesions: characterization at multiparametric MR
imaging--influence of a computer-aided diagnosis system. Radiology, 2014. 271: 761.
https://pubmed.ncbi.nlm.nih.gov/24592959
227. Litjens, G.J., et al. Clinical evaluation of a computer-aided diagnosis system for determining cancer
aggressiveness in prostate MRI. Eur Radiol, 2015. 25: 3187.
https://pubmed.ncbi.nlm.nih.gov/26060063
228. Hoang Dinh, A., et al. Quantitative Analysis of Prostate Multiparametric MR Images for Detection of
Aggressive Prostate Cancer in the Peripheral Zone: A Multiple Imager Study. Radiology, 2016. 280:
117.
https://pubmed.ncbi.nlm.nih.gov/26859255
229. Bryk, D.J., et al. The Role of Ipsilateral and Contralateral Transrectal Ultrasound-guided Systematic
Prostate Biopsy in Men With Unilateral Magnetic Resonance Imaging Lesion Undergoing Magnetic
Resonance Imaging-ultrasound Fusion-targeted Prostate Biopsy. Urology, 2017. 102: 178.
https://pubmed.ncbi.nlm.nih.gov/27871829
230. Freifeld, Y., et al. Optimal sampling scheme in men with abnormal multiparametric MRI undergoing
MRI-TRUS fusion prostate biopsy. Urol Oncol, 2019. 37: 57.
https://pubmed.ncbi.nlm.nih.gov/30446460
231. Kenigsberg, A.P., et al. Optimizing the Number of Cores Targeted During Prostate Magnetic
Resonance Imaging Fusion Target Biopsy. Eur Urol Oncol, 2018. 1: 418.
https://pubmed.ncbi.nlm.nih.gov/31158081

PROSTATE CANCER - LIMITED UPDATE 2021 147


232. Zhang, M., et al. Value of Increasing Biopsy Cores per Target with Cognitive MRI-targeted
Transrectal US Prostate Biopsy. Radiology, 2019. 291: 83.
https://pubmed.ncbi.nlm.nih.gov/30694165
233. Lu, A.J., et al. Role of Core Number and Location in Targeted Magnetic Resonance Imaging-
Ultrasound Fusion Prostate Biopsy. Eur Urol, 2019. 76: 14.
https://pubmed.ncbi.nlm.nih.gov/31047733
234. Distler, F.A., et al. The Value of PSA Density in Combination with PI-RADS for the Accuracy of
Prostate Cancer Prediction. J Urol, 2017. 198: 575.
https://pubmed.ncbi.nlm.nih.gov/28373135
235. Washino, S., et al. Combination of prostate imaging reporting and data system (PI-RADS) score and
prostate-specific antigen (PSA) density predicts biopsy outcome in prostate biopsy naive patients.
BJU Int, 2017. 119: 225.
https://pubmed.ncbi.nlm.nih.gov/26935594
236. Pagniez, M.A., et al. Predictive Factors of Missed Clinically Significant Prostate Cancers in Men
with Negative Magnetic Resonance Imaging: A Systematic Review and Meta-Analysis. J Urol, 2020.
204: 24.
https://pubmed.ncbi.nlm.nih.gov/31967522
237. Hansen, N.L., et al. The influence of prostate-specific antigen density on positive and negative
predictive values of multiparametric magnetic resonance imaging to detect Gleason score 7-10
prostate cancer in a repeat biopsy setting. BJU Int, 2017. 119: 724.
https://pubmed.ncbi.nlm.nih.gov/27488931
238. Hansen, N.L., et al. Multicentre evaluation of magnetic resonance imaging supported transperineal
prostate biopsy in biopsy-naive men with suspicion of prostate cancer. BJU Int, 2018. 122: 40.
https://pubmed.ncbi.nlm.nih.gov/29024425
239. Oishi, M., et al. Which Patients with Negative Magnetic Resonance Imaging Can Safely Avoid
Biopsy for Prostate Cancer? J Urol, 2019. 201: 268.
https://pubmed.ncbi.nlm.nih.gov/30189186
240. Boesen, L., et al. Prebiopsy Biparametric Magnetic Resonance Imaging Combined with Prostate-
specific Antigen Density in Detecting and Ruling out Gleason 7-10 Prostate Cancer in Biopsy-naive
Men. Eur Urol Oncol, 2019. 2: 311.
https://pubmed.ncbi.nlm.nih.gov/31200846
241. Ploussard, G., et al. The role of prostate cancer antigen 3 (PCA3) in prostate cancer detection.
Expert Rev Anticancer Ther, 2018. 18: 1013.
https://pubmed.ncbi.nlm.nih.gov/30016891
242. Schoots, I.G., et al. Multivariate risk prediction tools including MRI for individualized biopsy decision
in prostate cancer diagnosis: current status and future directions. World J Urol, 2020. 38: 517.
https://pubmed.ncbi.nlm.nih.gov/30868240
243. Saba, K., et al. External Validation and Comparison of Prostate Cancer Risk Calculators
Incorporating Multiparametric Magnetic Resonance Imaging for Prediction of Clinically Significant
Prostate Cancer. J Urol, 2020. 203: 719.
https://pubmed.ncbi.nlm.nih.gov/31651228
244. Radtke, J.P., et al. Prediction of significant prostate cancer in biopsy-naïve men: Validation of
a novel risk model combining MRI and clinical parameters and comparison to an ERSPC risk
calculator and PI-RADS. PLoS One, 2019. 14: e0221350.
https://pubmed.ncbi.nlm.nih.gov/31450235
245. Mannaerts, C.K., et al. Prostate Cancer Risk Assessment in Biopsy-naïve Patients: The Rotterdam
Prostate Cancer Risk Calculator in Multiparametric Magnetic Resonance Imaging-Transrectal
Ultrasound (TRUS) Fusion Biopsy and Systematic TRUS Biopsy. Eur Urol Oncol, 2018. 1: 109.
https://pubmed.ncbi.nlm.nih.gov/31100233
246. Kim, L., et al. Clinical utility and cost modelling of the phi test to triage referrals into image-based
diagnostic services for suspected prostate cancer: the PRIM (Phi to RefIne Mri) study. BMC Med,
2020. 18: 95.
https://pubmed.ncbi.nlm.nih.gov/32299423
247. Gronberg, H., et al. Prostate Cancer Diagnostics Using a Combination of the Stockholm3 Blood Test
and Multiparametric Magnetic Resonance Imaging. Eur Urol, 2018. 74: 722.
https://pubmed.ncbi.nlm.nih.gov/30001824
248. Dell’Oglio, P., et al. Impact of multiparametric MRI and MRI-targeted biopsy on pre-therapeutic risk
assessment in prostate cancer patients candidate for radical prostatectomy. World J Urol, 2019.
37: 221.
https://pubmed.ncbi.nlm.nih.gov/29948044

148 PROSTATE CANCER - LIMITED UPDATE 2021


249. Woo, S., et al. Prognostic Value of Pretreatment MRI in Patients With Prostate Cancer Treated With
Radiation Therapy: A Systematic Review and Meta-Analysis. AJR Am J Roentgenol, 2020. 214: 597.
https://pubmed.ncbi.nlm.nih.gov/31799874
250. Faiena, I., et al. PI-RADS Version 2 Category on 3 Tesla Multiparametric Prostate Magnetic
Resonance Imaging Predicts Oncologic Outcomes in Gleason 3 + 4 Prostate Cancer on Biopsy.
J Urol, 2019. 201: 91.
https://pubmed.ncbi.nlm.nih.gov/30142318
251. Houlahan, K.E., et al. Molecular Hallmarks of Multiparametric Magnetic Resonance Imaging Visibility
in Prostate Cancer. Eur Urol, 2019. 76: 18.
https://pubmed.ncbi.nlm.nih.gov/30685078
252. Lam, T.B.L., et al. EAU-EANM-ESTRO-ESUR-SIOG Prostate Cancer Guideline Panel Consensus
Statements for Deferred Treatment with Curative Intent for Localised Prostate Cancer from an
International Collaborative Study (DETECTIVE Study). Eur Urol, 2019. 76: 790.
https://pubmed.ncbi.nlm.nih.gov/31587989
253. Roobol, M.J., et al. A risk-based strategy improves prostate-specific antigen-driven detection of
prostate cancer. Eur Urol, 2010. 57: 79.
https://pubmed.ncbi.nlm.nih.gov/19733959
254. Eastham, J.A., et al. Variation of serum prostate-specific antigen levels: an evaluation of year-to-year
fluctuations. JAMA, 2003. 289: 2695.
https://pubmed.ncbi.nlm.nih.gov/12771116
255. Stephan, C., et al. Interchangeability of measurements of total and free prostate-specific antigen in
serum with 5 frequently used assay combinations: an update. Clin Chem, 2006. 52: 59.
https://pubmed.ncbi.nlm.nih.gov/16391327
256. Eggener, S.E., et al. Empiric antibiotics for an elevated prostate-specific antigen (PSA) level: a
randomised, prospective, controlled multi-institutional trial. BJU Int, 2013. 112: 925.
https://pubmed.ncbi.nlm.nih.gov/23890317
257. Xue, J., et al. Comparison between transrectal and transperineal prostate biopsy for detection of
prostate cancer: a meta-analysis and trial sequential analysis. Oncotarget, 2017. 8: 23322.
https://pubmed.ncbi.nlm.nih.gov/28177897
258. Roberts, M.J., et al. Prostate Biopsy-related Infection: A Systematic Review of Risk Factors,
Prevention Strategies, and Management Approaches. Urology, 2017. 104: 11.
https://pubmed.ncbi.nlm.nih.gov/28007492
259. Pilatz, A., et al. Update on Strategies to Reduce Infectious Complications After Prostate Biopsy. Eur
Urol Focus, 2019. 5: 20.
https://pubmed.ncbi.nlm.nih.gov/30503175
260. Zigeuner, R., et al. Detection of prostate cancer by TURP or open surgery in patients with previously
negative transrectal prostate biopsies. Urology, 2003. 62: 883.
https://pubmed.ncbi.nlm.nih.gov/14624913
261. Guo, C.C., et al. Intraductal carcinoma of the prostate on needle biopsy: Histologic features and
clinical significance. Mod Pathol, 2006. 19: 1528.
https://pubmed.ncbi.nlm.nih.gov/16980940
262. Ericson, K.J., et al. Prostate cancer detection following diagnosis of atypical small acinar
proliferation. Can J Urol, 2017. 24: 8714.
https://pubmed.ncbi.nlm.nih.gov/28436357
263. Walz, J., et al. High incidence of prostate cancer detected by saturation biopsy after previous
negative biopsy series. Eur Urol, 2006. 50: 498.
https://pubmed.ncbi.nlm.nih.gov/16631303
264. Moran, B.J., et al. Re-biopsy of the prostate using a stereotactic transperineal technique. J Urol,
2006. 176: 1376.
https://pubmed.ncbi.nlm.nih.gov/16952636
265. Nakai, Y., et al. Transperineal template-guided saturation biopsy aimed at sampling one core for each
milliliter of prostate volume: 103 cases requiring repeat prostate biopsy. BMC Urol, 2017. 17: 28.
https://pubmed.ncbi.nlm.nih.gov/28381267
266. Ekwueme, K., et al. Transperineal template-guided saturation biopsy using a modified technique:
outcome of 270 cases requiring repeat prostate biopsy. BJU Int, 2013. 111: E365.
https://pubmed.ncbi.nlm.nih.gov/23714648
267. Pepdjonovic, L., et al. Zero hospital admissions for infection after 577 transperineal prostate
biopsies using single-dose cephazolin prophylaxis. World J Urol, 2017. 35: 1199.
https://pubmed.ncbi.nlm.nih.gov/27987032

PROSTATE CANCER - LIMITED UPDATE 2021 149


268. Donovan, J., et al. Prostate Testing for Cancer and Treatment (ProtecT) feasibility study. Health
Technol Assess, 2003. 7: 1.
https://pubmed.ncbi.nlm.nih.gov/12709289
269. Eichler, K., et al. Diagnostic value of systematic biopsy methods in the investigation of prostate
cancer: a systematic review. J Urol, 2006. 175: 1605.
https://pubmed.ncbi.nlm.nih.gov/16600713
270. Shariat, S.F., et al. Using biopsy to detect prostate cancer. Rev Urol, 2008. 10: 262.
https://pubmed.ncbi.nlm.nih.gov/19145270
271. Kuru, T.H., et al. Definitions of terms, processes and a minimum dataset for transperineal
prostate biopsies: a standardization approach of the Ginsburg Study Group for Enhanced Prostate
Diagnostics. BJU Int, 2013. 112: 568.
https://pubmed.ncbi.nlm.nih.gov/23773772
272. Vyas, L., et al. Indications, results and safety profile of transperineal sector biopsies (TPSB) of the
prostate: a single centre experience of 634 cases. BJU Int, 2014. 114: 32.
https://pubmed.ncbi.nlm.nih.gov/24053629
273. Abdollah, F., et al. Trans-rectal versus trans-perineal saturation rebiopsy of the prostate: is there a
difference in cancer detection rate? Urology, 2011. 77: 921.
https://pubmed.ncbi.nlm.nih.gov/21131034
274. Ahmed, H.U., et al. Diagnostic accuracy of multi-parametric MRI and TRUS biopsy in prostate
cancer (PROMIS): a paired validating confirmatory study. Lancet, 2017. 389: 815.
https://pubmed.ncbi.nlm.nih.gov/28110982
275. Tschirdewahn, S., et al. Detection of Significant Prostate Cancer Using Target Saturation in
Transperineal Magnetic Resonance Imaging/Transrectal Ultrasonography-fusion Biopsy. Eur Urol
Focus, 2020.
https://pubmed.ncbi.nlm.nih.gov/32660838
276. Wegelin, O., et al. Comparing Three Different Techniques for Magnetic Resonance Imaging-targeted
Prostate Biopsies: A Systematic Review of In-bore versus Magnetic Resonance Imaging-transrectal
Ultrasound fusion versus Cognitive Registration. Is There a Preferred Technique? Eur Urol, 2017.
71: 517.
https://pubmed.ncbi.nlm.nih.gov/27568655
277. Hamid, S., et al. The SmartTarget Biopsy Trial: A Prospective, Within-person Randomised, Blinded
Trial Comparing the Accuracy of Visual-registration and Magnetic Resonance Imaging/Ultrasound
Image-fusion Targeted Biopsies for Prostate Cancer Risk Stratification. Eur Urol, 2019. 75: 733.
https://pubmed.ncbi.nlm.nih.gov/30527787
278. Simmons, L.A.M., et al. Accuracy of Transperineal Targeted Prostate Biopsies, Visual Estimation and
Image Fusion in Men Needing Repeat Biopsy in the PICTURE Trial. J Urol, 2018. 200: 1227.
https://pubmed.ncbi.nlm.nih.gov/30017964
279. Watts, K.L., et al. Systematic review and meta-analysis comparing cognitive vs. image-guided
fusion prostate biopsy for the detection of prostate cancer. Urol Oncol, 2020. 38: 734.e19.
https://pubmed.ncbi.nlm.nih.gov/32321689
280. Tu, X., et al. Transperineal Magnetic Resonance Imaging-Targeted Biopsy May Perform Better Than
Transrectal Route in the Detection of Clinically Significant Prostate Cancer: Systematic Review and
Meta-analysis. Clin Genitourin Cancer, 2019. 17: e860.
https://pubmed.ncbi.nlm.nih.gov/31281065
281. Cerruto, M.A., et al. Transrectal versus transperineal 14-core prostate biopsy in detection of prostate
cancer: a comparative evaluation at the same institution. Arch Ital Urol Androl, 2014. 86: 284.
https://pubmed.ncbi.nlm.nih.gov/25641452
282. Chae, Y., et al. The comparison between transperineal and transrectal ultrasound-guided prostate
needle biopsy. Korean J Urol, 2009. 50: 119.
https://www.cochranelibrary.com/central/doi/10.1002/central/CN-00753703/full
283. Guo, L.H., et al. Comparison between Ultrasound Guided Transperineal and Transrectal Prostate
Biopsy: A Prospective, Randomized, and Controlled Trial. Sci Rep, 2015. 5: 16089.
https://pubmed.ncbi.nlm.nih.gov/26526558
284. Hara, R., et al. Prostatic biopsy at Kawasaki Medical School: A prospective study of the results of
transperineal biopsy over the past 13 years and the results of systematic 12-site biopsy using the
transperineal and transrectal methods. Nishinihon J Urol, 2006. 68: 403.
https://www.researchgate.net/publication/289682762
285. Singh, S., A., et al. Comparison of infective complications in Transperineal versus Transrectal
Ultrasound Guided Prostatic Biopsy in patients suspected to have prostate cancer. Indian J Urol,
2017. 33: S43. [No abstract available].

150 PROSTATE CANCER - LIMITED UPDATE 2021


286. Udeh, E.I., et al. Transperineal versus transrectal prostate biopsy: our findings in a tertiary health
institution. Niger J Clin Pract, 2015. 18: 110.
https://pubmed.ncbi.nlm.nih.gov/25511354
287. Wegelin, O., et al. Complications and Adverse Events of Three Magnetic Resonance Imaging-based
Target Biopsy Techniques in the Diagnosis of Prostate Cancer Among Men with Prior Negative
Biopsies: Results from the FUTURE Trial, a Multicentre Randomised Controlled Trial. Eur Urol Oncol,
2019. 2: 617.
https://pubmed.ncbi.nlm.nih.gov/31519516
288. Pradere, B., et al. Non-antibiotic Strategies for the Prevention of Infectious Complications
following Prostate Biopsy: A Systematic Review and Meta-Analysis. J Urol, 2020:
101097ju0000000000001399.
https://pubmed.ncbi.nlm.nih.gov/33026903
289. Bennett, H.Y., et al. The global burden of major infectious complications following prostate biopsy.
Epidemiol Infect, 2016. 144: 1784.
https://pubmed.ncbi.nlm.nih.gov/26645476
290. Berry, B., et al. Comparison of complications after transrectal and transperineal prostate biopsy: a
national population-based study. BJU Int, 2020. 126: 97.
https://pubmed.ncbi.nlm.nih.gov/32124525
291. Pepe, P., et al. Morbidity after transperineal prostate biopsy in 3000 patients undergoing 12 vs 18 vs
more than 24 needle cores. Urology, 2013. 81: 1142.
https://pubmed.ncbi.nlm.nih.gov/23726443
292. Stefanova, V., et al. Transperineal Prostate Biopsies Using Local Anesthesia: Experience with 1,287
Patients. Prostate Cancer Detection Rate, Complications and Patient Tolerability. J Urol, 2019.
201: 1121.
https://pubmed.ncbi.nlm.nih.gov/30835607
293. Baba, K., et al. Assessment of antimicrobiral prophylaxis in transperineal prostate biopsy: A single-
center retrospective study of 485 cases. J Infect Chemother, 2018. 24: 637.
https://pubmed.ncbi.nlm.nih.gov/29685852
294. Abughosh, Z., et al. A prospective randomized trial of povidone-iodine prophylactic cleansing of the
rectum before transrectal ultrasound guided prostate biopsy. J Urol, 2013. 189: 1326.
https://pubmed.ncbi.nlm.nih.gov/23041343
295. Ghafoori, M., et al. Decrease in infection rate following use of povidone-iodine during transrectal
ultrasound guided biopsy of the prostate: a double blind randomized clinical trial. Iran J Radiol,
2012. 9: 67.
https://pubmed.ncbi.nlm.nih.gov/23329966
296. Kanjanawongdeengam, P., et al. Reduction in bacteremia rates after rectum sterilization before
transrectal, ultrasound-guided prostate biopsy: a randomized controlled trial. J Med Assoc Thai,
2009. 92: 1621.
https://pubmed.ncbi.nlm.nih.gov/20043564
297. Melekos, M.D. Efficacy of prophylactic antimicrobial regimens in preventing infectious complications
after transrectal biopsy of the prostate. Int Urol Nephrol, 1990. 22: 257.
https://pubmed.ncbi.nlm.nih.gov/2210982
298. Sharpe, J.R., et al. Urinary tract infection after transrectal needle biopsy of the prostate. J Urol,
1982. 127: 255.
https://pubmed.ncbi.nlm.nih.gov/7062377
299. Brown, R.W., et al. Bacteremia and bacteriuria after transrectal prostatic biopsy. Urology, 1981.
18: 145.
https://pubmed.ncbi.nlm.nih.gov/7269016
300. Taher, Y., et al. A. Prospective randomized controlled study to assess the effect of perineal region
cleansing with povidone iodine before transrectal needle biopsy of the prostate on infectious
complications. Urology 2014. 84: S171.
https://www.cochranelibrary.com/central/doi/10.1002/central/CN-01023606/full
301. Yu, L., et al. [Impact of insertion timing of iodophor cotton ball on the control of infection
complications after transrectal ultrasound guided prostate biopsy]. Zhonghua Yi Xue Za Zhi, 2014.
94: 609.
https://pubmed.ncbi.nlm.nih.gov/24762693
302. Tekdoǧan, Ü., et al. The efficiency of prophylactic antibiotic treatment in patients without risk factor
who underwent transrectal. Turk Uroloji Dergisi, 2006. 32: 261.
https://www.researchgate.net/publication/289651865

PROSTATE CANCER - LIMITED UPDATE 2021 151


303. Wang, H., et al. [Investigation of infection risk and the value of antibiotic prophylaxis during
transrectal biopsy of the prostate by endotoxin determination]. Zhonghua Nan Ke Xue, 2004.
10: 496.
https://pubmed.ncbi.nlm.nih.gov/15354517
304. Lindert, K.A., et al. Bacteremia and bacteriuria after transrectal ultrasound guided prostate biopsy.
J Urol, 2000. 164: 76.
https://pubmed.ncbi.nlm.nih.gov/10840428
305. Pilatz, A., et al. Antibiotic Prophylaxis for the Prevention of Infectious Complications following
Prostate Biopsy: A Systematic Review and Meta-Analysis. J Urol, 2020. 204: 224.
https://pubmed.ncbi.nlm.nih.gov/32105195
306. European Medicine Agency. Disabling and potentially permanent side effects lead to suspension or
restrictions of quinolone and fluoroquinolone antibiotics. 2019 [access date March 2021].
https://www.ema.europa.eu/en/documents/referral/quinolone-fluoroquinolone-article-31-referral-
disabling-potentially-permanent-side-effects-lead_en.pdf
307. Carignan, A., et al. Effectiveness of fosfomycin tromethamine prophylaxis in preventing infection
following transrectal ultrasound-guided prostate needle biopsy: Results from a large Canadian
cohort. J Glob Antimicrob Resist, 2019. 17: 112.
https://pubmed.ncbi.nlm.nih.gov/30553114
308. Pilatz, A., et al. European Association of Urology Position Paper on the Prevention of Infectious
Complications Following Prostate Biopsy. Eur Urol, 2021. 79: 11.
309. von Knobloch, R., et al. Bilateral fine-needle administered local anaesthetic nerve block for pain
control during TRUS-guided multi-core prostate biopsy: a prospective randomised trial. Eur Urol,
2002. 41: 508.
https://pubmed.ncbi.nlm.nih.gov/12074792
310. Adamakis, I., et al. Pain during transrectal ultrasonography guided prostate biopsy: a randomized
prospective trial comparing periprostatic infiltration with lidocaine with the intrarectal instillation of
lidocaine-prilocain cream. World J Urol, 2004. 22: 281.
https://pubmed.ncbi.nlm.nih.gov/14689224
311. Bass, E.J., et al. Magnetic resonance imaging targeted transperineal prostate biopsy: a local
anaesthetic approach. Prostate Cancer Prostatic Dis, 2017. 20: 311.
https://pubmed.ncbi.nlm.nih.gov/28485391
312. Xiang, J., et al. Transperineal versus transrectal prostate biopsy in the diagnosis of prostate cancer:
a systematic review and meta-analysis. World J Surg Oncol, 2019. 17: 31.
https://wjso.biomedcentral.com/articles/10.1186/s12957-019-1573-0
313. Iremashvili, V.V., et al. Periprostatic local anesthesia with pudendal block for transperineal
ultrasound-guided prostate biopsy: a randomized trial. Urology, 2010. 75: 1023.
https://pubmed.ncbi.nlm.nih.gov/20080288
314. Meyer, A.R., et al. Initial Experience Performing In-office Ultrasound-guided Transperineal Prostate
Biopsy Under Local Anesthesia Using the PrecisionPoint Transperineal Access System. Urology,
2018. 115: 8.
https://pubmed.ncbi.nlm.nih.gov/29409845
315. Kum, F., et al. Initial outcomes of local anaesthetic freehand transperineal prostate biopsies in the
outpatient setting. BJU Int, 2020. 125: 244.
https://pubmed.ncbi.nlm.nih.gov/30431694
316. NCCN Clinical practice Guidelines in Oncology®: Prostate Cancer Early Detection, Version 1. 2021
https://www.nccn.org/professionals/physician_gls/pdf/prostate_detection.pdf
317. Giannarini, G., et al. Continuing or discontinuing low-dose aspirin before transrectal prostate biopsy:
results of a prospective randomized trial. Urology, 2007. 70: 501.
https://pubmed.ncbi.nlm.nih.gov/17688919
318. Garcia C., et al. Does transperineal prostate biopsy reduce complications compared with transrectal
biopsy? a systematic review and meta-analysis of randomised controlled trials. 2016. 195:4 SUPPL.
1 p. e328.
https://www.jurology.com/article/S0022-5347(16)03167-0/pdf
319. Linzer, D.G., et al. Seminal vesicle biopsy: accuracy and implications for staging of prostate cancer.
Urology, 1996. 48: 757.
https://www.auajournals.org/doi/full/10.1016/j.juro.2016.02.2879
320. Pelzer, A.E., et al. Are transition zone biopsies still necessary to improve prostate cancer detection?
Results from the tyrol screening project. Eur Urol, 2005. 48: 916.
https://pubmed.ncbi.nlm.nih.gov/16126324

152 PROSTATE CANCER - LIMITED UPDATE 2021


321. Iczkowski, K.A., et al. Needle core length in sextant biopsy influences prostate cancer detection
rate. Urology, 2002. 59: 698.
https://pubmed.ncbi.nlm.nih.gov/11992843
322. Van der Kwast, T., et al. Guidelines on processing and reporting of prostate biopsies: the 2013
update of the pathology committee of the European Randomized Study of Screening for Prostate
Cancer (ERSPC). Virchows Arch, 2013. 463: 367.
https://pubmed.ncbi.nlm.nih.gov/23918245
323. Rogatsch, H., et al. Diagnostic effect of an improved preembedding method of prostate needle
biopsy specimens. Hum Pathol, 2000. 31: 1102.
https://pubmed.ncbi.nlm.nih.gov/11014578
324. Novis, D.A., et al. Diagnostic uncertainty expressed in prostate needle biopsies. A College of
American Pathologists Q-probes Study of 15,753 prostate needle biopsies in 332 institutions. Arch
Pathol Lab Med, 1999. 123: 687.
https://pubmed.ncbi.nlm.nih.gov/10420224
325. Iczkowski, K.A. Current prostate biopsy interpretation: criteria for cancer, atypical small acinar
proliferation, high-grade prostatic intraepithelial neoplasia, and use of immunostains. Arch Pathol
Lab Med, 2006. 130: 835.
https://pubmed.ncbi.nlm.nih.gov/16740037
326. Reyes, A.O., et al. Diagnostic effect of complete histologic sampling of prostate needle biopsy
specimens. Am J Clin Pathol, 1998. 109: 416.
https://pubmed.ncbi.nlm.nih.gov/9535395
327. Gordetsky, J.B., et al. Histologic findings associated with false-positive multiparametric magnetic
resonance imaging performed for prostate cancer detection. Hum Pathol, 2019. 83: 159.
https://pubmed.ncbi.nlm.nih.gov/30179687
328. Sauter, G., et al. Clinical Utility of Quantitative Gleason Grading in Prostate Biopsies and
Prostatectomy Specimens. Eur Urol, 2016. 69: 592.
https://pubmed.ncbi.nlm.nih.gov/26542947
329. Cole, A.I., et al. Prognostic Value of Percent Gleason Grade 4 at Prostate Biopsy in Predicting
Prostatectomy Pathology and Recurrence. J Urol, 2016. 196: 405.
https://pubmed.ncbi.nlm.nih.gov/26920466
330. Kweldam, C.F., et al. Disease-specific survival of patients with invasive cribriform and intraductal
prostate cancer at diagnostic biopsy. Mod Pathol, 2016. 29: 630.
https://pubmed.ncbi.nlm.nih.gov/26939875
331. Saeter, T., et al. Intraductal Carcinoma of the Prostate on Diagnostic Needle Biopsy Predicts
Prostate Cancer Mortality: A Population-Based Study. Prostate, 2017. 77: 859.
https://pubmed.ncbi.nlm.nih.gov/28240424
332. Sebo, T.J., et al. Predicting prostate carcinoma volume and stage at radical prostatectomy by
assessing needle biopsy specimens for percent surface area and cores positive for carcinoma,
perineural invasion, Gleason score, DNA ploidy and proliferation, and preoperative serum prostate
specific antigen: a report of 454 cases. Cancer, 2001. 91: 2196.
https://pubmed.ncbi.nlm.nih.gov/11391602
333. Grossklaus, D.J., et al. Percent of cancer in the biopsy set predicts pathological findings after
prostatectomy. J Urol, 2002. 167: 2032.
https://pubmed.ncbi.nlm.nih.gov/11956432
334. Freedland, S.J., et al. Preoperative model for predicting prostate specific antigen recurrence after
radical prostatectomy using percent of biopsy tissue with cancer, biopsy Gleason grade and serum
prostate specific antigen. J Urol, 2004. 171: 2215.
https://pubmed.ncbi.nlm.nih.gov/15126788
335. Brimo, F., et al. Prognostic value of various morphometric measurements of tumour extent in
prostate needle core tissue. Histopathology, 2008. 53: 177.
https://pubmed.ncbi.nlm.nih.gov/18752501
336. Bangma, C.H., et al. Active surveillance for low-risk prostate cancer. Crit Rev Oncol Hematol, 2013.
85: 295.
https://pubmed.ncbi.nlm.nih.gov/22878262
337. Eggener, S.E., et al. Molecular Biomarkers in Localized Prostate Cancer: ASCO Guideline. J Clin
Oncol, 2019: JCO1902768.
https://pubmed.ncbi.nlm.nih.gov/31829902
338. Sehdev, A.E., et al. Comparative analysis of sampling methods for grossing radical prostatectomy
specimens performed for nonpalpable (stage T1c) prostatic adenocarcinoma. Hum Pathol, 2001.
32: 494.
https://pubmed.ncbi.nlm.nih.gov/11381367

PROSTATE CANCER - LIMITED UPDATE 2021 153


339. Ruijter, E.T., et al. Rapid microwave-stimulated fixation of entire prostatectomy specimens. Biomed-
II MPC Study Group. J Pathol, 1997. 183: 369.
https://pubmed.ncbi.nlm.nih.gov/9422995
340. Chan, N.G., et al. Pathological reporting of colorectal cancer specimens: a retrospective survey in
an academic Canadian pathology department. Can J Surg, 2008. 51: 284.
https://pubmed.ncbi.nlm.nih.gov/18815652
341. Partin, A.W., et al. Contemporary update of prostate cancer staging nomograms (Partin Tables) for
the new millennium. Urology, 2001. 58: 843.
https://pubmed.ncbi.nlm.nih.gov/11744442
342. Harnden, P., et al. Should the Gleason grading system for prostate cancer be modified to account for
high-grade tertiary components? A systematic review and meta-analysis. Lancet Oncol, 2007. 8: 411.
https://pubmed.ncbi.nlm.nih.gov/17466898
343. Magi-Galluzzi, C., et al. International Society of Urological Pathology (ISUP) Consensus Conference
on Handling and Staging of Radical Prostatectomy Specimens. Working group 3: extraprostatic
extension, lymphovascular invasion and locally advanced disease. Mod Pathol, 2011. 24: 26.
https://pubmed.ncbi.nlm.nih.gov/20802467
344. Epstein, J.I., et al. Influence of capsular penetration on progression following radical prostatectomy:
a study of 196 cases with long-term followup. J Urol, 1993. 150: 135.
https://pubmed.ncbi.nlm.nih.gov/7685422
345. Marks, R.A., et al. The relationship between the extent of surgical margin positivity and prostate
specific antigen recurrence in radical prostatectomy specimens. Hum Pathol, 2007. 38: 1207.
https://pubmed.ncbi.nlm.nih.gov/17490720
346. Sung, M.T., et al. Radial distance of extraprostatic extension measured by ocular micrometer is an
independent predictor of prostate-specific antigen recurrence: A new proposal for the substaging of
pT3a prostate cancer. Am J Surg Pathol, 2007. 31: 311.
https://pubmed.ncbi.nlm.nih.gov/17255778
347. Aydin, H., et al. Positive proximal (bladder neck) margin at radical prostatectomy confers greater risk
of biochemical progression. Urology, 2004. 64: 551.
https://pubmed.ncbi.nlm.nih.gov/15351591
348. Ploussard, G., et al. The prognostic significance of bladder neck invasion in prostate cancer: is
microscopic involvement truly a T4 disease? BJU Int, 2010. 105: 776.
https://pubmed.ncbi.nlm.nih.gov/19863529
349. Hoedemaeker, R.F., et al. Staging prostate cancer. Microsc Res Tech, 2000. 51: 423.
https://pubmed.ncbi.nlm.nih.gov/11074612
350. Stamey, T.A., et al. Prostate cancer is highly predictable: a prognostic equation based on all
morphological variables in radical prostatectomy specimens. J Urol, 2000. 163: 1155.
https://pubmed.ncbi.nlm.nih.gov/10737486
351. Epstein, J.I., et al. Prognostic factors and reporting of prostate carcinoma in radical prostatectomy
and pelvic lymphadenectomy specimens. Scand J Urol Nephrol Suppl, 2005: 34.
https://pubmed.ncbi.nlm.nih.gov/16019758
352. Kikuchi, E., et al. Is tumor volume an independent prognostic factor in clinically localized prostate
cancer? J Urol, 2004. 172: 508.
https://pubmed.ncbi.nlm.nih.gov/15247716
353. van Oort, I.M., et al. Maximum tumor diameter is not an independent prognostic factor in high-risk
localized prostate cancer. World J Urol, 2008. 26: 237.
https://pubmed.ncbi.nlm.nih.gov/18265988
354. van der Kwast, T.H., et al. International Society of Urological Pathology (ISUP) Consensus
Conference on Handling and Staging of Radical Prostatectomy Specimens. Working group 2: T2
substaging and prostate cancer volume. Mod Pathol, 2011. 24: 16.
https://pubmed.ncbi.nlm.nih.gov/20818340
355. Evans, A.J., et al. Interobserver variability between expert urologic pathologists for extraprostatic
extension and surgical margin status in radical prostatectomy specimens. Am J Surg Pathol, 2008.
32: 1503.
https://pubmed.ncbi.nlm.nih.gov/18708939
356. Chuang, A.Y., et al. Positive surgical margins in areas of capsular incision in otherwise organ-
confined disease at radical prostatectomy: histologic features and pitfalls. Am J Surg Pathol, 2008.
32: 1201.
https://pubmed.ncbi.nlm.nih.gov/18580493

154 PROSTATE CANCER - LIMITED UPDATE 2021


357. Sammon, J.D., et al. Risk factors for biochemical recurrence following radical perineal
prostatectomy in a large contemporary series: a detailed assessment of margin extent and location.
Urol Oncol, 2013. 31: 1470.
https://pubmed.ncbi.nlm.nih.gov/22534086
358. Paner, G.P., et al. Updates in the Eighth Edition of the Tumor-Node-Metastasis Staging Classification
for Urologic Cancers. Eur Urol, 2018. 73: 560.
https://pubmed.ncbi.nlm.nih.gov/29325693
359. Smith, J.A., Jr., et al. Transrectal ultrasound versus digital rectal examination for the staging of
carcinoma of the prostate: results of a prospective, multi-institutional trial. J Urol, 1997. 157: 902.
https://pubmed.ncbi.nlm.nih.gov/9072596
360. Mitterberger, M., et al. The value of three-dimensional transrectal ultrasonography in staging
prostate cancer. BJU Int, 2007. 100: 47.
https://pubmed.ncbi.nlm.nih.gov/17433033
361. Sauvain, J.L., et al. Value of power doppler and 3D vascular sonography as a method for diagnosis
and staging of prostate cancer. Eur Urol, 2003. 44: 21.
https://pubmed.ncbi.nlm.nih.gov/12814671
362. de Rooij, M., et al. Accuracy of Magnetic Resonance Imaging for Local Staging of Prostate Cancer:
A Diagnostic Meta-analysis. Eur Urol, 2016. 70: 233.
https://pubmed.ncbi.nlm.nih.gov/26215604
363. Jager, G.J., et al. Local staging of prostate cancer with endorectal MR imaging: correlation with
histopathology. AJR Am J Roentgenol, 1996. 166: 845.
https://pubmed.ncbi.nlm.nih.gov/8610561
364. Cornud, F., et al. Extraprostatic spread of clinically localized prostate cancer: factors predictive of
pT3 tumor and of positive endorectal MR imaging examination results. Radiology, 2002. 224: 203.
https://pubmed.ncbi.nlm.nih.gov/12091684
365. Heijmink, S.W., et al. Prostate cancer: body-array versus endorectal coil MR imaging at
3 T--comparison of image quality, localization, and staging performance. Radiology, 2007. 244: 184.
https://pubmed.ncbi.nlm.nih.gov/17495178
366. Futterer, J.J., et al. Staging prostate cancer with dynamic contrast-enhanced endorectal MR
imaging prior to radical prostatectomy: experienced versus less experienced readers. Radiology,
2005. 237: 541.
https://pubmed.ncbi.nlm.nih.gov/16244263
367. Gandaglia, G., et al. The Key Combined Value of Multiparametric Magnetic Resonance Imaging, and
Magnetic Resonance Imaging-targeted and Concomitant Systematic Biopsies for the Prediction of
Adverse Pathological Features in Prostate Cancer Patients Undergoing Radical Prostatectomy. Eur
Urol, 2020. 77: 733.
https://pubmed.ncbi.nlm.nih.gov/31547938
368. Wang, L., et al. Prostate cancer: incremental value of endorectal MR imaging findings for prediction
of extracapsular extension. Radiology, 2004. 232: 133.
https://pubmed.ncbi.nlm.nih.gov/15166321
369. D’Amico, A.V., et al. Endorectal magnetic resonance imaging as a predictor of biochemical outcome
after radical prostatectomy in men with clinically localized prostate cancer. J Urol, 2000. 164: 759.
https://pubmed.ncbi.nlm.nih.gov/10953141
370. Engelbrecht, M.R., et al. Patient selection for magnetic resonance imaging of prostate cancer. Eur
Urol, 2001. 40: 300.
https://pubmed.ncbi.nlm.nih.gov/11684846
371. Abuzallouf, S., et al. Baseline staging of newly diagnosed prostate cancer: a summary of the
literature. J Urol, 2004. 171: 2122.
https://pubmed.ncbi.nlm.nih.gov/15126770
372. Kiss, B., et al. Current Status of Lymph Node Imaging in Bladder and Prostate Cancer. Urology,
2016. 96: 1.
https://pubmed.ncbi.nlm.nih.gov/26966038
373. Harisinghani, M.G., et al. Noninvasive detection of clinically occult lymph-node metastases in
prostate cancer. N Engl J Med, 2003. 348: 2491.
https://pubmed.ncbi.nlm.nih.gov/12815134
374. Hovels, A.M., et al. The diagnostic accuracy of CT and MRI in the staging of pelvic lymph nodes in
patients with prostate cancer: a meta-analysis. Clin Radiol, 2008. 63: 387.
https://pubmed.ncbi.nlm.nih.gov/18325358
375. Gabriele, D., et al. Is there still a role for computed tomography and bone scintigraphy in prostate
cancer staging? An analysis from the EUREKA-1 database. World J Urol, 2016. 34: 517.
https://pubmed.ncbi.nlm.nih.gov/26276152

PROSTATE CANCER - LIMITED UPDATE 2021 155


376. Flanigan, R.C., et al. Limited efficacy of preoperative computed tomographic scanning for the
evaluation of lymph node metastasis in patients before radical prostatectomy. Urology, 1996. 48: 428.
https://pubmed.ncbi.nlm.nih.gov/8804497
377. Tiguert, R., et al. Lymph node size does not correlate with the presence of prostate cancer
metastasis. Urology, 1999. 53: 367.
https://pubmed.ncbi.nlm.nih.gov/9933056
378. Spevack, L., et al. Predicting the patient at low risk for lymph node metastasis with localized
prostate cancer: an analysis of four statistical models. Int J Radiat Oncol Biol Phys, 1996. 34: 543.
https://pubmed.ncbi.nlm.nih.gov/8621276
379. Thoeny, H.C., et al. Metastases in normal-sized pelvic lymph nodes: detection with diffusion-
weighted MR imaging. Radiology, 2014. 273: 125.
https://pubmed.ncbi.nlm.nih.gov/24893049
380. Memorial Sloan Kettering Cancer Center. Dynamic Prostate Cancer Nomogram: Coefficients. 2020.
https://www.mskcc.org/nomograms/prostate/pre_op/coefficients
381. Briganti, A., et al. Updated nomogram predicting lymph node invasion in patients with prostate
cancer undergoing extended pelvic lymph node dissection: the essential importance of percentage
of positive cores. Eur Urol, 2012. 61: 480.
https://pubmed.ncbi.nlm.nih.gov/22078338
382. Gandaglia, G., et al. Development and Internal Validation of a Novel Model to Identify the
Candidates for Extended Pelvic Lymph Node Dissection in Prostate Cancer. Eur Urol, 2017. 72: 632.
https://pubmed.ncbi.nlm.nih.gov/28412062
383. Draulans, C., et al. Development and External Validation of a Multiparametric Magnetic Resonance
Imaging and International Society of Urological Pathology Based Add-On Prediction Tool to Identify
Prostate Cancer Candidates for Pelvic Lymph Node Dissection. J Urol, 2020. 203: 713.
https://pubmed.ncbi.nlm.nih.gov/31718396
384. von Eyben, F.E., et al. Meta-analysis of (11)C-choline and (18)F-choline PET/CT for management of
patients with prostate cancer. Nucl Med Commun, 2014. 35: 221.
https://pubmed.ncbi.nlm.nih.gov/24240194
385. Van den Bergh, L., et al. Final analysis of a prospective trial on functional imaging for nodal staging in
patients with prostate cancer at high risk for lymph node involvement. Urol Oncol, 2015. 33: 109 e23.
https://pubmed.ncbi.nlm.nih.gov/25655681
386. Schiavina, R., et al. Preoperative Staging With (11)C-Choline PET/CT Is Adequately Accurate in
Patients With Very High-Risk Prostate Cancer. Clin Genitourin Cancer, 2018. 16: 305.
https://pubmed.ncbi.nlm.nih.gov/29859737
387. Pinaquy, J.B., et al. Comparative effectiveness of [(18) F]-fluorocholine PET-CT and pelvic MRI with
diffusion-weighted imaging for staging in patients with high-risk prostate cancer. Prostate, 2015.
75: 323.
https://pubmed.ncbi.nlm.nih.gov/25393215
388. Heck, M.M., et al. Prospective comparison of computed tomography, diffusion-weighted magnetic
resonance imaging and [11C]choline positron emission tomography/computed tomography for
preoperative lymph node staging in prostate cancer patients. Eur J Nucl Med Mol Imaging, 2014.
41: 694.
https://pubmed.ncbi.nlm.nih.gov/24297503
389. Budiharto, T., et al. Prospective evaluation of 11C-choline positron emission tomography/computed
tomography and diffusion-weighted magnetic resonance imaging for the nodal staging of prostate
cancer with a high risk of lymph node metastases. Eur Urol, 2011. 60: 125.
https://pubmed.ncbi.nlm.nih.gov/21292388
390. Maurer, T., et al. Current use of PSMA-PET in prostate cancer management. Nat Rev Urol, 2016.
13: 226.
https://pubmed.ncbi.nlm.nih.gov/26902337
391. Dias, A.H., et al. Prostate-Specific Membrane Antigen PET/CT: Uptake in Lymph Nodes With Active
Sarcoidosis. Clin Nucl Med, 2017. 42: e175.
https://pubmed.ncbi.nlm.nih.gov/28045734
392. Froehner, M., et al. PSMA-PET/CT-Positive Paget Disease in a Patient with Newly Diagnosed
Prostate Cancer: Imaging and Bone Biopsy Findings. Case Rep Urol, 2017. 2017: 1654231.
https://pubmed.ncbi.nlm.nih.gov/28396816
393. Jochumsen, M.R., et al. Benign Traumatic Rib Fracture: A Potential Pitfall on 68Ga-Prostate-Specific
Membrane Antigen PET/CT for Prostate Cancer. Clin Nucl Med, 2018. 43: 38.
https://pubmed.ncbi.nlm.nih.gov/29076907

156 PROSTATE CANCER - LIMITED UPDATE 2021


394. Werner, R.A., et al. (18)F-Labeled, PSMA-Targeted Radiotracers: Leveraging the Advantages of
Radiofluorination for Prostate Cancer Molecular Imaging. Theranostics, 2020. 10: 1.
https://pubmed.ncbi.nlm.nih.gov/31903102
395. van Kalmthout, L.W.M., et al. Prospective Validation of Gallium-68 Prostate Specific Membrane
Antigen-Positron Emission Tomography/Computerized Tomography for Primary Staging of Prostate
Cancer. J Urol, 2020. 203: 537.
https://pubmed.ncbi.nlm.nih.gov/31487220
396. Jansen, B.H.E., et al. Pelvic lymph-node staging with (18)F-DCFPyL PET/CT prior to extended
pelvic lymph-node dissection in primary prostate cancer - the SALT trial. Eur J Nucl Med Mol
Imaging, 2021. 48: 509.
https://pubmed.ncbi.nlm.nih.gov/32789599
397. Perera, M., et al. Gallium-68 Prostate-specific Membrane Antigen Positron Emission Tomography
in Advanced Prostate Cancer-Updated Diagnostic Utility, Sensitivity, Specificity, and Distribution
of Prostate-specific Membrane Antigen-avid Lesions: A Systematic Review and Meta-analysis. Eur
Urol, 2020. 77: 403.
https://pubmed.ncbi.nlm.nih.gov/27363387
398. Uprimny, C., et al. 68Ga-PSMA-11 PET/CT in primary staging of prostate cancer: PSA and Gleason
score predict the intensity of tracer accumulation in the primary tumour. Eur J Nucl Med Mol
Imaging, 2017. 44: 941.
https://pubmed.ncbi.nlm.nih.gov/28138747
399. Wu, H., et al. Diagnostic Performance of (68)Gallium Labelled Prostate-Specific Membrane Antigen
Positron Emission Tomography/Computed Tomography and Magnetic Resonance Imaging for
Staging the Prostate Cancer with Intermediate or High Risk Prior to Radical Prostatectomy: A
Systematic Review and Meta-analysis. World J Mens Health, 2020. 38: 208.
https://pubmed.ncbi.nlm.nih.gov/31081294
400. Tulsyan, S., et al. Comparison of 68Ga-PSMA PET/CT and multiparametric MRI for staging of
high-risk prostate cancer68Ga-PSMA PET and MRI in prostate cancer. Nucl Med Commun, 2017.
38: 1094.
https://pubmed.ncbi.nlm.nih.gov/28957842
401. Shen, G., et al. Comparison of choline-PET/CT, MRI, SPECT, and bone scintigraphy in the diagnosis
of bone metastases in patients with prostate cancer: a meta-analysis. Skeletal Radiol, 2014.
43: 1503.
https://pubmed.ncbi.nlm.nih.gov/24841276
402. Briganti, A., et al. When to perform bone scan in patients with newly diagnosed prostate cancer:
external validation of the currently available guidelines and proposal of a novel risk stratification tool.
Eur Urol, 2010. 57: 551.
https://pubmed.ncbi.nlm.nih.gov/20034730
403. O’Sullivan, J.M., et al. Broadening the criteria for avoiding staging bone scans in prostate cancer: a
retrospective study of patients at the Royal Marsden Hospital. BJU Int, 2003. 92: 685.
https://pubmed.ncbi.nlm.nih.gov/14616446
404. Ayyathurai, R., et al. A study on staging bone scans in newly diagnosed prostate cancer. Urol Int,
2006. 76: 209.
https://pubmed.ncbi.nlm.nih.gov/16601380
405. Tateishi, U., et al. A meta-analysis of (18)F-Fluoride positron emission tomography for assessment of
metastatic bone tumor. Ann Nucl Med, 2010. 24: 523.
https://pubmed.ncbi.nlm.nih.gov/20559896
406. Evangelista, L., et al. Diagnostic imaging to detect and evaluate response to therapy in bone
metastases from prostate cancer: current modalities and new horizons. Eur J Nucl Med Mol
Imaging, 2016. 43: 1546.
https://pubmed.ncbi.nlm.nih.gov/26956538
407. Zacho, H.D., et al. No Added Value of (18)F-Sodium Fluoride PET/CT for the Detection of Bone
Metastases in Patients with Newly Diagnosed Prostate Cancer with Normal Bone Scintigraphy.
J Nucl Med, 2019. 60: 1713.
https://pubmed.ncbi.nlm.nih.gov/31147402
408. Zacho, H.D., et al. Observer Agreement and Accuracy of (18)F-Sodium Fluoride PET/CT in the
Diagnosis of Bone Metastases in Prostate Cancer. J Nucl Med, 2020. 61: 344.
https://pubmed.ncbi.nlm.nih.gov/31481577
409. Brogsitter, C., et al. 18F-Choline, 11C-choline and 11C-acetate PET/CT: comparative analysis for
imaging prostate cancer patients. Eur J Nucl Med Mol Imaging, 2013. 40 Suppl 1: S18.
https://pubmed.ncbi.nlm.nih.gov/23579863

PROSTATE CANCER - LIMITED UPDATE 2021 157


410. Picchio, M., et al. [11C]Choline PET/CT detection of bone metastases in patients with PSA
progression after primary treatment for prostate cancer: comparison with bone scintigraphy. Eur J
Nucl Med Mol Imaging, 2012. 39: 13.
https://pubmed.ncbi.nlm.nih.gov/21932120
411. Gutzeit, A., et al. Comparison of diffusion-weighted whole body MRI and skeletal scintigraphy for
the detection of bone metastases in patients with prostate or breast carcinoma. Skeletal Radiol,
2010. 39: 333.
https://pubmed.ncbi.nlm.nih.gov/20205350
412. Lecouvet, F.E., et al. Can whole-body magnetic resonance imaging with diffusion-weighted imaging
replace Tc 99m bone scanning and computed tomography for single-step detection of metastases
in patients with high-risk prostate cancer? Eur Urol, 2012. 62: 68.
https://pubmed.ncbi.nlm.nih.gov/22366187
413. Pasoglou, V., et al. One-step TNM staging of high-risk prostate cancer using magnetic resonance
imaging (MRI): toward an upfront simplified “all-in-one” imaging approach? Prostate, 2014. 74: 469.
https://pubmed.ncbi.nlm.nih.gov/24375774
414. Corfield, J., et al. (68)Ga-prostate specific membrane antigen (PSMA) positron emission tomography
(PET) for primary staging of high-risk prostate cancer: a systematic review. World J Urol, 2018.
36: 519.
https://pubmed.ncbi.nlm.nih.gov/29344682
415. Roach, P.J., et al. The Impact of (68)Ga-PSMA PET/CT on Management Intent in Prostate Cancer:
Results of an Australian Prospective Multicenter Study. J Nucl Med, 2018. 59: 82.
https://pubmed.ncbi.nlm.nih.gov/28646014
416. Yaxley, J.W., et al. Risk of metastatic disease on (68) gallium-prostate-specific membrane antigen
positron emission tomography/computed tomography scan for primary staging of 1253 men at the
diagnosis of prostate cancer. BJU Int, 2019. 124: 401.
https://pubmed.ncbi.nlm.nih.gov/31141284
417. Hofman, M.S., et al. Prostate-specific membrane antigen PET-CT in patients with high-risk prostate
cancer before curative-intent surgery or radiotherapy (proPSMA): a prospective, randomised,
multicentre study. Lancet, 2020. 395: 1208.
https://pubmed.ncbi.nlm.nih.gov/32209449
418. Cornford, P., et al. Prostate-specific Membrane Antigen Positron Emission Tomography Scans
Before Curative Treatment: Ready for Prime Time? Eur Urol, 2020. 78: e125.
https://pubmed.ncbi.nlm.nih.gov/32624287
419. Hicks, R.J., et al. Seduction by Sensitivity: Reality, Illusion, or Delusion? The Challenge of Assessing
Outcomes after PSMA Imaging Selection of Patients for Treatment. J Nucl Med, 2017. 58: 1969.
https://pubmed.ncbi.nlm.nih.gov/28935839
420. Hofman, M.S., et al. A prospective randomized multicentre study of the impact of gallium-68
prostate-specific membrane antigen (PSMA) PET/CT imaging for staging high-risk prostate cancer
prior to curative-intent surgery or radiotherapy (proPSMA study): clinical trial protocol. BJU Int,
2018. 122: 783.
https://pubmed.ncbi.nlm.nih.gov/29726071
421. Smith, B.D., et al. Future of cancer incidence in the United States: burdens upon an aging, changing
nation. J Clin Oncol, 2009. 27: 2758.
https://pubmed.ncbi.nlm.nih.gov/19403886
422. Arnold, M., et al. Recent trends in incidence of five common cancers in 26 European countries since
1988: Analysis of the European Cancer Observatory. Eur J Cancer, 2015. 51: 1164.
https://pubmed.ncbi.nlm.nih.gov/24120180
423. Liu, D., et al. Active surveillance versus surgery for low risk prostate cancer: a clinical decision
analysis. J Urol, 2012. 187: 1241.
https://pubmed.ncbi.nlm.nih.gov/22335873
424. Bill-Axelson, A., et al. Radical prostatectomy or watchful waiting in early prostate cancer. N Engl J
Med, 2014. 370: 932.
https://pubmed.ncbi.nlm.nih.gov/24597866
425. Kupelian, P.A., et al. Comparison of the efficacy of local therapies for localized prostate cancer in the
prostate-specific antigen era: a large single-institution experience with radical prostatectomy and
external-beam radiotherapy. J Clin Oncol, 2002. 20: 3376.
https://pubmed.ncbi.nlm.nih.gov/12177097
426. Bubolz, T., et al. Treatments for prostate cancer in older men: 1984-1997. Urology, 2001. 58: 977.
https://pubmed.ncbi.nlm.nih.gov/11744472

158 PROSTATE CANCER - LIMITED UPDATE 2021


427. Houterman, S., et al. Impact of comorbidity on treatment and prognosis of prostate cancer patients:
a population-based study. Crit Rev Oncol Hematol, 2006. 58: 60.
https://pubmed.ncbi.nlm.nih.gov/16213153
428. Ries L.A.G., et al. eds. SEER cancer Statistics Review, 1975-2005. 2008.
https://seer.cancer.gov/archive/csr/1975_2005/
429. Scosyrev, E., et al. Prostate cancer in the elderly: frequency of advanced disease at presentation
and disease-specific mortality. Cancer, 2012. 118: 3062.
https://pubmed.ncbi.nlm.nih.gov/22006014
430. Richstone, L., et al. Radical prostatectomy in men aged >or=70 years: effect of age on upgrading,
upstaging, and the accuracy of a preoperative nomogram. BJU Int, 2008. 101: 541.
https://pubmed.ncbi.nlm.nih.gov/18257855
431. Sun, L., et al. Men older than 70 years have higher risk prostate cancer and poorer survival in the
early and late prostate specific antigen eras. J Urol, 2009. 182: 2242.
https://pubmed.ncbi.nlm.nih.gov/19758616
432. Hamilton, A.S., et al. Trends in the treatment of localized prostate cancer using supplemented
cancer registry data. BJU Int, 2011. 107: 576.
https://pubmed.ncbi.nlm.nih.gov/20735387
433. Studenski, S., et al. Gait speed and survival in older adults. JAMA, 2011. 305: 50.
https://pubmed.ncbi.nlm.nih.gov/21205966
434. Ethun, C.G., et al. Frailty and cancer: Implications for oncology surgery, medical oncology, and
radiation oncology. CA Cancer J Clin, 2017. 67: 362.
https://pubmed.ncbi.nlm.nih.gov/28731537
435. Bellera, C.A., et al. Screening older cancer patients: first evaluation of the G-8 geriatric screening
tool. Ann Oncol, 2012. 23: 2166.
https://pubmed.ncbi.nlm.nih.gov/22250183
436. Hamaker, M.E., et al. The effect of a geriatric evaluation on treatment decisions and outcome for
older cancer patients - A systematic review. J Geriatr Oncol, 2018. 9: 430.
https://pubmed.ncbi.nlm.nih.gov/29631898
437. Rockwood, K., et al. Using the Clinical Frailty Scale in Allocating Scarce Health Care Resources.
Can Geriatr J, 2020. 23: 210.
https://pubmed.ncbi.nlm.nih.gov/32904824
438. McIsaac, D.I., et al. Frailty as a Predictor of Death or New Disability After Surgery: A Prospective
Cohort Study. Ann Surg, 2020. 271: 283.
https://pubmed.ncbi.nlm.nih.gov/30048320
439. van Walree, I.C., et al. Clinical judgment versus geriatric assessment for frailty in older patients with
cancer. J Geriatr Oncol, 2020. 11: 1138.
https://pubmed.ncbi.nlm.nih.gov/32576520
440. Albertsen, P.C., et al. Impact of comorbidity on survival among men with localized prostate cancer.
J Clin Oncol, 2011. 29: 1335.
https://pubmed.ncbi.nlm.nih.gov/21357791
441. Tewari, A., et al. Long-term survival probability in men with clinically localized prostate cancer: a
case-control, propensity modeling study stratified by race, age, treatment and comorbidities. J Urol,
2004. 171: 1513.
https://pubmed.ncbi.nlm.nih.gov/15017210
442. Parmelee, P.A., et al. Validation of the Cumulative Illness Rating Scale in a geriatric residential
population. J Am Geriatr Soc, 1995. 43: 130.
https://pubmed.ncbi.nlm.nih.gov/7836636
443. Groome, P.A., et al. Assessing the impact of comorbid illnesses on death within 10 years in prostate
cancer treatment candidates. Cancer, 2011. 117: 3943.
https://pubmed.ncbi.nlm.nih.gov/21858801
444. Charlson, M.E., et al. A new method of classifying prognostic comorbidity in longitudinal studies:
development and validation. J Chronic Dis, 1987. 40: 373.
https://pubmed.ncbi.nlm.nih.gov/3558716
445. Blanc-Bisson, C., et al. Undernutrition in elderly patients with cancer: target for diagnosis and
intervention. Crit Rev Oncol Hematol, 2008. 67: 243.
https://pubmed.ncbi.nlm.nih.gov/18554922
446. Sachs, G.A., et al. Cognitive impairment: an independent predictor of excess mortality: a cohort
study. Ann Intern Med, 2011. 155: 300.
https://pubmed.ncbi.nlm.nih.gov/21893623

PROSTATE CANCER - LIMITED UPDATE 2021 159


447. Robinson, T.N., et al. Preoperative cognitive dysfunction is related to adverse postoperative
outcomes in the elderly. J Am Coll Surg, 2012. 215: 12.
https://pubmed.ncbi.nlm.nih.gov/22626912
448. Borson, S., et al. The Mini-Cog as a screen for dementia: validation in a population-based sample.
J Am Geriatr Soc, 2003. 51: 1451.
https://pubmed.ncbi.nlm.nih.gov/14511167
449. Korc-Grodzicki, B., et al. Prevention of post-operative delirium in older patients with cancer
undergoing surgery. J Geriatr Oncol, 2015. 6: 60.
https://pubmed.ncbi.nlm.nih.gov/25454768
450. Oken, M.M., et al. Toxicity and response criteria of the Eastern Cooperative Oncology Group. Am J
Clin Oncol, 1982. 5: 649.
https://pubmed.ncbi.nlm.nih.gov/7165009
451. Katz, S., et al. Studies of illness in the aged. The index of ADL: a standardized measure of biological
and psychological function. JAMA, 1963. 185: 914.
https://pubmed.ncbi.nlm.nih.gov/14044222
452. Lawton, M.P., et al. Assessment of older people: self-maintaining and instrumental activities of daily
living. Gerontologist, 1969. 9: 179.
https://pubmed.ncbi.nlm.nih.gov/5349366
453. Stineman, M.G., et al. All-cause 1-, 5-, and 10-year mortality in elderly people according to activities
of daily living stage. J Am Geriatr Soc, 2012. 60: 485.
https://pubmed.ncbi.nlm.nih.gov/22352414
454. Paladino, J., et al. Communication Strategies for Sharing Prognostic Information With Patients:
Beyond Survival Statistics. JAMA, 2019.
https://pubmed.ncbi.nlm.nih.gov/31415085
455. Rostoft, S., et al. Shared decision-making in older patients with cancer - What does the patient
want? J Geriatr Oncol, 2020.
https://pubmed.ncbi.nlm.nih.gov/32839118
456. Soubeyran, P., et al. Screening for vulnerability in older cancer patients: the ONCODAGE
Prospective Multicenter Cohort Study. PLoS One, 2014. 9: e115060.
https://pubmed.ncbi.nlm.nih.gov/25503576
457. Chodak, G.W., et al. Results of conservative management of clinically localized prostate cancer.
N Engl J Med, 1994. 330: 242.
https://pubmed.ncbi.nlm.nih.gov/8272085
458. Sandblom, G., et al. Long-term survival in a Swedish population-based cohort of men with prostate
cancer. Urology, 2000. 56: 442.
https://pubmed.ncbi.nlm.nih.gov/10962312
459. Johansson, J.E., et al. Natural history of localised prostatic cancer. A population-based study in 223
untreated patients. Lancet, 1989. 1: 799.
https://pubmed.ncbi.nlm.nih.gov/2564901
460. Bill-Axelson, A., et al. Radical prostatectomy versus watchful waiting in early prostate cancer. N Engl
J Med, 2005. 352: 1977.
https://pubmed.ncbi.nlm.nih.gov/15888698
461. Adolfsson, J., et al. The 20-Yr outcome in patients with well- or moderately differentiated clinically
localized prostate cancer diagnosed in the pre-PSA era: the prognostic value of tumour ploidy and
comorbidity. Eur Urol, 2007. 52: 1028.
https://pubmed.ncbi.nlm.nih.gov/17467883
462. Jonsson, E., et al. Adenocarcinoma of the prostate in Iceland: a population-based study of stage,
Gleason grade, treatment and long-term survival in males diagnosed between 1983 and 1987.
Scand J Urol Nephrol, 2006. 40: 265.
https://pubmed.ncbi.nlm.nih.gov/16916765
463. Lu-Yao, G.L., et al. Outcomes of localized prostate cancer following conservative management.
Jama, 2009. 302: 1202.
https://pubmed.ncbi.nlm.nih.gov/19755699
464. Hayes, J.H., et al. Observation versus initial treatment for men with localized, low-risk prostate
cancer: a cost-effectiveness analysis. Ann Intern Med, 2013. 158: 853.
https://pubmed.ncbi.nlm.nih.gov/23778902
465. Albertsen, P.C. Observational studies and the natural history of screen-detected prostate cancer.
Curr Opin Urol, 2015. 25: 232.
https://pubmed.ncbi.nlm.nih.gov/25692723

160 PROSTATE CANCER - LIMITED UPDATE 2021


466. Bruinsma, S.M., et al. Expert consensus document: Semantics in active surveillance for men with
localized prostate cancer - results of a modified Delphi consensus procedure. Nat Rev Urol, 2017.
14: 312.
https://pubmed.ncbi.nlm.nih.gov/28290462
467. Hamdy, F.C., et al. 10-Year Outcomes after Monitoring, Surgery, or Radiotherapy for Localized
Prostate Cancer. N Engl J Med, 2016. 375: 1415.
https://pubmed.ncbi.nlm.nih.gov/27626136
468. Thomsen, F.B., et al. Active surveillance for clinically localized prostate cancer--a systematic review.
J Surg Oncol, 2014. 109: 830.
https://pubmed.ncbi.nlm.nih.gov/24610744
469. Tosoian, J.J., et al. Active Surveillance of Grade Group 1 Prostate Cancer: Long-term Outcomes
from a Large Prospective Cohort. Eur Urol, 2020. 77: 675.
https://pubmed.ncbi.nlm.nih.gov/26324359
470. van As, N.J., et al. Predicting the probability of deferred radical treatment for localised prostate
cancer managed by active surveillance. Eur Urol, 2008. 54: 1297.
https://pubmed.ncbi.nlm.nih.gov/18342430
471. Carter, H.B., et al. Expectant management of prostate cancer with curative intent: an update of the
Johns Hopkins experience. J Urol, 2007. 178: 2359.
https://pubmed.ncbi.nlm.nih.gov/17936806
472. Adamy, A., et al. Role of prostate specific antigen and immediate confirmatory biopsy in predicting
progression during active surveillance for low risk prostate cancer. J Urol, 2011. 185: 477.
https://pubmed.ncbi.nlm.nih.gov/21167529
473. Soloway, M.S., et al. Careful selection and close monitoring of low-risk prostate cancer patients on
active surveillance minimizes the need for treatment. Eur Urol, 2010. 58: 831.
https://pubmed.ncbi.nlm.nih.gov/20800964
474. Roemeling, S., et al. Active surveillance for prostate cancers detected in three subsequent rounds of
a screening trial: characteristics, PSA doubling times, and outcome. Eur Urol, 2007. 51: 1244.
https://pubmed.ncbi.nlm.nih.gov/17161520
475. Khatami, A., et al. PSA doubling time predicts the outcome after active surveillance in screening-
detected prostate cancer: results from the European randomized study of screening for prostate
cancer, Sweden section. Int J Cancer, 2007. 120: 170.
https://pubmed.ncbi.nlm.nih.gov/17013897
476. Klotz, L., et al. Long-term follow-up of a large active surveillance cohort of patients with prostate
cancer. J Clin Oncol, 2015. 33: 272.
https://pubmed.ncbi.nlm.nih.gov/25512465
477. Bill-Axelson, A., et al. Radical Prostatectomy or Watchful Waiting in Prostate Cancer - 29-Year
Follow-up. N Engl J Med, 2018. 379: 2319.
https://pubmed.ncbi.nlm.nih.gov/30575473
478. Wilt, T.J., et al. Radical Prostatectomy or Observation for Clinically Localized Prostate Cancer:
Extended Follow-up of the Prostate Cancer Intervention Versus Observation Trial (PIVOT). Eur Urol,
2020. 77: 713.
https://pubmed.ncbi.nlm.nih.gov/32089359
479. Steineck, G., et al. Quality of life after radical prostatectomy or watchful waiting. N Engl J Med,
2002. 347: 790.
https://pubmed.ncbi.nlm.nih.gov/12226149
480. Adolfsson, J. Watchful waiting and active surveillance: the current position. BJU Int, 2008. 102: 10.
https://pubmed.ncbi.nlm.nih.gov/18422774
481. Hatzinger, M., et al. [The history of prostate cancer from the beginning to DaVinci]. Aktuelle Urol,
2012. 43: 228.
https://pubmed.ncbi.nlm.nih.gov/23035261
482. Wilt, T.J., et al. Follow-up of Prostatectomy versus Observation for Early Prostate Cancer. N Engl
J Med, 2017. 377: 132.
https://pubmed.ncbi.nlm.nih.gov/28700844
483. Kretschmer, A., et al. Perioperative patient education improves long-term satisfaction rates of low-
risk prostate cancer patients after radical prostatectomy. World J Urol, 2017. 35: 1205.
https://pubmed.ncbi.nlm.nih.gov/28093628
484. Gyomber, D., et al. Improving informed consent for patients undergoing radical prostatectomy using
multimedia techniques: a prospective randomized crossover study. BJU Int, 2010. 106: 1152.
https://pubmed.ncbi.nlm.nih.gov/20346048

PROSTATE CANCER - LIMITED UPDATE 2021 161


485. Huber, J., et al. Multimedia support for improving preoperative patient education: a randomized
controlled trial using the example of radical prostatectomy. Ann Surg Oncol, 2013. 20: 15.
https://pubmed.ncbi.nlm.nih.gov/22851045
486. Wake, N., et al. Patient-specific 3D printed and augmented reality kidney and prostate cancer
models: impact on patient education. 3D Print Med, 2019. 5: 4.
https://pubmed.ncbi.nlm.nih.gov/30783869
487. De Nunzio, C., et al. The EORTC quality of life questionnaire predicts early and long-term
incontinence in patients treated with robotic assisted radical prostatectomy: Analysis of a large
single center cohort. Urol Oncol, 2019. 37: 1006.
https://pubmed.ncbi.nlm.nih.gov/31326315
488. Chang, J.I., et al. Preoperative Pelvic Floor Muscle Exercise and Postprostatectomy Incontinence: A
Systematic Review and Meta-analysis. Eur Urol, 2016. 69: 460.
https://pubmed.ncbi.nlm.nih.gov/26610857
489. Kumar, S., et al. Neo-adjuvant and adjuvant hormone therapy for localised and locally advanced
prostate cancer. Cochrane Database Syst Rev, 2006: CD006019.
https://pubmed.ncbi.nlm.nih.gov/17054269
490. Efstathiou, E., et al. Clinical and Biological Characterisation of Localised High-risk Prostate Cancer:
Results of a Randomised Preoperative Study of a Luteinising Hormone-releasing Hormone Agonist
with or Without Abiraterone Acetate plus Prednisone. Eur Urol, 2019. 76: 418.
https://pubmed.ncbi.nlm.nih.gov/31176622
491. Walsh, P.C., et al. Impotence following radical prostatectomy: insight into etiology and prevention.
J Urol, 1982. 128: 492.
https://pubmed.ncbi.nlm.nih.gov/7120554
492. Schuessler, W.W., et al. Laparoscopic radical prostatectomy: initial short-term experience. Urology,
1997. 50: 854.
https://pubmed.ncbi.nlm.nih.gov/9426713
493. Binder, J., et al. [Robot-assisted laparoscopy in urology. Radical prostatectomy and reconstructive
retroperitoneal interventions]. Urologe A, 2002. 41: 144.
https://pubmed.ncbi.nlm.nih.gov/11993092
494. Yaxley, J.W., et al. Robot-assisted laparoscopic prostatectomy versus open radical retropubic
prostatectomy: early outcomes from a randomised controlled phase 3 study. Lancet, 2016.
388: 1057.
https://pubmed.ncbi.nlm.nih.gov/27474375
495. Coughlin, G.D., et al. Robot-assisted laparoscopic prostatectomy versus open radical retropubic
prostatectomy: 24-month outcomes from a randomised controlled study. Lancet Oncol, 2018.
19: 1051.
https://pubmed.ncbi.nlm.nih.gov/30017351
496. Albertsen, P.C., et al. Competing risk analysis of men aged 55 to 74 years at diagnosis managed
conservatively for clinically localized prostate cancer. Jama, 1998. 280: 975.
https://pubmed.ncbi.nlm.nih.gov/9749479
497. Albertsen, P.C., et al. Statistical considerations when assessing outcomes following treatment for
prostate cancer. J Urol, 1999. 162: 439.
https://pubmed.ncbi.nlm.nih.gov/10411053
498. Iversen, P., et al. Bicalutamide (150 mg) versus placebo as immediate therapy alone or as adjuvant
to therapy with curative intent for early nonmetastatic prostate cancer: 5.3-year median followup
from the Scandinavian Prostate Cancer Group Study Number 6. J Urol, 2004. 172: 1871.
https://pubmed.ncbi.nlm.nih.gov/15540741
499. Jacobs, B.L., et al. Use of advanced treatment technologies among men at low risk of dying from
prostate cancer. Jama, 2013. 309: 2587.
https://pubmed.ncbi.nlm.nih.gov/23800935
500. Ramsay, C., et al. Systematic review and economic modelling of the relative clinical benefit and
cost-effectiveness of laparoscopic surgery and robotic surgery for removal of the prostate in men
with localised prostate cancer. Health Technol Assess, 2012. 16: 1.
https://pubmed.ncbi.nlm.nih.gov/23127367
501. Allan, C., et al. Laparoscopic versus Robotic-Assisted Radical Prostatectomy for the Treatment of
Localised Prostate Cancer: A Systematic Review. Urol Int, 2016. 96: 373.
https://pubmed.ncbi.nlm.nih.gov/26201500
502. Ilic, D., et al. Laparoscopic and robotic-assisted versus open radical prostatectomy for the
treatment of localised prostate cancer. Cochrane Database Syst Rev, 2017. 9: CD009625.
https://pubmed.ncbi.nlm.nih.gov/28895658

162 PROSTATE CANCER - LIMITED UPDATE 2021


503. Begg, C.B., et al. Variations in morbidity after radical prostatectomy. N Engl J Med, 2002. 346: 1138.
https://pubmed.ncbi.nlm.nih.gov/11948274
504. Gershman, B., et al. Redefining and Contextualizing the Hospital Volume-Outcome Relationship
for Robot-Assisted Radical Prostatectomy: Implications for Centralization of Care. J Urol, 2017.
198: 92.
https://pubmed.ncbi.nlm.nih.gov/28153509
505. Galfano, A., et al. A new anatomic approach for robot-assisted laparoscopic prostatectomy: a
feasibility study for completely intrafascial surgery. Eur Urol, 2010. 58: 457.
https://pubmed.ncbi.nlm.nih.gov/20566236
506. Checcucci, E., et al. Retzius-sparing robot-assisted radical prostatectomy vs the standard
approach: a systematic review and analysis of comparative outcomes. BJU Int, 2020. 125: 8.
https://pubmed.ncbi.nlm.nih.gov/31373142
507. Phukan, C., et al. Retzius sparing robotic assisted radical prostatectomy vs. conventional robotic
assisted radical prostatectomy: a systematic review and meta-analysis. World J Urol, 2020.
38: 1123.
https://pubmed.ncbi.nlm.nih.gov/31089802
508. Tai, T.E., et al. Effects of Retzius sparing on robot-assisted laparoscopic prostatectomy: a
systematic review with meta-analysis. Surg Endosc, 2020. 34: 4020.
https://pubmed.ncbi.nlm.nih.gov/31617093
509. Rosenberg, J.E., et al. Retzius-sparing versus standard robotic-assisted laparoscopic
prostatectomy for the treatment of clinically localized prostate cancer. Cochrane Database Syst Rev,
2020. 8: Cd013641.
https://pubmed.ncbi.nlm.nih.gov/32813279
510. Lee, J., et al. Retzius Sparing Robot-Assisted Radical Prostatectomy Conveys Early Regain of
Continence over Conventional Robot-Assisted Radical Prostatectomy: A Propensity Score Matched
Analysis of 1,863 Patients. J Urol, 2020. 203: 137.
https://pubmed.ncbi.nlm.nih.gov/31347951
511. Stonier, T., et al. Retzius-sparing robot-assisted radical prostatectomy (RS-RARP) vs standard
RARP: it’s time for critical appraisal. BJU Int, 2019. 123: 5.
https://pubmed.ncbi.nlm.nih.gov/29959814
512. Fossati, N., et al. The Benefits and Harms of Different Extents of Lymph Node Dissection During
Radical Prostatectomy for Prostate Cancer: A Systematic Review. Eur Urol, 2017. 72: 84.
https://pubmed.ncbi.nlm.nih.gov/28126351
513. Lestingi, J.F.P., et al. Extended Versus Limited Pelvic Lymph Node Dissection During Radical
Prostatectomy for Intermediate- and High-risk Prostate Cancer: Early Oncological Outcomes from a
Randomized Phase 3 Trial. Eur Urol, 2020.
https://pubmed.ncbi.nlm.nih.gov/33293077
514. Mattei, A., et al. The template of the primary lymphatic landing sites of the prostate should be
revisited: results of a multimodality mapping study. Eur Urol, 2008. 53: 118.
https://pubmed.ncbi.nlm.nih.gov/17709171
515. Roach, M., 3rd, et al. Predicting the risk of lymph node involvement using the pre-treatment prostate
specific antigen and Gleason score in men with clinically localized prostate cancer. Int J Radiat
Oncol Biol Phys, 1994. 28: 33.
https://pubmed.ncbi.nlm.nih.gov/7505775
516. Cimino, S., et al. Comparison between Briganti, Partin and MSKCC tools in predicting positive
lymph nodes in prostate cancer: a systematic review and meta-analysis. Scand J Urol, 2017.
51: 345.
https://pubmed.ncbi.nlm.nih.gov/28644701
517. Abdollah, F., et al. Indications for pelvic nodal treatment in prostate cancer should change.
Validation of the Roach formula in a large extended nodal dissection series. Int J Radiat Oncol Biol
Phys, 2012. 83: 624.
https://pubmed.ncbi.nlm.nih.gov/22099031
518. Dell’Oglio, P., et al. External validation of the European association of urology recommendations
for pelvic lymph node dissection in patients treated with robot-assisted radical prostatectomy.
J Endourol, 2014. 28: 416.
https://pubmed.ncbi.nlm.nih.gov/24188052
519. Hinev, A.I., et al. Validation of nomograms predicting lymph node involvement in patients with
prostate cancer undergoing extended pelvic lymph node dissection. Urol Int, 2014. 92: 300.
https://pubmed.ncbi.nlm.nih.gov/24480972

PROSTATE CANCER - LIMITED UPDATE 2021 163


520. Gandaglia, G., et al. External Validation of the 2019 Briganti Nomogram for the Identification
of Prostate Cancer Patients Who Should Be Considered for an Extended Pelvic Lymph Node
Dissection. Eur Urol, 2020. 78: 138.
https://pubmed.ncbi.nlm.nih.gov/32268944
521. Gandaglia, G., et al. A Novel Nomogram to Identify Candidates for Extended Pelvic Lymph Node
Dissection Among Patients with Clinically Localized Prostate Cancer Diagnosed with Magnetic
Resonance Imaging-targeted and Systematic Biopsies. Eur Urol, 2019. 75: 506.
https://pubmed.ncbi.nlm.nih.gov/30342844
522. van der Poel, H.G., et al. Sentinel node biopsy for prostate cancer: report from a consensus panel
meeting. BJU Int, 2017. 120: 204.
https://pubmed.ncbi.nlm.nih.gov/28188689
523. Harke, N.N., et al. Fluorescence-supported lymphography and extended pelvic lymph node
dissection in robot-assisted radical prostatectomy: a prospective, randomized trial. World J Urol,
2018. 36: 1817.
https://pubmed.ncbi.nlm.nih.gov/29767326
524. Wit, E.M.K., et al. Sentinel Node Procedure in Prostate Cancer: A Systematic Review to Assess
Diagnostic Accuracy. Eur Urol, 2017. 71: 596.
https://pubmed.ncbi.nlm.nih.gov/27639533
525. Weng, W.C., et al. Impact of prostatic anterior fat pads with lymph node staging in prostate cancer.
J Cancer, 2018. 9: 3361.
https://pubmed.ncbi.nlm.nih.gov/30271497
526. Hosny, M., et al. Can Anterior Prostatic Fat Harbor Prostate Cancer Metastasis? A Prospective
Cohort Study. Curr Urol, 2017. 10: 182.
https://pubmed.ncbi.nlm.nih.gov/29234260
527. Ball, M.W., et al. Pathological analysis of the prostatic anterior fat pad at radical prostatectomy:
insights from a prospective series. BJU Int, 2017. 119: 444.
https://pubmed.ncbi.nlm.nih.gov/27611825
528. Kwon, Y.S., et al. Oncologic outcomes in men with metastasis to the prostatic anterior fat pad lymph
nodes: a multi-institution international study. BMC Urol, 2015. 15: 79.
https://pubmed.ncbi.nlm.nih.gov/26231860
529. Ozkan, B., et al. Role of anterior prostatic fat pad dissection for extended lymphadenectomy in
prostate cancer: a non-randomized study of 100 patients. Int Urol Nephrol, 2015. 47: 959.
https://pubmed.ncbi.nlm.nih.gov/25899767
530. Kim, I.Y., et al. Detailed analysis of patients with metastasis to the prostatic anterior fat pad lymph
nodes: a multi-institutional study. J Urol, 2013. 190: 527.
https://pubmed.ncbi.nlm.nih.gov/23485503
531. Hansen, J., et al. Assessment of rates of lymph nodes and lymph node metastases in periprostatic
fat pads in a consecutive cohort treated with retropubic radical prostatectomy. Urology, 2012.
80: 877.
https://pubmed.ncbi.nlm.nih.gov/22950996
532. Rainwater, L.M., et al. Technical consideration in radical retropubic prostatectomy: blood loss after
ligation of dorsal venous complex. J Urol, 1990. 143: 1163.
https://pubmed.ncbi.nlm.nih.gov/2342176
533. Woldu, S.L., et al. Outcomes with delayed dorsal vein complex ligation during robotic assisted
laparoscopic prostatectomy. Can J Urol, 2013. 20: 7079.
https://pubmed.ncbi.nlm.nih.gov/24331354
534. Lei, Y., et al. Athermal division and selective suture ligation of the dorsal vein complex during robot-
assisted laparoscopic radical prostatectomy: description of technique and outcomes. Eur Urol,
2011. 59: 235.
https://pubmed.ncbi.nlm.nih.gov/20863611
535. Wu, S.D., et al. Suture versus staple ligation of the dorsal venous complex during robot-assisted
laparoscopic radical prostatectomy. BJU Int, 2010. 106: 385.
https://pubmed.ncbi.nlm.nih.gov/20067457
536. Walsh, P.C., et al. Radical prostatectomy and cystoprostatectomy with preservation of potency.
Results using a new nerve-sparing technique. Br J Urol, 1984. 56: 694.
https://pubmed.ncbi.nlm.nih.gov/6534493
537. Walz, J., et al. A Critical Analysis of the Current Knowledge of Surgical Anatomy of the Prostate
Related to Optimisation of Cancer Control and Preservation of Continence and Erection in
Candidates for Radical Prostatectomy: An Update. Eur Urol, 2016. 70: 301.
https://pubmed.ncbi.nlm.nih.gov/26850969

164 PROSTATE CANCER - LIMITED UPDATE 2021


538. Michl, U., et al. Nerve-sparing Surgery Technique, Not the Preservation of the Neurovascular
Bundles, Leads to Improved Long-term Continence Rates After Radical Prostatectomy. Eur Urol,
2016. 69: 584.
https://pubmed.ncbi.nlm.nih.gov/26277303
539. Avulova, S., et al. The Effect of Nerve Sparing Status on Sexual and Urinary Function: 3-Year
Results from the CEASAR Study. J Urol, 2018. 199: 1202.
https://pubmed.ncbi.nlm.nih.gov/29253578
540. Stolzenburg, J.U., et al. A comparison of outcomes for interfascial and intrafascial nerve-sparing
radical prostatectomy. Urology, 2010. 76: 743.
https://pubmed.ncbi.nlm.nih.gov/20573384
541. Steineck, G., et al. Degree of preservation of the neurovascular bundles during radical
prostatectomy and urinary continence 1 year after surgery. Eur Urol, 2015. 67: 559.
https://pubmed.ncbi.nlm.nih.gov/25457018
542. Shikanov, S., et al. Extrafascial versus interfascial nerve-sparing technique for robotic-assisted
laparoscopic prostatectomy: comparison of functional outcomes and positive surgical margins
characteristics. Urology, 2009. 74: 611.
https://pubmed.ncbi.nlm.nih.gov/19616830
543. Tewari, A.K., et al. Anatomical grades of nerve sparing: a risk-stratified approach to neural-
hammock sparing during robot-assisted radical prostatectomy (RARP). BJU Int, 2011. 108: 984.
https://pubmed.ncbi.nlm.nih.gov/21917101
544. Nielsen, M.E., et al. High anterior release of the levator fascia improves sexual function following
open radical retropubic prostatectomy. J Urol, 2008. 180: 2557.
https://pubmed.ncbi.nlm.nih.gov/18930504
545. Ko, Y.H., et al. Retrograde versus antegrade nerve sparing during robot-assisted radical
prostatectomy: which is better for achieving early functional recovery? Eur Urol, 2013. 63: 169.
https://pubmed.ncbi.nlm.nih.gov/23092543
546. Tewari, A.K., et al. Functional outcomes following robotic prostatectomy using athermal, traction
free risk-stratified grades of nerve sparing. World J Urol, 2013. 31: 471.
https://pubmed.ncbi.nlm.nih.gov/23354288
547. Catalona, W.J., et al. Nerve-sparing radical prostatectomy: evaluation of results after 250 patients.
J Urol, 1990. 143: 538.
https://pubmed.ncbi.nlm.nih.gov/2304166
548. Neill, M.G., et al. Does intrafascial dissection during nerve-sparing laparoscopic radical
prostatectomy compromise cancer control? BJU Int, 2009. 104: 1730.
https://pubmed.ncbi.nlm.nih.gov/20063449
549. Ward, J.F., et al. The impact of surgical approach (nerve bundle preservation versus wide local
excision) on surgical margins and biochemical recurrence following radical prostatectomy. J Urol,
2004. 172: 1328.
https://pubmed.ncbi.nlm.nih.gov/15371834
550. Engel, J., et al. Survival benefit of radical prostatectomy in lymph node-positive patients with
prostate cancer. Eur Urol, 2010. 57: 754.
https://pubmed.ncbi.nlm.nih.gov/20106588
551. Beulens, A.J.W., et al. Linking surgical skills to postoperative outcomes: a Delphi study on the
robot-assisted radical prostatectomy. J Robot Surg, 2019. 13: 675.
https://pubmed.ncbi.nlm.nih.gov/30610535
552. Gilbert, S.M., et al. Functional Outcomes Following Nerve Sparing Prostatectomy Augmented with
Seminal Vesicle Sparing Compared to Standard Nerve Sparing Prostatectomy: Results from a
Randomized Controlled Trial. J Urol, 2017. 198: 600.
https://pubmed.ncbi.nlm.nih.gov/28392393
553. Korman, H.J., et al. Radical prostatectomy: is complete resection of the seminal vesicles really
necessary? J Urol, 1996. 156: 1081.
https://pubmed.ncbi.nlm.nih.gov/8709312
554. Steiner, M.S., et al. Impact of anatomical radical prostatectomy on urinary continence. J Urol, 1991.
145: 512.
https://pubmed.ncbi.nlm.nih.gov/1997701
555. Li, H., et al. The Use of Unidirectional Barbed Suture for Urethrovesical Anastomosis during Robot-
Assisted Radical Prostatectomy: A Systematic Review and Meta-Analysis of Efficacy and Safety.
PLoS One, 2015. 10: e0131167.
https://pubmed.ncbi.nlm.nih.gov/26135310

PROSTATE CANCER - LIMITED UPDATE 2021 165


556. Kowalewski, K.F., et al. Interrupted versus Continuous Suturing for Vesicourethral Anastomosis During
Radical Prostatectomy: A Systematic Review and Meta-analysis. Eur Urol Focus, 2019. 5: 980.
https://pubmed.ncbi.nlm.nih.gov/29907547
557. Matsuyama, H., et al. Running suture versus interrupted suture for vesicourethral anastomosis in
retropubic radical prostatectomy: a randomized study. Int J Urol, 2015. 22: 271.
https://pubmed.ncbi.nlm.nih.gov/25400263
558. Wiatr, T., et al. Single Running Suture versus Single-Knot Running Suture for Vesicourethral
Anastomosis in Laparoscopic Radical Prostatectomy: A Prospective Randomised Comparative
Study. Urol Int, 2015. 95: 445.
https://pubmed.ncbi.nlm.nih.gov/26655169
559. Van Velthoven, R.F., et al. Technique for laparoscopic running urethrovesical anastomosis:the single
knot method. Urology, 2003. 61: 699.
https://pubmed.ncbi.nlm.nih.gov/12670546
560. Vest, S.A. Radical penineal prostatectomy. Surg Gynecol Obstet, 1940. 70: 935. [No abstract
available].
561. Igel, T.C., et al. Comparison of techniques for vesicourethral anastomosis: simple direct versus
modified Vest traction sutures. Urology, 1988. 31: 474.
https://pubmed.ncbi.nlm.nih.gov/3287741
562. Berlin, J.W., et al. Voiding cystourethrography after radical prostatectomy: normal findings and
correlation between contrast extravasation and anastomotic strictures. AJR Am J Roentgenol, 1994.
162: 87.
https://pubmed.ncbi.nlm.nih.gov/8273697
563. Levy, J.B., et al. Vesicourethral healing following radical prostatectomy: is it related to surgical
approach? Urology, 1994. 44: 888.
https://pubmed.ncbi.nlm.nih.gov/7985317
564. Novicki, D.E., et al. Comparison of the modified vest and the direct anastomosis for radical
retropubic prostatectomy. Urology, 1997. 49: 732.
https://pubmed.ncbi.nlm.nih.gov/9145979
565. Atherton, L., et al. Radical retropubic prostatectomy for carcinoma. J Urol, 1956. 75: 111.
https://pubmed.ncbi.nlm.nih.gov/13286806
566. Schoeppler, G.M., et al. The impact of bladder neck mucosal eversion during open radical
prostatectomy on bladder neck stricture and urinary extravasation. Int Urol Nephrol, 2012. 44: 1403.
https://pubmed.ncbi.nlm.nih.gov/22585294
567. Borboroglu, P.G., et al. Risk factors for vesicourethral anastomotic stricture after radical
prostatectomy. Urology, 2000. 56: 96.
https://pubmed.ncbi.nlm.nih.gov/10869633
568. Bellangino, M., et al. Systematic Review of Studies Reporting Positive Surgical Margins After
Bladder Neck Sparing Radical Prostatectomy. Curr Urol Rep, 2017. 18: 99.
https://pubmed.ncbi.nlm.nih.gov/29116405
569. Nyarangi-Dix, J.N., et al. Complete bladder neck preservation promotes long-term post-
prostatectomy continence without compromising midterm oncological outcome: analysis of a
randomised controlled cohort. World J Urol, 2018. 36: 349.
https://pubmed.ncbi.nlm.nih.gov/29214353
570. Ma, X., et al. Bladder neck preservation improves time to continence after radical prostatectomy: a
systematic review and meta-analysis. Oncotarget, 2016. 7: 67463.
https://pubmed.ncbi.nlm.nih.gov/27634899
571. Mungovan, S.F., et al. Preoperative Membranous Urethral Length Measurement and Continence
Recovery Following Radical Prostatectomy: A Systematic Review and Meta-analysis. Eur Urol,
2017. 71: 368.
https://pubmed.ncbi.nlm.nih.gov/27394644
572. Guru, K.A., et al. Is a cystogram necessary after robot-assisted radical prostatectomy? Urol Oncol,
2007. 25: 465.
https://pubmed.ncbi.nlm.nih.gov/18047953
573. Tillier, C., et al. Vesico-urethral anastomosis (VUA) evaluation of short- and long-term outcome
after robot-assisted laparoscopic radical prostatectomy (RARP): selective cystogram to improve
outcome. J Robot Surg, 2017. 11: 441.
https://pubmed.ncbi.nlm.nih.gov/28078524
574. Yadav, R., et al. Selective indication for check cystogram before catheter removal following robot
assisted radical prostatectomy. Indian J Urol, 2016. 32: 120.
https://pubmed.ncbi.nlm.nih.gov/27127354

166 PROSTATE CANCER - LIMITED UPDATE 2021


575. Schoeppler, G.M., et al. Detection of urinary leakage after radical retropubic prostatectomy by
contrast enhanced ultrasound - do we still need conventional retrograde cystography? BJU Int,
2010. 106: 1632.
https://pubmed.ncbi.nlm.nih.gov/20590540
576. Gratzke, C., et al. Early Catheter Removal after Robot-assisted Radical Prostatectomy: Surgical
Technique and Outcomes for the Aalst Technique (ECaRemA Study). Eur Urol, 2016. 69: 917.
https://pubmed.ncbi.nlm.nih.gov/26578444
577. James, P., et al. Safe removal of the urethral catheter 2 days following laparoscopic radical
prostatectomy. ISRN Oncol, 2012. 2012: 912642.
https://pubmed.ncbi.nlm.nih.gov/22957273
578. Lista, G., et al. Early Catheter Removal After Robot-assisted Radical Prostatectomy: Results from a
Prospective Single-institutional Randomized Trial (Ripreca Study). Eur Urol Focus, 2020. 6: 259.
https://pubmed.ncbi.nlm.nih.gov/30413390
579. Brassetti, A., et al. Removing the urinary catheter on post-operative day 2 after robot-assisted
laparoscopic radical prostatectomy: a feasibility study from a single high-volume referral centre.
J Robot Surg, 2018. 12: 467.
https://pubmed.ncbi.nlm.nih.gov/29177945
580. Tilki, D., et al. The impact of time to catheter removal on short-, intermediate- and long-term urinary
continence after radical prostatectomy. World J Urol, 2018. 36: 1247.
https://pubmed.ncbi.nlm.nih.gov/29582100
581. Berrondo, C., et al. Antibiotic prophylaxis at the time of catheter removal after radical
prostatectomy: A prospective randomized clinical trial. Urol Oncol, 2019. 37: 181 e7.
https://pubmed.ncbi.nlm.nih.gov/30558984
582. Martinschek, A., et al. Transurethral versus suprapubic catheter at robot-assisted radical
prostatectomy: a prospective randomized trial with 1-year follow-up. World J Urol, 2016. 34: 407.
https://pubmed.ncbi.nlm.nih.gov/26337521
583. Harke, N., et al. Postoperative patient comfort in suprapubic drainage versus transurethral
catheterization following robot-assisted radical prostatectomy: a prospective randomized clinical
trial. World J Urol, 2017. 35: 389.
https://pubmed.ncbi.nlm.nih.gov/27334135
584. Krane, L.S., et al. Impact of percutaneous suprapubic tube drainage on patient discomfort after
radical prostatectomy. Eur Urol, 2009. 56: 325.
https://pubmed.ncbi.nlm.nih.gov/19394131
585. Morgan, M.S., et al. An Assessment of Patient Comfort and Morbidity After Robot-Assisted Radical
Prostatectomy with Suprapubic Tube Versus Urethral Catheter Drainage. J Endourol, 2016. 30: 300.
https://pubmed.ncbi.nlm.nih.gov/26472083
586. Galfano, A., et al. Pain and discomfort after Retzius-sparing robot-assisted radical prostatectomy:
a comparative study between suprapubic cystostomy and urethral catheter as urinary drainage.
Minerva Urol Nefrol, 2019. 71: 381.
https://pubmed.ncbi.nlm.nih.gov/31144484
587. Prasad, S.M., et al. Early removal of urethral catheter with suprapubic tube drainage versus urethral
catheter drainage alone after robot-assisted laparoscopic radical prostatectomy. J Urol, 2014.
192: 89.
https://pubmed.ncbi.nlm.nih.gov/24440236
588. Afzal, M.Z., et al. Modification of Technique for Suprapubic Catheter Placement After Robot-
assisted Radical Prostatectomy Reduces Catheter-associated Complications. Urology, 2015.
86: 401.
https://pubmed.ncbi.nlm.nih.gov/26189333
589. Porcaro, A.B., et al. Is a Drain Needed After Robotic Radical Prostatectomy With or Without Pelvic
Lymph Node Dissection? Results of a Single-Center Randomized Clinical Trial. J Endourol, 2019.
https://pubmed.ncbi.nlm.nih.gov/30398382
590. Chenam, A., et al. Prospective randomised non-inferiority trial of pelvic drain placement vs no pelvic
drain placement after robot-assisted radical prostatectomy. BJU Int, 2018. 121: 357.
https://pubmed.ncbi.nlm.nih.gov/28872774
591. Novara, G., et al. Systematic review and meta-analysis of studies reporting oncologic outcome after
robot-assisted radical prostatectomy. Eur Urol, 2012. 62: 382.
https://pubmed.ncbi.nlm.nih.gov/22749851
592. Novara, G., et al. Systematic review and meta-analysis of perioperative outcomes and
complications after robot-assisted radical prostatectomy. Eur Urol, 2012. 62: 431.
https://pubmed.ncbi.nlm.nih.gov/22749853

PROSTATE CANCER - LIMITED UPDATE 2021 167


593. Ficarra, V., et al. Systematic review and meta-analysis of studies reporting potency rates after robot-
assisted radical prostatectomy. Eur Urol, 2012. 62: 418.
https://pubmed.ncbi.nlm.nih.gov/22749850
594. Ficarra, V., et al. Systematic review and meta-analysis of studies reporting urinary continence
recovery after robot-assisted radical prostatectomy. Eur Urol, 2012. 62: 405.
https://pubmed.ncbi.nlm.nih.gov/22749852
595. Maffezzini, M., et al. Evaluation of complications and results in a contemporary series of 300
consecutive radical retropubic prostatectomies with the anatomic approach at a single institution.
Urology, 2003. 61: 982.
https://pubmed.ncbi.nlm.nih.gov/12736020
596. Haglind, E., et al. Urinary Incontinence and Erectile Dysfunction After Robotic Versus Open Radical
Prostatectomy: A Prospective, Controlled, Nonrandomised Trial. Eur Urol, 2015. 68: 216.
https://pubmed.ncbi.nlm.nih.gov/25770484
597. Joshi, N., et al. Impact of posterior musculofascial reconstruction on early continence after robot-
assisted laparoscopic radical prostatectomy: results of a prospective parallel group trial. Eur Urol,
2010. 58: 84.
https://pubmed.ncbi.nlm.nih.gov/20362386
598. Sutherland, D.E., et al. Posterior rhabdosphincter reconstruction during robotic assisted radical
prostatectomy: results from a phase II randomized clinical trial. J Urol, 2011. 185: 1262.
https://pubmed.ncbi.nlm.nih.gov/21334025
599. Jeong, C.W., et al. Effects of new 1-step posterior reconstruction method on recovery of continence
after robot-assisted laparoscopic prostatectomy: results of a prospective, single-blind, parallel
group, randomized, controlled trial. J Urol, 2015. 193: 935.
https://pubmed.ncbi.nlm.nih.gov/25315960
600. Menon, M., et al. Assessment of early continence after reconstruction of the periprostatic tissues
in patients undergoing computer assisted (robotic) prostatectomy: results of a 2 group parallel
randomized controlled trial. J Urol, 2008. 180: 1018.
https://pubmed.ncbi.nlm.nih.gov/18639300
601. Stolzenburg, J.U., et al. Influence of bladder neck suspension stitches on early continence after
radical prostatectomy: a prospective randomized study of 180 patients. Asian J Androl, 2011.
13: 806.
https://pubmed.ncbi.nlm.nih.gov/21909121
602. Hurtes, X., et al. Anterior suspension combined with posterior reconstruction during robot-assisted
laparoscopic prostatectomy improves early return of urinary continence: a prospective randomized
multicentre trial. BJU Int, 2012. 110: 875.
https://pubmed.ncbi.nlm.nih.gov/22260307
603. Student, V., Jr., et al. Advanced Reconstruction of Vesicourethral Support (ARVUS) during Robot-
assisted Radical Prostatectomy: One-year Functional Outcomes in a Two-group Randomised
Controlled Trial. Eur Urol, 2017. 71: 822.
https://pubmed.ncbi.nlm.nih.gov/27283216
604. Noguchi, M., et al. A randomized clinical trial of suspension technique for improving early recovery
of urinary continence after radical retropubic prostatectomy. BJU Int, 2008. 102: 958.
https://pubmed.ncbi.nlm.nih.gov/18485031
605. Tikkinen, K.A.O., et al. EAU Guidelines on Thromboprophylaxis in Urological Surgery, Edn. presented
at the 32nd Annual Congress, London 2017. EAU Guidelines Office, Arnhem, The Netherlands.
https://uroweb.org/guideline/thromboprophylaxis/
606. Burkhard, F.C., et al. The role of lymphadenectomy in prostate cancer. Nat Clin Pract Urol, 2005.
2: 336.
https://pubmed.ncbi.nlm.nih.gov/16474786
607. Ploussard, G., et al. Pelvic lymph node dissection during robot-assisted radical prostatectomy:
efficacy, limitations, and complications-a systematic review of the literature. Eur Urol, 2014. 65: 7.
https://pubmed.ncbi.nlm.nih.gov/23582879
608. Davis, J.W., et al. Robot-assisted extended pelvic lymph node dissection (PLND) at the time of
radical prostatectomy (RP): a video-based illustration of technique, results, and unmet patient
selection needs. BJU Int, 2011. 108: 993.
https://pubmed.ncbi.nlm.nih.gov/21917102
609. Briganti, A., et al. Complications and other surgical outcomes associated with extended pelvic
lymphadenectomy in men with localized prostate cancer. Eur Urol, 2006. 50: 1006.
https://pubmed.ncbi.nlm.nih.gov/16959399

168 PROSTATE CANCER - LIMITED UPDATE 2021


610. Viani, G.A., et al. Intensity-modulated radiotherapy reduces toxicity with similar biochemical control
compared with 3-dimensional conformal radiotherapy for prostate cancer: A randomized clinical
trial. Cancer, 2016. 122: 2004.
https://pubmed.ncbi.nlm.nih.gov/27028170
611. Yu, T., et al. The Effectiveness of Intensity Modulated Radiation Therapy versus Three-Dimensional
Radiation Therapy in Prostate Cancer: A Meta-Analysis of the Literatures. PLoS One, 2016.
11: e0154499.
https://pubmed.ncbi.nlm.nih.gov/27171271
612. Wortel, R.C., et al. Late Side Effects After Image Guided Intensity Modulated Radiation Therapy
Compared to 3D-Conformal Radiation Therapy for Prostate Cancer: Results From 2 Prospective
Cohorts. Int J Radiat Oncol Biol Phys, 2016. 95: 680.
https://pubmed.ncbi.nlm.nih.gov/27055398
613. Zapatero, A., et al. Reduced late urinary toxicity with high-dose intensity-modulated radiotherapy
using intra-prostate fiducial markers for localized prostate cancer. Clin Transl Oncol, 2017. 19: 1161.
https://pubmed.ncbi.nlm.nih.gov/28374321
614. de Crevoisier, R., et al. Daily Versus Weekly Prostate Cancer Image Guided Radiation Therapy:
Phase 3 Multicenter Randomized Trial. Int J Radiat Oncol Biol Phys, 2018. 102: 1420.
https://pubmed.ncbi.nlm.nih.gov/30071296
615. Kishan, A.U., et al. Local Failure and Survival After Definitive Radiotherapy for Aggressive Prostate
Cancer: An Individual Patient-level Meta-analysis of Six Randomized Trials. Eur Urol, 2020. 77: 201.
https://pubmed.ncbi.nlm.nih.gov/31718822
616. Michalski, J.M., et al. Effect of Standard vs Dose-Escalated Radiation Therapy for Patients With
Intermediate-Risk Prostate Cancer: The NRG Oncology RTOG 0126 Randomized Clinical Trial.
JAMA Oncol, 2018. 4: e180039.
https://pubmed.ncbi.nlm.nih.gov/29543933
617. Zietman, A.L., et al. Randomized trial comparing conventional-dose with high-dose conformal
radiation therapy in early-stage adenocarcinoma of the prostate: long-term results from proton
radiation oncology group/american college of radiology 95-09. J Clin Oncol, 2010. 28: 1106.
https://pubmed.ncbi.nlm.nih.gov/20124169
618. Viani, G.A., et al. Higher-than-conventional radiation doses in localized prostate cancer treatment: a
meta-analysis of randomized, controlled trials. Int J Radiat Oncol Biol Phys, 2009. 74: 1405.
https://pubmed.ncbi.nlm.nih.gov/19616743
619. Peeters, S.T., et al. Dose-response in radiotherapy for localized prostate cancer: results of the Dutch
multicenter randomized phase III trial comparing 68 Gy of radiotherapy with 78 Gy. J Clin Oncol,
2006. 24: 1990.
https://pubmed.ncbi.nlm.nih.gov/16648499
620 Beckendorf, V., et al. 70 Gy versus 80 Gy in localized prostate cancer: 5-year results of GETUG 06
randomized trial. Int J Radiat Oncol Biol Phys, 2011. 80: 1056.
https://pubmed.ncbi.nlm.nih.gov/21147514
621. Heemsbergen, W.D., et al. Long-term results of the Dutch randomized prostate cancer trial: impact of
dose-escalation on local, biochemical, clinical failure, and survival. Radiother Oncol, 2014. 110: 104.
https://pubmed.ncbi.nlm.nih.gov/24246414
622. Dearnaley, D.P., et al. Escalated-dose versus control-dose conformal radiotherapy for prostate cancer:
long-term results from the MRC RT01 randomised controlled trial. Lancet Oncol, 2014. 15: 464.
https://pubmed.ncbi.nlm.nih.gov/24581940
623. Pasalic, D., et al. Dose Escalation for Prostate Adenocarcinoma: A Long-Term Update on the
Outcomes of a Phase 3, Single Institution Randomized Clinical Trial. Int J Radiat Oncol Biol Phys,
2019. 104: 790.
https://pubmed.ncbi.nlm.nih.gov/30836166
624. Kalbasi, A., et al. Dose-Escalated Irradiation and Overall Survival in Men With Nonmetastatic
Prostate Cancer. JAMA Oncol, 2015. 1: 897.
https://pubmed.ncbi.nlm.nih.gov/26181727
625. Zietman, A.L., et al. Comparison of conventional-dose vs high-dose conformal radiation therapy
in clinically localized adenocarcinoma of the prostate: a randomized controlled trial. JAMA, 2005.
294: 1233.
https://pubmed.ncbi.nlm.nih.gov/16160131
626. Peeters, S.T., et al. Acute and late complications after radiotherapy for prostate cancer: results
of a multicenter randomized trial comparing 68 Gy to 78 Gy. Int J Radiat Oncol Biol Phys, 2005.
61: 1019.
https://pubmed.ncbi.nlm.nih.gov/15752881

PROSTATE CANCER - LIMITED UPDATE 2021 169


627. Dearnaley, D.P., et al. The early toxicity of escalated versus standard dose conformal radiotherapy
with neo-adjuvant androgen suppression for patients with localised prostate cancer: results from the
MRC RT01 trial (ISRCTN47772397). Radiother Oncol, 2007. 83: 31.
https://pubmed.ncbi.nlm.nih.gov/17391791
628. Kuban, D.A., et al. Long-term results of the M. D. Anderson randomized dose-escalation trial for
prostate cancer. Int J Radiat Oncol Biol Phys, 2008. 70: 67.
https://pubmed.ncbi.nlm.nih.gov/17765406
629. Matzinger, O., et al. Acute toxicity of curative radiotherapy for intermediate- and high-risk localised
prostate cancer in the EORTC trial 22991. Eur J Cancer, 2009. 45: 2825.
https://pubmed.ncbi.nlm.nih.gov/19682889
630. Zapatero, A., et al. Risk-Adapted Androgen Deprivation and Escalated Three-Dimensional
Conformal Radiotherapy for Prostate Cancer: Does Radiation Dose Influence Outcome of Patients
Treated With Adjuvant Androgen Deprivation? A GICOR Study. J Clin Oncol, 2005. 23: 6561.
https://pubmed.ncbi.nlm.nih.gov/16170164
631. Zelefsky, M.J., et al. Long-term outcome of high dose intensity modulated radiation therapy for
patients with clinically localized prostate cancer. J Urol, 2006. 176: 1415.
https://pubmed.ncbi.nlm.nih.gov/16952647
632. Vora, S.A., et al. Analysis of biochemical control and prognostic factors in patients treated with
either low-dose three-dimensional conformal radiation therapy or high-dose intensity-modulated
radiotherapy for localized prostate cancer. Int J Radiat Oncol Biol Phys, 2007. 68: 1053.
https://pubmed.ncbi.nlm.nih.gov/17398023
633. Kupelian, P.A., et al. Hypofractionated intensity-modulated radiotherapy (70 Gy at 2.5 Gy per
fraction) for localized prostate cancer: Cleveland Clinic experience. Int J Radiat Oncol Biol Phys,
2007. 68: 1424.
https://pubmed.ncbi.nlm.nih.gov/17544601
634. Gomez-Iturriaga Pina, A., et al. Median 5 year follow-up of 125iodine brachytherapy as
monotherapy in men aged<or=55 years with favorable prostate cancer. Urology, 2010. 75: 1412.
https://pubmed.ncbi.nlm.nih.gov/20035986
635. Ishiyama, H., et al. Genitourinary toxicity after high-dose-rate (HDR) brachytherapy combined
with Hypofractionated External beam radiotherapy for localized prostate cancer: an analysis to
determine the correlation between dose-volume histogram parameters in HDR brachytherapy and
severity of toxicity. Int J Radiat Oncol Biol Phys, 2009. 75: 23.
https://pubmed.ncbi.nlm.nih.gov/19243900
636. Gelblum, D.Y., et al. Rectal complications associated with transperineal interstitial brachytherapy for
prostate cancer. Int J Radiat Oncol Biol Phys, 2000. 48: 119.
https://pubmed.ncbi.nlm.nih.gov/10924980
637. Lee, W.R., et al. Late toxicity and biochemical recurrence after external-beam radiotherapy
combined with permanent-source prostate brachytherapy. Cancer, 2007. 109: 1506.
https://pubmed.ncbi.nlm.nih.gov/17340591
638. Zelefsky, M.J., et al. Five-year biochemical outcome and toxicity with transperineal CT-planned
permanent I-125 prostate implantation for patients with localized prostate cancer. Int J Radiat Oncol
Biol Phys, 2000. 47: 1261.
https://pubmed.ncbi.nlm.nih.gov/10889379
639. Fowler, J.F. The radiobiology of prostate cancer including new aspects of fractionated radiotherapy.
Acta Oncol, 2005. 44: 265.
https://pubmed.ncbi.nlm.nih.gov/16076699
640. Dasu, A., et al. Prostate alpha/beta revisited -- an analysis of clinical results from 14 168 patients.
Acta Oncol, 2012. 51: 963.
https://pubmed.ncbi.nlm.nih.gov/22966812
641. Dearnaley, D., et al. Conventional versus hypofractionated high-dose intensity-modulated
radiotherapy for prostate cancer: preliminary safety results from the CHHiP randomised controlled
trial. Lancet Oncol, 2012. 13: 43.
https://pubmed.ncbi.nlm.nih.gov/22169269
642. Kuban, D.A., et al. Preliminary Report of a Randomized Dose Escalation Trial for Prostate Cancer
using Hypofractionation. Int J Radiat Oncol Biol Phys, 2010. 78: S58.
https://www.redjournal.org/article/S0360-3016(10)01144-2/abstract
643. Pollack, A., et al. Randomized trial of hypofractionated external-beam radiotherapy for prostate
cancer. J Clin Oncol, 2013. 31: 3860.
https://pubmed.ncbi.nlm.nih.gov/24101042

170 PROSTATE CANCER - LIMITED UPDATE 2021


644. Aluwini, S., et al. Hypofractionated versus conventionally fractionated radiotherapy for patients with
prostate cancer (HYPRO): acute toxicity results from a randomised non-inferiority phase 3 trial.
Lancet Oncol, 2015. 16: 274.
https://pubmed.ncbi.nlm.nih.gov/25656287
645. Lee, W.R., et al. Randomized Phase III Noninferiority Study Comparing Two Radiotherapy
Fractionation Schedules in Patients With Low-Risk Prostate Cancer. J Clin Oncol, 2016. 34: 2325.
https://pubmed.ncbi.nlm.nih.gov/27044935
646. Dearnaley, D., et al. Conventional versus hypofractionated high-dose intensity-modulated
radiotherapy for prostate cancer: 5-year outcomes of the randomised, non-inferiority, phase 3
CHHiP trial. Lancet Oncol, 2016. 17: 1047.
https://pubmed.ncbi.nlm.nih.gov/27339115
647. Aluwini, S., et al. Hypofractionated versus conventionally fractionated radiotherapy for patients
with prostate cancer (HYPRO): late toxicity results from a randomised, non-inferiority, phase 3 trial.
Lancet Oncol, 2016. 17: 464.
https://pubmed.ncbi.nlm.nih.gov/26968359
648. Incrocci, L., et al. Hypofractionated versus conventionally fractionated radiotherapy for patients with
localised prostate cancer (HYPRO): final efficacy results from a randomised, multicentre, open-label,
phase 3 trial. Lancet Oncol, 2016. 17: 1061.
https://pubmed.ncbi.nlm.nih.gov/27339116
649. Catton, C.N., et al. Randomized Trial of a Hypofractionated Radiation Regimen for the Treatment of
Localized Prostate Cancer. J Clin Oncol, 2017. 35: 1884.
https://pubmed.ncbi.nlm.nih.gov/28296582
650. Koontz, B.F., et al. A systematic review of hypofractionation for primary management of prostate
cancer. Eur Urol, 2015. 68: 683.
https://pubmed.ncbi.nlm.nih.gov/25171903
651. Hocht, S., et al. Hypofractionated radiotherapy for localized prostate cancer. Strahlenther Onkol,
2017. 193: 1.
https://pubmed.ncbi.nlm.nih.gov/27628966
652. Hickey, B.E., et al. Hypofractionation for clinically localized prostate cancer. Cochrane Database
Syst Rev, 2019. 9: CD011462.
https://pubmed.ncbi.nlm.nih.gov/31476800
653. de Vries, K.C., et al. Hyprofractionated Versus Conventionally Fractionated Radiation Therapy for
Patients with Intermediate- or High-Risk, Localized, Prostate Cancer: 7-Year Outcomes From the
Randomized, Multicenter, Open-Label, Phase 3 HYPRO Trial. Int J Radiat Oncol Biol Phys, 2020.
106: 108.
https://pubmed.ncbi.nlm.nih.gov/31593756
654. Aluwini, S., et al. Stereotactic body radiotherapy with a focal boost to the MRI-visible tumor as
monotherapy for low- and intermediate-risk prostate cancer: early results. Radiat Oncol, 2013. 8: 84.
https://pubmed.ncbi.nlm.nih.gov/23570391
655. Katz, A., et al. Stereotactic Body Radiation Therapy for Low-, Intermediate-, and High-Risk Prostate
Cancer: Disease Control and Quality of Life at 6 Years. Int J Radiat Oncol Biol Phys, 2013. 87: S24.
https://www.redjournal.org/article/S0360-3016(13)00738-4/fulltext
656. Widmark, A., et al. Ultra-hypofractionated versus conventionally fractionated radiotherapy for
prostate cancer: 5-year outcomes of the HYPO-RT-PC randomised, non-inferiority, phase 3 trial.
Lancet, 2019. 394: 385.
https://pubmed.ncbi.nlm.nih.gov/31227373
657. Jackson, W.C., et al. Stereotactic Body Radiation Therapy for Localized Prostate Cancer: A
Systematic Review and Meta-Analysis of Over 6,000 Patients Treated On Prospective Studies. Int J
Radiat Oncol Biol Phys, 2019. 104: 778.
https://pubmed.ncbi.nlm.nih.gov/30959121
658. Cushman, T.R., et al. Stereotactic body radiation therapy for prostate cancer: systematic review and
meta-analysis of prospective trials. Oncotarget, 2019. 10: 5660.
https://pubmed.ncbi.nlm.nih.gov/31608141
659. Brand, D.H., et al. Intensity-modulated fractionated radiotherapy versus stereotactic body
radiotherapy for prostate cancer (PACE-B): acute toxicity findings from an international, randomised,
open-label, phase 3, non-inferiority trial. Lancet Oncol, 2019. 20: 1531.
https://pubmed.ncbi.nlm.nih.gov/31540791
660. Bolla, M., et al. External irradiation with or without long-term androgen suppression for prostate
cancer with high metastatic risk: 10-year results of an EORTC randomised study. Lancet Oncol,
2010. 11: 1066.
https://pubmed.ncbi.nlm.nih.gov/20933466

PROSTATE CANCER - LIMITED UPDATE 2021 171


661. Pilepich, M.V., et al. Androgen suppression adjuvant to definitive radiotherapy in prostate
carcinoma--long-term results of phase III RTOG 85-31. Int J Radiat Oncol Biol Phys, 2005. 61: 1285.
https://pubmed.ncbi.nlm.nih.gov/15817329
662. Roach, M., 3rd, et al. Short-term neoadjuvant androgen deprivation therapy and external-beam
radiotherapy for locally advanced prostate cancer: long-term results of RTOG 8610. J Clin Oncol,
2008. 26: 585.
https://pubmed.ncbi.nlm.nih.gov/18172188
663. D’Amico, A.V., et al. Androgen suppression and radiation vs radiation alone for prostate cancer: a
randomized trial. JAMA, 2008. 299: 289.
https://pubmed.ncbi.nlm.nih.gov/18212313
664. Denham, J.W., et al. Short-term neoadjuvant androgen deprivation and radiotherapy for locally
advanced prostate cancer: 10-year data from the TROG 96.01 randomised trial. Lancet Oncol,
2011. 12: 451.
https://pubmed.ncbi.nlm.nih.gov/21440505
665. Spratt, D.E., et al. Prostate Radiotherapy With Adjuvant Androgen Deprivation Therapy (ADT)
Improves Metastasis-Free Survival Compared to Neoadjuvant ADT: An Individual Patient Meta-
Analysis. J Clin Oncol, 2021. 39: 136.
https://pubmed.ncbi.nlm.nih.gov/33275486
666. Lawton, C.A., et al. An update of the phase III trial comparing whole pelvic to prostate only
radiotherapy and neoadjuvant to adjuvant total androgen suppression: updated analysis of RTOG
94-13, with emphasis on unexpected hormone/radiation interactions. Int J Radiat Oncol Biol Phys,
2007. 69: 646.
https://pubmed.ncbi.nlm.nih.gov/17531401
667. Horwitz, E.M., et al. Ten-year follow-up of radiation therapy oncology group protocol 92-02: a phase
III trial of the duration of elective androgen deprivation in locally advanced prostate cancer. J Clin
Oncol, 2008. 26: 2497.
https://pubmed.ncbi.nlm.nih.gov/18413638
668. Bolla, M., et al. Duration of androgen suppression in the treatment of prostate cancer. N Engl J Med,
2009. 360: 2516.
https://pubmed.ncbi.nlm.nih.gov/19516032
669. Pisansky, T.M., et al. Duration of androgen suppression before radiotherapy for localized prostate
cancer: radiation therapy oncology group randomized clinical trial 9910. J Clin Oncol, 2015. 33: 332.
https://pubmed.ncbi.nlm.nih.gov/25534388
670. Fossa, S.D., et al. Ten- and 15-yr Prostate Cancer-specific Mortality in Patients with Nonmetastatic
Locally Advanced or Aggressive Intermediate Prostate Cancer, Randomized to Lifelong Endocrine
Treatment Alone or Combined with Radiotherapy: Final Results of The Scandinavian Prostate
Cancer Group-7. Eur Urol, 2016. 70: 684.
https://pubmed.ncbi.nlm.nih.gov/27025586
671. Warde, P., et al. Combined androgen deprivation therapy and radiation therapy for locally advanced
prostate cancer: a randomised, phase 3 trial. Lancet, 2011. 378: 2104.
https://pubmed.ncbi.nlm.nih.gov/22056152
672. Mason, M.D., et al. Final Report of the Intergroup Randomized Study of Combined Androgen-
Deprivation Therapy Plus Radiotherapy Versus Androgen-Deprivation Therapy Alone in Locally
Advanced Prostate Cancer. J Clin Oncol, 2015. 33: 2143.
https://pubmed.ncbi.nlm.nih.gov/25691677
673. Sargos, P., et al. Long-term androgen deprivation, with or without radiotherapy, in locally advanced
prostate cancer: updated results from a phase III randomised trial. BJU Int, 2020. 125: 810.
https://pubmed.ncbi.nlm.nih.gov/30946523
674. Zelefsky, M.J., et al. Dose escalation for prostate cancer radiotherapy: predictors of long-term
biochemical tumor control and distant metastases-free survival outcomes. Eur Urol, 2011. 60: 1133.
https://pubmed.ncbi.nlm.nih.gov/21889832
675. Zapatero, A., et al. High-dose radiotherapy with short-term or long-term androgen deprivation in
localised prostate cancer (DART01/05 GICOR): a randomised, controlled, phase 3 trial. Lancet
Oncol, 2015. 16: 320.
https://pubmed.ncbi.nlm.nih.gov/25702876
676. Bolla, M., et al. Short Androgen Suppression and Radiation Dose Escalation for Intermediate- and
High-Risk Localized Prostate Cancer: Results of EORTC Trial 22991. J Clin Oncol, 2016. 34: 1748.
https://pubmed.ncbi.nlm.nih.gov/26976418
677. Gray, P.J., et al. Patient-reported outcomes after 3-dimensional conformal, intensity-modulated, or
proton beam radiotherapy for localized prostate cancer. Cancer, 2013. 119: 1729.
https://pubmed.ncbi.nlm.nih.gov/23436283

172 PROSTATE CANCER - LIMITED UPDATE 2021


678. Sheets, N.C., et al. Intensity-modulated radiation therapy, proton therapy, or conformal radiation
therapy and morbidity and disease control in localized prostate cancer. JAMA, 2012. 307: 1611.
https://pubmed.ncbi.nlm.nih.gov/22511689
679. National Institute for Health and Care Excellence (NICE). Biodegradable spacer insertion to reduce
rectal toxicity during radiotherapy for prostate cancer. Interventional procedures guidance [IPG590].
2017. [access data March 2021].
https://www.nice.org.uk/guidance/ipg590
680. Miller, L.E., et al. Association of the Placement of a Perirectal Hydrogel Spacer With the Clinical
Outcomes of Men Receiving Radiotherapy for Prostate Cancer: A Systematic Review and Meta-
analysis. JAMA Netw Open, 2020. 3: e208221.
https://pubmed.ncbi.nlm.nih.gov/32585020
681. Hamstra, D.A., et al. Continued Benefit to Rectal Separation for Prostate Radiation Therapy: Final
Results of a Phase III Trial. Int J Radiat Oncol Biol Phys, 2017. 97: 976.
https://pubmed.ncbi.nlm.nih.gov/28209443
682. Aminsharifi, A., et al. Major Complications and Adverse Events Related to the Injection of the
SpaceOAR Hydrogel System Before Radiotherapy for Prostate Cancer: Review of the Manufacturer
and User Facility Device Experience Database. J Endourol, 2019. 33: 868.
https://pubmed.ncbi.nlm.nih.gov/31452385
683. Pinkawa, M., et al. Learning curve in the application of a hydrogel spacer to protect the rectal wall
during radiotherapy of localized prostate cancer. Urology, 2013. 82: 963.
https://pubmed.ncbi.nlm.nih.gov/24074991
684. Ash, D., et al. ESTRO/EAU/EORTC recommendations on permanent seed implantation for localized
prostate cancer. Radiother Oncol, 2000. 57: 315.
https://pubmed.ncbi.nlm.nih.gov/11104892
685. Martens, C., et al. Relationship of the International Prostate Symptom score with urinary flow
studies, and catheterization rates following 125I prostate brachytherapy. Brachytherapy, 2006. 5: 9.
https://pubmed.ncbi.nlm.nih.gov/16563992
686. Le, H., et al. The influence of prostate volume on outcome after high-dose-rate brachytherapy alone
for localized prostate cancer. Int J Radiat Oncol Biol Phys, 2013. 87: 270.
https://pubmed.ncbi.nlm.nih.gov/23849693
687. Salembier, C., et al. A history of transurethral resection of the prostate should not be a contra-
indication for low-dose-rate (125)I prostate brachytherapy: results of a prospective Uro-GEC phase-
II trial. J Contemp Brachytherapy, 2020. 12: 1.
https://pubmed.ncbi.nlm.nih.gov/32190063
688. Salembier, C., et al. Prospective multi-center dosimetry study of low-dose Iodine-125 prostate
brachytherapy performed after transurethral resection. J Contemp Brachytherapy, 2013. 5: 63.
https://pubmed.ncbi.nlm.nih.gov/23878549
689. Stone, N.N., et al. Prostate brachytherapy in men with gland volume of 100cc or greater: Technique,
cancer control, and morbidity. Brachytherapy, 2013. 12: 217.
https://pubmed.ncbi.nlm.nih.gov/23384439
690. Crook, J.M., et al. Comparison of health-related quality of life 5 years after SPIRIT: Surgical
Prostatectomy Versus Interstitial Radiation Intervention Trial. J Clin Oncol, 2011. 29: 362.
https://pubmed.ncbi.nlm.nih.gov/21149658
691. Machtens, S., et al. Long-term results of interstitial brachytherapy (LDR-Brachytherapy) in the
treatment of patients with prostate cancer. World J Urol, 2006. 24: 289.
https://pubmed.ncbi.nlm.nih.gov/16645877
692. Grimm, P., et al. Comparative analysis of prostate-specific antigen free survival outcomes for
patients with low, intermediate and high risk prostate cancer treatment by radical therapy. Results
from the Prostate Cancer Results Study Group. BJU Int, 2012. 109 Suppl 1: 22.
https://pubmed.ncbi.nlm.nih.gov/22239226
693. Potters, L., et al. Monotherapy for stage T1-T2 prostate cancer: radical prostatectomy, external
beam radiotherapy, or permanent seed implantation. Radiother Oncol, 2004. 71: 29.
https://pubmed.ncbi.nlm.nih.gov/15066293
694. Sylvester, J.E., et al. Fifteen-year biochemical relapse-free survival, cause-specific survival, and
overall survival following I(125) prostate brachytherapy in clinically localized prostate cancer: Seattle
experience. Int J Radiat Oncol Biol Phys, 2011. 81: 376.
https://pubmed.ncbi.nlm.nih.gov/20864269
695. Potters, L., et al. 12-year outcomes following permanent prostate brachytherapy in patients with
clinically localized prostate cancer. J Urol, 2005. 173: 1562.
https://pubmed.ncbi.nlm.nih.gov/15821486

PROSTATE CANCER - LIMITED UPDATE 2021 173


696. Stone, N.N., et al. Intermediate term biochemical-free progression and local control following
125iodine brachytherapy for prostate cancer. J Urol, 2005. 173: 803.
https://pubmed.ncbi.nlm.nih.gov/15711273
697. Zelefsky, M.J., et al. Multi-institutional analysis of long-term outcome for stages T1-T2 prostate
cancer treated with permanent seed implantation. Int J Radiat Oncol Biol Phys, 2007. 67: 327.
https://pubmed.ncbi.nlm.nih.gov/17084558
698. Lawton, C.A., et al. Results of a phase II trial of transrectal ultrasound-guided permanent radioactive
implantation of the prostate for definitive management of localized adenocarcinoma of the prostate
(radiation therapy oncology group 98-05). Int J Radiat Oncol Biol Phys, 2007. 67: 39.
https://pubmed.ncbi.nlm.nih.gov/17084551
699. Stock, R.G., et al. Importance of post-implant dosimetry in permanent prostate brachytherapy. Eur
Urol, 2002. 41: 434.
https://pubmed.ncbi.nlm.nih.gov/12074816
700. Hoskin, P.J., et al. GEC/ESTRO recommendations on high dose rate afterloading brachytherapy for
localised prostate cancer: an update. Radiother Oncol, 2013. 107: 325.
https://pubmed.ncbi.nlm.nih.gov/23773409
701. Galalae, R.M., et al. Long-term outcome after elective irradiation of the pelvic lymphatics and
local dose escalation using high-dose-rate brachytherapy for locally advanced prostate cancer. Int
J Radiat Oncol Biol Phys, 2002. 52: 81.
https://pubmed.ncbi.nlm.nih.gov/11777625
702. Hoskin, P.J., et al. Randomised trial of external beam radiotherapy alone or combined with high-
dose-rate brachytherapy boost for localised prostate cancer. Radiother Oncol, 2012. 103: 217.
https://pubmed.ncbi.nlm.nih.gov/22341794
703. Pieters, B.R., et al. Comparison of three radiotherapy modalities on biochemical control and overall
survival for the treatment of prostate cancer: a systematic review. Radiother Oncol, 2009. 93: 168.
https://pubmed.ncbi.nlm.nih.gov/19748692
704. Hauswald, H., et al. High-Dose-Rate Monotherapy for Localized Prostate Cancer: 10-Year Results.
Int J Radiat Oncol Biol Phys, 2016. 94: 667.
https://pubmed.ncbi.nlm.nih.gov/26443877
705. Zamboglou, N., et al. High-dose-rate interstitial brachytherapy as monotherapy for clinically
localized prostate cancer: treatment evolution and mature results. Int J Radiat Oncol Biol Phys,
2013. 85: 672.
https://pubmed.ncbi.nlm.nih.gov/22929859
706. Hoskin, P.J., et al. High dose rate brachytherapy in combination with external beam radiotherapy in
the radical treatment of prostate cancer: initial results of a randomised phase three trial. Radiother
Oncol, 2007. 84: 114.
https://pubmed.ncbi.nlm.nih.gov/17531335
707. Hoskin, P., et al. High-dose-rate brachytherapy alone given as two or one fraction to patients for
locally advanced prostate cancer: acute toxicity. Radiother Oncol, 2014. 110: 268.
https://pubmed.ncbi.nlm.nih.gov/24231242
708. King, C.R., et al. Health-related quality of life after stereotactic body radiation therapy for localized
prostate cancer: results from a multi-institutional consortium of prospective trials. Int J Radiat Oncol
Biol Phys, 2013. 87: 939.
https://pubmed.ncbi.nlm.nih.gov/24119836
709. Pagliarulo, V., et al. Contemporary role of androgen deprivation therapy for prostate cancer. Eur
Urol, 2012. 61: 11.
https://pubmed.ncbi.nlm.nih.gov/21871711
710. Oefelein, M.G., et al. Reassessment of the definition of castrate levels of testosterone: implications
for clinical decision making. Urology, 2000. 56: 1021.
https://pubmed.ncbi.nlm.nih.gov/11113751
711. Morote, J., et al. Individual variations of serum testosterone in patients with prostate cancer
receiving androgen deprivation therapy. BJU Int, 2009. 103: 332.
https://pubmed.ncbi.nlm.nih.gov/19007366
712. Pickles, T., et al. Incomplete testosterone suppression with luteinizing hormone-releasing hormone
agonists: does it happen and does it matter? BJU Int, 2012. 110: E500.
https://pubmed.ncbi.nlm.nih.gov/22564197
713. Klotz, L., et al. MP74-01 Nadir testosterone within first year of androgen-deprivation therapy (ADT)
predicts for time to castration-resistant progression: a secondary analysis of the PR-7 trial of
intermittent versus continuous ADT. J Urol. 191: e855.
https://pubmed.ncbi.nlm.nih.gov/25732157

174 PROSTATE CANCER - LIMITED UPDATE 2021


714. Desmond, A.D., et al. Subcapsular orchiectomy under local anaesthesia. Technique, results and
implications. Br J Urol, 1988. 61: 143.
https://pubmed.ncbi.nlm.nih.gov/3349279
715. Scherr, D.S., et al. The nonsteroidal effects of diethylstilbestrol: the rationale for androgen deprivation
therapy without estrogen deprivation in the treatment of prostate cancer. J Urol, 2003. 170: 1703.
https://pubmed.ncbi.nlm.nih.gov/14532759
716. Klotz, L., et al. A phase 1-2 trial of diethylstilbestrol plus low dose warfarin in advanced prostate
carcinoma. J Urol, 1999. 161: 169.
https://pubmed.ncbi.nlm.nih.gov/10037391
717. Farrugia, D., et al. Stilboestrol plus adrenal suppression as salvage treatment for patients failing
treatment with luteinizing hormone-releasing hormone analogues and orchidectomy. BJU Int, 2000.
85: 1069.
https://pubmed.ncbi.nlm.nih.gov/10848697
718. Hedlund, P.O., et al. Parenteral estrogen versus combined androgen deprivation in the treatment
of metastatic prostatic cancer: part 2. Final evaluation of the Scandinavian Prostatic Cancer Group
(SPCG) Study No. 5. Scand J Urol Nephrol, 2008. 42: 220.
https://pubmed.ncbi.nlm.nih.gov/18432528
719. Bubley, G.J. Is the flare phenomenon clinically significant? Urology, 2001. 58: 5.
https://pubmed.ncbi.nlm.nih.gov/11502435
720. Collette, L., et al. Why phase III trials of maximal androgen blockade versus castration in M1
prostate cancer rarely show statistically significant differences. Prostate, 2001. 48: 29.
https://pubmed.ncbi.nlm.nih.gov/11391684
721. Krakowsky, Y., et al. Risk of Testosterone Flare in the Era of the Saturation Model: One More
Historical Myth. Eur Urol Focus, 2019. 5: 81.
https://pubmed.ncbi.nlm.nih.gov/28753828
722. Klotz, L., et al. The efficacy and safety of degarelix: a 12-month, comparative, randomized, open-
label, parallel-group phase III study in patients with prostate cancer. BJU Int, 2008. 102: 1531.
https://pubmed.ncbi.nlm.nih.gov/19035858
723. Seidenfeld, J., et al. Single-therapy androgen suppression in men with advanced prostate cancer: a
systematic review and meta-analysis. Ann Intern Med, 2000. 132: 566.
https://pubmed.ncbi.nlm.nih.gov/10744594
724. Ostergren, P.B., et al. Luteinizing Hormone-Releasing Hormone Agonists are Superior to
Subcapsular Orchiectomy in Lowering Testosterone Levels of Men with Prostate Cancer: Results
from a Randomized Clinical Trial. J Urol, 2017. 197: 1441.
https://pubmed.ncbi.nlm.nih.gov/27939836
725. Shore, N.D. Experience with degarelix in the treatment of prostate cancer. Ther Adv Urol, 2013. 5: 11.
https://pubmed.ncbi.nlm.nih.gov/23372607
726. Sciarra, A., et al. A meta-analysis and systematic review of randomized controlled trials with
degarelix versus gonadotropin-releasing hormone agonists for advanced prostate cancer. Medicine
(Baltimore), 2016. 95: e3845.
https://pubmed.ncbi.nlm.nih.gov/27399062
727. Shore, N.D., et al. Oral Relugolix for Androgen-Deprivation Therapy in Advanced Prostate Cancer.
N Engl J Med, 2020. 382: 2187.
https://pubmed.ncbi.nlm.nih.gov/32469183
728. U.S. Food and Drug Adminstration approves relugolix for advanced prostate cancer 2020 [access
date March 2021].
https://www.fda.gov/drugs/drug-approvals-and-databases/fda-approves-relugolix-advanced-
prostate-cancer
729. Moffat, L.E. Comparison of Zoladex, diethylstilbestrol and cyproterone acetate treatment in
advanced prostate cancer. Eur Urol, 1990. 18 Suppl 3: 26.
https://pubmed.ncbi.nlm.nih.gov/2151272
730. Schroder, F.H., et al. Metastatic prostate cancer treated by flutamide versus cyproterone acetate.
Final analysis of the “European Organization for Research and Treatment of Cancer” (EORTC)
Protocol 30892. Eur Urol, 2004. 45: 457.
https://pubmed.ncbi.nlm.nih.gov/15041109
731. Smith, M.R., et al. Bicalutamide monotherapy versus leuprolide monotherapy for prostate cancer:
effects on bone mineral density and body composition. J Clin Oncol, 2004. 22: 2546.
https://pubmed.ncbi.nlm.nih.gov/15226323
732. Iversen, P. Antiandrogen monotherapy: indications and results. Urology, 2002. 60: 64.
https://pubmed.ncbi.nlm.nih.gov/12231053

PROSTATE CANCER - LIMITED UPDATE 2021 175


733. Wadhwa, V.K., et al. Long-term changes in bone mineral density and predicted fracture risk in
patients receiving androgen-deprivation therapy for prostate cancer, with stratification of treatment
based on presenting values. BJU Int, 2009. 104: 800.
https://pubmed.ncbi.nlm.nih.gov/19338564
734. Montgomery, R.B., et al. Maintenance of intratumoral androgens in metastatic prostate cancer: a
mechanism for castration-resistant tumor growth. Cancer Res, 2008. 68: 4447.
https://pubmed.ncbi.nlm.nih.gov/18519708
735. U.S. Food and Drug Adminstration approves abiraterone acetate in combination with prednisone for
high-risk metastatic castration-sensitive prostate cancer. 2018 [access date March 2021].
https://www.fda.gov/drugs/resources-information-approved-drugs/fda-approves-abiraterone-
acetate-combination-prednisone-high-risk-metastatic-castration-sensitive
736. U.S. Food and Drug Adminstration approves enzalutamide for metastatic castration-sensitive
prostate cancer 2019 [access date March 2021].
https://www.fda.gov/drugs/resources-information-approved-drugs/fda-approves-enzalutamide-
metastatic-castration-sensitive-prostate-cancer
737. U.S. Food and Drug Adminstration approves apalutamide for metastatic castration-sensitive
prostate cancer. 2019 [access date March 2021].
https://www.fda.gov/drugs/resources-information-approved-drugs/fda-approves-apalutamide-
metastatic-castration-sensitive-prostate-cancer
738. European Medicines Agency. Zytiga. 2011 [access date March 2021].
https://www.ema.europa.eu/en/medicines/human/EPAR/zytiga
739. European Medicines Agency. Erleada (apalutamide). 2019 [access date March 2021].
https://www.ema.europa.eu/en/medicines/human/EPAR/erleada
740. European Medicines Agency. Nubeqa (darolutamide). 2020 [access date March 2021].
https://www.ema.europa.eu/en/medicines/human/EPAR/nubeqa
741. European Medicines Agency. Xtandi (enzalutamide). 2013 [access date March 2021].
https://www.ema.europa.eu/en/medicines/human/EPAR/xtandi
742. Chi, K.N., et al. Apalutamide for Metastatic, Castration-Sensitive Prostate Cancer. N Engl J Med,
2019. 381: 13.
https://pubmed.ncbi.nlm.nih.gov/31150574
743. Armstrong, A.J., et al. ARCHES: A Randomized, Phase III Study of Androgen Deprivation Therapy
With Enzalutamide or Placebo in Men With Metastatic Hormone-Sensitive Prostate Cancer. J Clin
Oncol, 2019. 37: 2974.
https://pubmed.ncbi.nlm.nih.gov/31329516
744. Fizazi, K., et al. Abiraterone plus Prednisone in Metastatic, Castration-Sensitive Prostate Cancer.
N Engl J Med, 2017. 377: 352.
https://pubmed.ncbi.nlm.nih.gov/28578607
745. Moilanen, A.M., et al. Discovery of ODM-201, a new-generation androgen receptor inhibitor
targeting resistance mechanisms to androgen signaling-directed prostate cancer therapies. Sci Rep,
2015. 5: 12007.
https://pubmed.ncbi.nlm.nih.gov/26137992
746. Zurth, C., et al. Blood-brain barrier penetration of [14C]darolutamide compared with [14C]
enzalutamide in rats using whole body autoradiography. J Clin Oncol, 2018. 36: 345.
https://ascopubs.org/doi/abs/10.1200/JCO.2018.36.6_suppl.345
747. Lord, C.J., et al. PARP inhibitors: Synthetic lethality in the clinic. Science, 2017. 355: 1152.
https://pubmed.ncbi.nlm.nih.gov/28302823
748. Darvin, P., et al. Immune checkpoint inhibitors: recent progress and potential biomarkers. Exp
Molecul Med, 2018. 50: 1.
https://www.researchgate.net/publication/329617084
749. Hargadon, K.M., et al. Immune checkpoint blockade therapy for cancer: An overview of FDA-
approved immune checkpoint inhibitors. Int Immunopharmacol, 2018. 62: 29.
https://pubmed.ncbi.nlm.nih.gov/29990692
750. de Bono, J.S., et al. Randomized Phase II Study Evaluating Akt Blockade with Ipatasertib, in
Combination with Abiraterone, in Patients with Metastatic Prostate Cancer with and without PTEN
Loss. Clin Cancer Res, 2019. 25: 928.
https://pubmed.ncbi.nlm.nih.gov/30037818
751. Sarker, D., et al. Targeting the PI3K/AKT pathway for the treatment of prostate cancer. Clin Cancer
Res, 2009. 15: 4799.
https://pubmed.ncbi.nlm.nih.gov/19638457
752. Fahmy, W.E., et al. Cryosurgery for prostate cancer. Arch Androl, 2003. 49: 397.
https://pubmed.ncbi.nlm.nih.gov/12893518

176 PROSTATE CANCER - LIMITED UPDATE 2021


753. Rees, J., et al. Cryosurgery for prostate cancer. BJU Int, 2004. 93: 710.
https://pubmed.ncbi.nlm.nih.gov/15049977
754. Han, K.R., et al. Third-generation cryosurgery for primary and recurrent prostate cancer. BJU Int,
2004. 93: 14.
https://pubmed.ncbi.nlm.nih.gov/14678360
755. Beerlage, H.P., et al. Current status of minimally invasive treatment options for localized prostate
carcinoma. Eur Urol, 2000. 37: 2.
https://pubmed.ncbi.nlm.nih.gov/10671777
756. van der Poel, H.G., et al. Focal Therapy in Primary Localised Prostate Cancer: The European
Association of Urology Position in 2018. Eur Urol, 2018. 74: 84.
https://pubmed.ncbi.nlm.nih.gov/29373215
757. Valerio, M., et al. New and Established Technology in Focal Ablation of the Prostate: A Systematic
Review. Eur Urol, 2017. 71: 17.
https://pubmed.ncbi.nlm.nih.gov/27595377
758. Ramsay, C.R., et al. Ablative therapy for people with localised prostate cancer: a systematic review
and economic evaluation. Health Technol Assess, 2015. 19: 1.
https://pubmed.ncbi.nlm.nih.gov/26140518
759. Madersbacher, S., et al. High-energy shockwaves and extracorporeal high-intensity focused
ultrasound. J Endourol, 2003. 17: 667.
https://pubmed.ncbi.nlm.nih.gov/14622487
760. Mouraviev, V., et al. Pathologic basis of focal therapy for early-stage prostate cancer. Nat Rev Urol,
2009. 6: 205.
https://pubmed.ncbi.nlm.nih.gov/19352395
761. Cooperberg, M.R., et al. Contemporary trends in low risk prostate cancer: risk assessment and
treatment. J Urol, 2007. 178: S14.
https://pubmed.ncbi.nlm.nih.gov/17644125
762. Polascik, T.J., et al. Pathologic stage T2a and T2b prostate cancer in the recent prostate-specific
antigen era: implications for unilateral ablative therapy. Prostate, 2008. 68: 1380.
https://pubmed.ncbi.nlm.nih.gov/18543281
763. Ahmed, H.U., et al. Will focal therapy become a standard of care for men with localized prostate
cancer? Nat Clin Pract Oncol, 2007. 4: 632.
https://pubmed.ncbi.nlm.nih.gov/17965641
764. Eggener, S.E., et al. Focal therapy for localized prostate cancer: a critical appraisal of rationale and
modalities. J Urol, 2007. 178: 2260.
https://pubmed.ncbi.nlm.nih.gov/17936815
765. Crawford, E.D., et al. Targeted focal therapy: a minimally invasive ablation technique for early
prostate cancer. Oncology (Williston Park), 2007. 21: 27.
https://pubmed.ncbi.nlm.nih.gov/17313155
766. Bates, A.S., Ayers, J., et al. A Systematic Review of Focal Ablative Therapy for Clinically Localised
Prostate Cancer in Comparison with Standard Management Options: Limitations of the Available
Evidence and Recommendations for Clinical Practice and Further Research. Eur Urol Focus, 2021.
https://pubmed.ncbi.nlm.nih.gov/33423943
767. Azzouzi, A.R., et al. Padeliporfin vascular-targeted photodynamic therapy versus active surveillance
in men with low-risk prostate cancer (CLIN1001 PCM301): an open-label, phase 3, randomised
controlled trial. Lancet Oncol, 2017. 18: 181.
https://pubmed.ncbi.nlm.nih.gov/28007457
768. Gill, I.S., et al. Randomized Trial of Partial Gland Ablation with Vascular Targeted Phototherapy
versus Active Surveillance for Low Risk Prostate Cancer: Extended Followup and Analyses of
Effectiveness. J Urol, 2018. 200: 786.
https://pubmed.ncbi.nlm.nih.gov/29864437
769. Albisinni, S., et al. Comparing High-Intensity Focal Ultrasound Hemiablation to Robotic Radical
Prostatectomy in the Management of Unilateral Prostate Cancer: A Matched-Pair Analysis.
J Endourol, 2017. 31: 14.
https://pubmed.ncbi.nlm.nih.gov/27799004
770. Zheng, X., et al. Focal Laser Ablation Versus Radical Prostatectomy for Localized Prostate Cancer:
Survival Outcomes From a Matched Cohort. Clin Genitourin Cancer, 2019. 17: 464.
https://pubmed.ncbi.nlm.nih.gov/31594734
771. Zhou, X., et al. Comparative Effectiveness of Radiotherapy versus Focal Laser Ablation in Patients
with Low and Intermediate Risk Localized Prostate Cancer. Sci Rep, 2020. 10: 9112.
https://pubmed.ncbi.nlm.nih.gov/32499484

PROSTATE CANCER - LIMITED UPDATE 2021 177


772. MacLennan, S., et al. A core outcome set for localised prostate cancer effectiveness trials. BJU Int,
2017. 120: E64.
https://pubmed.ncbi.nlm.nih.gov/28346770
773. Guillaumier, S., et al. A Multicentre Study of 5-year Outcomes Following Focal Therapy in Treating
Clinically Significant Nonmetastatic Prostate Cancer. Eur Urol, 2018. 74: 422.
https://pubmed.ncbi.nlm.nih.gov/29960750
774. Loeb, S., et al. Active surveillance for prostate cancer: a systematic review of clinicopathologic
variables and biomarkers for risk stratification. Eur Urol, 2015. 67: 619.
https://pubmed.ncbi.nlm.nih.gov/25457014
775. Ha, Y.S., et al. Prostate-specific antigen density toward a better cutoff to identify better candidates
for active surveillance. Urology, 2014. 84: 365.
https://pubmed.ncbi.nlm.nih.gov/24925834
776. Morash, C., et al. Active surveillance for the management of localized prostate cancer: Guideline
recommendations. Can Urol Assoc J, 2015. 9: 171.
https://pubmed.ncbi.nlm.nih.gov/26225165
777. Satasivam, P., et al. Can Confirmatory Biopsy be Omitted in Patients with Prostate Cancer
Favorable Diagnostic Features on Active Surveillance? J Urol, 2016. 195: 74.
https://pubmed.ncbi.nlm.nih.gov/26192258
778. Petrelli, F., et al. Predictive Factors for Reclassification and Relapse in Prostate Cancer Eligible for
Active Surveillance: A Systematic Review and Meta-analysis. Urology, 2016. 91: 136.
https://pubmed.ncbi.nlm.nih.gov/26896733
779. Moreira, D.M., et al. Baseline Perineural Invasion is Associated with Shorter Time to Progression in
Men with Prostate Cancer Undergoing Active Surveillance: Results from the REDEEM Study. J Urol,
2015. 194: 1258.
https://pubmed.ncbi.nlm.nih.gov/25988518
780. Dieffenbacher, S., et al. Standardized Magnetic Resonance Imaging Reporting Using the Prostate
Cancer Radiological Estimation of Change in Sequential Evaluation Criteria and Magnetic
Resonance Imaging/Transrectal Ultrasound Fusion with Transperineal Saturation Biopsy to Select
Men on Active Surveillance. Eur Urol Focus, 2020. 7: 102.
https://pubmed.ncbi.nlm.nih.gov/30878348
781. Gallagher, K.M., et al. Four-year outcomes from a multiparametric magnetic resonance imaging
(MRI)-based active surveillance programme: PSA dynamics and serial MRI scans allow omission of
protocol biopsies. BJU Int, 2019. 123: 429.
https://pubmed.ncbi.nlm.nih.gov/30113755
782. Schoots, I.G., et al. Is magnetic resonance imaging-targeted biopsy a useful addition to systematic
confirmatory biopsy in men on active surveillance for low-risk prostate cancer? A systematic review
and meta-analysis. BJU Int, 2018. 122: 946.
https://pubmed.ncbi.nlm.nih.gov/29679430
783. Mamawala, M.K., et al. Utility of multiparametric magnetic resonance imaging in the risk
stratification of men with Grade Group 1 prostate cancer on active surveillance. BJU Int, 2020.
125: 861.
https://pubmed.ncbi.nlm.nih.gov/32039537
784. Klotz, L., et al. Active Surveillance Magnetic Resonance Imaging Study (ASIST): Results of a
Randomized Multicenter Prospective Trial. Eur Urol, 2019. 75: 300.
https://pubmed.ncbi.nlm.nih.gov/30017404
785. Klotz, L., et al. Randomized Study of Systematic Biopsy Versus Magnetic Resonance Imaging and
Targeted and Systematic Biopsy in Men on Active Surveillance (ASIST): 2-year Postbiopsy Follow-
up. Eur Urol, 2020. 77: 311.
https://pubmed.ncbi.nlm.nih.gov/31708295
786. Schoots, I.G., et al. Role of MRI in low-risk prostate cancer: finding the wolf in sheep’s clothing or
the sheep in wolf’s clothing? Curr Opin Urol, 2017. 27: 238.
https://pubmed.ncbi.nlm.nih.gov/28306604
787. Hsiang, W., et al. Outcomes of Serial Multiparametric Magnetic Resonance Imaging and
Subsequent Biopsy in Men with Low-risk Prostate Cancer Managed with Active Surveillance. Eur
Urol Focus, 2019.
https://pubmed.ncbi.nlm.nih.gov/31147263
788. Osses, D.F., et al. Prostate cancer upgrading with serial prostate magnetic resonance imaging and
repeat biopsy in men on active surveillance: are confirmatory biopsies still necessary? BJU Int,
2020. 126: 124.
https://pubmed.ncbi.nlm.nih.gov/32232921

178 PROSTATE CANCER - LIMITED UPDATE 2021


789. Chesnut, G.T., et al. Role of Changes in Magnetic Resonance Imaging or Clinical Stage in Evaluation
of Disease Progression for Men with Prostate Cancer on Active Surveillance. Eur Urol, 2020. 77: 501.
https://pubmed.ncbi.nlm.nih.gov/31874726
790. Hamoen, E.H.J., et al. Value of Serial Multiparametric Magnetic Resonance Imaging and Magnetic
Resonance Imaging-guided Biopsies in Men with Low-risk Prostate Cancer on Active Surveillance
After 1 Yr Follow-up. Eur Urol Focus, 2019. 5: 407.
https://pubmed.ncbi.nlm.nih.gov/29331622
791. Thurtle, D., et al. Progression and treatment rates using an active surveillance protocol incorporating
image-guided baseline biopsies and multiparametric magnetic resonance imaging monitoring for
men with favourable-risk prostate cancer. BJU Int, 2018. 122: 59.
https://pubmed.ncbi.nlm.nih.gov/29438586
792. Olivier, J., et al. Low-risk prostate cancer selected for active surveillance with negative MRI at entry:
can repeat biopsies at 1 year be avoided? A pilot study. World J Urol, 2019. 37: 253.
https://pubmed.ncbi.nlm.nih.gov/30039385
793. Klotz, L., et al. Clinical results of long-term follow-up of a large, active surveillance cohort with
localized prostate cancer. J Clin Oncol, 2010. 28: 126.
https://pubmed.ncbi.nlm.nih.gov/19917860
794. Ross, A.E., et al. Prostate-specific antigen kinetics during follow-up are an unreliable trigger for
intervention in a prostate cancer surveillance program. J Clin Oncol, 2010. 28: 2810.
https://pubmed.ncbi.nlm.nih.gov/20439642
795. Thomsen, F.B., et al. Association between PSA kinetics and cancer-specific mortality in patients
with localised prostate cancer: analysis of the placebo arm of the SPCG-6 study. Ann Oncol, 2016.
27: 460.
https://pubmed.ncbi.nlm.nih.gov/26681677
796. Immediate versus deferred treatment for advanced prostatic cancer: initial results of the Medical
Research Council Trial. The Medical Research Council Prostate Cancer Working Party Investigators
Group. Br J Urol, 1997. 79: 235.
https://pubmed.ncbi.nlm.nih.gov/9052476
797. Chen, R.C., et al. Active Surveillance for the Management of Localized Prostate Cancer (Cancer
Care Ontario Guideline): American Society of Clinical Oncology Clinical Practice Guideline
Endorsement. J Clin Oncol, 2016. 34: 2182.
https://pubmed.ncbi.nlm.nih.gov/26884580
798. Musunuru, H.B., et al. Active Surveillance for Intermediate Risk Prostate Cancer: Survival Outcomes
in the Sunnybrook Experience. J Urol, 2016. 196: 1651.
https://pubmed.ncbi.nlm.nih.gov/27569437
799. Raldow, A.C., et al. Risk Group and Death From Prostate Cancer: Implications for Active
Surveillance in Men With Favorable Intermediate-Risk Prostate Cancer. JAMA Oncol, 2015. 1: 334.
https://pubmed.ncbi.nlm.nih.gov/26181182
800. Studer, U.E., et al. Using PSA to guide timing of androgen deprivation in patients with T0-4 N0-2 M0
prostate cancer not suitable for local curative treatment (EORTC 30891). Eur Urol, 2008. 53: 941.
https://pubmed.ncbi.nlm.nih.gov/18191322
801. James, N.D., et al. Addition of docetaxel, zoledronic acid, or both to first-line long-term hormone
therapy in prostate cancer (STAMPEDE): survival results from an adaptive, multiarm, multistage,
platform randomised controlled trial. Lancet, 2016. 387: 1163.
https://pubmed.ncbi.nlm.nih.gov/26719232
802. Krauss, D., et al. Lack of benefit for the addition of androgen deprivation therapy to dose-escalated
radiotherapy in the treatment of intermediate- and high-risk prostate cancer. Int J Radiat Oncol Biol
Phys, 2011. 80: 1064.
https://pubmed.ncbi.nlm.nih.gov/20584576
803. Kupelian, P.A., et al. Effect of increasing radiation doses on local and distant failures in patients with
localized prostate cancer. Int J Radiat Oncol Biol Phys, 2008. 71: 16.
https://pubmed.ncbi.nlm.nih.gov/17996382
804. Chang, K., et al. Comparison of two adjuvant hormone therapy regimens in patients with high-
risk localized prostate cancer after radical prostatectomy: primary results of study CU1005. Asian
J Androl, 2016. 18: 452.
https://pubmed.ncbi.nlm.nih.gov/26323560
805. Walsh, P.C. Immediate versus deferred treatment for advanced prostatic cancer: initial results of
the Medical Research Council trial. The Medical Research Council Prostate Cancer Working Party
Investigators Group. J Urol, 1997. 158: 1623.
https://pubmed.ncbi.nlm.nih.gov/9302187

PROSTATE CANCER - LIMITED UPDATE 2021 179


806. Donohue, J.F., et al. Poorly differentiated prostate cancer treated with radical prostatectomy: long-
term outcome and incidence of pathological downgrading. J Urol, 2006. 176: 991.
https://pubmed.ncbi.nlm.nih.gov/16890678
807. Bianco, F.J., Jr., et al. Radical prostatectomy: long-term cancer control and recovery of sexual and
urinary function (“trifecta”). Urology, 2005. 66: 83.
https://pubmed.ncbi.nlm.nih.gov/16194712
808. Walz, J., et al. A nomogram predicting 10-year life expectancy in candidates for radical
prostatectomy or radiotherapy for prostate cancer. J Clin Oncol, 2007. 25: 3576.
https://pubmed.ncbi.nlm.nih.gov/17704404
809. Eastham, J.A., et al. Variations among individual surgeons in the rate of positive surgical margins in
radical prostatectomy specimens. J Urol, 2003. 170: 2292.
https://pubmed.ncbi.nlm.nih.gov/14634399
810. Vickers, A.J., et al. The surgical learning curve for laparoscopic radical prostatectomy: a
retrospective cohort study. Lancet Oncol, 2009. 10: 475.
https://pubmed.ncbi.nlm.nih.gov/19342300
811. Trinh, Q.D., et al. A systematic review of the volume-outcome relationship for radical prostatectomy.
Eur Urol, 2013. 64: 786.
https://pubmed.ncbi.nlm.nih.gov/23664423
812. Pisansky, T.M., et al. Correlation of pretherapy prostate cancer characteristics with histologic
findings from pelvic lymphadenectomy specimens. Int J Radiat Oncol Biol Phys, 1996. 34: 33.
https://pubmed.ncbi.nlm.nih.gov/12118563
813. Joniau, S., et al. Pretreatment tables predicting pathologic stage of locally advanced prostate
cancer. Eur Urol, 2015. 67: 319.
https://pubmed.ncbi.nlm.nih.gov/24684960
814. Loeb, S., et al. Intermediate-term potency, continence, and survival outcomes of radical
prostatectomy for clinically high-risk or locally advanced prostate cancer. Urology, 2007. 69: 1170.
https://pubmed.ncbi.nlm.nih.gov/17572209
815. Carver, B.S., et al. Long-term outcome following radical prostatectomy in men with clinical stage T3
prostate cancer. J Urol, 2006. 176: 564.
https://pubmed.ncbi.nlm.nih.gov/16813890
816. Freedland, S.J., et al. Radical prostatectomy for clinical stage T3a disease. Cancer, 2007. 109: 1273.
https://pubmed.ncbi.nlm.nih.gov/17315165
817. Joniau, S., et al. Radical prostatectomy in very high-risk localized prostate cancer: long-term
outcomes and outcome predictors. Scand J Urol Nephrol, 2012. 46: 164.
https://pubmed.ncbi.nlm.nih.gov/22364377
818. Johnstone, P.A., et al. Radical prostatectomy for clinical T4 prostate cancer. Cancer, 2006. 106: 2603.
https://pubmed.ncbi.nlm.nih.gov/16700037
819. Leibel, S.A., et al. The effects of local and regional treatment on the metastatic outcome in prostatic
carcinoma with pelvic lymph node involvement. Int J Radiat Oncol Biol Phys, 1994. 28: 7.
https://pubmed.ncbi.nlm.nih.gov/8270461
820. Asbell, S.O., et al. Elective pelvic irradiation in stage A2, B carcinoma of the prostate: analysis of
RTOG 77-06. Int J Radiat Oncol Biol Phys, 1988. 15: 1307.
https://pubmed.ncbi.nlm.nih.gov/3058656
821. Pommier, P., et al. Is there a role for pelvic irradiation in localized prostate adenocarcinoma?
Preliminary results of GETUG-01. J Clin Oncol, 2007. 25: 5366.
https://pubmed.ncbi.nlm.nih.gov/18048817
822. Joseph, D., et al. Radiation Dose Escalation or Longer Androgen Suppression to Prevent Distant
Progression in Men With Locally Advanced Prostate Cancer: 10-Year Data From the TROG 03.04
RADAR Trial. Int J Radiat Oncol Biol Phys, 2020. 106: 693.
https://pubmed.ncbi.nlm.nih.gov/32092343
823. Morris, W.J., et al. Androgen Suppression Combined with Elective Nodal and Dose Escalated
Radiation Therapy (the ASCENDE-RT Trial): An Analysis of Survival Endpoints for a Randomized Trial
Comparing a Low-Dose-Rate Brachytherapy Boost to a Dose-Escalated External Beam Boost for
High- and Intermediate-risk Prostate Cancer. Int J Radiat Oncol Biol Phys, 2017. 98: 275.
https://pubmed.ncbi.nlm.nih.gov/28262473
824. Rodda, S., et al. ASCENDE-RT: An Analysis of Treatment-Related Morbidity for a Randomized Trial
Comparing a Low-Dose-Rate Brachytherapy Boost with a Dose-Escalated External Beam Boost for
High- and Intermediate-Risk Prostate Cancer. Int J Radiat Oncol Biol Phys, 2017. 98: 286.
https://pubmed.ncbi.nlm.nih.gov/28433432

180 PROSTATE CANCER - LIMITED UPDATE 2021


825. Jackson, W.C., et al. Addition of Androgen-Deprivation Therapy or Brachytherapy Boost to External
Beam Radiotherapy for Localized Prostate Cancer: A Network Meta-Analysis of Randomized Trials.
J Clin Oncol, 2020. 38: 3024.
https://pubmed.ncbi.nlm.nih.gov/32396488
826. Keyes, M., et al. American Brachytherapy Society Task Group Report: Use of androgen deprivation
therapy with prostate brachytherapy-A systematic literature review. Brachytherapy, 2017. 16: 245.
https://pubmed.ncbi.nlm.nih.gov/28110898
827. Moris, L., et al. Benefits and Risks of Primary Treatments for High-risk Localized and Locally Advanced
Prostate Cancer: An International Multidisciplinary Systematic Review. Eur Urol, 2020. 77: 614.
https://pubmed.ncbi.nlm.nih.gov/32146018
828. Yossepowitch, O., et al. Radical prostatectomy for clinically localized, high risk prostate cancer:
critical analysis of risk assessment methods. J Urol, 2007. 178: 493.
https://pubmed.ncbi.nlm.nih.gov/17561152
829. Bastian, P.J., et al. Clinical and pathologic outcome after radical prostatectomy for prostate cancer
patients with a preoperative Gleason sum of 8 to 10. Cancer, 2006. 107: 1265.
https://pubmed.ncbi.nlm.nih.gov/16900523
830. Surgery Versus Radiotherapy for Locally Advanced Prostate Cancer (SPCG-15). 2014. 2019.
https://clinicaltrials.gov/ct2/show/NCT02102477
831. Walz, J., et al. Pathological results and rates of treatment failure in high-risk prostate cancer patients
after radical prostatectomy. BJU Int, 2011. 107: 765.
https://pubmed.ncbi.nlm.nih.gov/20875089
832. D’Amico, A.V., et al. Pretreatment nomogram for prostate-specific antigen recurrence after radical
prostatectomy or external-beam radiation therapy for clinically localized prostate cancer. J Clin
Oncol, 1999. 17: 168.
https://pubmed.ncbi.nlm.nih.gov/10458230
833. Spahn, M., et al. Outcome predictors of radical prostatectomy in patients with prostate-specific
antigen greater than 20 ng/ml: a European multi-institutional study of 712 patients. Eur Urol, 2010.
58: 1.
https://pubmed.ncbi.nlm.nih.gov/20299147
834. Zwergel, U., et al. Outcome of prostate cancer patients with initial PSA> or =20 ng/ml undergoing
radical prostatectomy. Eur Urol, 2007. 52: 1058.
https://pubmed.ncbi.nlm.nih.gov/17418938
835. Magheli, A., et al. Importance of tumor location in patients with high preoperative prostate specific
antigen levels (greater than 20 ng/ml) treated with radical prostatectomy. J Urol, 2007. 178: 1311.
https://pubmed.ncbi.nlm.nih.gov/17698095
836. Gerber, G.S., et al. Results of radical prostatectomy in men with locally advanced prostate cancer:
multi-institutional pooled analysis. Eur Urol, 1997. 32: 385.
https://pubmed.ncbi.nlm.nih.gov/9412793
837. Ward, J.F., et al. Radical prostatectomy for clinically advanced (cT3) prostate cancer since the
advent of prostate-specific antigen testing: 15-year outcome. BJU Int, 2005. 95: 751.
https://pubmed.ncbi.nlm.nih.gov/15794776
838. Moschini, M., et al. Outcomes for Patients with Clinical Lymphadenopathy Treated with Radical
Prostatectomy. Eur Urol, 2016. 69: 193.
https://pubmed.ncbi.nlm.nih.gov/26264160
839. Ventimiglia, E., et al. A Systematic Review of the Role of Definitive Local Treatment in Patients with
Clinically Lymph Node-positive Prostate Cancer. Eur Urol Oncol, 2019. 2: 294.
https://pubmed.ncbi.nlm.nih.gov/31200844
840. Tward, J.D., et al. Radiation therapy for clinically node-positive prostate adenocarcinoma is correlated
with improved overall and prostate cancer-specific survival. Pract Radiat Oncol, 2013. 3: 234.
https://pubmed.ncbi.nlm.nih.gov/24674370
841. Lin, C.C., et al. Androgen deprivation with or without radiation therapy for clinically node-positive
prostate cancer. J Natl Cancer Inst, 2015. 107.
https://pubmed.ncbi.nlm.nih.gov/25957435
842. Seisen, T., et al. Efficacy of Local Treatment in Prostate Cancer Patients with Clinically Pelvic Lymph
Node-positive Disease at Initial Diagnosis. Eur Urol, 2017. 73: 452.
https://pubmed.ncbi.nlm.nih.gov/28890245
843. James, N.D., et al. Failure-Free Survival and Radiotherapy in Patients With Newly Diagnosed
Nonmetastatic Prostate Cancer: Data From Patients in the Control Arm of the STAMPEDE Trial.
JAMA Oncol, 2016. 2: 348.
https://pubmed.ncbi.nlm.nih.gov/26606329

PROSTATE CANCER - LIMITED UPDATE 2021 181


844. Rusthoven, C.G., et al. The impact of definitive local therapy for lymph node-positive prostate
cancer: a population-based study. Int J Radiat Oncol Biol Phys, 2014. 88: 1064.
https://pubmed.ncbi.nlm.nih.gov/24661660
845. Bryant, A.K., et al. Definitive Radiation Therapy and Survival in Clinically Node-Positive Prostate
Cancer. Int J Radiat Oncol Biol Phys, 2018. 101: 1188.
https://pubmed.ncbi.nlm.nih.gov/29891203
846. Sarkar, R.R., et al. Association between Radical Prostatectomy and Survival in Men with Clinically
Node-positive Prostate Cancer. Eur Urol Oncol, 2019. 2: 584.
https://pubmed.ncbi.nlm.nih.gov/31411995
847. Studer, U.E., et al. Immediate or deferred androgen deprivation for patients with prostate cancer not
suitable for local treatment with curative intent: European Organisation for Research and Treatment
of Cancer (EORTC) Trial 30891. J Clin Oncol, 2006. 24: 1868.
https://pubmed.ncbi.nlm.nih.gov/16622261
848. Ghavamian, R., et al. Radical retropubic prostatectomy plus orchiectomy versus orchiectomy alone
for pTxN+ prostate cancer: a matched comparison. J Urol, 1999. 161: 1223.
https://pubmed.ncbi.nlm.nih.gov/10081874
849. Messing, E.M., et al. Immediate versus deferred androgen deprivation treatment in patients with
node-positive prostate cancer after radical prostatectomy and pelvic lymphadenectomy. Lancet
Oncol, 2006. 7: 472.
https://pubmed.ncbi.nlm.nih.gov/16750497
850. Hanks, G.E. External-beam radiation therapy for clinically localized prostate cancer: patterns of care
studies in the United States. NCI Monogr, 1988: 75.
https://pubmed.ncbi.nlm.nih.gov/3050542
851. Bader, P., et al. Is a limited lymph node dissection an adequate staging procedure for prostate
cancer? J Urol, 2002. 168: 514.
https://pubmed.ncbi.nlm.nih.gov/12131300
852. Briganti, A., et al. Two positive nodes represent a significant cut-off value for cancer specific
survival in patients with node positive prostate cancer. A new proposal based on a two-institution
experience on 703 consecutive N+ patients treated with radical prostatectomy, extended pelvic
lymph node dissection and adjuvant therapy. Eur Urol, 2009. 55: 261.
https://pubmed.ncbi.nlm.nih.gov/18838212
853. Schumacher, M.C., et al. Good outcome for patients with few lymph node metastases after radical
retropubic prostatectomy. Eur Urol, 2008. 54: 344.
https://pubmed.ncbi.nlm.nih.gov/18511183
854. Abdollah, F., et al. More extensive pelvic lymph node dissection improves survival in patients with
node-positive prostate cancer. Eur Urol, 2015. 67: 212.
https://pubmed.ncbi.nlm.nih.gov/24882672
855. Pound, C.R., et al. Natural history of progression after PSA elevation following radical
prostatectomy. JAMA, 1999. 281: 1591.
https://pubmed.ncbi.nlm.nih.gov/10235151
856. Aus, G., et al. Prognostic factors and survival in node-positive (N1) prostate cancer-a prospective
study based on data from a Swedish population-based cohort. Eur Urol, 2003. 43: 627.
https://pubmed.ncbi.nlm.nih.gov/12767363
857. Cheng, L., et al. Risk of prostate carcinoma death in patients with lymph node metastasis. Cancer,
2001. 91: 66.
https://pubmed.ncbi.nlm.nih.gov/11148561
858. Seiler, R., et al. Removal of limited nodal disease in patients undergoing radical prostatectomy:
long-term results confirm a chance for cure. J Urol, 2014. 191: 1280.
https://pubmed.ncbi.nlm.nih.gov/24262495
859. Passoni, N.M., et al. Prognosis of patients with pelvic lymph node (LN) metastasis after radical
prostatectomy: value of extranodal extension and size of the largest LN metastasis. BJU Int, 2014.
114: 503.
https://pubmed.ncbi.nlm.nih.gov/24053552
860. Daneshmand, S., et al. Prognosis of patients with lymph node positive prostate cancer following
radical prostatectomy: long-term results. J Urol. 172: 2252.
https://pubmed.ncbi.nlm.nih.gov/15538242
861. Touijer, K.A., et al. Long-term outcomes of patients with lymph node metastasis treated with radical
prostatectomy without adjuvant androgen-deprivation therapy. Eur Urol, 2014. 65: 20.
https://pubmed.ncbi.nlm.nih.gov/23619390

182 PROSTATE CANCER - LIMITED UPDATE 2021


862. Spratt, D.E., et al. Individual Patient-Level Meta-Analysis of the Performance of the Decipher
Genomic Classifier in High-Risk Men After Prostatectomy to Predict Development of Metastatic
Disease. J Clin Oncol, 2017. 35: 1991.
https://pubmed.ncbi.nlm.nih.gov/28358655
863. Thompson, I.M., et al. Adjuvant radiotherapy for pathological T3N0M0 prostate cancer significantly
reduces risk of metastases and improves survival: long-term followup of a randomized clinical trial.
J Urol, 2009. 181: 956.
https://pubmed.ncbi.nlm.nih.gov/19167731
864. Bolla, M., et al. Postoperative radiotherapy after radical prostatectomy for high-risk prostate cancer:
long-term results of a randomised controlled trial (EORTC trial 22911). Lancet, 2012. 380: 2018.
https://pubmed.ncbi.nlm.nih.gov/23084481
865. Wiegel, T., et al. Adjuvant Radiotherapy Versus Wait-and-See After Radical Prostatectomy: 10-year
Follow-up of the ARO 96-02/AUO AP 09/95 Trial. Eur Urol, 2014. 66: 243.
https://pubmed.ncbi.nlm.nih.gov/24680359
866. Hackman, G., et al. Randomised Trial of Adjuvant Radiotherapy Following Radical Prostatectomy
Versus Radical Prostatectomy Alone in Prostate Cancer Patients with Positive Margins or
Extracapsular Extension. Eur Urol, 2019. 76: 586.
https://pubmed.ncbi.nlm.nih.gov/31375279
867. Fossati, N., et al. Long-term Impact of Adjuvant Versus Early Salvage Radiation Therapy in pT3N0
Prostate Cancer Patients Treated with Radical Prostatectomy: Results from a Multi-institutional
Series. Eur Urol, 2017. 71: 886.
https://pubmed.ncbi.nlm.nih.gov/27484843
868. Buscariollo, D.L., et al. Long-term results of adjuvant versus early salvage postprostatectomy
radiation: A large single-institutional experience. Pract Radiat Oncol, 2017. 7: e125.
https://pubmed.ncbi.nlm.nih.gov/28274403
869. Hwang, W.L., et al. Comparison Between Adjuvant and Early-Salvage Postprostatectomy
Radiotherapy for Prostate Cancer With Adverse Pathological Features. JAMA Oncol, 2018.
4: e175230.
https://pubmed.ncbi.nlm.nih.gov/29372236
870. Parker, C.C., et al. Timing of radiotherapy after radical prostatectomy (RADICALS-RT): a
randomised, controlled phase 3 trial. Lancet, 2020. 396: 1413.
https://pubmed.ncbi.nlm.nih.gov/33002429
871. Kneebone, A., et al. Adjuvant radiotherapy versus early salvage radiotherapy following radical
prostatectomy (TROG 08.03/ANZUP RAVES): a randomised, controlled, phase 3, non-inferiority trial.
Lancet Oncol, 2020. 21: 1331.
https://pubmed.ncbi.nlm.nih.gov/33002437
872. Sargos, P., et al. Adjuvant radiotherapy versus early salvage radiotherapy plus short-term androgen
deprivation therapy in men with localised prostate cancer after radical prostatectomy (GETUG-AFU
17): a randomised, phase 3 trial. Lancet Oncol, 2020. 21: 1341.
https://pubmed.ncbi.nlm.nih.gov/33002438
873. Vale, C.L., et al. Adjuvant or early salvage radiotherapy for the treatment of localised and locally
advanced prostate cancer: a prospectively planned systematic review and meta-analysis of
aggregate data. Lancet, 2020. 396: 1422.
https://pubmed.ncbi.nlm.nih.gov/33002431
874. Tilki, D., et al. Timing of radiotherapy after radical prostatectomy. Lancet, 2020. 396: 1374.
https://pubmed.ncbi.nlm.nih.gov/33002430
875. Iversen, P., et al. Antiandrogen monotherapy in patients with localized or locally advanced prostate
cancer: final results from the bicalutamide Early Prostate Cancer programme at a median follow-up
of 9.7 years. BJU Int, 2010. 105: 1074.
https://pubmed.ncbi.nlm.nih.gov/22129214
876. Ahlgren, G.M., et al. Docetaxel Versus Surveillance After Radical Prostatectomy for High-risk
Prostate Cancer: Results from the Prospective Randomised, Open-label Phase 3 Scandinavian
Prostate Cancer Group 12 Trial. Eur Urol, 2018. 73: 870.
https://pubmed.ncbi.nlm.nih.gov/29395502
877. Schweizer, M.T., et al. Adjuvant leuprolide with or without docetaxel in patients with high-risk
prostate cancer after radical prostatectomy (TAX-3501): important lessons for future trials. Cancer,
2013. 119: 3610.
https://pubmed.ncbi.nlm.nih.gov/23943299
878. Abdollah, F., et al. Impact of adjuvant radiotherapy on survival of patients with node-positive
prostate cancer. J Clin Oncol, 2014. 32: 3939.
https://pubmed.ncbi.nlm.nih.gov/25245445

PROSTATE CANCER - LIMITED UPDATE 2021 183


879. Abdollah, F., et al. Impact of Adjuvant Radiotherapy in Node-positive Prostate Cancer Patients: The
Importance of Patient Selection. Eur Urol, 2018. 74: 253.
https://pubmed.ncbi.nlm.nih.gov/29720348
880. Gupta, M., et al. Adjuvant radiation with androgen-deprivation therapy for men with lymph node
metastases after radical prostatectomy: identifying men who benefit. BJU Int, 2019. 123: 252.
https://pubmed.ncbi.nlm.nih.gov/29626845
881. Jegadeesh, N., et al. The role of adjuvant radiotherapy in pathologically lymph node-positive
prostate cancer. Cancer, 2017. 123: 512.
https://pubmed.ncbi.nlm.nih.gov/27859018
882. Briganti, A., et al. Combination of adjuvant hormonal and radiation therapy significantly prolongs
survival of patients with pT2-4 pN+ prostate cancer: results of a matched analysis. Eur Urol, 2011.
59: 832.
https://pubmed.ncbi.nlm.nih.gov/21354694
883. Moghanaki, D., et al. Elective irradiation of pelvic lymph nodes during postprostatectomy salvage
radiotherapy. Cancer, 2013. 119: 52.
https://pubmed.ncbi.nlm.nih.gov/22736478
884. Tilki, D., et al. Adjuvant radiation therapy is associated with better oncological outcome
compared with salvage radiation therapy in patients with pN1 prostate cancer treated with radical
prostatectomy. BJU Int, 2017. 119: 717.
https://pubmed.ncbi.nlm.nih.gov/27743493
885. Mandel, P., et al. Long-term oncological outcomes in patients with limited nodal disease undergoing
radical prostatectomy and pelvic lymph node dissection without adjuvant treatment. World J Urol,
2017. 35: 1833.
https://pubmed.ncbi.nlm.nih.gov/28828530
886. Marra, G., et al. Management of Patients with Node-positive Prostate Cancer at Radical
Prostatectomy and Pelvic Lymph Node Dissection: A Systematic Review. Eur Urol Oncol, 2020.
3: 565.
https://pubmed.ncbi.nlm.nih.gov/32933887
887. Ploussard, G., et al. Predictive factors of oncologic outcomes in patients who do not achieve
undetectable prostate specific antigen after radical prostatectomy. J Urol, 2013. 190: 1750.
https://pubmed.ncbi.nlm.nih.gov/23643600
888. Wiegel, T., et al. Prostate-specific antigen persistence after radical prostatectomy as a predictive
factor of clinical relapse-free survival and overall survival: 10-year data of the ARO 96-02 trial. Int
J Radiat Oncol Biol Phys, 2015. 91: 288.
https://pubmed.ncbi.nlm.nih.gov/25445556
889. Moreira, D.M., et al. Natural history of persistently elevated prostate specific antigen after radical
prostatectomy: results from the SEARCH database. J Urol, 2009. 182: 2250.
https://pubmed.ncbi.nlm.nih.gov/19758614
890. Moreira, D.M., et al. Definition and preoperative predictors of persistently elevated prostate-specific
antigen after radical prostatectomy: results from the Shared Equal Access Regional Cancer Hospital
(SEARCH) database. BJU Int, 2010. 105: 1541.
https://pubmed.ncbi.nlm.nih.gov/19912191
891. Spratt, D.E., et al. Performance of a Prostate Cancer Genomic Classifier in Predicting Metastasis in
Men with Prostate-specific Antigen Persistence Postprostatectomy. Eur Urol, 2018. 74: 107.
https://pubmed.ncbi.nlm.nih.gov/29233664
892. Preisser, F., et al. Persistent Prostate-Specific Antigen After Radical Prostatectomy and Its Impact
on Oncologic Outcomes. Eur Urol, 2019. 76: 106.
https://pubmed.ncbi.nlm.nih.gov/30772034
893. Xiang, C., et al. Prediction of Biochemical Recurrence Following Radiotherapy among Patients with
Persistent PSA after Radical Prostatectomy: A Single-Center Experience. Urol Int, 2018. 101: 47.
https://pubmed.ncbi.nlm.nih.gov/29627830
894. Rogers, C.G., et al. Natural history of disease progression in patients who fail to achieve an
undetectable prostate-specific antigen level after undergoing radical prostatectomy. Cancer, 2004.
101: 2549.
https://pubmed.ncbi.nlm.nih.gov/15470681
895. Farolfi, A., et al. (68)Ga-PSMA-11 PET/CT in prostate cancer patients with biochemical recurrence
after radical prostatectomy and PSA <0.5 ng/ml. Efficacy and impact on treatment strategy. Eur
J Nucl Med Mol Imaging, 2019. 46: 11.
https://pubmed.ncbi.nlm.nih.gov/29905907

184 PROSTATE CANCER - LIMITED UPDATE 2021


896. Ceci, F., et al. (68)Ga-PSMA-11 PET/CT in recurrent prostate cancer: efficacy in different clinical
stages of PSA failure after radical therapy. Eur J Nucl Med Mol Imaging, 2019. 46: 31.
https://pubmed.ncbi.nlm.nih.gov/25975367
897. Rauscher, I., et al. Efficacy, Predictive Factors, and Prediction Nomograms for (68)Ga-labeled
Prostate-specific Membrane Antigen-ligand Positron-emission Tomography/Computed Tomography
in Early Biochemical Recurrent Prostate Cancer After Radical Prostatectomy. Eur Urol, 2018. 73: 656.
https://pubmed.ncbi.nlm.nih.gov/29358059
898. Wondergem, M., et al. Early lesion detection with (18)F-DCFPyL PET/CT in 248 patients with
biochemically recurrent prostate cancer. Eur J Nucl Med Mol Imaging, 2019. 46: 1911.
https://pubmed.ncbi.nlm.nih.gov/31230088
899. Mena, E., et al. Clinical impact of PSMA-based (18)F-DCFBC PET/CT imaging in patients with
biochemically recurrent prostate cancer after primary local therapy. Eur J Nucl Med Mol Imaging,
2018. 45: 4.
https://pubmed.ncbi.nlm.nih.gov/28894899
900. Habl, G., et al. (68) Ga-PSMA-PET for radiation treatment planning in prostate cancer recurrences
after surgery: Individualized medicine or new standard in salvage treatment. Prostate, 2017. 77: 920.
https://pubmed.ncbi.nlm.nih.gov/28317152
901. Schmidt-Hegemann, N.S., et al. Outcome after PSMA PET/CT based radiotherapy in patients with
biochemical persistence or recurrence after radical prostatectomy. Radiat Oncol, 2018. 13: 37.
https://pubmed.ncbi.nlm.nih.gov/29499730
902. Farolfi, A., et al. (68)Ga-PSMA-11 Positron Emission Tomography Detects Residual Prostate Cancer
after Prostatectomy in a Multicenter Retrospective Study. J Urol, 2019. 202: 1174.
https://pubmed.ncbi.nlm.nih.gov/31233369
903. Meijer, D., et al. Biochemical Persistence of Prostate-specific Antigen after Robot-assisted
Laparoscopic Radical Prostatectomy: Tumor localizations using PSMA PET/CT imaging. J Nucl
Med, 2020.
https://pubmed.ncbi.nlm.nih.gov/33158904
904. Choo, R., et al. Prospective study evaluating postoperative radiotherapy plus 2-year androgen
suppression for post-radical prostatectomy patients with pathologic T3 disease and/or positive
surgical margins. Int J Radiat Oncol Biol Phys, 2009. 75: 407.
https://pubmed.ncbi.nlm.nih.gov/19211197
905. Gandaglia, G., et al. Impact of Postoperative Radiotherapy in Men with Persistently Elevated
Prostate-specific Antigen After Radical Prostatectomy for Prostate Cancer: A Long-term Survival
Analysis. Eur Urol, 2017. 72: 910.
https://pubmed.ncbi.nlm.nih.gov/28622831
906. Garcia-Barreras, S., et al. Predictive factors and the important role of detectable prostate-specific
antigen for detection of clinical recurrence and cancer-specific mortality following robot-assisted
radical prostatectomy. Clin Transl Oncol, 2018. 20: 1004.
https://pubmed.ncbi.nlm.nih.gov/29243074
907. Lohm, G., et al. Salvage radiotherapy in patients with persistently detectable PSA or PSA rising from an
undetectable range after radical prostatectomy gives comparable results. World J Urol, 2013. 31: 423.
https://pubmed.ncbi.nlm.nih.gov/22460203
908. Ploussard, G., et al. Clinical outcomes after salvage radiotherapy without androgen deprivation
therapy in patients with persistently detectable PSA after radical prostatectomy: results from a
national multicentre study. World J Urol, 2014. 32: 1331.
https://pubmed.ncbi.nlm.nih.gov/24270970
909. Fossati, N., et al. Impact of Early Salvage Radiation Therapy in Patients with Persistently Elevated or
Rising Prostate-specific Antigen After Radical Prostatectomy. Eur Urol, 2018. 73: 436.
https://pubmed.ncbi.nlm.nih.gov/28779974
910. Guerif, S.G., et al. The acute toxicity results of the GETUG-AFU 22 study: A multicenter randomized
phase II trial comparing the efficacy of a short hormone therapy in combination with radiotherapy
to radiotherapy alone as a salvage treatment for patients with detectable PSA after radical
prostatectomy. J Clin Oncol, 2017. 35: 16.
https://ascopubs.org/doi/abs/10.1200/JCO.2017.35.6_suppl.16
911. Ploussard, G., et al. Management of Persistently Elevated PSA after Radical Prostatectomy: a
Systematic Review of the Literature. Eur Oncol Oncol, 2021. S2588: 00006.
https://pubmed.ncbi.nlm.nih.gov/33574012
912. Amling, C.L., et al. Defining prostate specific antigen progression after radical prostatectomy: what
is the most appropriate cut point? J Urol, 2001. 165: 1146.
https://pubmed.ncbi.nlm.nih.gov/11257657

PROSTATE CANCER - LIMITED UPDATE 2021 185


913. Toussi, A., et al. Standardizing the Definition of Biochemical Recurrence after Radical
Prostatectomy-What Prostate Specific Antigen Cut Point Best Predicts a Durable Increase and
Subsequent Systemic Progression? J Urol, 2016. 195: 1754.
https://pubmed.ncbi.nlm.nih.gov/26721226
914. Stephenson, A.J., et al. Defining biochemical recurrence of prostate cancer after radical
prostatectomy: a proposal for a standardized definition. J Clin Oncol, 2006. 24: 3973.
https://pubmed.ncbi.nlm.nih.gov/16921049
915. Roach, M., 3rd, et al. Defining biochemical failure following radiotherapy with or without hormonal
therapy in men with clinically localized prostate cancer: recommendations of the RTOG-ASTRO
Phoenix Consensus Conference. Int J Radiat Oncol Biol Phys, 2006. 65: 965.
https://pubmed.ncbi.nlm.nih.gov/16798415
916. Van den Broeck, T., et al. Prognostic Value of Biochemical Recurrence Following Treatment with
Curative Intent for Prostate Cancer: A Systematic Review. Eur Urol, 2019. 75: 967.
https://pubmed.ncbi.nlm.nih.gov/30342843
917. Jackson, W.C., et al. Intermediate Endpoints After Postprostatectomy Radiotherapy: 5-Year Distant
Metastasis to Predict Overall Survival. Eur Urol, 2018. 74: 413.
https://pubmed.ncbi.nlm.nih.gov/29306514
918. Choueiri, T.K., et al. Impact of postoperative prostate-specific antigen disease recurrence and the
use of salvage therapy on the risk of death. Cancer, 2010. 116: 1887.
https://pubmed.ncbi.nlm.nih.gov/20162710
919. Freiberger, C., et al. Long-term prognostic significance of rising PSA levels following radiotherapy
for localized prostate cancer - focus on overall survival. Radiat Oncol, 2017. 12: 98.
https://pubmed.ncbi.nlm.nih.gov/28615058
920. Royce, T.J., et al. Surrogate End Points for All-Cause Mortality in Men With Localized Unfavorable-
Risk Prostate Cancer Treated With Radiation Therapy vs Radiation Therapy Plus Androgen
Deprivation Therapy: A Secondary Analysis of a Randomized Clinical Trial. JAMA Oncol, 2017.
3: 652.
https://pubmed.ncbi.nlm.nih.gov/28097317
921. Tilki, D., et al. External Validation of the European Association of Urology Biochemical Recurrence
Risk Groups to Predict Metastasis and Mortality After Radical Prostatectomy in a European Cohort.
Eur Urol, 2019. 75: 896.
https://pubmed.ncbi.nlm.nih.gov/30955970
922. Zagars, G.K., et al. Kinetics of serum prostate-specific antigen after external beam radiation for
clinically localized prostate cancer. Radiother Oncol, 1997. 44: 213.
https://pubmed.ncbi.nlm.nih.gov/9380819
923. Rouviere, O., et al. Imaging of prostate cancer local recurrences: why and how? Eur Radiol, 2010.
20: 1254.
https://pubmed.ncbi.nlm.nih.gov/19921202
924. Beresford, M.J., et al. A systematic review of the role of imaging before salvage radiotherapy for
post-prostatectomy biochemical recurrence. Clin Oncol (R Coll Radiol), 2010. 22: 46.
https://pubmed.ncbi.nlm.nih.gov/19948393
925. Gomez, P., et al. Radionuclide bone scintigraphy in patients with biochemical recurrence after
radical prostatectomy: when is it indicated? BJU Int, 2004. 94: 299.
https://pubmed.ncbi.nlm.nih.gov/15291855
926. Kane, C.J., et al. Limited value of bone scintigraphy and computed tomography in assessing
biochemical failure after radical prostatectomy. Urology, 2003. 61: 607.
https://pubmed.ncbi.nlm.nih.gov/12639656
927. Evangelista, L., et al. Choline PET or PET/CT and biochemical relapse of prostate cancer: a
systematic review and meta-analysis. Clin Nucl Med, 2013. 38: 305.
https://pubmed.ncbi.nlm.nih.gov/23486334
928. Fanti, S., et al. PET/CT with (11)C-choline for evaluation of prostate cancer patients with
biochemical recurrence: meta-analysis and critical review of available data. Eur J Nucl Med Mol
Imaging, 2016. 43: 55.
https://pubmed.ncbi.nlm.nih.gov/26450693
929. Fuccio, C., et al. Role of 11C-choline PET/CT in the restaging of prostate cancer patients showing a
single lesion on bone scintigraphy. Ann Nucl Med, 2010. 24: 485.
https://pubmed.ncbi.nlm.nih.gov/20544323
930. Fuccio, C., et al. Role of 11C-choline PET/CT in the re-staging of prostate cancer patients with
biochemical relapse and negative results at bone scintigraphy. Eur J Radiol, 2012. 81: e893.
https://pubmed.ncbi.nlm.nih.gov/22621862

186 PROSTATE CANCER - LIMITED UPDATE 2021


931. Treglia, G., et al. Relationship between prostate-specific antigen kinetics and detection rate of
radiolabelled choline PET/CT in restaging prostate cancer patients: a meta-analysis. Clin Chem Lab
Med, 2014. 52: 725.
https://pubmed.ncbi.nlm.nih.gov/24310773
932. Castellucci, P., et al. Early Biochemical Relapse After Radical Prostatectomy: Which Prostate
Cancer Patients May Benefit from a Restaging 11C-Choline PET/CT Scan Before Salvage Radiation
Therapy? J Nucl Med, 2014. 55: 1424.
https://pubmed.ncbi.nlm.nih.gov/24935990
933. Mitchell, C.R., et al. Operational characteristics of (11)c-choline positron emission tomography/
computerized tomography for prostate cancer with biochemical recurrence after initial treatment.
J Urol, 2013. 189: 1308.
https://pubmed.ncbi.nlm.nih.gov/23123372
934. Soyka, J.D., et al. Clinical impact of 18F-choline PET/CT in patients with recurrent prostate cancer.
Eur J Nucl Med Mol Imaging, 2012. 39: 936.
https://pubmed.ncbi.nlm.nih.gov/22415598
935. Ceci, F., et al. Impact of 11C-choline PET/CT on clinical decision making in recurrent prostate
cancer: results from a retrospective two-centre trial. Eur J Nucl Med Mol Imaging, 2014. 41: 2222.
https://pubmed.ncbi.nlm.nih.gov/25182750
936. Beer, A.J., et al. Radionuclide and hybrid imaging of recurrent prostate cancer. Lancet Oncol, 2011.
12: 181.
https://pubmed.ncbi.nlm.nih.gov/20599424
937. Beheshti, M., et al. Detection of bone metastases in patients with prostate cancer by 18F
fluorocholine and 18F fluoride PET-CT: a comparative study. Eur J Nucl Med Mol Imaging, 2008.
35: 1766.
https://pubmed.ncbi.nlm.nih.gov/18465129
938. Nanni, C., et al. (18)F-FACBC (anti1-amino-3-(18)F-fluorocyclobutane-1-carboxylic acid) versus
(11)C-choline PET/CT in prostate cancer relapse: results of a prospective trial. Eur J Nucl Med Mol
Imaging, 2016. 43: 1601.
https://pubmed.ncbi.nlm.nih.gov/26960562
939. Bach-Gansmo, T., et al. Multisite Experience of the Safety, Detection Rate and Diagnostic
Performance of Fluciclovine ((18)F) Positron Emission Tomography/Computerized Tomography
Imaging in the Staging of Biochemically Recurrent Prostate Cancer. J Urol, 2017. 197: 676.
https://pubmed.ncbi.nlm.nih.gov/27746282
940. Chen, B., et al. Comparison of Diagnostic Utility of Fluciclovine PET/CT Versus Pelvic
Multiparametric MRI for Prostate Cancer in the Pelvis in the Setting of Rising PSA After Initial
Treatment. Clin Nucl Med, 2020. 45: 349.
https://pubmed.ncbi.nlm.nih.gov/31977495
941. Morigi, J.J., et al. Prospective Comparison of 18F-Fluoromethylcholine Versus 68Ga-PSMA PET/
CT in Prostate Cancer Patients Who Have Rising PSA After Curative Treatment and Are Being
Considered for Targeted Therapy. J Nucl Med, 2015. 56: 1185.
https://pubmed.ncbi.nlm.nih.gov/26112024
942. Afshar-Oromieh, A., et al. Comparison of PET imaging with a (68)Ga-labelled PSMA ligand and
(18)F-choline-based PET/CT for the diagnosis of recurrent prostate cancer. Eur J Nucl Med Mol
Imaging, 2014. 41: 11.
https://pubmed.ncbi.nlm.nih.gov/24072344
943. Caroli, P., et al. (68)Ga-PSMA PET/CT in patients with recurrent prostate cancer after radical
treatment: prospective results in 314 patients. Eur J Nucl Med Mol Imaging, 2018. 45: 2035.
https://pubmed.ncbi.nlm.nih.gov/29922948
944. Fendler, W.P., et al. Assessment of 68Ga-PSMA-11 PET Accuracy in Localizing Recurrent Prostate
Cancer: A Prospective Single-Arm Clinical Trial. JAMA Oncol, 2019. 5: 856.
https://pubmed.ncbi.nlm.nih.gov/30920593
945. Emmett, L., et al. Treatment Outcomes from (68)Ga-PSMA PET/CT-Informed Salvage Radiation
Treatment in Men with Rising PSA After Radical Prostatectomy: Prognostic Value of a Negative
PSMA PET. J Nucl Med, 2017. 58: 1972.
https://pubmed.ncbi.nlm.nih.gov/28747524
946. Bluemel, C., et al. 68Ga-PSMA-PET/CT in Patients With Biochemical Prostate Cancer Recurrence
and Negative 18F-Choline-PET/CT. Clin Nucl Med, 2016. 41: 515.
https://pubmed.ncbi.nlm.nih.gov/26975008

PROSTATE CANCER - LIMITED UPDATE 2021 187


947. Giesel, F.L., et al. Intraindividual Comparison of (18)F-PSMA-1007 and (18)F-DCFPyL PET/CT in the
Prospective Evaluation of Patients with Newly Diagnosed Prostate Carcinoma: A Pilot Study. J Nucl
Med, 2018. 59: 1076.
https://pubmed.ncbi.nlm.nih.gov/29269569
948. Eiber, M., et al. Whole-body MRI including diffusion-weighted imaging (DWI) for patients with
recurring prostate cancer: technical feasibility and assessment of lesion conspicuity in DWI. J Magn
Reson Imaging, 2011. 33: 1160.
https://pubmed.ncbi.nlm.nih.gov/21509875
949. Liauw, S.L., et al. Evaluation of the prostate bed for local recurrence after radical prostatectomy
using endorectal magnetic resonance imaging. Int J Radiat Oncol Biol Phys, 2013. 85: 378.
https://pubmed.ncbi.nlm.nih.gov/22717242
950. Linder, B.J., et al. Early localization of recurrent prostate cancer after prostatectomy by endorectal
coil magnetic resonance imaging. Can J Urol, 2014. 21: 7283.
https://pubmed.ncbi.nlm.nih.gov/24978358
951. Kitajima, K., et al. Detection of recurrent prostate cancer after radical prostatectomy: comparison
of 11C-choline PET/CT with pelvic multiparametric MR imaging with endorectal coil. J Nucl Med,
2014. 55: 223.
https://pubmed.ncbi.nlm.nih.gov/24434294
952. Achard, V., et al. Recurrent prostate cancer after radical prostatectomy: restaging performance of
18F-choline hybrid PET/MRI. Med Oncol, 2019. 36: 67.
https://pubmed.ncbi.nlm.nih.gov/31190232
953. Luiting, H.B., et al. Use of gallium-68 prostate-specific membrane antigen positron-emission
tomography for detecting lymph node metastases in primary and recurrent prostate cancer and
location of recurrence after radical prostatectomy: an overview of the current literature. BJU Int,
2020. 125: 206.
https://pubmed.ncbi.nlm.nih.gov/31680398
954. Boreta, L., et al. Location of Recurrence by Gallium-68 PSMA-11 PET Scan in Prostate Cancer
Patients Eligible for Salvage Radiotherapy. Urology, 2019. 129: 165.
https://pubmed.ncbi.nlm.nih.gov/30928607
955. Guberina, N., et al. Whole-Body Integrated [(68)Ga]PSMA-11-PET/MR Imaging in Patients with
Recurrent Prostate Cancer: Comparison with Whole-Body PET/CT as the Standard of Reference.
Mol Imaging Biol, 2020. 22: 788.
https://pubmed.ncbi.nlm.nih.gov/31482413
956. Donati, O.F., et al. Multiparametric prostate MR imaging with T2-weighted, diffusion-weighted,
and dynamic contrast-enhanced sequences: are all pulse sequences necessary to detect locally
recurrent prostate cancer after radiation therapy? Radiology, 2013. 268: 440.
https://pubmed.ncbi.nlm.nih.gov/23481164
957. Abd-Alazeez, M., et al. Multiparametric MRI for detection of radiorecurrent prostate cancer: added
value of apparent diffusion coefficient maps and dynamic contrast-enhanced images. Prostate
Cancer Prostatic Dis, 2015. 18: 128.
https://pubmed.ncbi.nlm.nih.gov/25644248
958. Alonzo, F., et al. Detection of locally radio-recurrent prostate cancer at multiparametric MRI: Can
dynamic contrast-enhanced imaging be omitted? Diagn Interv Imaging, 2016. 97: 433.
https://pubmed.ncbi.nlm.nih.gov/26928245
959. Dinis Fernandes, C., et al. Quantitative 3T multiparametric MRI of benign and malignant prostatic
tissue in patients with and without local recurrent prostate cancer after external-beam radiation
therapy. J Magn Reson Imaging, 2019. 50: 269.
https://pubmed.ncbi.nlm.nih.gov/30585368
960. Dinis Fernandes, C., et al. Quantitative 3-T multi-parametric MRI and step-section pathology of
recurrent prostate cancer patients after radiation therapy. Eur Radiol, 2019. 29: 4160.
https://pubmed.ncbi.nlm.nih.gov/30421016
961. Ceci, F., et al. 11C-choline PET/CT detects the site of relapse in the majority of prostate cancer
patients showing biochemical recurrence after EBRT. Eur J Nucl Med Mol Imaging, 2014. 41: 878.
https://pubmed.ncbi.nlm.nih.gov/24346416
962. Boorjian, S.A., et al. Radiation therapy after radical prostatectomy: impact on metastasis and
survival. J Urol, 2009. 182: 2708.
https://pubmed.ncbi.nlm.nih.gov/19836762
963. Kneebone, A., et al. A Phase III Multi-Centre Randomised Trial comparing adjuvant versus early
salvage Radiotherapy following a Radical Prostatectomy: Results of the TROG 08.03 and ANZUP
“RAVES” Trial. Int J Radiat Oncol Biol Phys, 2019. 105: S37.
https://www.redjournal.org/article/S0360-3016(19)31291-X/fulltext

188 PROSTATE CANCER - LIMITED UPDATE 2021


964. Stish, B.J., et al. Improved Metastasis-Free and Survival Outcomes With Early Salvage
Radiotherapy in Men With Detectable Prostate-Specific Antigen After Prostatectomy for Prostate
Cancer. J Clin Oncol, 2016. 34: 3864.
https://pubmed.ncbi.nlm.nih.gov/27480153
965. Pfister, D., et al. Early salvage radiotherapy following radical prostatectomy. Eur Urol, 2014. 65: 1034.
https://pubmed.ncbi.nlm.nih.gov/23972524
966. Siegmann, A., et al. Salvage radiotherapy after prostatectomy - what is the best time to treat?
Radiother Oncol, 2012. 103: 239.
https://pubmed.ncbi.nlm.nih.gov/22119375
967. Ohri, N., et al. Can early implementation of salvage radiotherapy for prostate cancer improve the
therapeutic ratio? A systematic review and regression meta-analysis with radiobiological modelling.
Eur J Cancer, 2012. 48: 837.
https://pubmed.ncbi.nlm.nih.gov/21945099
968. Wiegel, T., et al. Achieving an undetectable PSA after radiotherapy for biochemical progression
after radical prostatectomy is an independent predictor of biochemical outcome--results of a
retrospective study. Int J Radiat Oncol Biol Phys, 2009. 73: 1009.
https://pubmed.ncbi.nlm.nih.gov/18963539
969. Trock, B.J., et al. Prostate cancer-specific survival following salvage radiotherapy vs observation in
men with biochemical recurrence after radical prostatectomy. Jama, 2008. 299: 2760.
https://pubmed.ncbi.nlm.nih.gov/18560003
970. Boorjian, S.A., et al. Long-term risk of clinical progression after biochemical recurrence following
radical prostatectomy: the impact of time from surgery to recurrence. Eur Urol, 2011. 59: 893.
https://pubmed.ncbi.nlm.nih.gov/21388736
971. Sweeney, C., et al. The Development of Intermediate Clinical Endpoints in Cancer of the Prostate
(ICECaP). J Natl Cancer Inst, 2015. 107: djv261.
https://pubmed.ncbi.nlm.nih.gov/26409187
972. Xie, W., et al. Metastasis-Free Survival Is a Strong Surrogate of Overall Survival in Localized Prostate
Cancer. J Clin Oncol, 2017. 35: 3097.
https://pubmed.ncbi.nlm.nih.gov/28796587
973. Bartkowiak, D., et al. Prostate-specific antigen after salvage radiotherapy for postprostatectomy
biochemical recurrence predicts long-term outcome including overall survival. Acta Oncol, 2018.
57: 362.
https://pubmed.ncbi.nlm.nih.gov/28816074
974. Soto, D.E., et al. Concurrent androgen deprivation therapy during salvage prostate radiotherapy
improves treatment outcomes in high-risk patients. Int J Radiat Oncol Biol Phys, 2012. 82: 1227.
https://pubmed.ncbi.nlm.nih.gov/21549519
975. Tendulkar, R.D., et al. Contemporary Update of a Multi-Institutional Predictive Nomogram for
Salvage Radiotherapy After Radical Prostatectomy. J Clin Oncol, 2016. 34: 3648.
https://pubmed.ncbi.nlm.nih.gov/27528718
976. Jackson, W.C., et al. Combining prostate-specific antigen nadir and time to nadir allows for early
identification of patients at highest risk for development of metastasis and death following salvage
radiation therapy. Pract Radiat Oncol, 2014. 4: 99.
https://pubmed.ncbi.nlm.nih.gov/24890350
977. Shipley, W., et al. Radiation with or without Antiandrogen Therapy in Recurrent Prostate Cancer.
N Eng J Med, 2017. 376: 417.
https://pubmed.ncbi.nlm.nih.gov/28146658
978. Carrie, C., et al. Short-term androgen deprivation therapy combined with radiotherapy as salvage
treatment after radical prostatectomy for prostate cancer (GETUG-AFU 16): a 112-month follow-up
of a phase 3, randomised trial. Lancet Oncol, 2019. 20: 1740.
https://pubmed.ncbi.nlm.nih.gov/27160475
979. Dess, R.T., et al. Association of Presalvage Radiotherapy PSA Levels After Prostatectomy With
Outcomes of Long-term Antiandrogen Therapy in Men With Prostate Cancer. JAMA Oncol, 2020.
6: 735.
https://pubmed.ncbi.nlm.nih.gov/32215583
980. Spratt, D.E., et al. A Systematic Review and Framework for the Use of Hormone Therapy with
Salvage Radiation Therapy for Recurrent Prostate Cancer. Eur Urol, 2018. 73: 156.
https://pubmed.ncbi.nlm.nih.gov/28716370
981. Gandaglia, G., et al. Use of Concomitant Androgen Deprivation Therapy in Patients Treated with
Early Salvage Radiotherapy for Biochemical Recurrence After Radical Prostatectomy: Long-term
Results from a Large, Multi-institutional Series. Eur Urol, 2018. 73: 512.
https://pubmed.ncbi.nlm.nih.gov/29229176

PROSTATE CANCER - LIMITED UPDATE 2021 189


982. Fossati, N., et al. Assessing the Role and Optimal Duration of Hormonal Treatment in Association
with Salvage Radiation Therapy After Radical Prostatectomy: Results from a Multi-Institutional
Study. Eur Urol, 2019. 76: 443.
https://pubmed.ncbi.nlm.nih.gov/30799187
983. Michalski, J.M., et al. Development of RTOG consensus guidelines for the definition of the clinical
target volume for postoperative conformal radiation therapy for prostate cancer. Int J Radiat Oncol
Biol Phys, 2010. 76: 361.
https://pubmed.ncbi.nlm.nih.gov/19394158
984. Poortmans, P., et al. Guidelines for target volume definition in post-operative radiotherapy for
prostate cancer, on behalf of the EORTC Radiation Oncology Group. Radiother Oncol, 2007. 84: 121.
https://pubmed.ncbi.nlm.nih.gov/17706307
985. Wiltshire, K.L., et al. Anatomic boundaries of the clinical target volume (prostate bed) after radical
prostatectomy. Int J Radiat Oncol Biol Phys, 2007. 69: 1090.
https://pubmed.ncbi.nlm.nih.gov/17967303
986. Sassowsky, M., et al. Use of EORTC target definition guidelines for dose-intensified salvage
radiation therapy for recurrent prostate cancer: results of the quality assurance program of the
randomized trial SAKK 09/10. Int J Radiat Oncol Biol Phys, 2013. 87: 534.
https://pubmed.ncbi.nlm.nih.gov/23972722
987. Gay, H.A., et al. Pelvic normal tissue contouring guidelines for radiation therapy: a Radiation
Therapy Oncology Group consensus panel atlas. Int J Radiat Oncol Biol Phys, 2012. 83: e353.
https://pubmed.ncbi.nlm.nih.gov/22483697
988. Ramey, S.J., et al. Multi-institutional Evaluation of Elective Nodal Irradiation and/or Androgen
Deprivation Therapy with Postprostatectomy Salvage Radiotherapy for Prostate Cancer. Eur Urol,
2018. 74: 99.
https://pubmed.ncbi.nlm.nih.gov/29128208
989. Pisansky, T.M., et al. Salvage Radiation Therapy Dose Response for Biochemical Failure of Prostate
Cancer After Prostatectomy-A Multi-Institutional Observational Study. Int J Radiat Oncol Biol Phys,
2016. 96: 1046.
https://pubmed.ncbi.nlm.nih.gov/27745980
990. King, C.R. The dose-response of salvage radiotherapy following radical prostatectomy: A
systematic review and meta-analysis. Radiother Oncol, 2016. 121: 199.
https://pubmed.ncbi.nlm.nih.gov/27863963
991. Fossati, N., et al. Assessing the Optimal Timing for Early Salvage Radiation Therapy in Patients with
Prostate-specific Antigen Rise After Radical Prostatectomy. Eur Urol, 2016. 69: 728.
https://pubmed.ncbi.nlm.nih.gov/26497924
992. Abugharib, A., et al. Very Early Salvage Radiotherapy Improves Distant Metastasis-Free Survival.
J Urol, 2017. 197: 662.
https://pubmed.ncbi.nlm.nih.gov/27614333
993. Fiorino, C., et al. Predicting the 5-Year Risk of Biochemical Relapse After Postprostatectomy
Radiation Therapy in >/=PT2, pN0 Patients With a Comprehensive Tumor Control Probability Model.
Int J Radiat Oncol Biol Phys, 2016. 96: 333.
https://pubmed.ncbi.nlm.nih.gov/27497691
994. Ghadjar, P., et al. Acute Toxicity and Quality of Life After Dose-Intensified Salvage Radiation Therapy
for Biochemically Recurrent Prostate Cancer After Prostatectomy: First Results of the Randomized
Trial SAKK 09/10. J Clin Oncol, 2015. 33: 4158.
https://pubmed.ncbi.nlm.nih.gov/26527774
995. Ghadjar, P., et al. Impact of dose intensified salvage radiation therapy on urinary continence
recovery after radical prostatectomy: Results of the randomized trial SAKK 09/10. Radiother Oncol,
2018. 126: 257.
https://pubmed.ncbi.nlm.nih.gov/29103826
996. Goenka, A., et al. Improved toxicity profile following high-dose postprostatectomy salvage radiation
therapy with intensity-modulated radiation therapy. Eur Urol, 2011. 60: 1142.
https://pubmed.ncbi.nlm.nih.gov/21855208
997. Ost, P., et al. High-dose salvage intensity-modulated radiotherapy with or without androgen
deprivation after radical prostatectomy for rising or persisting prostate-specific antigen: 5-year
results. Eur Urol, 2011. 60: 842.
https://pubmed.ncbi.nlm.nih.gov/21514039
998. Steuber, T., et al. Standard of Care Versus Metastases-directed Therapy for PET-detected Nodal
Oligorecurrent Prostate Cancer Following Multimodality Treatment: A Multi-institutional Case-control
Study. Eur Urol Focus, 2019. 5: 1007.
https://pubmed.ncbi.nlm.nih.gov/29530632

190 PROSTATE CANCER - LIMITED UPDATE 2021


999. De Bleser, E., et al. Metastasis-directed Therapy in Treating Nodal Oligorecurrent Prostate Cancer: A
Multi-institutional Analysis Comparing the Outcome and Toxicity of Stereotactic Body Radiotherapy
and Elective Nodal Radiotherapy. Eur Urol, 2019. 76: 732.
https://pubmed.ncbi.nlm.nih.gov/31331782
1000. Suardi, N., et al. Long-term outcomes of salvage lymph node dissection for clinically recurrent
prostate cancer: results of a single-institution series with a minimum follow-up of 5 years. Eur Urol,
2015. 67: 299.
https://pubmed.ncbi.nlm.nih.gov/24571959
1001. Tilki, D., et al. Salvage lymph node dissection for nodal recurrence of prostate cancer after radical
prostatectomy. J Urol, 2015. 193: 484.
https://pubmed.ncbi.nlm.nih.gov/25180792
1002. Fossati, N., et al. Identifying the Optimal Candidate for Salvage Lymph Node Dissection for Nodal
Recurrence of Prostate Cancer: Results from a Large, Multi-institutional Analysis. Eur Urol, 2019.
75: 176.
https://pubmed.ncbi.nlm.nih.gov/30301694
1003. Ploussard, G., et al. Salvage Lymph Node Dissection for Nodal Recurrent Prostate Cancer: A
Systematic Review. Eur Urol, 2019. 76: 493.
https://pubmed.ncbi.nlm.nih.gov/30391078
1004. Ost, P., et al. Metastasis-directed therapy of regional and distant recurrences after curative
treatment of prostate cancer: a systematic review of the literature. Eur Urol, 2015. 67: 852.
https://pubmed.ncbi.nlm.nih.gov/25240974
1005. Rischke, H.C., et al. Adjuvant radiotherapy after salvage lymph node dissection because of nodal
relapse of prostate cancer versus salvage lymph node dissection only. Strahlenther Onkol, 2015.
191: 310.
https://pubmed.ncbi.nlm.nih.gov/25326142
1006. Bravi, C.A., et al. Long-term Outcomes of Salvage Lymph Node Dissection for Nodal Recurrence of
Prostate Cancer After Radical Prostatectomy: Not as Good as Previously Thought. Eur Urol, 2020.
78: 661.
https://pubmed.ncbi.nlm.nih.gov/32624288
1007. Valle, L.F., et al. A Systematic Review and Meta-analysis of Local Salvage Therapies After
Radiotherapy for Prostate Cancer (MASTER). Eur Urol, 2020.
https://pubmed.ncbi.nlm.nih.gov/33309278
1008. Gontero, P., et al. Salvage Radical Prostatectomy for Recurrent Prostate Cancer: Morbidity and
Functional Outcomes from a Large Multicenter Series of Open versus Robotic Approaches. J Urol,
2019. 202: 725.
https://pubmed.ncbi.nlm.nih.gov/31075058
1009. Chade, D.C., et al. Cancer control and functional outcomes of salvage radical prostatectomy for
radiation-recurrent prostate cancer: a systematic review of the literature. Eur Urol, 2012. 61: 961.
https://pubmed.ncbi.nlm.nih.gov/22280856
1010. Marra, G., et al. Oncological outcomes of salvage radical prostatectomy for recurrent prostate
cancer in the contemporary era: A multicenter retrospective study. Urol Oncol, 2021.
https://pubmed.ncbi.nlm.nih.gov/33436329
1011. Chade, D.C., et al. Salvage radical prostatectomy for radiation-recurrent prostate cancer: a multi-
institutional collaboration. Eur Urol, 2011. 60: 205.
https://pubmed.ncbi.nlm.nih.gov/21420229
1012. Mandel, P., et al. Salvage radical prostatectomy for recurrent prostate cancer: verification of
European Association of Urology guideline criteria. BJU Int, 2016. 117: 55.
https://pubmed.ncbi.nlm.nih.gov/25711672
1013. Ogaya-Pinies, G., et al. Salvage robotic-assisted radical prostatectomy: oncologic and functional
outcomes from two high-volume institutions. World J Urol, 2019. 37: 1499.
https://pubmed.ncbi.nlm.nih.gov/30006908
1014. Gotto, G.T., et al. Impact of prior prostate radiation on complications after radical prostatectomy.
J Urol, 2010. 184: 136.
https://pubmed.ncbi.nlm.nih.gov/20478594
1015. Ward, J.F., et al. Salvage surgery for radiorecurrent prostate cancer: contemporary outcomes.
J Urol, 2005. 173: 1156.
https://pubmed.ncbi.nlm.nih.gov/15758726
1016. Sanderson, K.M., et al. Salvage radical prostatectomy: quality of life outcomes and long-term
oncological control of radiorecurrent prostate cancer. J Urol, 2006. 176: 2025.
https://pubmed.ncbi.nlm.nih.gov/17070244

PROSTATE CANCER - LIMITED UPDATE 2021 191


1017. Ginsburg, K.B., et al. Avoidance of androgen deprivation therapy in radiorecurrent prostate cancer
as a clinically meaningful endpoint for salvage cryoablation. Prostate, 2017. 77: 1446.
https://pubmed.ncbi.nlm.nih.gov/28856702
1018. Spiess, P.E., et al. A pretreatment nomogram predicting biochemical failure after salvage
cryotherapy for locally recurrent prostate cancer. BJU Int, 2010. 106: 194.
https://pubmed.ncbi.nlm.nih.gov/19922545
1019. Li, R., et al. The Effect of Androgen Deprivation Therapy Before Salvage Whole-gland Cryoablation
After Primary Radiation Failure in Prostate Cancer Treatment. Urology, 2015. 85: 1137.
https://pubmed.ncbi.nlm.nih.gov/25799176
1020. Kovac, E., et al. Five-Year Biochemical Progression-Free Survival Following Salvage Whole-Gland
Prostate Cryoablation: Defining Success with Nadir Prostate-Specific Antigen. J Endourol, 2016.
30: 624.
https://pubmed.ncbi.nlm.nih.gov/26915721
1021. Ahmad, I., et al. Prostate gland lengths and iceball dimensions predict micturition functional
outcome following salvage prostate cryotherapy in men with radiation recurrent prostate cancer.
PLoS One, 2013. 8: e69243.
https://pubmed.ncbi.nlm.nih.gov/23950886
1022. Pisters, L.L., et al. Salvage prostate cryoablation: initial results from the cryo on-line data registry.
J Urol, 2008. 180: 559.
https://pubmed.ncbi.nlm.nih.gov/19524984
1023. Cespedes, R.D., et al. Long-term followup of incontinence and obstruction after salvage
cryosurgical ablation of the prostate: results in 143 patients. J Urol, 1997. 157: 237.
https://pubmed.ncbi.nlm.nih.gov/8976261
1024. Chin, J.L., et al. Results of salvage cryoablation of the prostate after radiation: identifying predictors
of treatment failure and complications. J Urol, 2001. 165: 1937.
https://pubmed.ncbi.nlm.nih.gov/11371885
1025. Henríquez López, I., et al. Salvage brachytherapy for locally-recurrent prostate cancer after radiation
therapy: A comparison of efficacy and toxicity outcomes with high-dose rate and low-dose rate
brachytherapy. Radiother Oncol, 2019. 141: 156.
https://pubmed.ncbi.nlm.nih.gov/31570236
1026. Crook, J.M., et al. A Prospective Phase 2 Trial of Transperineal Ultrasound-Guided Brachytherapy
for Locally Recurrent Prostate Cancer After External Beam Radiation Therapy (NRG Oncology/
RTOG-0526). Int J Radiat Oncol Biol Phys, 2019. 103: 335.
https://pubmed.ncbi.nlm.nih.gov/30312717
1027. Smith, W.H., et al. Salvage low dose rate brachytherapy for prostate cancer recurrence following
definitive external beam radiation therapy. Radiother Oncol, 2020. 155: 42.
https://pubmed.ncbi.nlm.nih.gov/33075391
1028. Łyczek, J., et al. HDR brachytherapy as a solution in recurrences of locally advanced prostate
cancer. J Contemp Brachytherapy, 2009. 1: 105.
https://pubmed.ncbi.nlm.nih.gov/27795720
1029. Pasquier, D., et al. Salvage Stereotactic Body Radiation Therapy for Local Prostate Cancer
Recurrence After Radiation Therapy: A Retrospective Multicenter Study of the GETUG. Int J Radiat
Oncol Biol Phys, 2019. 105: 727.
https://pubmed.ncbi.nlm.nih.gov/31344433
1030. Fuller, D., et al. Retreatment for Local Recurrence of Prostatic Carcinoma After Prior Therapeutic
Irradiation: Efficacy and Toxicity of HDR-Like SBRT. Int J Radiat Oncol Biol Phys, 2020. 106: 291.
https://pubmed.ncbi.nlm.nih.gov/31629838
1031. Crouzet, S., et al. Salvage high-intensity focused ultrasound (HIFU) for locally recurrent prostate cancer
after failed radiation therapy: Multi-institutional analysis of 418 patients. BJU Int, 2017. 119: 896.
https://pubmed.ncbi.nlm.nih.gov/28063191
1032. Murat, F.J., et al. Mid-term results demonstrate salvage high-intensity focused ultrasound (HIFU)
as an effective and acceptably morbid salvage treatment option for locally radiorecurrent prostate
cancer. Eur Urol, 2009. 55: 640.
https://pubmed.ncbi.nlm.nih.gov/18508188
1033. Kanthabalan, A., et al. Focal salvage high-intensity focused ultrasound in radiorecurrent prostate
cancer. BJU Int, 2017. 120: 246.
https://pubmed.ncbi.nlm.nih.gov/28258616
1034. Jones, T.A., et al. High Intensity Focused Ultrasound for Radiorecurrent Prostate Cancer: A North
American Clinical Trial. J Urol, 2018. 199: 133.
https://pubmed.ncbi.nlm.nih.gov/28652121

192 PROSTATE CANCER - LIMITED UPDATE 2021


1035. van den Bergh, R.C., et al. Role of Hormonal Treatment in Prostate Cancer Patients with
Nonmetastatic Disease Recurrence After Local Curative Treatment: A Systematic Review. Eur Urol,
2016. 69: 802.
https://pubmed.ncbi.nlm.nih.gov/26691493
1036. Duchesne, G.M., et al. Timing of androgen-deprivation therapy in patients with prostate cancer with
a rising PSA (TROG 03.06 and VCOG PR 01-03 [TOAD]): a randomised, multicentre, non-blinded,
phase 3 trial. Lancet Oncol, 2016. 17: 727.
https://pubmed.ncbi.nlm.nih.gov/27155740
1037. Siddiqui, S.A., et al. Timing of androgen deprivation therapy and its impact on survival after radical
prostatectomy: a matched cohort study. J Urol, 2008. 179: 1830.
https://pubmed.ncbi.nlm.nih.gov/18353378
1038. Crook, J.M., et al. Intermittent androgen suppression for rising PSA level after radiotherapy. N Engl
J Med, 2012. 367: 895.
https://pubmed.ncbi.nlm.nih.gov/22931259
1039. Levine, G.N., et al. Androgen-deprivation therapy in prostate cancer and cardiovascular risk: a
science advisory from the American Heart Association, American Cancer Society, and American
Urological Association: endorsed by the American Society for Radiation Oncology. Circulation,
2010. 121: 833.
https://pubmed.ncbi.nlm.nih.gov/20124128
1040. O’Farrell, S., et al. Risk and Timing of Cardiovascular Disease After Androgen-Deprivation Therapy
in Men With Prostate Cancer. J Clin Oncol, 2015. 33: 1243.
https://pubmed.ncbi.nlm.nih.gov/25732167
1041. James, N.D., et al. Survival with Newly Diagnosed Metastatic Prostate Cancer in the “Docetaxel
Era”: Data from 917 Patients in the Control Arm of the STAMPEDE Trial (MRC PR08, CRUK/06/019).
Eur Urol, 2015. 67: 1028.
https://pubmed.ncbi.nlm.nih.gov/25301760
1042. Glass, T.R., et al. Metastatic carcinoma of the prostate: identifying prognostic groups using
recursive partitioning. J Urol, 2003. 169: 164.
https://pubmed.ncbi.nlm.nih.gov/12478127
1043. Gravis, G., et al. Prognostic Factors for Survival in Noncastrate Metastatic Prostate Cancer:
Validation of the Glass Model and Development of a Novel Simplified Prognostic Model. Eur Urol,
2015. 68: 196.
https://pubmed.ncbi.nlm.nih.gov/25277272
1044. Gravis, G., et al. Androgen Deprivation Therapy (ADT) Plus Docetaxel Versus ADT Alone in
Metastatic Non castrate Prostate Cancer: Impact of Metastatic Burden and Long-term Survival
Analysis of the Randomized Phase 3 GETUG-AFU15 Trial. Eur Urol, 2016. 70: 256.
https://pubmed.ncbi.nlm.nih.gov/26610858
1045. Sweeney, C.J., et al. Chemohormonal Therapy in Metastatic Hormone-Sensitive Prostate Cancer. N
Engl J Med, 2015. 373: 737.
https://pubmed.ncbi.nlm.nih.gov/26244877
1046. Kyriakopoulos, C.E., et al. Chemohormonal Therapy in Metastatic Hormone-Sensitive Prostate
Cancer: Long-Term Survival Analysis of the Randomized Phase III E3805 CHAARTED Trial. J Clin
Oncol, 2018. 36: 1080.
https://pubmed.ncbi.nlm.nih.gov/29384722
1047. Gravis, G., et al. Burden of Metastatic Castrate Naive Prostate Cancer Patients, to Identify Men
More Likely to Benefit from Early Docetaxel: Further Analyses of CHAARTED and GETUG-AFU15
Studies. Eur Urol, 2018. 73: 847.
https://pubmed.ncbi.nlm.nih.gov/29475737
1048. Parker, C.C., et al. Radiotherapy to the primary tumour for newly diagnosed, metastatic prostate
cancer (STAMPEDE): a randomised controlled phase 3 trial. Lancet, 2018. 392: 2353.
https://pubmed.ncbi.nlm.nih.gov/30355464
1049. Hussain, M., et al. Absolute prostate-specific antigen value after androgen deprivation is a strong
independent predictor of survival in new metastatic prostate cancer: data from Southwest Oncology
Group Trial 9346 (INT-0162). J Clin Oncol, 2006. 24: 3984.
https://pubmed.ncbi.nlm.nih.gov/16921051
1050. Harshman, L.C., et al. Seven-Month Prostate-Specific Antigen Is Prognostic in Metastatic Hormone-
Sensitive Prostate Cancer Treated With Androgen Deprivation With or Without Docetaxel. J Clin
Oncol, 2018. 36: 376.
https://pubmed.ncbi.nlm.nih.gov/29261442

PROSTATE CANCER - LIMITED UPDATE 2021 193


1051. Kunath, F., et al. Non-steroidal antiandrogen monotherapy compared with luteinising hormone-
releasing hormone agonists or surgical castration monotherapy for advanced prostate cancer.
Cochrane Database Syst Rev, 2014. 6: CD009266.
https://pubmed.ncbi.nlm.nih.gov/24979481
1052. Niraula, S., et al. Treatment of prostate cancer with intermittent versus continuous androgen
deprivation: a systematic review of randomized trials. J Clin Oncol, 2013. 31: 2029.
https://pubmed.ncbi.nlm.nih.gov/23630216
1053. Botrel, T.E., et al. Intermittent versus continuous androgen deprivation for locally advanced,
recurrent or metastatic prostate cancer: a systematic review and meta-analysis. BMC Urol, 2014.
14: 9.
https://pubmed.ncbi.nlm.nih.gov/24460605
1054. Tsai, H.T., et al. Efficacy of intermittent androgen deprivation therapy vs conventional continuous
androgen deprivation therapy for advanced prostate cancer: a meta-analysis. Urology, 2013. 82: 327.
https://pubmed.ncbi.nlm.nih.gov/23896094
1055. Brungs, D., et al. Intermittent androgen deprivation is a rational standard-of-care treatment for all
stages of progressive prostate cancer: results from a systematic review and meta-analysis. Prostate
Cancer Prostatic Dis, 2014. 17: 105.
https://pubmed.ncbi.nlm.nih.gov/24686773
1056. Magnan, S., et al. Intermittent vs Continuous Androgen Deprivation Therapy for Prostate Cancer: A
Systematic Review and Meta-analysis. JAMA Oncol, 2015. 1: 1261.
https://pubmed.ncbi.nlm.nih.gov/26378418
1057. Hussain, M., et al. Intermittent versus continuous androgen deprivation in prostate cancer. N Engl
J Med, 2013. 368: 1314.
https://pubmed.ncbi.nlm.nih.gov/23550669
1058. Hussain, M., et al. Evaluating Intermittent Androgen-Deprivation Therapy Phase III Clinical Trials: The
Devil Is in the Details. J Clin Oncol, 2016. 34: 280.
https://pubmed.ncbi.nlm.nih.gov/26552421
1059. Verhagen, P.C., et al. Intermittent versus continuous cyproterone acetate in bone metastatic prostate
cancer: results of a randomized trial. World J Urol, 2014. 32: 1287.
https://pubmed.ncbi.nlm.nih.gov/24258313
1060. Calais da Silva, F., et al. Locally advanced and metastatic prostate cancer treated with intermittent
androgen monotherapy or maximal androgen blockade: results from a randomised phase 3 study by
the South European Uroncological Group. Eur Urol, 2014. 66: 232.
https://pubmed.ncbi.nlm.nih.gov/23582949
1061. Nair, B., et al. Early versus deferred androgen suppression in the treatment of advanced prostatic
cancer. Cochrane Database Syst Rev, 2002: Cd003506.
https://pubmed.ncbi.nlm.nih.gov/11869665
1062. Kunath, F., et al. Early versus deferred standard androgen suppression therapy for advanced
hormone-sensitive prostate cancer. Cochrane Database Syst Rev, 2019. 6: CD003506.
https://pubmed.ncbi.nlm.nih.gov/31194882
1063. Eisenberger, M.A., et al. Bilateral orchiectomy with or without flutamide for metastatic prostate
cancer. N Engl J Med, 1998. 339: 1036.
https://pubmed.ncbi.nlm.nih.gov/9761805
1064. Maximum androgen blockade in advanced prostate cancer: an overview of the randomised trials.
Prostate Cancer Trialists’ Collaborative Group. Lancet, 2000. 355: 1491.
https://pubmed.ncbi.nlm.nih.gov/10801170
1065. Schmitt, B., et al. Maximal androgen blockade for advanced prostate cancer. Cochrane Database
Syst Rev, 2000: Cd001526.
https://pubmed.ncbi.nlm.nih.gov/10796804
1066. Akaza, H., et al. Combined androgen blockade with bicalutamide for advanced prostate cancer: long-
term follow-up of a phase 3, double-blind, randomized study for survival. Cancer, 2009. 115: 3437.
https://pubmed.ncbi.nlm.nih.gov/19536889
1067. Gravis, G., et al. Androgen-deprivation therapy alone or with docetaxel in non-castrate metastatic
prostate cancer (GETUG-AFU 15): a randomised, open-label, phase 3 trial. Lancet Oncol, 2013.
14: 149.
https://pubmed.ncbi.nlm.nih.gov/23306100
1068. Clarke, N.W., et al. Addition of docetaxel to hormonal therapy in low- and high-burden metastatic
hormone sensitive prostate cancer: long-term survival results from the STAMPEDE trial. Ann Oncol,
2019. 30: 1992.
https://pubmed.ncbi.nlm.nih.gov/31560068

194 PROSTATE CANCER - LIMITED UPDATE 2021


1069. Smith, T.J., et al. Recommendations for the Use of WBC Growth Factors: American Society of
Clinical Oncology Clinical Practice Guideline Update. J Clin Oncol, 2015. 33: 3199.
https://pubmed.ncbi.nlm.nih.gov/26169616
1070. Sathianathen, N.J., et al. Taxane-based chemohormonal therapy for metastatic hormone-sensitive
prostate cancer. Cochrane Database Syst Rev, 2018. 10: CD012816.
https://pubmed.ncbi.nlm.nih.gov/30320443
1071. Rydzewska, L.H.M., et al. Adding abiraterone to androgen deprivation therapy in men with
metastatic hormone-sensitive prostate cancer: A systematic review and meta-analysis. Eur J
Cancer, 2017. 84: 88.
https://pubmed.ncbi.nlm.nih.gov/28800492
1072. Hoyle, A.P., et al. Abiraterone in “High-” and “Low-risk” Metastatic Hormone-sensitive Prostate
Cancer. Eur Urol, 2019. 76: 719.
https://pubmed.ncbi.nlm.nih.gov/31447077
1073. Davis, I.D., et al. Enzalutamide with Standard First-Line Therapy in Metastatic Prostate Cancer.
N Engl J Med, 2019. 381: 121.
https://pubmed.ncbi.nlm.nih.gov/31157964
1074. Sydes, M.R., et al. Adding abiraterone or docetaxel to long-term hormone therapy for prostate
cancer: directly randomised data from the STAMPEDE multi-arm, multi-stage platform protocol. Ann
Oncol, 2018. 29: 1235.
https://pubmed.ncbi.nlm.nih.gov/29529169
1075. Wallis, C.J.D., et al. Comparison of Abiraterone Acetate and Docetaxel with Androgen Deprivation
Therapy in High-risk and Metastatic Hormone-naive Prostate Cancer: A Systematic Review and
Network Meta-analysis. Eur Urol, 2018. 73: 834.
https://pubmed.ncbi.nlm.nih.gov/29037513
1076. Vale, C.L., et al. What is the optimal systemic treatment of men with metastatic, hormone-naive
prostate cancer? A STOPCAP systematic review and network meta-analysis. Ann Oncol, 2018.
29: 1249.
https://pubmed.ncbi.nlm.nih.gov/29788164
1077. Marchioni, M., et al. New Antiandrogen Compounds Compared to Docetaxel for Metastatic
Hormone Sensitive Prostate Cancer: Results from a Network Meta-Analysis. J Urol, 2020. 203: 751.
https://pubmed.ncbi.nlm.nih.gov/31689158
1078. Sathianathen, N.J., et al. Indirect Comparisons of Efficacy between Combination Approaches in
Metastatic Hormone-sensitive Prostate Cancer: A Systematic Review and Network Meta-analysis.
Eur Urol, 2020. 77: 365.
https://pubmed.ncbi.nlm.nih.gov/31679970
1079. Boeve, L.M.S., et al. Effect on Survival of Androgen Deprivation Therapy Alone Compared to
Androgen Deprivation Therapy Combined with Concurrent Radiation Therapy to the Prostate in
Patients with Primary Bone Metastatic Prostate Cancer in a Prospective Randomised Clinical Trial:
Data from the HORRAD Trial. Eur Urol, 2019. 75: 410.
https://pubmed.ncbi.nlm.nih.gov/30266309
1080. Burdett, S., et al. Prostate Radiotherapy for Metastatic Hormone-sensitive Prostate Cancer: A
STOPCAP Systematic Review and Meta-analysis. Eur Urol, 2019. 76: 115.
https://pubmed.ncbi.nlm.nih.gov/30826218
1081. Ost, P., et al. Surveillance or Metastasis-Directed Therapy for Oligometastatic Prostate Cancer
Recurrence: A Prospective, Randomized, Multicenter Phase II Trial. J Clin Oncol, 2018. 36: 446.
https://pubmed.ncbi.nlm.nih.gov/29240541
1082. Phillips, R., et al. Outcomes of Observation vs Stereotactic Ablative Radiation for Oligometastatic
Prostate Cancer: The ORIOLE Phase 2 Randomized Clinical Trial. JAMA Oncol, 2020. 6: 650.
https://pubmed.ncbi.nlm.nih.gov/32215577
1083. Battaglia, A., et al. Novel Insights into the Management of Oligometastatic Prostate Cancer: A
Comprehensive Review. Eur Urol Oncol, 2019. 2: 174.
https://pubmed.ncbi.nlm.nih.gov/31017094
1084. Connor, M.J., et al. Targeting Oligometastasis with Stereotactic Ablative Radiation Therapy or
Surgery in Metastatic Hormone-sensitive Prostate Cancer: A Systematic Review of Prospective
Clinical Trials. Eur Urol Oncol, 2020. 3: 582.
https://pubmed.ncbi.nlm.nih.gov/32891600
1085. Eisenhauer, E.A., et al. New response evaluation criteria in solid tumours: revised RECIST guideline
(version 1.1). Eur J Cancer, 2009. 45: 228.
https://pubmed.ncbi.nlm.nih.gov/19097774

PROSTATE CANCER - LIMITED UPDATE 2021 195


1086. U.S. Food and Drug Adminstration approves liquid biopsy NGS companion diagnostic test for
multiple cancers and biomarkers 2020 [access date March 2021].
https://www.fda.gov/drugs/fda-approves-liquid-biopsy-ngs-companion-diagnostic-test-multiple-
cancers-and-biomarkers
1087. Lotan, T.L., et al. Report From the International Society of Urological Pathology (ISUP) Consultation
Conference on Molecular Pathology of Urogenital Cancers. I. Molecular Biomarkers in Prostate
Cancer. Am J Surg Pathol, 2020. 44: e15.
https://pubmed.ncbi.nlm.nih.gov/32044806
1088. Dienstmann, R., et al. Standardized decision support in next generation sequencing reports of
somatic cancer variants. Mol Oncol, 2014. 8: 859.
https://pubmed.ncbi.nlm.nih.gov/24768039
1089. Li, M.M., et al. Standards and Guidelines for the Interpretation and Reporting of Sequence Variants in
Cancer: A Joint Consensus Recommendation of the Association for Molecular Pathology, American
Society of Clinical Oncology, and College of American Pathologists. J Mol Diagn, 2017. 19: 4.
https://pubmed.ncbi.nlm.nih.gov/27993330
1090. Hussain, M., et al. LBA12_PRPROfound: Phase III study of olaparib versus enzalutamide or
abiraterone for metastatic castration-resistant prostate cancer (mCRPC) with homologous
recombination repair (HRR) gene alterations. Ann Oncol, 2019. 30.
https://www.researchgate.net/publication/336195431
1091. de Bono, J., et al. Olaparib for Metastatic Castration-Resistant Prostate Cancer. N Engl J Med,
2020. 382: 2091.
https://pubmed.ncbi.nlm.nih.gov/32343890
1092. Hussain, M., et al. Survival with Olaparib in Metastatic Castration-Resistant Prostate Cancer. N Engl
J Med, 2020. 383: 2345.
https://pubmed.ncbi.nlm.nih.gov/32955174
1093. U.S. Food and Drug Adminstration. Pembrolizumab (KEYTRUDA) 2016 [access date March 2021].
2020.
https://www.ema.europa.eu/en/medicines/human/summaries-opinion/keytruda-2
1094. Le, D.T., et al. PD-1 Blockade in Tumors with Mismatch-Repair Deficiency. N Engl J Med, 2015.
372: 2509.
https://pubmed.ncbi.nlm.nih.gov/26028255
1095. de Wit, R., et al. Cabazitaxel versus Abiraterone or Enzalutamide in Metastatic Prostate Cancer.
N Engl J Med, 2019. 381: 2506.
https://pubmed.ncbi.nlm.nih.gov/31566937
1096. Loriot, Y., et al. Prior long response to androgen deprivation predicts response to next-generation
androgen receptor axis targeted drugs in castration resistant prostate cancer. Eur J Cancer, 2015.
51: 1946.
https://pubmed.ncbi.nlm.nih.gov/26208462
1097. Smith, M.R., et al. Natural history of rising serum prostate-specific antigen in men with castrate
nonmetastatic prostate cancer. J Clin Oncol, 2005. 23: 2918.
https://pubmed.ncbi.nlm.nih.gov/15860850
1098. Smith, M.R., et al. Disease and host characteristics as predictors of time to first bone metastasis
and death in men with progressive castration-resistant nonmetastatic prostate cancer. Cancer,
2011. 117: 2077.
https://pubmed.ncbi.nlm.nih.gov/21523719
1099. Crawford, E.D., et al. Challenges and recommendations for early identification of metastatic disease
in prostate cancer. Urology, 2014. 83: 664.
https://pubmed.ncbi.nlm.nih.gov/24411213
1100. Fendler, W.P., et al. Prostate-Specific Membrane Antigen Ligand Positron Emission Tomography in
Men with Nonmetastatic Castration-Resistant Prostate Cancer. Clin Cancer Res, 2019. 25: 7448.
https://pubmed.ncbi.nlm.nih.gov/31511295
1101. Hussain, M., et al. Enzalutamide in Men with Nonmetastatic, Castration-Resistant Prostate Cancer.
N Engl J Med, 2018. 378: 2465.
https://pubmed.ncbi.nlm.nih.gov/29949494
1102. Smith, M.R., et al. Apalutamide Treatment and Metastasis-free Survival in Prostate Cancer. N Engl
J Med, 2018. 378: 1408.
https://pubmed.ncbi.nlm.nih.gov/29420164
1103. Fizazi, K., et al. Darolutamide in Nonmetastatic, Castration-Resistant Prostate Cancer. N Engl
J Med, 2019. 380: 1235.
https://pubmed.ncbi.nlm.nih.gov/30763142

196 PROSTATE CANCER - LIMITED UPDATE 2021


1104. Hussain, M., et al. Effects of continued androgen-deprivation therapy and other prognostic factors
on response and survival in phase II chemotherapy trials for hormone-refractory prostate cancer: a
Southwest Oncology Group report. J Clin Oncol, 1994. 12: 1868.
https://pubmed.ncbi.nlm.nih.gov/8083710
1105. Taylor, C.D., et al. Importance of continued testicular suppression in hormone-refractory prostate
cancer. J Clin Oncol, 1993. 11: 2167.
https://pubmed.ncbi.nlm.nih.gov/8229130
1106. Petrylak, D.P., et al. Docetaxel and estramustine compared with mitoxantrone and prednisone for
advanced refractory prostate cancer. N Engl J Med, 2004. 351: 1513.
https://pubmed.ncbi.nlm.nih.gov/15470214
1107. Berthold, D.R., et al. Docetaxel plus prednisone or mitoxantrone plus prednisone for advanced
prostate cancer: updated survival in the TAX 327 study. J Clin Oncol, 2008. 26: 242.
https://pubmed.ncbi.nlm.nih.gov/18182665
1108. Tannock, I.F., et al. Docetaxel plus prednisone or mitoxantrone plus prednisone for advanced
prostate cancer. N Engl J Med, 2004. 351: 1502.
https://pubmed.ncbi.nlm.nih.gov/15470213
1109. Ryan, C.J., et al. Abiraterone in metastatic prostate cancer without previous chemotherapy. N Engl
J Med, 2013. 368: 138.
https://pubmed.ncbi.nlm.nih.gov/23228172
1110. Rathkopf, D.E., et al. Updated interim efficacy analysis and long-term safety of abiraterone acetate
in metastatic castration-resistant prostate cancer patients without prior chemotherapy (COU-
AA-302). Eur Urol, 2014. 66: 815.
https://pubmed.ncbi.nlm.nih.gov/24647231
1111. Ryan, C.J., et al. Abiraterone acetate plus prednisone versus placebo plus prednisone in
chemotherapy-naive men with metastatic castration-resistant prostate cancer (COU-AA-302): final
overall survival analysis of a randomised, double-blind, placebo-controlled phase 3 study. Lancet
Oncol, 2015. 16: 152.
https://pubmed.ncbi.nlm.nih.gov/25601341
1112. Beer, T.M., et al. Enzalutamide in metastatic prostate cancer before chemotherapy. N Engl J Med,
2014. 371: 424.
https://pubmed.ncbi.nlm.nih.gov/24881730
1113. Kantoff, P.W., et al. Sipuleucel-T immunotherapy for castration-resistant prostate cancer. N Engl
J Med, 2010. 363: 411.
https://pubmed.ncbi.nlm.nih.gov/20818862
1114. Small, E.J., et al. Placebo-controlled phase III trial of immunologic therapy with sipuleucel-T
(APC8015) in patients with metastatic, asymptomatic hormone refractory prostate cancer. J Clin
Oncol, 2006. 24: 3089.
https://pubmed.ncbi.nlm.nih.gov/16809734
1115. De Bono, J.S., et al. IPATential150: Phase III study of ipatasertib (ipat) plus abiraterone (abi) vs
placebo (pbo) plus abi in metastatic castration-resistant prostate cancer (mCRPC). Ann Oncol 2020.
31: S1142.
https://www.annalsofoncology.org/article/S0923-7534(20)42332-4/abstract
1116. Roviello, G., et al. Targeting the androgenic pathway in elderly patients with castration-resistant
prostate cancer: A meta-analysis of randomized trials. Medicine (Baltimore), 2016. 95: e4636.
https://pubmed.ncbi.nlm.nih.gov/27787354
1117. Graff, J.N., et al. Efficacy and safety of enzalutamide in patients 75 years or older with
chemotherapy-naive metastatic castration-resistant prostate cancer: results from PREVAIL. Ann
Oncol, 2016. 27: 286.
https://pubmed.ncbi.nlm.nih.gov/26578735
1118. Evans, C.P., et al. The PREVAIL Study: Primary Outcomes by Site and Extent of Baseline Disease
for Enzalutamide-treated Men with Chemotherapy-naive Metastatic Castration-resistant Prostate
Cancer. Eur Urol, 2016. 70: 675.
https://pubmed.ncbi.nlm.nih.gov/27006332
1119. Alumkal, J.J., et al. Effect of Visceral Disease Site on Outcomes in Patients With Metastatic
Castration-resistant Prostate Cancer Treated With Enzalutamide in the PREVAIL Trial. Clin Genitourin
Cancer, 2017. 15: 610.
https://pubmed.ncbi.nlm.nih.gov/28344102
1120. Shore, N.D., et al. Efficacy and safety of enzalutamide versus bicalutamide for patients with
metastatic prostate cancer (TERRAIN): a randomised, double-blind, phase 2 study. Lancet Oncol,
2016. 17: 153.
https://pubmed.ncbi.nlm.nih.gov/26774508

PROSTATE CANCER - LIMITED UPDATE 2021 197


1121. Beer, T.M., et al. Enzalutamide in Men with Chemotherapy-naive Metastatic Castration-resistant
Prostate Cancer: Extended Analysis of the Phase 3 PREVAIL Study. Eur Urol, 2017. 71: 151.
https://pubmed.ncbi.nlm.nih.gov/27477525
1122. Scher, H.I., et al. Trial Design and Objectives for Castration-Resistant Prostate Cancer: Updated
Recommendations From the Prostate Cancer Clinical Trials Working Group 3. J Clin Oncol, 2016.
34: 1402.
https://pubmed.ncbi.nlm.nih.gov/26903579
1123. Armstrong, A.J., et al. Prediction of survival following first-line chemotherapy in men with castration-
resistant metastatic prostate cancer. Clin Cancer Res, 2010. 16: 203.
https://pubmed.ncbi.nlm.nih.gov/20008841
1124. Italiano, A., et al. Docetaxel-based chemotherapy in elderly patients (age 75 and older) with
castration-resistant prostate cancer. Eur Urol, 2009. 55: 1368.
https://pubmed.ncbi.nlm.nih.gov/18706755
1125. Horgan, A.M., et al. Tolerability and efficacy of docetaxel in older men with metastatic castrate-
resistant prostate cancer (mCRPC) in the TAX 327 trial. J Geriatr Oncol, 2014. 5: 119.
https://pubmed.ncbi.nlm.nih.gov/24495703
1126. Kellokumpu-Lehtinen, P.L., et al. 2-Weekly versus 3-weekly docetaxel to treat castration-resistant
advanced prostate cancer: a randomised, phase 3 trial. Lancet Oncol, 2013. 14: 117.
https://pubmed.ncbi.nlm.nih.gov/23294853
1127. Fizazi, K., et al. Abiraterone acetate for treatment of metastatic castration-resistant prostate cancer:
final overall survival analysis of the COU-AA-301 randomised, double-blind, placebo-controlled
phase 3 study. Lancet Oncol, 2012. 13: 983.
https://pubmed.ncbi.nlm.nih.gov/22995653
1128. de Bono, J.S., et al. Abiraterone and increased survival in metastatic prostate cancer. N Engl J Med,
2011. 364: 1995.
https://pubmed.ncbi.nlm.nih.gov/21612468
1129. Parker, C., et al. Alpha emitter radium-223 and survival in metastatic prostate cancer. N Engl J Med,
2013. 369: 213.
https://pubmed.ncbi.nlm.nih.gov/23863050
1130. Bahl, A., et al. Impact of cabazitaxel on 2-year survival and palliation of tumour-related pain in men with
metastatic castration-resistant prostate cancer treated in the TROPIC trial. Ann Oncol, 2013. 24: 2402.
https://pubmed.ncbi.nlm.nih.gov/23723295
1131. de Bono, J.S., et al. Prednisone plus cabazitaxel or mitoxantrone for metastatic castration-resistant
prostate cancer progressing after docetaxel treatment: a randomised open-label trial. Lancet, 2010.
376: 1147.
https://pubmed.ncbi.nlm.nih.gov/20888992
1132. Scher, H.I., et al. Increased survival with enzalutamide in prostate cancer after chemotherapy. N Engl
J Med, 2012. 367: 1187.
https://pubmed.ncbi.nlm.nih.gov/22894553
1133. de Bono, J.S., et al. Final overall survival (OS) analysis of PROfound: Olaparib vs physician’s
choice of enzalutamide or abiraterone in patients (pts) with metastatic castration-resistant prostate
cancer (mCRPC) and homologous recombination repair (HRR) gene alterations. Ann Oncol 2020.
31: S507.
https://oncologypro.esmo.org/meeting-resources/esmo-virtual-congress-2020/final-overall-survival-
os-analysis-of-profound-olaparib-vs-physician-s-choice-of-enzalutamide-or-abiraterone-in-patients-
pts-with-metastatic-c
1134. Scher, H.I., et al. Clinical trials in relapsed prostate cancer: defining the target. J Natl Cancer Inst,
1996. 88: 1623.
https://pubmed.ncbi.nlm.nih.gov/8931606
1135. Sartor, A., et al. Cabazitaxel vs docetaxel in chemotherapy-naive (CN) patients with metastatic
castration-resistant prostate cancer (mCRPC): A three-arm phase III study (FIRSTANA). J Clin Oncol
2016. 34: Abstract 5006.
https://ascopubs.org/doi/10.1200/JCO.2016.34.15_suppl.5006
1136. Eisenberger, M., et al. Phase III Study Comparing a Reduced Dose of Cabazitaxel (20 mg/m(2)) and
the Currently Approved Dose (25 mg/m(2)) in Postdocetaxel Patients With Metastatic Castration-
Resistant Prostate Cancer-PROSELICA. J Clin Oncol, 2017. 35: 3198.
https://pubmed.ncbi.nlm.nih.gov/28809610
1137. Di Lorenzo, G., et al. Peg-filgrastim and cabazitaxel in prostate cancer patients. Anticancer Drugs,
2013. 24: 84.
https://pubmed.ncbi.nlm.nih.gov/23044721

198 PROSTATE CANCER - LIMITED UPDATE 2021


1138. Hoskin, P., et al. Efficacy and safety of radium-223 dichloride in patients with castration-resistant
prostate cancer and symptomatic bone metastases, with or without previous docetaxel use: a
prespecified subgroup analysis from the randomised, double-blind, phase 3 ALSYMPCA trial.
Lancet Oncol, 2014. 15: 1397.
https://pubmed.ncbi.nlm.nih.gov/25439694
1139. European Medicines Agency (EMA). EMA restricts use of prostate cancer medicine Xofigo. 2018.
https://www.ema.europa.eu/en/news/ema-restricts-use-prostate-cancer-medicine-xofigo
1140. Smith, M., et al. Addition of radium-223 to abiraterone acetate and prednisone or prednisolone in
patients with castration-resistant prostate cancer and bone metastases (ERA 223): a randomised,
double-blind, placebo-controlled, phase 3 trial. Lancet Oncol, 2019. 20: 408.
https://pubmed.ncbi.nlm.nih.gov/30738780
1141. de Bono, J.S., et al. Subsequent Chemotherapy and Treatment Patterns After Abiraterone Acetate in
Patients with Metastatic Castration-resistant Prostate Cancer: Post Hoc Analysis of COU-AA-302.
Eur Urol, 2017. 71: 656.
https://pubmed.ncbi.nlm.nih.gov/27402060
1142. Badrising, S., et al. Clinical activity and tolerability of enzalutamide (MDV3100) in patients with
metastatic, castration-resistant prostate cancer who progress after docetaxel and abiraterone
treatment. Cancer, 2014. 120: 968.
https://pubmed.ncbi.nlm.nih.gov/24382803
1143. Zhang, T., et al. Enzalutamide versus abiraterone acetate for the treatment of men with metastatic
castration-resistant prostate cancer. Expert Opin Pharmacother, 2015. 16: 473.
https://pubmed.ncbi.nlm.nih.gov/25534660
1144. Antonarakis, E.S., et al. AR-V7 and resistance to enzalutamide and abiraterone in prostate cancer.
N Engl J Med, 2014. 371: 1028.
https://pubmed.ncbi.nlm.nih.gov/25184630
1145. Attard, G., et al. Abiraterone Alone or in Combination With Enzalutamide in Metastatic Castration-
Resistant Prostate Cancer With Rising Prostate-Specific Antigen During Enzalutamide Treatment.
J Clin Oncol, 2018. 36: 2639.
https://pubmed.ncbi.nlm.nih.gov/30028657
1146. Mateo, J., et al. DNA-Repair Defects and Olaparib in Metastatic Prostate Cancer. N Engl J Med,
2015. 373: 1697.
https://pubmed.ncbi.nlm.nih.gov/26510020
1147. Mateo, J., et al. Olaparib in patients with metastatic castration-resistant prostate cancer with DNA
repair gene aberrations (TOPARP-B): a multicentre, open-label, randomised, phase 2 trial. Lancet
Oncol, 2020. 21: 162.
https://pubmed.ncbi.nlm.nih.gov/31806540
1148. European Medicines Agency. Lynparza (olaparib). 2014 [access date March 2021].
https://www.ema.europa.eu/en/medicines/human/EPAR/lynparza
1149. Abida, W., et al. Rucaparib in Men With Metastatic Castration-Resistant Prostate Cancer Harboring
a BRCA1 or BRCA2 Gene Alteration. J Clin Oncol, 2020. 38: 3763.
https://pubmed.ncbi.nlm.nih.gov/32795228
1150. U.S. Food and Drug Adminstration grants accelerated approval to rucaparib for BRCA-mutated
metastatic castration-resistant prostate cancer. 2020 [access date March 2021].
https://www.fda.gov/drugs/fda-grants-accelerated-approval-rucaparib-brca-mutated-metastatic-
castration-resistant-prostate
1151. Khalaf, D.J., et al. Optimal sequencing of enzalutamide and abiraterone acetate plus prednisone
in metastatic castration-resistant prostate cancer: a multicentre, randomised, open-label, phase 2,
crossover trial. Lancet Oncol, 2019. 20: 1730.
https://www.researchgate.net/publication/337174368
1152. Miyake, H., et al. Comparative Assessment of Efficacies Between 2 Alternative Therapeutic
Sequences With Novel Androgen Receptor-Axis-Targeted Agents in Patients With Chemotherapy-
Naïve Metastatic Castration-Resistant Prostate Cancer. Clin Genitourin Cancer, 2017. 15: e591.
https://pubmed.ncbi.nlm.nih.gov/28063845
1153. Terada, N., et al. Exploring the optimal sequence of abiraterone and enzalutamide in patients with
chemotherapy-naïve castration-resistant prostate cancer: The Kyoto-Baltimore collaboration. Int
J Urol, 2017. 24: 441.
https://pubmed.ncbi.nlm.nih.gov/28455853
1154. Azad, A.A., et al. Efficacy of enzalutamide following abiraterone acetate in chemotherapy-naive
metastatic castration-resistant prostate cancer patients. Eur Urol, 2015. 67: 23.
https://pubmed.ncbi.nlm.nih.gov/25018038

PROSTATE CANCER - LIMITED UPDATE 2021 199


1155. Kobayashi, T., et al. Sequential Use of Androgen Receptor Axis-targeted Agents in Chemotherapy-
naive Castration-resistant Prostate Cancer: A Multicenter Retrospective Analysis With 3-Year
Follow-up. Clin Genitourin Cancer, 2020. 18: e46.
https://pubmed.ncbi.nlm.nih.gov/31759831
1156. Komura, K., et al. Comparison of Radiographic Progression-Free Survival and PSA Response
on Sequential Treatment Using Abiraterone and Enzalutamide for Newly Diagnosed Castration-
Resistant Prostate Cancer: A Propensity Score Matched Analysis from Multicenter Cohort. J Clin
Med, 2019. 8.
https://pubmed.ncbi.nlm.nih.gov/31430900
1157. Matsubara, N., et al. Abiraterone Followed by Enzalutamide Versus Enzalutamide Followed by
Abiraterone in Chemotherapy-naive Patients With Metastatic Castration-resistant Prostate Cancer.
Clin Genitourin Cancer, 2018. 16: 142.
https://pubmed.ncbi.nlm.nih.gov/29042308
1158. Maughan, B.L., et al. Comparing Sequencing of Abiraterone and Enzalutamide in Men With
Metastatic Castration-Resistant Prostate Cancer: A Retrospective Study. Prostate, 2017. 77: 33.
https://pubmed.ncbi.nlm.nih.gov/27527643
1159. de Bono, J.S., et al. Antitumour Activity and Safety of Enzalutamide in Patients with Metastatic
Castration-resistant Prostate Cancer Previously Treated with Abiraterone Acetate Plus Prednisone
for ≥24 weeks in Europe. Eur Urol, 2018. 74: 37.
https://pubmed.ncbi.nlm.nih.gov/28844372
1160. Mori, K., et al. Sequential therapy of abiraterone and enzalutamide in castration-resistant prostate
cancer: a systematic review and meta-analysis. Prostate Cancer Prostatic Dis, 2020. 23: 539.
https://www.nature.com/articles/s41391-020-0222-6
1161. Lavaud, P., et al. Anticancer Activity and Tolerance of Treatments Received Beyond Progression in
Men Treated Upfront with Androgen Deprivation Therapy With or Without Docetaxel for Metastatic
Castration-naïve Prostate Cancer in the GETUG-AFU 15 Phase 3 Trial. Eur Urol, 2018. 73: 696.
https://pubmed.ncbi.nlm.nih.gov/29074061
1162. Serafini, A.N. Current status of systemic intravenous radiopharmaceuticals for the treatment of
painful metastatic bone disease. Int J Radiat Oncol Biol Phys, 1994. 30: 1187.
https://pubmed.ncbi.nlm.nih.gov/7525518
1163. Antonarakis, E.S., et al. Pembrolizumab for Treatment-Refractory Metastatic Castration-Resistant
Prostate Cancer: Multicohort, Open-Label Phase II KEYNOTE-199 Study. J Clin Oncol, 2020.
38: 395.
https://pubmed.ncbi.nlm.nih.gov/31774688
1164. Ballinger, J.R. Theranostic radiopharmaceuticals: established agents in current use. Br J Radiol,
2018. 91: 20170969.
https://pubmed.ncbi.nlm.nih.gov/29474096
1165. Emmett, L., et al. Lutetium (177) PSMA radionuclide therapy for men with prostate cancer: a review of
the current literature and discussion of practical aspects of therapy. J Med Radiat Sci, 2017. 64: 52.
https://pubmed.ncbi.nlm.nih.gov/28303694
1166. Calopedos, R.J.S., et al. Lutetium-177-labelled anti-prostate-specific membrane antigen antibody
and ligands for the treatment of metastatic castrate-resistant prostate cancer: a systematic review
and meta-analysis. Prostate Cancer Prostatic Dis, 2017. 20: 352.
https://pubmed.ncbi.nlm.nih.gov/28440324
1167. Hofman, M.S., et al. [(177)Lu]-PSMA-617 radionuclide treatment in patients with metastatic
castration-resistant prostate cancer (LuPSMA trial): a single-centre, single-arm, phase 2 study.
Lancet Oncol, 2018. 19: 825.
https://pubmed.ncbi.nlm.nih.gov/29752180
1168. Emmett, L., et al. Results of a Prospective Phase 2 Pilot Trial of (177)Lu-PSMA-617 Therapy
for Metastatic Castration-Resistant Prostate Cancer Including Imaging Predictors of Treatment
Response and Patterns of Progression. Clin Genitourin Cancer, 2019. 17: 15.
https://pubmed.ncbi.nlm.nih.gov/30425003
1169. Hofman, M.S., et al. TheraP: a randomized phase 2 trial of (177) Lu-PSMA-617 theranostic
treatment vs cabazitaxel in progressive metastatic castration-resistant prostate cancer (Clinical Trial
Protocol ANZUP 1603). BJU Int, 2019. 124 Suppl 1: 5.
https://pubmed.ncbi.nlm.nih.gov/31638341
1170. Gillessen, S., et al. Management of patients with advanced prostate cancer: recommendations of
the St Gallen Advanced Prostate Cancer Consensus Conference (APCCC) 2015. Ann Oncol, 2015.
26: 1589.
https://pubmed.ncbi.nlm.nih.gov/27141017

200 PROSTATE CANCER - LIMITED UPDATE 2021


1171. Saad, F., et al. Prostate-specific Antigen Progression in Enzalutamide-treated Men with
Nonmetastatic Castration-resistant Prostate Cancer: Any Rise in Prostate-specific Antigen May
Require Closer Monitoring. Eur Urol, 2020. 78: 847.
https://pubmed.ncbi.nlm.nih.gov/33010985
1172. Aggarwal, R., et al. Heterogeneous Flare in Prostate-specific Membrane Antigen Positron Emission
Tomography Tracer Uptake with Initiation of Androgen Pathway Blockade in Metastatic Prostate
Cancer. Eur Urol Oncol, 2018. 1: 78.
https://pubmed.ncbi.nlm.nih.gov/31100231
1173. Payne, H., et al. Prostate-specific antigen: an evolving role in diagnosis, monitoring, and treatment
evaluation in prostate cancer. Urol Oncol, 2011. 29: 593.
https://pubmed.ncbi.nlm.nih.gov/20060331
1174. Pezaro, C.J., et al. Visceral disease in castration-resistant prostate cancer. Eur Urol, 2014. 65: 270.
https://pubmed.ncbi.nlm.nih.gov/24295792
1175. Ohlmann C, et al. Second-line chemotherapy with docetaxel for prostate-specific antigen
relapse in men with hormone refractory prostate cancer previously treated with docetaxel based
chemotherapy. Eur Urol Suppl 2006. 5: abstract #289.
https://ascopubs.org/doi/abs/10.1200/jco.2005.23.16_suppl.4682
1176. Gillessen, S., et al. Management of Patients with Advanced Prostate Cancer: Report of the
Advanced Prostate Cancer Consensus Conference 2019. Eur Urol, 2020. 77: 508.
https://pubmed.ncbi.nlm.nih.gov/32001144
1177. Rao, K., et al. Uro-oncology multidisciplinary meetings at an Australian tertiary referral centre--
impact on clinical decision-making and implications for patient inclusion. BJU Int, 2014. 114 Suppl
1: 50.
https://pubmed.ncbi.nlm.nih.gov/25070295
1178. Cereceda, L.E., et al. Management of vertebral metastases in prostate cancer: a retrospective
analysis in 119 patients. Clin Prostate Cancer, 2003. 2: 34.
https://pubmed.ncbi.nlm.nih.gov/15046682
1179. Chaichana, K.L., et al. Outcome following decompressive surgery for different histological types of
metastatic tumors causing epidural spinal cord compression. Clinical article. J Neurosurg Spine,
2009. 11: 56.
https://pubmed.ncbi.nlm.nih.gov/19569942
1180. Hoskin, P., et al. A Multicenter Randomized Trial of Ibandronate Compared With Single-Dose
Radiotherapy for Localized Metastatic Bone Pain in Prostate Cancer. J Natl Cancer Inst, 2015. 107.
https://pubmed.ncbi.nlm.nih.gov/26242893
1181. Frankel, B.M., et al. Percutaneous vertebral augmentation: an elevation in adjacent-level fracture risk
in kyphoplasty as compared with vertebroplasty. Spine J, 2007. 7: 575.
https://pubmed.ncbi.nlm.nih.gov/17905320
1182. Dutka, J., et al. Time of survival and quality of life of the patients operatively treated due to
pathological fractures due to bone metastases. Ortop Traumatol Rehabil, 2003. 5: 276.
https://pubmed.ncbi.nlm.nih.gov/18034018
1183. Frankel, B.M., et al. Segmental polymethylmethacrylate-augmented pedicle screw fixation in
patients with bone softening caused by osteoporosis and metastatic tumor involvement: a clinical
evaluation. Neurosurgery, 2007. 61: 531.
https://pubmed.ncbi.nlm.nih.gov/17881965
1184. Lawton, A.J., et al. Assessment and Management of Patients With Metastatic Spinal Cord
Compression: A Multidisciplinary Review. J Clin Oncol, 2019. 37: 61.
https://pubmed.ncbi.nlm.nih.gov/30395488
1185. Saad, F., et al. A randomized, placebo-controlled trial of zoledronic acid in patients with hormone-
refractory metastatic prostate carcinoma. J Natl Cancer Inst, 2002. 94: 1458.
https://pubmed.ncbi.nlm.nih.gov/12359855
1186. Fizazi, K., et al. Denosumab versus zoledronic acid for treatment of bone metastases in men with
castration-resistant prostate cancer: a randomised, double-blind study. Lancet, 2011. 377: 813.
https://pubmed.ncbi.nlm.nih.gov/21353695
1187. Smith, M.R., et al. Denosumab and bone-metastasis-free survival in men with castration-resistant
prostate cancer: results of a phase 3, randomised, placebo-controlled trial. Lancet, 2012. 379: 39.
https://pubmed.ncbi.nlm.nih.gov/22093187
1188. Marco, R.A., et al. Functional and oncological outcome of acetabular reconstruction for the
treatment of metastatic disease. J Bone Joint Surg Am, 2000. 82: 642.
https://pubmed.ncbi.nlm.nih.gov/10819275

PROSTATE CANCER - LIMITED UPDATE 2021 201


1189. Stopeck, A.T., et al. Safety of long-term denosumab therapy: results from the open label extension
phase of two phase 3 studies in patients with metastatic breast and prostate cancer. Support Care
Cancer, 2016. 24: 447.
https://pubmed.ncbi.nlm.nih.gov/26335402
1190. Aapro, M., et al. Guidance on the use of bisphosphonates in solid tumours: recommendations of an
international expert panel. Ann Oncol, 2008. 19: 420.
https://pubmed.ncbi.nlm.nih.gov/17906299
1191. Medication-Related Osteonecrosis of the Jaws, eds. S. Otto. 2015, Berlin Heidelberg.
1192. European Medicines Agency. Xgeva. 2019 [access date March 2021].
https://www.ema.europa.eu/en/medicines/human/EPAR/xgeva
1193. Stopeck, A.T., et al. Denosumab compared with zoledronic acid for the treatment of bone
metastases in patients with advanced breast cancer: a randomized, double-blind study. J Clin
Oncol, 2010. 28: 5132.
https://pubmed.ncbi.nlm.nih.gov/21060033
1194. Body, J.J., et al. Hypocalcaemia in patients with metastatic bone disease treated with denosumab.
Eur J Cancer, 2015. 51: 1812.
https://pubmed.ncbi.nlm.nih.gov/26093811
1195. Rice, S.M., et al. Depression and Prostate Cancer: Examining Comorbidity and Male-Specific
Symptoms. Am J Mens Health, 2018. 12: 1864.
https://pubmed.ncbi.nlm.nih.gov/29957106
1196. van Stam, M.A., et al. Prevalence and correlates of mental health problems in prostate cancer
survivors: A case-control study comparing survivors with general population peers. Urol Oncol,
2017. 35: 531 e1.
https://pubmed.ncbi.nlm.nih.gov/28457651
1197. Horwitz, E.M., et al. Definitions of biochemical failure that best predict clinical failure in patients with
prostate cancer treated with external beam radiation alone: a multi-institutional pooled analysis.
J Urol, 2005. 173: 797.
https://pubmed.ncbi.nlm.nih.gov/15711272
1198. Jackson, W.C., et al. Impact of Biochemical Failure After Salvage Radiation Therapy on Prostate
Cancer-specific Mortality: Competition Between Age and Time to Biochemical Failure. Eur Urol
Oncol, 2018. 1: 276.
https://pubmed.ncbi.nlm.nih.gov/31100248
1199. Grivas, N., et al. Ultrasensitive prostate-specific antigen level as a predictor of biochemical
progression after robot-assisted radical prostatectomy: Towards risk adapted follow-up. J Clin Lab
Anal, 2019. 33: e22693.
https://pubmed.ncbi.nlm.nih.gov/30365194
1200. Stamey, T.A., et al. Prostate specific antigen in the diagnosis and treatment of adenocarcinoma of
the prostate. II. Radical prostatectomy treated patients. J Urol, 1989. 141: 1076.
https://pubmed.ncbi.nlm.nih.gov/2468795
1201. Shen, S., et al. Ultrasensitive serum prostate specific antigen nadir accurately predicts the risk of
early relapse after radical prostatectomy. J Urol, 2005. 173: 777.
https://pubmed.ncbi.nlm.nih.gov/15711268
1202. Teeter, A.E., et al. Does Early Prostate Specific Antigen Doubling Time after Radical Prostatectomy,
Calculated Prior to Prostate Specific Antigen Recurrence, Correlate with Prostate Cancer
Outcomes? A Report from the SEARCH Database Group. J Urol, 2018. 199: 713.
https://pubmed.ncbi.nlm.nih.gov/28870860
1203. Seikkula, H., et al. Role of ultrasensitive prostate-specific antigen in the follow-up of prostate cancer
after radical prostatectomy. Urol Oncol, 2015. 33: 16.e1.
https://pubmed.ncbi.nlm.nih.gov/25456996
1204. Ray, M.E., et al. PSA nadir predicts biochemical and distant failures after external beam
radiotherapy for prostate cancer: a multi-institutional analysis. Int J Radiat Oncol Biol Phys, 2006.
64: 1140.
https://pubmed.ncbi.nlm.nih.gov/16198506
1205. Oefelein, M.G., et al. The incidence of prostate cancer progression with undetectable serum
prostate specific antigen in a series of 394 radical prostatectomies. J Urol, 1995. 154: 2128.
https://pubmed.ncbi.nlm.nih.gov/7500474
1206. Doneux, A., et al. The utility of digital rectal examination after radical radiotherapy for prostate
cancer. Clin Oncol (R Coll Radiol), 2005. 17: 172.
https://pubmed.ncbi.nlm.nih.gov/15901001

202 PROSTATE CANCER - LIMITED UPDATE 2021


1207. Chaplin, B.J., et al. Digital rectal examination is no longer necessary in the routine follow-up of men
with undetectable prostate specific antigen after radical prostatectomy: the implications for follow-
up. Eur Urol, 2005. 48: 906.
https://pubmed.ncbi.nlm.nih.gov/16126322
1208. Warren, K.S., et al. Is routine digital rectal examination required for the followup of prostate cancer?
J Urol, 2007. 178: 115.
https://pubmed.ncbi.nlm.nih.gov/17499293
1209. Saad, F., et al. Testosterone Breakthrough Rates during Androgen Deprivation Therapy for
Castration Sensitive Prostate Cancer. J Urol, 2020. 204: 416.
https://pubmed.ncbi.nlm.nih.gov/32096678
1210. Rouleau, M., et al. Discordance between testosterone measurement methods in castrated prostate
cancer patients. Endocr Connect, 2019. 8: 132.
https://pubmed.ncbi.nlm.nih.gov/30673630
1211. Morote, J., et al. Serum Testosterone Levels in Prostate Cancer Patients Undergoing Luteinizing
Hormone-Releasing Hormone Agonist Therapy. Clin Genitourin Cancer, 2018. 16: e491.
https://pubmed.ncbi.nlm.nih.gov/29198640
1212. Daniell, H.W. Osteoporosis due to androgen deprivation therapy in men with prostate cancer.
Urology, 2001. 58: 101.
https://pubmed.ncbi.nlm.nih.gov/11502461
1213. Beer, T.M., et al. The prognostic value of hemoglobin change after initiating androgen-deprivation
therapy for newly diagnosed metastatic prostate cancer: A multivariate analysis of Southwest
Oncology Group Study 8894. Cancer, 2006. 107: 489.
https://pubmed.ncbi.nlm.nih.gov/16804926
1214. Iacovelli, R., et al. The Cardiovascular Toxicity of Abiraterone and Enzalutamide in Prostate Cancer.
Clin Genitourin Cancer, 2018. 16: e645.
https://pubmed.ncbi.nlm.nih.gov/29339044
1215. Ng, H.S., et al. Development of comorbidities in men with prostate cancer treated with androgen
deprivation therapy: an Australian population-based cohort study. Prostate Cancer Prostatic Dis,
2018. 21: 403.
https://pubmed.ncbi.nlm.nih.gov/29720722
1216. Kanis, J.A., et al. Case finding for the management of osteoporosis with FRAX--assessment and
intervention thresholds for the UK. Osteoporos Int, 2008. 19: 1395.
https://pubmed.ncbi.nlm.nih.gov/18751937
1217. Cianferotti, L., et al. The prevention of fragility fractures in patients with non-metastatic prostate
cancer: a position statement by the international osteoporosis foundation. Oncotarget, 2017.
8: 75646.
https://pubmed.ncbi.nlm.nih.gov/29088899
1218. Rubin, K.H., et al. Comparison of different screening tools (FRAX®, OST, ORAI, OSIRIS, SCORE and
age alone) to identify women with increased risk of fracture. A population-based prospective study.
Bone, 2013. 56: 16.
https://pubmed.ncbi.nlm.nih.gov/23669650
1219. Conde, F.A., et al. Risk factors for male osteoporosis. Urol Oncol, 2003. 21: 380.
https://pubmed.ncbi.nlm.nih.gov/14670549
1220. Hamdy, R.C., et al. Algorithm for the management of osteoporosis. South Med J, 2010. 103: 1009.
https://pubmed.ncbi.nlm.nih.gov/20818296
1221. Higano, C.S. Bone loss and the evolving role of bisphosphonate therapy in prostate cancer. Urol
Oncol, 2003. 21: 392.
https://pubmed.ncbi.nlm.nih.gov/14670551
1222. Coleman, R., et al. Bone health in cancer: ESMO Clinical Practice Guidelines. Ann Oncol, 2020.
31: 1650.
https://pubmed.ncbi.nlm.nih.gov/32801018
1223. Edmunds, K., et al. Incidence of the adverse effects of androgen deprivation therapy for prostate
cancer: a systematic literature review. Support Care Cancer, 2020. 28: 2079.
https://pubmed.ncbi.nlm.nih.gov/31912360
1224. Edmunds, K., et al. The role of exercise in the management of adverse effects of androgen
deprivation therapy for prostate cancer: a rapid review. Support Care Cancer, 2020. 28: 5661.
https://pubmed.ncbi.nlm.nih.gov/32699997
1225. Thomas, H.R., et al. Association Between Androgen Deprivation Therapy and Patient-reported
Depression in Men With Recurrent Prostate Cancer. Clin Genitourin Cancer, 2018. 16: 313.
https://pubmed.ncbi.nlm.nih.gov/29866496

PROSTATE CANCER - LIMITED UPDATE 2021 203


1226. Miller, P.D., et al. Prostate specific antigen and bone scan correlation in the staging and monitoring
of patients with prostatic cancer. Br J Urol, 1992. 70: 295.
https://pubmed.ncbi.nlm.nih.gov/1384920
1227. Bryce, A.H., et al. Patterns of Cancer Progression of Metastatic Hormone-sensitive Prostate Cancer
in the ECOG3805 CHAARTED Trial. Eur Urol Oncol, 2020. 3: 717.
https://pubmed.ncbi.nlm.nih.gov/32807727
1228. Padhani, A.R., et al. Rationale for Modernising Imaging in Advanced Prostate Cancer. Eur Urol
Focus, 2017. 3: 223.
https://pubmed.ncbi.nlm.nih.gov/28753774
1229. Lecouvet, F.E., et al. Monitoring the response of bone metastases to treatment with Magnetic
Resonance Imaging and nuclear medicine techniques: a review and position statement by the
European Organisation for Research and Treatment of Cancer imaging group. Eur J Cancer, 2014.
50: 2519.
https://pubmed.ncbi.nlm.nih.gov/25139492
1230. Ulmert, D., et al. A novel automated platform for quantifying the extent of skeletal tumour
involvement in prostate cancer patients using the Bone Scan Index. Eur Urol, 2012. 62: 78.
https://pubmed.ncbi.nlm.nih.gov/22306323
1231. Padhani, A.R., et al. METastasis Reporting and Data System for Prostate Cancer: Practical
Guidelines for Acquisition, Interpretation, and Reporting of Whole-body Magnetic Resonance
Imaging-based Evaluations of Multiorgan Involvement in Advanced Prostate Cancer. Eur Urol, 2017.
71: 81.
https://pubmed.ncbi.nlm.nih.gov/27317091
1232. Trabulsi, E.J., et al. Optimum Imaging Strategies for Advanced Prostate Cancer: ASCO Guideline.
J Clin Oncol, 2020. 38: 1963.
https://pubmed.ncbi.nlm.nih.gov/31940221
1233. Bourke, L., et al. Survivorship and Improving Quality of Life in Men with Prostate Cancer. Eur Urol,
2015. 68: 374.
https://pubmed.ncbi.nlm.nih.gov/25941049
1234. Resnick, M.J., et al. Prostate cancer survivorship care guideline: American Society of Clinical
Oncology Clinical Practice Guideline endorsement. J Clin Oncol, 2015. 33: 1078.
https://pubmed.ncbi.nlm.nih.gov/25667275
1235. Carlsson, S., et al. Surgery-related complications in 1253 robot-assisted and 485 open retropubic
radical prostatectomies at the Karolinska University Hospital, Sweden. Urology, 2010. 75: 1092.
https://pubmed.ncbi.nlm.nih.gov/20022085
1236. Ficarra, V., et al. Retropubic, laparoscopic, and robot-assisted radical prostatectomy: a systematic
review and cumulative analysis of comparative studies. Eur Urol, 2009. 55: 1037.
https://pubmed.ncbi.nlm.nih.gov/19185977
1237. Rabbani, F., et al. Comprehensive standardized report of complications of retropubic and
laparoscopic radical prostatectomy. Eur Urol, 2010. 57: 371.
https://pubmed.ncbi.nlm.nih.gov/19945779
1238. Resnick, M.J., et al. Long-term functional outcomes after treatment for localized prostate cancer.
N Engl J Med, 2013. 368: 436.
https://pubmed.ncbi.nlm.nih.gov/23363497
1239. Parekh, A., et al. Reduced penile size and treatment regret in men with recurrent prostate cancer
after surgery, radiotherapy plus androgen deprivation, or radiotherapy alone. Urology, 2013. 81: 130.
https://pubmed.ncbi.nlm.nih.gov/23273077
1240. Msezane, L.P., et al. Bladder neck contracture after robot-assisted laparoscopic radical
prostatectomy: evaluation of incidence and risk factors and impact on urinary function. J Endourol,
2008. 22: 377.
https://pubmed.ncbi.nlm.nih.gov/18095861
1241. Chiong, E., et al. Port-site hernias occurring after the use of bladeless radially expanding trocars.
Urology, 2010. 75: 574.
https://pubmed.ncbi.nlm.nih.gov/19854489
1242. Haglind, E., et al. Corrigendum re: “Urinary Incontinence and Erectile Dysfunction After Robotic
Versus Open Radical Prostatectomy: A Prospective, Controlled, Nonrandomised Trial” [Eur Urol
2015;68:216-25]. Eur Urol, 2017. 72: e81.
https://pubmed.ncbi.nlm.nih.gov/28552613
1243. Park, B., et al. Comparison of oncological and functional outcomes of pure versus robotic-assisted
laparoscopic radical prostatectomy performed by a single surgeon. Scand J Urol, 2013. 47: 10.
https://pubmed.ncbi.nlm.nih.gov/22835035

204 PROSTATE CANCER - LIMITED UPDATE 2021


1244. Donovan, J.L., et al. Patient-Reported Outcomes after Monitoring, Surgery, or Radiotherapy for
Prostate Cancer. N Engl J Med, 2016. 375: 1425.
https://pubmed.ncbi.nlm.nih.gov/27626365
1245. Barocas, D.A., et al. Association Between Radiation Therapy, Surgery, or Observation for Localized
Prostate Cancer and Patient-Reported Outcomes After 3 Years. JAMA, 2017. 317: 1126.
https://pubmed.ncbi.nlm.nih.gov/28324093
1246. Wallis, C.J., et al. Second malignancies after radiotherapy for prostate cancer: systematic review
and meta-analysis. BMJ, 2016. 352: i851.
https://pubmed.ncbi.nlm.nih.gov/26936410
1247. Budaus, L., et al. Functional outcomes and complications following radiation therapy for prostate
cancer: a critical analysis of the literature. Eur Urol, 2012. 61: 112.
https://pubmed.ncbi.nlm.nih.gov/22001105
1248. Nguyen, P.L., et al. Adverse Effects of Androgen Deprivation Therapy and Strategies to Mitigate
Them. Eur Urol, 2014: 67(5):825.
https://pubmed.ncbi.nlm.nih.gov/25097095
1249. Donovan, K.A., et al. Psychological effects of androgen-deprivation therapy on men with prostate
cancer and their partners. Cancer, 2015. 121: 4286.
https://pubmed.ncbi.nlm.nih.gov/26372364
1250. Cherrier, M.M., et al. Cognitive and mood changes in men undergoing intermittent combined
androgen blockade for non-metastatic prostate cancer. Psychooncology, 2009. 18: 237.
https://pubmed.ncbi.nlm.nih.gov/18636420
1251. Alibhai, S.M., et al. Effects of long-term androgen deprivation therapy on cognitive function over
36 months in men with prostate cancer. Cancer, 2017. 123: 237.
https://pubmed.ncbi.nlm.nih.gov/27583806
1252. Herr, H.W., et al. Quality of life of asymptomatic men with nonmetastatic prostate cancer on
androgen deprivation therapy. J Urol, 2000. 163: 1743.
https://pubmed.ncbi.nlm.nih.gov/10799173
1253. Potosky, A.L., et al. Quality-of-life outcomes after primary androgen deprivation therapy: results
from the Prostate Cancer Outcomes Study. J Clin Oncol, 2001. 19: 3750.
https://pubmed.ncbi.nlm.nih.gov/11533098
1254. Iversen, P., et al. Bicalutamide monotherapy compared with castration in patients with
nonmetastatic locally advanced prostate cancer: 6.3 years of followup. J Urol, 2000. 164: 1579.
https://pubmed.ncbi.nlm.nih.gov/11025708
1255. Iversen, P., et al. Nonsteroidal antiandrogens: a therapeutic option for patients with advanced
prostate cancer who wish to retain sexual interest and function. BJU Int, 2001. 87: 47.
https://pubmed.ncbi.nlm.nih.gov/11121992
1256. Boccardo, F., et al. Bicalutamide monotherapy versus flutamide plus goserelin in prostate cancer
patients: results of an Italian Prostate Cancer Project study. J Clin Oncol, 1999. 17: 2027.
https://pubmed.ncbi.nlm.nih.gov/10561254
1257. Walker, L.M., et al. Luteinizing hormone--releasing hormone agonists: a quick reference for
prevalence rates of potential adverse effects. Clin Genitourin Cancer, 2013. 11: 375.
https://pubmed.ncbi.nlm.nih.gov/23891497
1258. Elliott, S., et al. Androgen deprivation therapy for prostate cancer: recommendations to improve
patient and partner quality of life. J Sex Med, 2010. 7: 2996.
https://pubmed.ncbi.nlm.nih.gov/20626600
1259. de Voogt, H.J., et al. Cardiovascular side effects of diethylstilbestrol, cyproterone acetate,
medroxyprogesterone acetate and estramustine phosphate used for the treatment of advanced
prostatic cancer: results from European Organization for Research on Treatment of Cancer trials
30761 and 30762. J Urol, 1986. 135: 303.
https://pubmed.ncbi.nlm.nih.gov/2935644
1260. Irani, J., et al. Efficacy of venlafaxine, medroxyprogesterone acetate, and cyproterone acetate for
the treatment of vasomotor hot flushes in men taking gonadotropin-releasing hormone analogues
for prostate cancer: a double-blind, randomised trial. Lancet Oncol, 2010. 11: 147.
https://pubmed.ncbi.nlm.nih.gov/19963436
1261. Sloan, J.A., et al. Methodologic lessons learned from hot flash studies. J Clin Oncol, 2001. 19: 4280.
https://pubmed.ncbi.nlm.nih.gov/11731510
1262. Moraska, A.R., et al. Gabapentin for the management of hot flashes in prostate cancer survivors: a
longitudinal continuation Study-NCCTG Trial N00CB. J Support Oncol, 2010. 8: 128.
https://pubmed.ncbi.nlm.nih.gov/20552926

PROSTATE CANCER - LIMITED UPDATE 2021 205


1263. Frisk, J., et al. Two modes of acupuncture as a treatment for hot flushes in men with prostate
cancer--a prospective multicenter study with long-term follow-up. Eur Urol, 2009. 55: 156.
https://pubmed.ncbi.nlm.nih.gov/18294761
1264. Smith, M.R., et al. Risk of clinical fractures after gonadotropin-releasing hormone agonist therapy for
prostate cancer. J Urol, 2006. 175: 136.
https://pubmed.ncbi.nlm.nih.gov/16406890
1265. Cree, M., et al. Mortality and institutionalization following hip fracture. J Am Geriatr Soc, 2000. 48: 283.
https://pubmed.ncbi.nlm.nih.gov/10733054
1266. Compston, J.E., et al. Osteoporosis. Lancet, 2019. 393: 364.
https://pubmed.ncbi.nlm.nih.gov/30696576
1267. Saylor, P.J., et al. Metabolic complications of androgen deprivation therapy for prostate cancer.
J Urol, 2009. 181: 1998.
https://pubmed.ncbi.nlm.nih.gov/19286225
1268. Gonnelli, S., et al. Obesity and fracture risk. Clin Cases Miner Bone Metab, 2014. 11: 9.
https://pubmed.ncbi.nlm.nih.gov/25002873
1269. Sieber, P.R., et al. Bicalutamide 150 mg maintains bone mineral density during monotherapy for
localized or locally advanced prostate cancer. J Urol, 2004. 171: 2272.
https://pubmed.ncbi.nlm.nih.gov/15126801
1270. Wadhwa, V.K., et al. Bicalutamide monotherapy preserves bone mineral density, muscle strength
and has significant health-related quality of life benefits for osteoporotic men with prostate cancer.
BJU Int, 2011. 107: 1923.
https://pubmed.ncbi.nlm.nih.gov/20950306
1271. Higano, C., et al. Bone mineral density in patients with prostate cancer without bone metastases
treated with intermittent androgen suppression. Urology, 2004. 64: 1182.
https://pubmed.ncbi.nlm.nih.gov/15596194
1272. Nobes, J.P., et al. A prospective, randomized pilot study evaluating the effects of metformin and
lifestyle intervention on patients with prostate cancer receiving androgen deprivation therapy. BJU
Int, 2012. 109: 1495.
https://pubmed.ncbi.nlm.nih.gov/21933330
1273. Grundy, S.M., et al. Diagnosis and management of the metabolic syndrome: an American Heart
Association/National Heart, Lung, and Blood Institute Scientific Statement. Circulation, 2005.
112: 2735.
https://pubmed.ncbi.nlm.nih.gov/16157765
1274. Braga-Basaria, M., et al. Metabolic syndrome in men with prostate cancer undergoing long-term
androgen-deprivation therapy. J Clin Oncol, 2006. 24: 3979.
https://pubmed.ncbi.nlm.nih.gov/16921050
1275. Cheung, A.S., et al. Muscle and bone effects of androgen deprivation therapy: current and emerging
therapies. Endocr Relat Cancer, 2014. 21: R371.
https://pubmed.ncbi.nlm.nih.gov/25056176
1276. Smith, M.R., et al. Sarcopenia during androgen-deprivation therapy for prostate cancer. J Clin
Oncol, 2012. 30: 3271.
https://pubmed.ncbi.nlm.nih.gov/22649143
1277. Saigal, C.S., et al. Androgen deprivation therapy increases cardiovascular morbidity in men with
prostate cancer. Cancer, 2007. 110: 1493.
https://pubmed.ncbi.nlm.nih.gov/17657815
1278. Lu-Yao, G., et al. Changing patterns in competing causes of death in men with prostate cancer: a
population based study. J Urol, 2004. 171: 2285.
https://pubmed.ncbi.nlm.nih.gov/15126804
1279. Keating, N.L., et al. Diabetes and cardiovascular disease during androgen deprivation therapy:
observational study of veterans with prostate cancer. J Natl Cancer Inst, 2010. 102: 39.
https://pubmed.ncbi.nlm.nih.gov/19996060
1280. Efstathiou, J.A., et al. Cardiovascular mortality and duration of androgen deprivation for locally
advanced prostate cancer: analysis of RTOG 92-02. Eur Urol, 2008. 54: 816.
https://pubmed.ncbi.nlm.nih.gov/18243498
1281. Jones, C.U., et al. Radiotherapy and short-term androgen deprivation for localized prostate cancer.
N Engl J Med, 2011. 365: 107.
https://pubmed.ncbi.nlm.nih.gov/21751904
1282. Nguyen, P.L., et al. Association of androgen deprivation therapy with cardiovascular death in
patients with prostate cancer: a meta-analysis of randomized trials. Jama, 2011. 306: 2359.
https://pubmed.ncbi.nlm.nih.gov/22147380

206 PROSTATE CANCER - LIMITED UPDATE 2021


1283. Bourke, L., et al. Endocrine therapy in prostate cancer: time for reappraisal of risks, benefits and
cost-effectiveness? Br J Cancer, 2013. 108: 9.
https://pubmed.ncbi.nlm.nih.gov/23321508
1284. Blankfield, R.P. Androgen deprivation therapy for prostate cancer and cardiovascular death. JAMA,
2012. 307: 1252; author reply 1252.
https://pubmed.ncbi.nlm.nih.gov/22453560
1285. Bosco, C., et al. Quantifying observational evidence for risk of fatal and nonfatal cardiovascular
disease following androgen deprivation therapy for prostate cancer: a meta-analysis. Eur Urol, 2015.
68: 386.
https://pubmed.ncbi.nlm.nih.gov/25484142
1286. Nguyen, P.L., et al. Influence of androgen deprivation therapy on all-cause mortality in men with
high-risk prostate cancer and a history of congestive heart failure or myocardial infarction. Int
J Radiat Oncol Biol Phys, 2012. 82: 1411.
https://pubmed.ncbi.nlm.nih.gov/21708431
1287. Tsai, H.K., et al. Androgen deprivation therapy for localized prostate cancer and the risk of
cardiovascular mortality. J Natl Cancer Inst, 2007. 99: 1516.
https://pubmed.ncbi.nlm.nih.gov/17925537
1288. Albertsen, P.C., et al. Cardiovascular morbidity associated with gonadotropin releasing hormone
agonists and an antagonist. Eur Urol, 2014. 65: 565.
https://pubmed.ncbi.nlm.nih.gov/24210090
1289. Gilbert, S.E., et al. Effects of a lifestyle intervention on endothelial function in men on long-term
androgen deprivation therapy for prostate cancer. Br J Cancer, 2016. 114: 401.
https://pubmed.ncbi.nlm.nih.gov/26766737
1290. Bourke, L., et al. Exercise for Men with Prostate Cancer: A Systematic Review and Meta-analysis.
Eur Urol, 2015: 69(4):693.
https://pubmed.ncbi.nlm.nih.gov/26632144
1291. Ahmadi, H., et al. Androgen deprivation therapy: evidence-based management of side effects. BJU
Int, 2013. 111: 543.
https://pubmed.ncbi.nlm.nih.gov/23351025
1292. Meng, F., et al. Stroke related to androgen deprivation therapy for prostate cancer: a meta-analysis
and systematic review. BMC Cancer, 2016. 16: 180.
https://pubmed.ncbi.nlm.nih.gov/26940836
1293. Nead, K.T., et al. Androgen Deprivation Therapy and Future Alzheimer’s Disease Risk. J Clin Oncol,
2016. 34: 566.
https://pubmed.ncbi.nlm.nih.gov/26644522
1294. Bennett, D., et al. Factors influencing job loss and early retirement in working men with prostate
cancer-findings from the population-based Life After Prostate Cancer Diagnosis (LAPCD) study.
J Cancer Surviv, 2018. 12: 669.
https://pubmed.ncbi.nlm.nih.gov/30058009
1295. Borji, M., et al. Positive Effects of Cognitive Behavioral Therapy on Depression, Anxiety and Stress
of Family Caregivers of Patients with Prostate Cancer: A Randomized Clinical Trial. Asian Pac
J Cancer Prev, 2017. 18: 3207.
https://pubmed.ncbi.nlm.nih.gov/29281868
1296. Bourke, L., et al. A qualitative study evaluating experiences of a lifestyle intervention in men with
prostate cancer undergoing androgen suppression therapy. Trials, 2012. 13: 208.
https://pubmed.ncbi.nlm.nih.gov/23151126
1297. Berruti, A., et al. Incidence of skeletal complications in patients with bone metastatic prostate
cancer and hormone refractory disease: predictive role of bone resorption and formation markers
evaluated at baseline. J Urol, 2000. 164: 1248.
https://pubmed.ncbi.nlm.nih.gov/10992374
1298. Carlin, B.I., et al. The natural history, skeletal complications, and management of bone metastases
in patients with prostate carcinoma. Cancer, 2000. 88: 2989.
https://pubmed.ncbi.nlm.nih.gov/10898342
1299. Smith, D.P., et al. Quality of life three years after diagnosis of localised prostate cancer: population
based cohort study. BMJ, 2009. 339: b4817.
https://pubmed.ncbi.nlm.nih.gov/19945997
1300. Taylor, K.L., et al. Long-term disease-specific functioning among prostate cancer survivors and
noncancer controls in the prostate, lung, colorectal, and ovarian cancer screening trial. J Clin Oncol,
2012. 30: 2768.
https://pubmed.ncbi.nlm.nih.gov/22734029

PROSTATE CANCER - LIMITED UPDATE 2021 207


1301. Cella, D.F., et al. The Functional Assessment of Cancer Therapy scale: development and validation
of the general measure. J Clin Oncol, 1993. 11: 570.
https://pubmed.ncbi.nlm.nih.gov/8445433
1302. Esper, P., et al. Measuring quality of life in men with prostate cancer using the functional assessment
of cancer therapy-prostate instrument. Urology, 1997. 50: 920.
https://pubmed.ncbi.nlm.nih.gov/9426724
1303. Groenvold, M., et al. Validation of the EORTC QLQ-C30 quality of life questionnaire through
combined qualitative and quantitative assessment of patient-observer agreement. J Clin Epidemiol,
1997. 50: 441.
https://pubmed.ncbi.nlm.nih.gov/9179103
1304. van Andel, G., et al. An international field study of the EORTC QLQ-PR25: a questionnaire for
assessing the health-related quality of life of patients with prostate cancer. Eur J Cancer, 2008.
44: 2418.
https://pubmed.ncbi.nlm.nih.gov/18774706
1305. Wei, J.T., et al. Development and validation of the expanded prostate cancer index composite (EPIC)
for comprehensive assessment of health-related quality of life in men with prostate cancer. Urology,
2000. 56: 899.
https://pubmed.ncbi.nlm.nih.gov/11113727
1306. Szymanski, K.M., et al. Development and validation of an abbreviated version of the expanded
prostate cancer index composite instrument for measuring health-related quality of life among
prostate cancer survivors. Urology, 2010. 76: 1245.
https://pubmed.ncbi.nlm.nih.gov/20350762
1307. Litwin, M.S., et al. The UCLA Prostate Cancer Index: development, reliability, and validity of a
health-related quality of life measure. Med Care, 1998. 36: 1002.
https://pubmed.ncbi.nlm.nih.gov/9674618
1308. Giesler, R.B., et al. Assessing quality of life in men with clinically localized prostate cancer:
development of a new instrument for use in multiple settings. Qual Life Res, 2000. 9: 645.
https://pubmed.ncbi.nlm.nih.gov/11236855
1309. Potosky, A.L., et al. Prostate cancer practice patterns and quality of life: the Prostate Cancer
Outcomes Study. J Natl Cancer Inst, 1999. 91: 1719.
https://pubmed.ncbi.nlm.nih.gov/10528021
1310. Hoffman, K.E., et al. Patient-Reported Outcomes Through 5 Years for Active Surveillance, Surgery,
Brachytherapy, or External Beam Radiation With or Without Androgen Deprivation Therapy for
Localized Prostate Cancer. Jama, 2020. 323: 149.
https://pubmed.ncbi.nlm.nih.gov/31935027
1311. Giberti, C., et al. Radical retropubic prostatectomy versus brachytherapy for low-risk prostatic
cancer: a prospective study. World J Urol, 2009. 27: 607.
https://pubmed.ncbi.nlm.nih.gov/19455340
1312. Giberti, C., et al. Robotic prostatectomy versus brachytherapy for the treatment of low risk prostate
cancer. Can J Urol, 2017. 24: 8728.
https://pubmed.ncbi.nlm.nih.gov/28436359
1313. Lardas, M., et al. Quality of Life Outcomes after Primary Treatment for Clinically Localised Prostate
Cancer: A Systematic Review. Eur Urol, 2017. 72: 869.
https://pubmed.ncbi.nlm.nih.gov/28757301
1314. Giesler, R.B., et al. Improving the quality of life of patients with prostate carcinoma: a randomized
trial testing the efficacy of a nurse-driven intervention. Cancer, 2005. 104: 752.
https://pubmed.ncbi.nlm.nih.gov/15986401
1315. Anderson, C.A., et al. Conservative management for postprostatectomy urinary incontinence.
Cochrane Database Syst Rev, 2015. 1: CD001843.
https://pubmed.ncbi.nlm.nih.gov/25602133
1316. Chen, Y.C., et al. Surgical treatment for urinary incontinence after prostatectomy: A meta-analysis
and systematic review. PLoS One, 2017. 12: e0130867.
https://pubmed.ncbi.nlm.nih.gov/28467435
1317. Pavlovich, C.P., et al. Nightly vs on-demand sildenafil for penile rehabilitation after minimally invasive
nerve-sparing radical prostatectomy: results of a randomized double-blind trial with placebo. BJU
Int, 2013. 112: 844.
https://pubmed.ncbi.nlm.nih.gov/23937708
1318. Philippou, Y.A., et al. Penile rehabilitation for postprostatectomy erectile dysfunction. Cochrane
Database Syst Rev, 2018. 10: CD012414.
https://pubmed.ncbi.nlm.nih.gov/30352488

208 PROSTATE CANCER - LIMITED UPDATE 2021


1319. Salonia, A., et al., EAU Guidelines on Sexual and Reproductive Health, Edn. presented at the 36th
Annual Congress, Milan. EAU Guidelines Office, Arnhem, The Netherlands.
https://uroweb.org/guideline/sexual-and-reproductive-health/
1320. Dieperink, K.B., et al. The effects of multidisciplinary rehabilitation: RePCa-a randomised study
among primary prostate cancer patients. Br J Cancer, 2013. 109: 3005.
https://pubmed.ncbi.nlm.nih.gov/3859951
1321. Galvao, D.A., et al. Combined resistance and aerobic exercise program reverses muscle loss in
men undergoing androgen suppression therapy for prostate cancer without bone metastases: a
randomized controlled trial. J Clin Oncol, 2010. 28: 340.
https://pubmed.ncbi.nlm.nih.gov/19949016
1322. Bourke, L., et al. Lifestyle changes for improving disease-specific quality of life in sedentary men
on long-term androgen-deprivation therapy for advanced prostate cancer: a randomised controlled
trial. European urology, 2014. 65: 865.
https://pubmed.ncbi.nlm.nih.gov/24119318
1323. Cella, D., et al. Estimating clinically meaningful changes for the Functional Assessment of Cancer
Therapy--Prostate: results from a clinical trial of patients with metastatic hormone-refractory
prostate cancer. Value Health, 2009. 12: 124.
https://pubmed.ncbi.nlm.nih.gov/18647260
1324. Skolarus, T.A., et al. Androgen-deprivation-associated bone disease. Curr Opin Urol, 2014. 24: 601.
https://pubmed.ncbi.nlm.nih.gov/25144145
1325. Nair-Shalliker, V., et al. Post-treatment levels of plasma 25- and 1,25-dihydroxy vitamin D and
mortality in men with aggressive prostate cancer. Sci Rep, 2020. 10: 7736.
https://pubmed.ncbi.nlm.nih.gov/32385370
1326. Grant, W.B. Review of Recent Advances in Understanding the Role of Vitamin D in Reducing Cancer
Risk: Breast, Colorectal, Prostate, and Overall Cancer. Anticancer Res, 2020. 40: 491.
https://pubmed.ncbi.nlm.nih.gov/31892604
1327. Shapiro, C.L., et al. Management of Osteoporosis in Survivors of Adult Cancers With Nonmetastatic
Disease: ASCO Clinical Practice Guideline. J Clin Oncol, 2019. 37: 2916.
https://pubmed.ncbi.nlm.nih.gov/31532726
1328. Briot, K., et al. French recommendations for osteoporosis prevention and treatment in patients with
prostate cancer treated by androgen deprivation. Joint Bone Spine, 2019. 86: 21.
https://pubmed.ncbi.nlm.nih.gov/30287350
1329. Saylor, P.J., et al. Bone Health and Bone-Targeted Therapies for Prostate Cancer: ASCO
Endorsement of a Cancer Care Ontario Guideline. J Clin Oncol, 2020. 38: 1736.
https://pubmed.ncbi.nlm.nih.gov/31990618
1330. Brown, J.E., et al. Guidance for the assessment and management of prostate cancer treatment-
induced bone loss. A consensus position statement from an expert group. J Bone Oncol, 2020.
25: 100311.
https://pubmed.ncbi.nlm.nih.gov/32995252
1331 Smith, M.R., et al. Randomized controlled trial of zoledronic acid to prevent bone loss in men
receiving androgen deprivation therapy for nonmetastatic prostate cancer. J Urol, 2003. 169: 2008.
https://pubmed.ncbi.nlm.nih.gov/12771706
1332. Michaelson, M.D., et al. Randomized controlled trial of annual zoledronic acid to prevent
gonadotropin-releasing hormone agonist-induced bone loss in men with prostate cancer. J Clin
Oncol, 2007. 25: 1038.
https://pubmed.ncbi.nlm.nih.gov/17369566
1333. Migliorati, C.A., et al. Bisphosphonate-associated osteonecrosis: a long-term complication of
bisphosphonate treatment. Lancet Oncol, 2006. 7: 508.
https://pubmed.ncbi.nlm.nih.gov/16750501
1334. Wadhwa, V.K., et al. Frequency of zoledronic acid to prevent further bone loss in osteoporotic
patients undergoing androgen deprivation therapy for prostate cancer. BJU Int, 2010. 105: 1082.
https://pubmed.ncbi.nlm.nih.gov/19912210
1335. Smith, M.R., et al. Denosumab in men receiving androgen-deprivation therapy for prostate cancer.
N Engl J Med, 2009. 361: 745.
https://pubmed.ncbi.nlm.nih.gov/19671656
1336. Chen, R.C., et al. Association Between Choice of Radical Prostatectomy, External Beam
Radiotherapy, Brachytherapy, or Active Surveillance and Patient-Reported Quality of Life Among
Men With Localized Prostate Cancer. Jama, 2017. 317: 1141.
https://pubmed.ncbi.nlm.nih.gov/28324092

PROSTATE CANCER - LIMITED UPDATE 2021 209


1337. Sanda, M.G., et al. Clinically Localized Prostate Cancer: AUA/ASTRO/SUO Guideline. Part I: Risk
Stratification, Shared Decision Making, and Care Options. J Urol, 2018. 99: 683.
https://pubmed.ncbi.nlm.nih.gov/29203269
1338. Makarov, D.V., et al. AUA white paper on implementation of shared decision making into urological
practice. Urol Pract, 2016. 3: 355.
https://www.sciencedirect.com/science/article/abs/pii/S2352077915002733
1339. Stiggelbout, A.M., et al. Shared decision making: Concepts, evidence, and practice. Patient Educ
Couns, 2015. 98: 1172.
https://pubmed.ncbi.nlm.nih.gov/26215573
1340. Violette, P.D., et al. Decision aids for localized prostate cancer treatment choice: Systematic review
and meta-analysis. CA Cancer J Clin, 2015. 65: 239.
https://pubmed.ncbi.nlm.nih.gov/25772796
1341. Ramsey, S.D., et al. Unanticipated and underappreciated outcomes during management of local
stage prostate cancer: a prospective survey. J Urol, 2010. 184: 120.
https://pubmed.ncbi.nlm.nih.gov/20478590
1342. Connolly, T., et al. Regret in decision making. Curr Direct Psycho Sci, 2002. 11: 212.
https://pubmed.ncbi.nlm.nih.gov/16045415
1343. Maguire, R., et al. Expecting the worst? The relationship between retrospective and prospective
appraisals of illness on quality of life in prostate cancer survivors. Psychooncology, 2018. 27: 1237.
https://pubmed.ncbi.nlm.nih.gov/29430755
1344. Schroeck, F.R., et al. Satisfaction and regret after open retropubic or robot-assisted laparoscopic
radical prostatectomy. Eur Urol, 2008. 54: 785.
https://pubmed.ncbi.nlm.nih.gov/18585849
1345. Steentjes, L., et al. Factors associated with current and severe physical side-effects after prostate
cancer treatment: What men report. Eur J Cancer Care (Engl), 2018. 27.
https://pubmed.ncbi.nlm.nih.gov/27726215
1346. Orom, H., et al. What Is a “Good” Treatment Decision? Decisional Control, Knowledge, Treatment
Decision Making, and Quality of Life in Men with Clinically Localized Prostate Cancer. Med Decis
Making, 2016. 36: 714.
https://pubmed.ncbi.nlm.nih.gov/26957566
1347. Davison, B.J., et al. Quality of life, sexual function and decisional regret at 1 year after surgical
treatment for localized prostate cancer. BJU Int, 2007. 100: 780.
https://pubmed.ncbi.nlm.nih.gov/17578466
1348. Martínez-González, N.A., et al. Shared decision making for men facing prostate cancer treatment: a
systematic review of randomized controlled trials. Patient Prefer Adherence, 2019. 13: 1153.
https://pubmed.ncbi.nlm.nih.gov/31413545
1349. Menichetti, J., et al. Quality of life in active surveillance and the associations with decision-making-a
literature review. Transl Androl Urol, 2018. 7: 160.
https://pubmed.ncbi.nlm.nih.gov/29594030
1350. Ivlev, I., et al. Prostate Cancer Screening Patient Decision Aids: A Systematic Review and Meta-
analysis. Am J Prev Med, 2018. 55: 896.
https://pubmed.ncbi.nlm.nih.gov/30337235
1351. Kinsella, N., et al. A Single Educational Seminar Increases Confidence and Decreases Dropout from
Active Surveillance by 5 Years After Diagnosis of Prostate Cancer. Eur Urol Oncol, 2019. 2: 464.
https://pubmed.ncbi.nlm.nih.gov/31277784
1352. Hoffman, R.M., et al. Selecting Active Surveillance: Decision Making Factors for Men with a Low-
Risk Prostate Cancer. Med Decis Making, 2019. 39: 962.
https://pubmed.ncbi.nlm.nih.gov/31631745
1353. Berry, D.L., et al. Decision Support with the Personal Patient Profile-Prostate: A Multicenter
Randomized Trial. J Urol, 2018. 199: 89.
https://pubmed.ncbi.nlm.nih.gov/28754540
1354. Campagna, J.P., et al. Prostate Cancer Survival Estimates by the General Public Using Unrestricted
Internet Searches and Online Nomograms. Eur Urol Focus, 2020. 6: 959.
https://pubmed.ncbi.nlm.nih.gov/30723050
1355. de Freitas, H.M., et al. Patient Preferences for Metastatic Hormone-Sensitive Prostate Cancer
Treatments: A Discrete Choice Experiment Among Men in Three European Countries. Adv Ther,
2019. 36: 318.
https://pubmed.ncbi.nlm.nih.gov/30617763

210 PROSTATE CANCER - LIMITED UPDATE 2021


1356. Lorent, M., et al. Meta-analysis of predictive models to assess the clinical validity and utility
for patient-centered medical decision making: application to the CAncer of the Prostate Risk
Assessment (CAPRA). BMC Med Inform Decis Mak, 2019. 19: 2.
https://pubmed.ncbi.nlm.nih.gov/30616621
1357. Riikonen, J.M., et al. Decision Aids for Prostate Cancer Screening Choice: A Systematic Review and
Meta-analysis. JAMA Intern Med, 2019. 179: 1072.
https://pubmed.ncbi.nlm.nih.gov/31233091
1358. Vromans, R.D., et al. Communicative aspects of decision aids for localized prostate cancer
treatment - A systematic review. Urol Oncol, 2019. 37: 409.
https://pubmed.ncbi.nlm.nih.gov/31053529

10. CONFLICT OF INTEREST


All members of the EAU-EANM-ESTRO-ESUR-SIOG Prostate Cancer Guidelines Panel have provided
disclosure statements of all relationships that they have that might be perceived as a potential source of
a conflict of interest. This information is publicly accessible through the European Association of Urology
website: https://uroweb.org/guideline/prostate-cancer/.
This guidelines document was developed with the financial support of the European Association of
Urology. No external sources of funding and support have been involved. The EAU is a non-profit organization
and funding is limited to administrative assistance and travel and meeting expenses. No honoraria or other
reimbursements have been provided.

11. CITATION INFORMATION


The format in which to cite the EAU Guidelines will vary depending on the style guide of the journal in which the
citation appears. Accordingly, the number of authors or whether, for instance, to include the publisher, location,
or an ISBN number may vary.

The compilation of the complete Guidelines should be referenced as:


EAU Guidelines. Edn. presented at the EAU Annual Congress Milan 2021. ISBN 978-94-92671-13-4.

If a publisher and/or location is required, include:


EAU Guidelines Office, Arnhem, The Netherlands. http://uroweb.org/guidelines/compilations-of-all-guidelines/

References to individual guidelines should be structured in the following way:


Contributors’ names. Title of resource. Publication type. ISBN. Publisher and publisher location, year.

PROSTATE CANCER - LIMITED UPDATE 2021 211


212 PROSTATE CANCER - LIMITED UPDATE 2021
EAU Guidelines on
Renal
Transplantation
A. Breda (Chair), K. Budde, A. Figueiredo, E. Lledó García,
J. Olsburgh (Vice-chair), H. Regele
Guidelines Associates: R. Boissier, V. Hevia,
O. Rodríguez Faba, R.H. Zakri

© European Association of Urology 2021


TABLE OF CONTENTS PAGE
Inhoud
1. INTRODUCTION 4
1.1 Aim and objectives 4
1.2 Panel Composition 4
1.3 Available publications 4
1.4 Publication history 4

2. METHODS 4
2.1 Introduction 4
2.2 Review and future goals 5

3. THE GUIDELINE 5
3.1 Organ retrieval and transplantation surgery 5
3.1.1 Living-donor nephrectomy 5
3.1.2 Organ preservation 6
3.1.2.1 Kidney storage solutions and cold storage 6
3.1.2.2 Duration of organ preservation 7
3.1.2.3 Methods of kidney preservation: static and dynamic preservation 7
3.1.3 Donor Kidney biopsies 9
3.1.3.1 Procurement Biopsies 9
3.1.3.1.1 Background and prognostic value 9
3.1.3.2 Type and size of biopsy 10
3.1.3.3 Summary of evidence and recommendations 10
3.1.3.4 Implantation biopsies 10
3.1.4 Living and deceased donor implantation surgery 11
3.1.4.1 Anaesthetic and peri-operative aspects 11
3.1.4.2 Immediate pre-op haemodialysis 11
3.1.4.3 Operating on patients taking anti-platelet and anti-coagulation agents 11
3.1.4.4 What measures should be taken to prevent venous thrombosis
including deep vein thrombosis during and after renal transplant? 11
3.1.4.5 Is there a role for peri-operative antibiotics in renal transplantation? 12
3.1.4.6 Is there a role for specific fluid regimes during renal transplantation
and central venous pressure measurement in kidney transplant
recipients? 12
3.1.4.7 Is there a role for dopaminergic drugs, furosemide or mannitol in
renal transplantation? 13
3.1.5 Surgical approaches for first, second, third and further transplants 13
3.1.5.1 Single kidney transplant - living and deceased donors 14
3.1.5.2 Robot-assisted kidney transplant surgery 16
3.1.5.3 Dual kidney transplants 16
3.1.5.4 Ureteric implantation in normal urinary tract 16
3.1.5.5 Transplantation/ureteric implantation in abnormal urogenital tract 18
3.1.6 Donor complications 18
3.1.6.1 Long-term complications 18
3.1.7 Recipient complications 19
3.1.7.1 General complications 19
3.1.7.2 Haemorrhage 19
3.1.7.3 Arterial thrombosis 19
3.1.7.4 Venous thrombosis 20
3.1.7.5 Transplant renal artery stenosis. 20
3.1.7.6 Arteriovenous fistulae and pseudo-aneurysms after renal biopsy 21
3.1.7.7 Lymphocele 21
3.1.7.8 Urinary leak 21
3.1.7.9 Ureteral stenosis 22
3.1.7.10 Haematuria 22
3.1.7.11 Reflux and acute pyelonephritis 22
3.1.7.12 Kidney stones 23
3.1.7.13 Wound infection 23

2 RENAL TRANSPLANTATION - TEXT UPDATE 2021


3.1.7.14 Incisional hernia 23
3.1.8 Urological malignancy and renal transplantation 23
3.1.8.1 Malignancy prior to renal transplantation 23
3.1.8.1.1 In the recipient 23
3.1.8.1.2 In the potential donor kidney 24
3.1.8.2 Malignancy after renal transplantation 25
3.1.9 Matching of donors and recipients 25
3.1.10 Immunosuppression after kidney transplantation 26
3.1.10.1 Calcineurin inhibitors 27
3.1.10.2 Mycophenolates 28
3.1.10.3 Azathioprine 29
3.1.10.4 Steroids 29
3.1.10.5 Inhibitors of the mammalian target of rapamycin 29
3.1.10.6 Induction with Interleukin-2 receptor antibodies 30
3.1.10.7 T-cell depleting induction therapy 31
3.1.10.8 Belatacept 31
3.1.11 Immunological complications 31
3.1.11.1 Hyper-acute rejection 32
3.1.11.2 Treatment of T-cell mediated acute rejection 33
3.1.11.3 Treatment of antibody mediated rejection (ABMR) 33
3.1.12 Follow-up after transplantation 34
3.1.12.1 Chronic allograft dysfunction/interstitial fibrosis and tubular atrophy 34

4. REFERENCES 35

5. CONFLICT OF INTEREST 55

6. CITATION INFORMATION 55

RENAL TRANSPLANTATION - TEXT UPDATE 2021 3


1. INTRODUCTION
1.1 Aim and objectives
The European Association of Urology (EAU) Renal Transplantation Guidelines aim to provide a comprehensive
overview of the medical and technical aspects relating to renal transplantation. It must be emphasised that
clinical guidelines present the best evidence available to the experts but following guideline recommendations
will not necessarily result in the best outcome. Guidelines can never replace clinical expertise when making
treatment decisions for individual patients, but rather help to focus decisions - also taking personal values and
preferences/individual circumstances of patients into account. Guidelines are not mandates and do not purport
to be a legal standard of care.

1.2 Panel Composition


The EAU Renal Transplantation Guidelines panel consists of an international multidisciplinary group
of urological surgeons, a nephrologist and a pathologist. All experts involved in the production of this
document have submitted potential conflict of interest statements which can be viewed on the EAU website:
http://www.uroweb.org/guideline/renal-transplantation/.

1.3 Available publications


A quick reference document, the Pocket Guidelines, is available in print and as an app for iOS and Android
devices. These are abridged versions which may require consultation together with the full text version. All are
available through the EAU website: http://www.uroweb.org/guideline/renal-transplantation/.

1.4 Publication history


The EAU published the first Renal Transplantation Guidelines in 2003 with updates in 2004 and 2009. A
comprehensive update of the 2009 document was published in 2017. This document is a full update of the
2017 Renal Transplantation Guidelines.

2. METHODS
2.1 Introduction
For the 2021 Renal Transplantation Guidelines, new and relevant evidence was identified, collated and
appraised through a structured assessment of the literature. Broad and comprehensive literature searches,
covering the Renal Transplantation Guidelines were performed, covering a time frame between May 31st
2018 and 1st April 2020. A total of 1,202 unique records were identified, retrieved and screened for relevance.
Databases searched included Medline, EMBASE, and the Cochrane Libraries. Detailed search strategies are
available online: http://www.uroweb.org/guideline/renal-transplantation/.

For each recommendation within the guidelines there is an accompanying online strength rating form, the
bases of which is a modified GRADE methodology [1, 2]. Each strength rating form addresses a number of key
elements namely:
1. the overall quality of the evidence which exists for the recommendation, references used in
this text are graded according to a classification system modified from the Oxford Centre for
Evidence-Based Medicine Levels of Evidence [3];
2. the magnitude of the effect (individual or combined effects);
3. the certainty of the results (precision, consistency, heterogeneity and other statistical or
study related factors);
4. the balance between desirable and undesirable outcomes;
5. the impact of patient values and preferences on the intervention;
6. the certainty of those patient values and preferences.

These key elements are the basis which panels use to define the strength rating of each recommendation.
The strength of each recommendation is represented by the words ‘strong’ or ‘weak’ [4]. The strength of each
recommendation is determined by the balance between desirable and undesirable consequences of alternative
management strategies, the quality of the evidence (including certainty of estimates), and nature and variability
of patient values and preferences.

4 RENAL TRANSPLANTATION - TEXT UPDATE 2021


Additional information can be found in the general Methodology section of this print, and online at
the EAU website; http://www.uroweb.org/guideline/. A list of associations endorsing the EAU Guidelines can
also be viewed online at the above address.

2.2 Review and future goals


This document was subject to independent peer review prior to publication in 2017. Publications ensuing from
systematic reviews have all been peer reviewed.

The results of ongoing systematic reviews will be included in the 2022 update of the Renal Transplantation
Guidelines. Ongoing systematic reviews include:
1. What is the best treatment for symptomatic obstructive benign prostatic enlargement in renal
transplantation patients [5]?
2. For patients with kidney graft stones, does surgical treatment provide better stone free rates
than external shock wave lithotripsy [6]?

3. THE GUIDELINE
3.1 Organ retrieval and transplantation surgery
3.1.1 Living-donor nephrectomy
The endoscopic (laparoscopic) approach is the preferred technique for living-donor nephrectomy in established
kidney transplant programmes [7]. Nevertheless, open surgery, preferably by a mini-incision approach, can still
be considered a valid option, despite increased pain in the post-operative period [8].

Endoscopic living-donor nephrectomy (ELDN) includes:


• Pure or hand-assisted transperitoneal laparoscopy;
• Pure or hand-assisted retroperitoneal approach;
• Laparo-Endoscopic Single Site Surgery (LESS);
• Natural Orifice Transluminal Endoscopic Surgery-assisted (NOTES);
• Laparo-Endoscopic Single Site Surgery and robotic-assisted transperitoneal or retroperitoneal approach.

There is strong evidence in support of laparoscopic living-donor nephrectomy (LLDN), including several
systematic reviews and meta-analysis, which have compared its safety and efficacy to open donor
nephrectomy. Laparoscopic living-donor nephrectomy is associated with similar rates of graft function and
rejection, urological complications and patient and graft survival. However, measures related to analgesic
requirements, pain, hospital stay, and time to return to work are significantly better for laparoscopic procedures
[9-12].
Standard LLDN is usually done through 5 and 12 mm ports, but has also been done with 3 or
3.5 mm ports [13]. According to a recent meta-analysis, hand-assisted LLDN is associated with shorter
operative time and warm ischaemia, but equivalent safety and overall results [14]. Laparoscopic living-donor
nephrectomy can also be performed with robotic assistance, with equivalent results according to a systematic
review [15]. However, the numbers are still low and recent studies, including a meta-analysis, have reported
higher complication rates for this approach [16, 17].

Laparo-endoscopic single site surgery nephrectomy allows the surgeon to work through a single incision
(usually the umbilicus) with a multi-entry port. The same or a separate incision is then used for kidney
withdrawal. Several retrospective and at least three prospective randomised trials demonstrated equivalent
safety and results, with a trend towards less pain and better cosmetic results [18, 19]. However, LESS is
considered a more technically demanding procedure when compared with classic LLDN and its role is yet to be
defined.
Natural orifice transluminal endoscopic surgery-assisted transvaginal nephrectomy avoids the
abdominal incision needed for kidney extraction, aimed at minimising scaring and pain. Initial reports suggest
that this approach is safe, however experience with this technique is still highly limited [20].
Right LLDN has been considered more difficult, yielding inferior results. However, both left and right
LLDN can be performed with equivalent safety and efficacy according to large retrospective studies, systematic
reviews and meta-analysis [21, 22].

RENAL TRANSPLANTATION - TEXT UPDATE 2021 5


Laparoscopic living-donor nephrectomy has brought attention to potential failures of different devices such
as, endoscopic staplers and locking and non-locking clips, used to secure the renal hilum [21]. There is no
scientific evidence that one device is safer than another for securing the renal artery [23-25]. However, the U.S.
Food and Drug Administration (FDA) and the manufacturers of locking clips have issued a contraindication
against their use in securing the artery during LLDN.

Summary of evidence LE
Laparoscopic living-donor nephrectomy is associated with similar rates of graft function and rejection, 1a
urological complications and patient and graft survival to open nephrectomy.
Measures related to analgesic requirements, pain, hospital stay, and time to return to work are 1a
significantly better for laparoscopic procedures.

Recommendations Strength rating


Offer pure or hand-assisted laparoscopic/retroperitoneoscopic surgery as the preferential Strong
technique for living-donor nephrectomy.
Perform open living-donor nephrectomy in centres where endoscopic techniques are not Strong
implemented.
Perform laparo-endoscopic single site surgery, robotic and natural orifice transluminal Strong
endoscopic surgery-assisted living-donor nephrectomy in highly-specialised centres only.

3.1.2 Organ preservation


In kidneys donated after cardiac death (DCD) evidence suggests that warm ischemia contributes to worse
graft outcome. Donor haemodynamic parameters (systolic blood pressure, oxygen saturation and shock index:
heart rate divided by systolic blood pressure) may be predictors of delayed graft function (DGF) and graft
failure; however, further studies are required to validate this [26]. The duration of asystolic warm ischaemia
during procurement in DCD donors is associated with increased risk of graft failure. Overall five year graft
failure (including primary graft non-function) was associated with longer asystolic warm ischaemia times
[27]. Extraction time (beginning with aortic cross-clamp and ending with placement of the kidneys on ice), is
an important factor for DGF. Incidents of DGF were 27.8% and 60% at up to 60 minutes and 120 minutes
extraction time, respectively [28].
A retrospective study of 64,024 living donor kidney transplants found that cold ischaemia time (CIT),
human leukocyte antigen (HLA) mismatch, donor age, panel reactive antibody, recipient diabetes, donor and
recipient body mass index (BMI), recipient race and gender, right nephrectomy, open nephrectomy, dialysis
status, ABO incompatibility, and previous transplants were independent predictors of DGF in living donor
kidney transplants [29]. Five-year graft survival among living donor kidney transplant recipients with DGF was
significantly lower than in those without DGF. Delayed graft function increased the risk of graft failure by more
than two-fold [29].

3.1.2.1 Kidney storage solutions and cold storage


There are two main sources for kidney graft injury: ischaemia (warm and cold), and reperfusion injury. The
aims of modern kidney storage solutions include: control of cell-swelling during hypothermic ischaemia;
maintenance of intra- and extra-cellular electrolyte gradient during ischaemia; buffering of acidosis; provision
of energy reserve; and minimisation of oxidative reperfusion injury. There is no agreement on which of the
mechanisms is most important for post-ischaemic renal graft function [30]. No storage solution seems to
combine all mechanisms. Previously, Euro-Collins was widely used, but is no longer recommended.

Presently, University of Wisconsin (UW), and histidine-tryptophan-ketoglutarate (HTK) solution are equally
effective and are standard for multi-organ or single kidney harvesting procedures. The characteristics of HTK
are its low viscosity, low potassium concentration and low cost. University of Wisconsin solution has been
the standard static cold preservation solution for the procurement of liver, kidney, pancreas, and intestine
[31]. University of Wisconsin, HTK, and Celsior solutions have provided similar allograft outcomes in most
clinical trials; however, some differences have become apparent in recent studies and registry reports [32,
33]. Marshall’s hypertonic citrate solution (MHCS) is also suitable for use in the preservation of human kidneys
before transplantation [34]. In experimental studies of kidney preservation, HTK and UW retained a greater
capacity to preserve endothelial structure and pH buffering function during warm ischaemia in comparison
to MHCS and Celsior, especially in uncontrolled DCD donors [35]. In the absence of a cost-utility analysis,
the results of the meta-analysis from the randomised controlled trials (RCTs) comparing UW with Celsior and
MHSC in standard cadaver donors, indicate that these cold storage solutions are equivalent [36].

6 RENAL TRANSPLANTATION - TEXT UPDATE 2021


For living donors, in whom immediate kidney transplantation is planned, perfusion with crystalloid solution
is sufficient. Kidneys coming from DCD donors, especially those uncontrolled, are high-risk marginal organs
due to prolonged warm ischaemia periods, and require specific measures in order to diminish the rate of non-
function or DGF. More than 60% of kidney grafts currently come from Expanded Criteria Donors (ECD) (any
donor aged > 65 years and/or donor aged > 55 years with any of the following: acute renal dysfunction, stroke
or arterial hypertension) [37].

Summary of evidence LE
University of Wisconsin and HTK solution are equally effective and are standard for multi-organ or 1b
single kidney harvesting procedures.
A meta-analysis of RCTs indicated that UW and Celsior solution are equivalent in standard cadaver 1a
donors.

Recommendations Strength rating


Use either University of Wisconsin or histidine tryptophane ketoglutarate preservation Strong
solutions for cold storage.
Use Celsior or Marshall’s solution for cold storage if University of Wisconsin or histidine Strong
tryptophane ketoglutarate solutions are not available.

3.1.2.2 Duration of organ preservation


Cold ischaemia time should be as short as possible. Kidneys from ECDs after brain death (DBD) and DCD
donors are more sensitive to ischaemia than standard criteria donors. Kidneys from DBD donors should ideally
be transplanted within a 18 to 21 hour time period; there is no significant influence on graft survival within a
18 hour CIT [36, 38, 39]. Kidneys from DCD donors should ideally be transplanted within 12 hours [40], whilst
kidneys from ECDs should ideally be transplanted within 12 to 15 hours [41, 42].

3.1.2.3 Methods of kidney preservation: static and dynamic preservation


Whichever method is used, cold storage is critical. The use of cold preservation as a therapeutic window
to deliver pharmacological or gene therapy treatments could, from an investigational point of view, improve
both short- and long-term graft outcomes [43]. Cooling reduces the metabolic rate of biological tissue
minimising continuous cellular processes that lead to depletion of ATP and accumulation of metabolic
products. Reperfusion with oxygenated blood invokes ischaemia-reperfusion injury. Hypothermic perfusion
does not enable normal cellular metabolic function or prevent depletion of energy stores [44]; however, it
prevents the deleterious effects of simple cooling, especially in the setting of prolonged warm-ischaemic time
in uncontrolled DCD donors. Two meta-analyses suggest that hypothermic machine perfusion (HMP) reduces
DGF compared with static cold storage [45, 46]. Outcomes for primary non-function (PNF) are less clear, but
one meta-analysis limited to high quality studies suggests a reduction in PNF rates with HMP [46]. A Cochrane
systematic review and meta-analysis showed that HMP reduced the risk of DGF when compared to static cold
storage (CS) for kidneys from both DCD and DBD donors [47].

The increased demand for organs has led to the increased use of “higher risk” kidney grafts. Kidneys from
DCD donors or grafts coming from ECDs are more susceptible to preservation injury and have a higher risk of
unfavourable outcomes [48, 49].
Dynamic, instead of static, preservation could allow for organ optimisation, offering a platform
for viability assessment, active organ repair and resuscitation. Ex situ machine perfusion and in situ regional
perfusion in the donor are emerging as potential tools to preserve vulnerable grafts. Preclinical findings
have driven clinical organ preservation research that investigates dynamic preservation, in various modes
(continuous, pre-implantation) and temperatures (hypo-, sub-, or normothermic) [44].

There are several methods of kidney preservation including:


• Initial flushing with cold preservation solution followed by ice storage. However, the limitations of static
CS in preserving marginal organs such as ECD kidneys has led to the increased use of dynamic methods.
• Current dynamic preservation strategies entering clinical practice and the different modalities of their
use are: HMP, hypothermic regional perfusion, normothermic machine perfusion, normothermic regional
perfusion, sub-normothermic machine perfusion and sub-normothermic regional perfusion [44].
• Continuous pulsatile HMP seems to be a good preservation method for marginal organs, either initially or
after a period of simple CS (shipping of suboptimal kidneys) [50].

RENAL TRANSPLANTATION - TEXT UPDATE 2021 7


• Some evidence shows that hypothermic dynamic preservation should be controlled by pressure and not
flow, using low pressures to avoid pressure-related injury. The perfusion solutions used are specific, and
are qualitatively different to CS solutions [33].
• Nonoxygenated HMP of the kidney at low perfusion pressures (20-30 mmHg) has been shown to reduce
DGF [45]. The largest RCT comparing simple CS with HMP of deceased donor kidneys showed an overall
reduced risk of DGF and a survival benefit, most pronounced in ECD kidneys [51]. Hypothermic machine
perfusion of kidneys from type III DCD donors decreased DGF with no impact on graft survival [48].
• Hypothermic machine-perfusion reduces the risk of DGF in standard criteria DBD donor kidneys
regardless of CIT [52].
• Increased vascular resistance and high perfusate injury marker concentrations are risk factors for DGF;
however, they do not justify discarding the kidney. The flow perfusion value seems to be an indicator of
graft viability in uncontrolled DCD donors, particularly donors with a high creatinine level [53]. However,
research is required to identify a strong and reliable measure for predicting kidney viability from machine
perfusion [36]. Perfusion parameters (renal flow and renal vascular resistance) have low predictive values
and should not be used as the sole criterion to assess viability of kidney grafts [54].
• The effect of oxygenated HMP was investigated in an RCT initiated by the Consortium on Organ
Preservation in Europe on type III DCD kidneys and ECD kidneys [44]. Graft loss was significantly lower
after oxygenated HMP compared to HMP [55]. No significant differences between the two groups were
shown for DGF, PNF and patient death. No difference in eGFR at one year was observed between HMPO
vs. HMP; however, sensitivity analysis, accounting for all-cause graft failure, showed a higher eGFR in
oxygenated HMP [55].
• A short period of normothermic machine perfusion (NMP) immediately prior to implantation has been
shown to improve kidney graft function, replenish ATP and reduce injury in experimental models [56, 57].
• A retrospective study of normothermic regional perfusion (NRP) in uncontrolled DCD donors concluded
that NRP appears to decrease graft failure when used as a preconditioning technique with subsequent
HMP preservation in these donors [58].
• Active research is being developed on preservation of prolonged warm-ischaemically damaged human
kidneys (types I and II DCD) by in situ normothermic extracorporal hemoperfusion with oxygenation and
leukocyte depletion before procurement [59]. Oxygen carriage is achieved by using blood depleted of
leukocytes. Potential advantages of this preservation technique are reduction in ischaemia-reperfusion
injury as well as the possibility of assessing organ viability.
• Currently there is one registered ongoing RCT on pre-implantation NMP using an oxygenated,
sanguineous normothermic perfusion solution (http://www.isrctn.com/ISRCTN15821205). However,
kidney function can be evaluated during NMP by assessing macroscopic appearance of blood perfusion,
renal blood flow and urine output [60].
• Continuous subnormothermic machine perfusion and controlled oxygenated rewarming has
demonstrated improved creatinine clearance and preservation of structural integrity compared with
continuous oxygenated HMP in a research setting [61].

Summary of evidence LE
A meta-analysis of RCTs comparing CS with HMP of deceased donor kidneys showed a reduced risk 1a
of DGF for HMP.
Hypothermic dynamic preservation should be controlled by pressure and not flow, using low pressures 2a
to avoid pressure-related injury.
Perfusion parameters (renal flow and renal vascular resistance) have low predictive values and should 2b
not be used as the sole criterion to assess viability of kidney grafts.

Recommendations Strength rating


Minimise ischaemia times. Strong
Use hypothermic machine-perfusion (where available) in deceased donor kidneys to reduce Strong
delayed graft function.
Hypothermic machine-perfusion may be used in standard criteria deceased donor kidneys. Strong
Use low pressure values in hypothermic machine perfusion preservation. Strong
Hypothermic machine-perfusion must be continuous and controlled by pressure and not Strong
flow.
Do not discard grafts based only on increased vascular resistance and high perfusate injury Weak
marker concentrations during hypothermic machine perfusion preservation.

8 RENAL TRANSPLANTATION - TEXT UPDATE 2021


3.1.3 Donor Kidney biopsies
Donor kidney biopsies can serve different purposes including:
• histological assessment of organ quality prior to transplantation (often referred to as procurement or
harvest biopsies);
• histological analysis of focal lesions, especially if there is a suspicion of neoplasia;
• detection of donor derived lesions as reference for subsequent post-transplant biopsies (often referred to
as baseline, zero-time or implantation biopsies).

3.1.3.1 Procurement Biopsies


3.1.3.1.1 Background and prognostic value
Procurement biopsies are used for the detection of tissue injury to aid the decision of whether or not a
deceased donor kidney is suitable for transplantation. These biopsies are most commonly performed in donors
with clinical suspicion of chronic kidney injury (ECDs) [62].
Kidney discard in Europe is rarely based on histology findings, as procurement biopsies are not
regularly performed for graft allocation in the Eurotransplant region [62]. However, since biopsy findings are
the most frequent cause for discarding donor organs in the United States [63-65], their prognostic value
has been analysed in numerous studies. A recently published systematic review of studies on donor kidney
biopsies revealed a lack of prospective studies and marked heterogeneity regarding the type of lesions being
assessed, their scoring, the definitions of post-transplant outcomes and the statistical methods employed [66].
Therefore, the published evidence suggests that the use of procurement biopsies for deciding on suitability for
transplantation of donor kidneys may have some important limitations including the following [62, 66, 67]:

• There is no consistent association between histological lesions observed in donor kidney biopsies and
post-transplant outcomes.
The concept of procurement biopsies in elderly donors was introduced by a study from Gaber et al., in 1995.
This study observed significantly worse outcomes in recipients of kidneys with > 20% globally sclerotic
glomeruli [68]. However, subsequent studies yielded highly variable results and it cannot be concluded that
glomerulosclerosis is independently associated with graft outcomes [66]. A similar variability was also observed
for other potentially relevant lesions like arterial injury, interstitial fibrosis and tubular atrophy with each showing
predictive value in some studies, but not in others [66].

• There is no agreement on prognostically relevant lesions and how they should be scored.
Specific grading systems for donor kidney biopsies have not yet been developed. Lesion scoring in pre-
transplant biopsies is mostly based on the Banff consensus for post-transplant renal allograft pathology, which
is supported by the 2007 Banff Conference report [69].
Many attempts have been made to use composite semi-quantitative scoring systems to express
the global extent of tissue injury in donor kidney biopsies. These scoring systems are mostly based on simple
addition of the Banff scores for individual lesions, most commonly glomerulosclerosis, arteriolar hyalinosis,
arterial intimal fibrosis, interstitial fibrosis and tubular atrophy and rarely include clinical parameters like donor
age [70], serum creatinine values and donor hypertension [71].
A limited number of histological scoring systems are based on modelling analysis [70-74]. Only the
Maryland Aggregate Pathology Index (MAPI) [74] scoring system and the Leuven donor risk score [70], use
graft failure as their endpoint and have been independently validated in a second cohort. Other studies used
surrogate clinical endpoints like DGF [72] and estimated glomerular filtration rate (eGFR) at three months [73]
to calculate histological models. In addition, these models were not validated in independent cohorts. The
variation in how the components are weighted to achieve the composite score and the different endpoints used
may explain the conflicting conclusions in the literature [62, 66, 67].

• Due to the time constraints of organ allocation procurement biopsies are mostly read on frozen sections
by on-call pathologists, which might affect the diagnostic reliability of reported findings.
This may have substantial impact on the diagnostic reliability of the procedure since frozen sections are prone
to morphological artefacts that can impair the detection and scoring of potentially important lesions such as
arteriolar hyalinosis and interstitial fibrosis [75, 76]. There is strong evidence that dedicated renal pathologists
should examine formalin-fixed paraffin-embedded core-needle biopsies. Paraffin histology employing special
stains is technically superior to frozen sections since morphological details are better preserved on paraffin
sections than on frozen sections and potentially confounding artefacts can be avoided. Rapid processing of
tissue for paraffin histology is technically feasible, but the respective protocols are not universally implemented
and are not available on a 24/7 basis in most departments. Another source of variability is the professional
experience of the pathologist in charge. Procurement biopsies are commonly read by the on-call general

RENAL TRANSPLANTATION - TEXT UPDATE 2021 9


pathologist who frequently has no specific training in renal pathology. A recent study specifically addressing
this issue found that the on-call pathologists tended to overestimate chronic injury in biopsies [77].

3.1.3.2 Type and size of biopsy


Many transplant centres obtain wedge biopsies of donor kidneys rather than needle biopsies due to the
presumed higher risk of bleeding complications with the latter. Wedge biopsies sample the cortex superficially
whereas needle biopsies reach deeper aspects of the cortex. Needle biopsies also allow sampling from
different areas of the kidney. Submit 14 or 16 G needle biopsies as obtaining adequate biopsies with 18 G
needles requires multiple cores [78]. Several studies comparing wedge with needle biopsies concluded that
needle biopsies perform much better in the evaluation of vascular lesions because interlobular arteries are
rarely sampled in wedge biopsies. Both methods were comparable for glomerular or tubulointerstitial lesions
[79-82]. It was also demonstrated that glomerulosclerosis is significantly more pronounced in the subcapsular
zone compared with deeper areas of the cortex [83]. The problem of insufficient sampling of arteries and over
representation of (subcapsular) glomerular scars in wedge biopsies, can only be avoided if particular attention
is paid to the correct performance of the biopsy, with a minimal depth of 5 mm [84]. The predictive value of
glomerulosclerosis increases significantly with higher numbers of glomeruli in the wedge biopsy, with ideally
at least 25 glomeruli required for evaluation [81]. There is limited evidence regarding complication rates in pre-
implantation biopsies.
Use of a skin punch biopsy device might be an attractive alternative. Skin punch biopsies measure
3 mm in diameter. They have a shorter length than needle biopsies therefore avoiding injury to large calibre
arteries at the corticomedullary junction whilst still sampling tissue from deeper areas of the cortex [85].

3.1.3.3 Summary of evidence and recommendations

Summary of evidence LE
Individual histologic lesions like glomerulosclerosis, arterial luminal narrowing or tubulointerstitial injury 3
observed in donor kidney biopsies have limited prognostic value for long-term allograft survival.
Composite histological scoring systems provide a more comprehensive measure of overall organ 3
damage. However, published scoring systems still lack independent validation and robust thresholds.
Size of the biopsy is of critical importance for its diagnostic value. An adequate biopsy reaches 3
beyond the immediate subcapsular area (≥ 5 mm) and contains ≥ 25 glomeruli and ≥ one artery.
Needle biopsies, wedge biopsies or specimens obtained with a skin punch biopsy device will result in
equally adequate biopsies if sampling is properly performed. Obtaining adequate biopsies with 18 G
needles is difficult and requires multiple cores.

Recommendations Strength rating


Do not base decisions on the acceptance of a donor organ on histological findings alone, Strong
since this might lead to an unnecessary high rate of discarded grafts. Interpret histology
in context with clinical parameters of donor and recipient including perfusion parameters
where available.
Use paraffin histology for histomorphology as it is superior to frozen sections; however, its Strong
diagnostic value has to be balanced against a potential delay of transplantation.
Procurement biopsies should be read by a renal pathologist or a general pathologist with Strong
specific training in kidney pathology.

3.1.3.4 Implantation biopsies


Implantation biopsies are used to provide baseline information on donor kidney injury for comparison
with subsequent post-transplant kidney biopsies. Baseline biopsies can be essential for clear distinction
between pre-existing damage and acquired lesions. They are particularly valuable in cases of thrombotic
microangiopathy, arteriolar hyalinosis or acute tubular injury. In contrast to procurement biopsies that are
obtained at the time of organ harvesting, implantation biopsies are usually taken before implantation in order
to cover potential effects of CIT. Their diagnostic contribution has not been formally quantified in the literature
which might be due to the difficulties of measuring the value of implantation biopsies for improving diagnoses.
Despite the lack of formal studies investigating their value it seems very reasonable to perform implantation
biopsies in deceased donor kidneys.

10 RENAL TRANSPLANTATION - TEXT UPDATE 2021


3.1.4 Living and deceased donor implantation surgery
3.1.4.1 Anaesthetic and peri-operative aspects
Good communication between nephrologists, anaesthetists and surgeons is required for optimal anaesthetic
and peri-operative care of the renal transplant patient. Anaesthetic care of the living kidney donor [86] and
renal transplant recipient [87] have been reviewed and recent guidelines from the European Renal Association-
European Dialysis and Transplantation Association (ERA-EDTA) [88] are cross referenced.

3.1.4.2 Immediate pre-op haemodialysis


Routine use of haemodialysis immediately prior to renal transplantation is not indicated [88]. Hyperkalaemia
is the most common indication for haemodialysis pre-operatively. The risks of haemodialysis compared with
medical therapy must be considered along with the risks of intra-operative fluid overload, electrolyte and acid-
base disturbances, particularly where a deceased donor kidney is transplanted with a significant risk of DGF.
Pre-operative haemodialysis may initiate a pro-inflammatory state, delay surgery, increase the CIT and increase
the risk of DGF [89].

Summary of evidence LE
Pre-operative haemodialysis has the potential to delay transplantation, increase CIT and increase the 2
risk of DGF.

Recommendation Strength rating


Use dialysis or conservative measures to manage fluid and electrolyte imbalance prior to Weak
transplant surgery taking into consideration the likelihood of immediate graft function.

3.1.4.3 Operating on patients taking anti-platelet and anti-coagulation agents


Many patients active on the transplant waiting list have vascular disease and/or a pro-thrombotic condition that
should be risk-assessed prior to transplantation. Dual anti-platelet therapy is commonly given to patients with
coronary artery stents for six to twelve months; peri-operative management plans for these patients should be
discussed with a cardiologist so that the risks of withdrawal of the anti-platelet agent can be fully considered.
Options for reversal of anti-coagulation and post-operative anti-coagulation should be discussed with a
haematologist prior to patient listing.

Some patients will be active on a transplant waiting list whilst continuing to take anti-platelet and/or anti-
coagulation agents. The indication for anti-platelet or anti-coagulation agents should be clearly documented
for each individual. Potential increased risk of peri-operative bleeding needs to be weighed against potential
harm from arterial or venous thrombosis. In accordance with the American College of Chest Physicians and
the European Society of Cardiology guidelines [90, 91], the literature suggests that continuing anti-platelet
therapy with aspirin, ticlopidine or clopidogrel does not confer a significantly greater risk of peri/post-operative
complications [92], however, the number of patients studied was low. If needed, the effect of anti-platelet
agents can be reduced with intra-operative platelet infusions.

Summary of evidence LE
A retrospective single-centre case-control study in patients undergoing kidney transplantation 3
concluded that continuing anti-platelet therapy with aspirin, ticlopidine or clopidogrel does not confer
a significantly greater risk of peri/post-operative complications.

Recommendations Strength rating


Consider continuing anti-platelet therapy in patients on the transplant waiting list. Weak
Discuss patients who take anti-platelet and anti-coagulation agents prior to transplant Weak
surgery with relevant cardiologist/haematologist/nephrologist.

3.1.4.4 What measures should be taken to prevent venous thrombosis including deep vein thrombosis
during and after renal transplant?
Peri-operative administration of short-acting anti-coagulation agents reduces peri-operative risk of venous
thrombosis (including in ileo-femoral and renal veins); however, due to associated increased blood loss
administration requires knowledge of individual patient risk factors. None of the current major thrombosis
prevention guidelines directly address thromboprophylaxis in the renal transplant peri-operative period. A small

RENAL TRANSPLANTATION - TEXT UPDATE 2021 11


RCT [93] showed no difference in early post-operative graft loss or thromboembolic complications with or
without prophylactic anti-coagulation. Those administered prophylactic anti-coagulation had significantly lower
haemoglobin whilst those administered prophylactic unfractionated heparin had prolonged lymph drainage.
Based on this study, routine pharmacological prophylaxis is not recommended in low-risk living donor
recipients. Mechanical measures to decrease ileo-femoral deep vein thrombosis (DVT) can be used where there
is no contraindication due to peripheral vascular disease particularly where there are concerns about bleeding
risks with pharmacological prophylaxis.

Summary of evidence LE
A small RCT (n=75) showed no difference in early post-operative graft loss or thromboembolic 1b
complications with or without prophylactic anti-coagulation.

Recommendation Strength rating


Do not routinely give post-operative prophylactic unfractionated or low-molecular-weight Weak
heparin to low-risk living donor transplant recipients.

3.1.4.5 Is there a role for peri-operative antibiotics in renal transplantation?


Prophylactic peri-operative antibiotics are generally used in renal transplant surgery but the optimal antibiotic
regimen is not known and increasing antibiotic resistance may hamper their effectiveness in this setting.
A multicentre, prospective RCT showed no difference at one month in surgical site, bacterial, fungal or
viral infection between those receiving a single dose broad spectrum antibiotic at induction of anaesthesia
compared to those receiving antibiotic 12 hourly for 3-5 days [94]. A retrospective comparison of peri-operative
intravenous cefazolin prophylaxis compared to no antibiotic showed no difference in infectious complications
(surgical site, urinary tract, bacteraemia or central catheter-related infection) in the first month after renal
transplantation [95].

Summary of evidence LE
A multicentre, prospective RCT showed that the incidents of surgical site infection and urinary tract 1b
infection were similar in those receiving a single dose broad spectrum antibiotic at induction of
anaesthesia and those receiving antibiotics 12 hourly for 3-5 days.

Recommendation Strength rating


Use single-dose, rather than multi-dose, peri-operative prophylactic antibiotics in routine Strong
renal transplant recipients.

3.1.4.6 Is there a role for specific fluid regimes during renal transplantation and central venous pressure
measurement in kidney transplant recipients?
Careful peri- and post-operative fluid balance is essential for optimal renal graft function. There is no evidence
determining if crystalloids or colloids are better for intravenous fluid management during renal transplant
surgery, however colloids may be immunogenic. If normal saline (0.9%) is used, monitoring for metabolic
acidosis is recommended in the peri-operative period. A prospective double-blind RCT compared normal
saline to lactated Ringer’s solution as intra-operative intravenous fluid therapy. Serum creatinine at day three
post-surgery did not differ between the two groups. However, Ringer’s lactate caused less hyperkalaemia
and metabolic acidosis than normal saline. Balanced solutions may be the optimal and safer option for intra-
operative intravenous fluid therapy [96].

Central venous pressure (CVP) measurement helps anaesthetists guide fluid management. A small prospective
non-blinded RCT compared two normal (0.9%) saline regimens: constant infusion (10-12 mL/kg-1/h-1 from start
of surgery until reperfusion) and CVP-based infusion (target CVP appropriate to stage of operation) [97]. Central
venous pressure directed infusion produced a more stable haemodynamic profile, better diuresis and early
graft function. Directed hydration may decrease DGF rates and CVP measurement may help optimise early
graft function.

12 RENAL TRANSPLANTATION - TEXT UPDATE 2021


Summary of evidence LE
A small (n=51) prospective RCT found that use of Ringer’s lactate solution was associated with 1b
less hyperkalaemia and acidosis compared with normal saline in patients undergoing kidney
transplantation.
A small (n=40) prospective RCT comparing constant infusion vs. CVP found that CVP produced a 1b
more stable haemodynamic profile, better diuresis and early graft function.

Recommendations Strength rating


Optimise pre-, peri- and post-operative hydration to improve renal graft function. Strong
Use balanced crystalloid solutions for intra-operative intravenous fluid therapy. Weak
Use target directed intra-operative hydration to decrease delayed graft function rates and Strong
optimise early graft function.

3.1.4.7 Is there a role for dopaminergic drugs, furosemide or mannitol in renal transplantation?
Low-dose dopamine (LDD) has been used in renal transplantation due to a perceived improvement in urine
output and early graft function. Use of LDD in kidney donors is outside of the scope of this section. Conflicting
results prevent a consensus statement on routine use of LDD in transplant recipients. A small (n=20) prospective
randomised cross-over study in deceased donor renal transplantation suggested significant improvements in
urine output and creatinine clearance in the first nine hours post-surgery without adverse events [98]. By contrast,
a retrospective comparison of LDD in the first twelve hours post-deceased donor renal transplantation showed
no difference in diuresis or kidney function, but those administered LDD (n=57) had increased heart rates, longer
intensive therapy unit stay and higher six-month mortality than those not treated with LDD (n=48) [99].

Considerable variation exists in the use of diuretics during renal transplant recipient surgery and there is
little evidence to suggest any benefit from their use [100]. No evidence on the use of mannitol during renal
transplant recipient surgery was found during the panel’s literature search. Use of mannitol in kidney donors is
outside the scope of this section.

Summary of evidence LE
A retrospective comparative study of LDD treated vs. non-treated renal transplantation patients 2b
concluded that LDD administration did not improve kidney function in the first twelve hours post renal
transplantation, but did result in increased heart rates, longer intensive therapy unit stay and higher
six-month mortality in those receiving LDD.

Recommendation Strength rating


Do not routinely use low-dose dopaminergic agents in the early post-operative period. Weak

3.1.5 Surgical approaches for first, second, third and further transplants
Transplant (bench/back-table) preparation is a crucial step in the transplantation process. The kidney must
be inspected whilst on a sterile ice slush, removing peri-nephric fat when possible to permit inspection of the
quality of the organ and to exclude exophytic renal tumours. Biopsy of the kidney on the back-table may be
performed to help in the multifactorial decision-making process regarding the quality and usage of the kidney
for both single and/or dual transplantation. Suspicious parenchymal lesions also require biopsy. Techniques for
intra-operative kidney biopsy are discussed in section 3.1.3.

The number, quality and integrity of renal vessels and ureter(s) should be established and lymphatics at the
renal hilum ligated. The quality of the intima of the donor renal artery should be evaluated. Branches of the
renal artery not going to the kidney or ureter(s) should be tied.
In deceased donor kidney transplantation the quality of the aortic patch should be determined.
If severe atheroma of the patch, ostium or distal renal artery is seen then the aortic patch and/or distal renal
artery can be removed to provide a better quality donor renal artery for implantation. Back table reconstruction
of multiple donor arteries is discussed in section 3.1.5.1. The length of the renal vein should be evaluated.
Renal vein branches should be secured/tied.
For a deceased donor right kidney, lengthening the renal vein on the back table may be performed
if needed with donor inferior vena cava (IVC) [101]. Techniques for lengthening a short living donor right renal
vein from donor gonadal vein or recipient saphenous vein require pre-operative planning and specific consent
(discussed in section 3.1.5.1).

RENAL TRANSPLANTATION - TEXT UPDATE 2021 13


The length, quality and number of the ureter(s) should be established. The peri-pelvic and proximal
peri-ureteral tissue in the ‘golden triangle’ should be preserved.

Recommendation Strength rating


Assess the utility (including inspection) of the kidney for transplantation before Strong
commencement of immunosuppression and induction of anaesthesia for deceased donor
kidney transplantation.

3.1.5.1 Single kidney transplant - living and deceased donors


The standard surgical approach for first or second single kidney transplant (SKT) operations remains open
kidney transplant (OKT). Emerging surgical technologies using minimal access surgical approaches have been
developed and the different surgical approaches (minimally invasive open, laparoscopic and robot-assisted)
were compared in a systematic review [102].
An extra-peritoneal approach to either iliac fossa should be used as the operative approach in most
first or second SKT operations. There is no evidence to prefer placement of a left or right kidney into either
iliac fossa [103]. Peri-iliac vessel lymphatics should be ligated to try and prevent post-operative lymphocele.
Appropriate segments of iliac artery and vein should be mobilised to facilitate appropriate tension-free vascular
anastomoses and the final positioning of the transplanted kidney. There is evidence supporting the benefits of
cooling the kidney surface during implantation [104].

Recommendations Strength rating


Choose either iliac fossa for placement of a first or second single kidney transplant. Weak
Ligate peri-iliac vessel lymphatics (lymphostasis) to reduce post-operative lymphocele. Weak

A variety of techniques have been described to help with the anastomosis of a short renal vein. This is most
commonly encountered with a right kidney, especially from a living donor. To achieve equivalent outcomes with
right kidneys appropriate surgical technical manoeuvres may be needed to optimise right kidney implantation.

A number of studies suggest marginally worse outcomes with use of the right compared to left kidney. two
large registry studies demonstrate a slightly higher risk of early graft failure using right compared to left kidneys
from living donors [105-107]. A registry study of 2,450 paired kidneys, donated after cardiac death, observed
with right kidneys: more early surgical complications; an increased risk in DGF (Odds Ratio [OD] 1.46); and
inferior one year graft survival (OD 1.62), but not at subsequent time points [105]. However, surgical techniques
used to compensate for a right kidney, anastomosis time and surgeon experience were not recorded. A recent
registry study of 87,112 deceased-donor kidney recipient pairs reported a modest increase in DGF (adjusted
OD 1.15) and all-cause graft failure (adjusted hazard ratio 1.07), within the first six months, associated with use
of the right kidney, but there was no association with recipient mortality [108]. Furthermore, data from cohort
studies [101, 103] and one registry study [104] suggest equivalent outcomes with either left or right deceased
donor kidneys. Meta-analysis of data from one RCT and fourteen cohort studies suggested equivalent graft
outcomes [109]. Overall, these findings do not support declining an organ for kidney transplantation based on
laterality of kidney offered.

Techniques to manage a short renal vein can be addressed in the donor and/or recipient. Ligation of internal
iliac vein(s) may be necessary to elevate the iliac vein and avoid tension on the renal vein anastomosis [103].
Transposition of the iliac artery and vein may enhance the position for the venous anastomosis [110]. The right
renal vein may be lengthened. With deceased donor kidneys this is usually done with donor IVC [111]. In living
donors, lengthening of the renal vein may be achieved with donor gonadal vein retrieved at donor nephrectomy
[112] or with recipient saphenous vein [113], although both require specific consent and in general the other
aforementioned techniques are preferred.

Summary of evidence LE
Prospective cohort studies demonstrated that: 3
• transposition of the recipient iliac vein is an appropriate technical solution to compensate for the
short length of the renal vein in right kidney LDN (n=43);
• the living donor right kidney renal vein can be successfully lengthened using donor gonadal vein
(n=17) or recipient saphenous vein (n=19).

14 RENAL TRANSPLANTATION - TEXT UPDATE 2021


Recommendation Strength rating
Assess the length of the donor renal vein and if it is short consider one of a variety of Weak
surgical techniques to optimise the venous anastomosis.

A history suggesting previous iliac or femoral vein thrombosis should initiate pre-operative imaging to establish
patency of one iliac vein and the IVC. An intra-operative finding of an unexpected iliac vein and/or vena cava
thrombosis may lead to abandonment of implantation. With pre-operative planning, native renal (orthotopic) or
superior mesenteric vein or gonadal vein collaterals can be used.

The external or common iliac arteries are equally good for arterial anastomosis. The internal iliac artery is more
frequently affected by atherosclerosis than the external or common iliac arteries. End-to-side anastomosis of
donor renal artery to recipient external and/or common iliac artery is recommended in general over an end-to-
end anastomosis to the internal iliac artery. The only RCT comparing these techniques suggests no difference
[114]; however, the study was limited by small numbers and a high (8%) overall renal artery thrombosis rate.
The sites of the vascular anastomosis should be chosen carefully according to the length of the
renal artery and vein to avoid kinking of the vessels when the kidney is placed into its final location, usually
in the iliac fossa. The site of the arterial anastomosis should avoid atheromatous plaques in the iliac artery to
decrease the risk of iliac artery dissection. The intima of the donor and recipient arteries should be checked
prior to commencing the arterial anastomosis to ensure that there is no intimal rupture/flap. If this is found it
must be repaired prior to, or as part of, the arterial anastomosis.
A Carrel patch is usually maintained on a deceased donor renal artery although it can be removed
if there is either severe ostial atheroma/stenosis (with good quality proximal renal artery) or if the length of the
renal artery is too long for the appropriate implantation site on the iliac artery (which is more common with the
right renal artery).

Multiple renal arteries supplying a deceased donor kidney can be maintained on a Carrel patch (of appropriate
length) and implanted as a single anastomosis. In living donor transplantation, multiple renal arteries require
a variety of strategies to achieve optimum re-perfusion [100]. Two arteries can be implanted separately
or to achieve a single anastomosis: a very small second artery (especially if supplying the upper pole)
may be sacrificed; the two arteries may be joined together (as a trouser graft); or the smaller artery can
be anastomosed onto the side of the main artery (end-to-side anastomoses). A lower polar artery may be
re-vascularised via anastomosis to the inferior epigastric artery [115]. In living donor transplantation where
three or more donor arteries exist, consideration should be given to alternate kidney donors. In circumstances
using a living donor kidney with three or more donor arteries, strategies include a combination of the above
techniques or, after appropriate consent, use an explanted (recipient’s own) internal iliac artery graft [116] or
saphenous vein graft [117].

In cases where an iliac artery prosthetic replacement has previously been carried out because of severe
symptomatic iliac atheroma, the renal artery should be implanted into the prosthesis. Administration of
systemic heparin should be considered prior to clamping of a vascular prosthesis [118].

A variety of sutures and suturing techniques for the vascular anastomosis are described, but in general
practice, a 5/0 and 6/0 non-absorbable mono-filament polypropylene suture(s) are used for the renal vein and
renal artery anastomosis. Despite this, there is no evidence to recommend one suturing technique over another
to prevent, for example, transplant artery stenosis. Use of an expanded polytetrafluroethylene suture compared
to standard polypropylene suture may reduce blood loss due to a better needle/thread ratio [119].

In third or further transplants the surgical approach must be planned pre-operatively so that appropriate arterial
inflow and venous outflow exists with adequate space to implant the new kidney [120, 121]. Nephrectomy
of an old transplant kidney may be required prior to transplantation or at the time of transplantation [120].
Mobilisation of the common or internal iliac artery, internal iliac vein or IVC may be required. An intra-peritoneal
approach (via the iliac fossa or midline) may be required [122]. Rarely orthotopic transplantation is needed
[120, 123].

Evidence suggests that minimising the anastomosis time and/or rewarming time results in reduced DGF [124].
The effect on long term graft function is uncertain, but may also be impacted by short anastomosis time [125].

RENAL TRANSPLANTATION - TEXT UPDATE 2021 15


Summary of evidence LE
A small RCT (n=38) comparing end-to-end anastomosis to the internal iliac artery vs. end-to-side 1b
anastomosis to the external iliac artery found that both techniques showed similar results in the post-
operative period and at three-years follow-up.
Cohort studies have demonstrated third or further transplants are a valid therapeutic option with 3
reasonable short- and long-term patient and graft survival.

Recommendations Strength rating


Use the external or common iliac arteries for an end-to-side arterial anastomosis to donor Weak
renal artery.
Use an end-to-end anastomosis to the internal iliac artery as an alternative to external or Weak
common iliac arteries.
Check the intima of the donor and recipient arteries prior to commencing the arterial Strong
anastomosis to ensure that there is no intimal rupture/flap. If this is found it must be
repaired prior to/as part of the arterial anastomosis.
Pre-operatively plan the surgical approach in third or further transplants, to ensure that Strong
appropriate arterial inflow and venous outflow exists with adequate space to implant the
new kidney.

3.1.5.2 Robot-assisted kidney transplant surgery


Robot-assisted kidney transplant (RAKT) surgery using living donor kidneys has been evaluated in multi-centre
prospective non-randomised studies (using IDEAL consortium principles) [126]. Single-centre prospective non-
randomised studies are on-going addressing RAKT with use of deceased donor kidneys. Both trans- and
extra-peritoneal approaches for RAKT are described. Potential advantages of RAKT may exist (decreased post-
operative pain, incision length and lymphocele rate). Potential issues with RAKT are the exclusion of recipients
with severe atherosclerosis or third (or further) kidney transplants, a higher than expected rate of DGF and a
small number of reported early arterial thromboses despite carefully selected cases [127]. The learning curve for
RAKT has been reported to be 35 cases for experienced surgeons in a retrospective multicentre series of 187
patients undergoing RAKT [128]. Complication and DGF rates decreased significantly and plateaued after the first
20 cases. The rate of Clavien-Dindo grade III/IV complications was 14% during the first ten RAKTs, but only 3%
after this [128]. The rate of arterial graft thrombosis (1.6%) was comparable with that for open kidney transplant
(0.5 - 3.5%) [128]. A ten year single-centre retrospective analysis of 239 obese RAKT patients concluded that
RAKT can be safely performed in obese patients with minimal risk of developing a surgical site infection [129]. A
graft failure rate of 7.1% was reported during follow-up mostly due to acute rejection. Patient and graft survival
was 95% and 93% at three years, respectively [129]. Evidence is too premature to recommend RAKT outside of
appropriately mentored prospective studies.

3.1.5.3 Dual kidney transplants


Dual kidney transplant (DKT) is performed when the quality of a single deceased donor kidney is thought to be
insufficient for appropriate long-term graft function and that the outcome with both kidneys would be better. A
variety of surgical techniques have been described to implant the pair of donor kidneys [130]. These include
unilateral extra-peritoneal (UEP) or intra-peritoneal (UIP) and bilateral extra-peritoneal (BEP) or intra-peritoneal
(BIP) that can be via a midline [131] or two lateral incisions.
The aim of a unilateral approach is to leave the contralateral iliac fossa intact for future
transplantation in the event of graft loss and to reduce CIT for the second kidney transplant [132]. The unilateral
approach may require mobilisation and division of the internal iliac vein to facilitate the two renal veins to iliac
vein anastomoses. Modifications of the unilateral technique include single renal artery and vein anastomoses
(with bench reconstruction) to further reduce CIT for the second kidney [133-135]. Dual kidney transplant takes
longer and has higher blood loss than SKT regardless of the technique used. Data suggest shorter operative
time and hospital stay with UEP compared to BEP [136], but other data suggest similar outcomes from all DKT
techniques. No RCT exists to recommend one technique for all patients or situations.
En-bloc retrieval is performed when kidneys are retrieved from children weighing < 15 kg. Depending
on the size of the donor kidney and size and weight of the adult recipient(s), en-bloc transplantation of the two
kidneys may be performed or, if appropriate, the aorta and IVC patch may be divided for SKT [137].

3.1.5.4 Ureteric implantation in normal urinary tract


Ureteric anastomotic techniques described for renal transplant recipients with no underlying urological
abnormality include: extra (Lich-Gregoir) or intra (Leadbetter-Politano) vesical uretero-neo-cystotomy and
uretero-ureterostomy using native ureter. A meta-analysis [138] of two RCTs and 24 observational studies

16 RENAL TRANSPLANTATION - TEXT UPDATE 2021


favoured the extra-vesical Lich-Gregoir technique to an intravesical approach leading to reduced overall
complications (specifically urine leak, stricture and post-operative haematuria). Fewer urinary tract infections
(UTIs) were observed with the extravesical approach when compared with the intra-vesical technique in one
RCT [139]. Pyelo- or uretero-ureterostomy to the ipsilateral native ureter has been described as a primary
technique in recipients with non-refluxing native ureters [140]. A meta-analysis suggested ureteric stricture,
obstruction, and stone formation were more common after uretero-ureterostomy whereas vesicoureteral reflux
and UTIs were more common after uretero-neo-cystostomy [141].
The donor ureter should be kept as short as possible with peri-ureteric fat preserved to ensure
adequate ureteric blood supply. The location on the bladder to position an extra-vesical anastomosis was
shown in one small RCT to be advantageous at the posterior bladder rather than anterior position to facilitate
future endoscopic manipulation if needed, and reported less hydronephrosis post stent removal [142]. In cases
where donor ureter has been damaged at retrieval then pyelo-native-ureterostomy or pyelo-neo-cystotomy can
be performed. Mono-filament absorbable sutures should be used for the urinary anastomosis to prevent stone
formation around the suture material [143].

Summary of evidence LE
A meta-analysis of two RCTs and 24 observational studies favoured the extra-vesical Lich-Gregoir 1a
technique for reduced overall complications.
A multi-centre prospective comparison study found the incidence of overall complications was similar 2b
for pyelo- and uretero-ureteral anastomosis and that for both procedures no graft was lost due to
urological complications.

Recommendations Strength rating


Perform Lich-Gregoir-like extra-vesical ureteric anastomosis technique to minimise urinary Strong
tract complications in renal transplant recipients with normal urological anatomy.
Pyelo/uretero-ureteral anastomosis is an alternative especially for a very short or poorly Strong
vascularised transplant ureter.

Transplant ureteric anastomosis can be performed with or without a ureteric stent. If a stent is placed a second
procedure is generally required for removal. A Cochrane review [144] concluded that stents are recommended
to reduce major urological complications, especially urinary leak. The optimal timing for stent removal has
yet to be defined [145]. A meta-analysis of five RCTS including 568 kidney transplantation patients showed
a significant reduction in UTIs for early (≤ 7 days) vs. late removal (≥ 14 days) [146]. No significant differences
where observed between the two groups in relation to post-operative complications such as ureteral stricture,
ureteral obstruction, and ureteral leakage [146]. A second meta-analysis including 3,612 patients also reported
a significant reduction in UTIs with early stent removal (< 3 weeks) vs. late removal (> 3 weeks) [147]. No
significant differences where observed between the two groups regarding the incidents of ureteral stenosis and
ureteral leakage [147].
Most commonly, stents are removed with local anaesthetic flexible cystoscopy unless there is a
need to combine with another procedure warranting general anaesthetic. Various techniques to reduce the
morbidity of a second procedure involve tying the stent to the catheter or use of percutaneous stents [148].

Recommendation Strength rating


Use transplant ureteric stents prophylactically to prevent major urinary complications. Strong

Duplex ureters are not infrequently identified at organ retrieval/kidney benching or during work-up for LDN [149,
150]. Duplex ureters can be anastomosed together and then joined to the bladder as one unit (double pant) or
kept as two separate anastomoses. This also applies to the two single ureters in DKT in adults or with en-bloc
transplantation from paediatric donors. The arguments for two separate ureteric anastomoses to the bladder
are that an already tenuous blood supply may be further compromised with added suturing and handling, and if
there is an issue with one ureter the other should remain unaffected. The advantages to forming one single (two
ureter) anastomosis to the bladder are that only one cystotomy is needed; it may be faster and complications
may be reduced. There is a lack of high-quality evidence relating to duplex ureters.

Recommendation Strength rating


Use the same surgical principals for single ureters to manage duplex ureters and Strong
anastomose them either separately or combined.

RENAL TRANSPLANTATION - TEXT UPDATE 2021 17


3.1.5.5 Transplantation/ureteric implantation in abnormal urogenital tract
The following points should be considered when performing kidney transplantation in the abnormal urogenital
tract:
• In patients with an ileal conduit, a kidney transplant may be placed upside down to align the ureter to the
conduit and avoid a redundant ureter [151].
• The technique used to implant transplant ureter(s) into an ileal conduit is the same as the method used
with native ureter(s) (Bricker; Wallace).
• In bladder augmentation or continent pouches, ureters should be implanted with a tunnel technique or
extra-vesically (Lich-Gregoir). The latter is favoured in most patients.
• In patients with a Mitrofanoff catheterisable stoma or continent ileo-caecal pouch with catheterisable
stoma, consideration should be given to the positioning of the catheterisable stoma (umbilical or iliac
fossa - usually right-side) with clear communication with the transplant surgeons so that the position
of any future transplant kidney is not compromised. If an intra-peritoneal placement of a future kidney
transplant is likely, then placement of a Mitrofanoff exiting in the iliac fossa is preferable at the umbilicus.
If a future kidney transplant is likely in the right iliac fossa then placement of a Mitrofanoff exiting at the
umbilicus or left iliac fossa may be preferable.

3.1.6 Donor complications


Living-donor nephrectomy, like any other intervention, is potentially associated with complications and
mortality. However, the fact that the operation is performed on a healthy individual amplifies the relevance of
any complications. Potential complications should be included in the process of informed consent.
Reported surgical mortality is 0.01% to 0.03% with no apparent alteration due to changes in
surgical techniques or donor selection in recent years [152, 153]. According to a recent systematic review
(190 studies) and meta-analysis (41 studies) on complications in minimally invasive LDN, reporting on a total
of 32,308 LDNs, intra-operative complications occur in 2.2% (the most common being bleeding in 1.5% and
injury to other organs in 0.8%) and post-operative complications occur in 7% (infectious complications in
2.6% and bleeding in 1%) [152]. Conversion to open surgery was reported in 1.1%, half due to bleeding and
half due to injury to other organs. Surgical re-interventions occurred in 0.6%; the majority due to bleeding or
to evacuate a haematoma [152]. A low trigger for conversion or re-operation should be observed in order to
minimise the risk of serious complications.
A recent review looked for complications in 14,964 LDNs performed in the U.S. from 2008-2012 and
found an overall peri-operative complication rate of 16.8%, gastrointestinal (4.4%), bleeding (3.0%), respiratory
(2.5%), surgical/anaesthesia-related injuries (2.4%), and “other” complications (6.6%). Among the sample,
2.4% required intensive care and in-hospital mortality was 0.007% [16].
Major Clavien Classification of Surgical Complications grade IV or higher affected 2.5% of donors.
Risk factors for Clavien grade IV or higher events included obesity (adjusted odds ratio [aOR] 1.55, p = 0.0005),
pre-donation haematologic (aOR 2.78, p = 0.0002), psychiatric conditions (aOR 1.45, p = 0.04) and robotic
nephrectomy (aOR 2.07, p = 0.002). An annual centre volume > 50 (aOR 0.55, p < 0.0001) was associated with
lower risk [16].

3.1.6.1 Long-term complications


Long-term complications are mostly related to the single-kidney condition. Renal function in living donors
decreases after donation before improving for many years; however, in the long run it shows signs of slight
deterioration [154-156]. There is a steady increase in the incidence of proteinuria; hypertension post-transplant
having been shown as the main cause of increased albumin excretion [157].

The overall incidence of end-stage renal disease (ESRD) (0.4-1.1%) does not differ from the general population
[154, 155, 158, 159]. According to a recent large retrospective study, the majority of ESRD developing after
living kidney donation is due to new-onset disease that would have affected both kidneys [160]. However, there
are some identified risk factors for deterioration of renal function after donation. According to a recent study
that evaluated 119,769 live kidney donors in the United States, obese (BMI > 30) living kidney donors have a
1.9-fold higher risk for ESRD compared to their non-obese counterparts [161]. Long-term risk of death is no
higher than for an age- and co-morbidity-matched population [153, 158].

Health related quality of life (HRQoL), including mental condition, remains on average better than the general
population after donation [158, 159, 162]. However, some donors experience significant deterioration in their
perceived QoL [162]. While global HRQoL is comparable or superior to population normative data, some
factors identifiable around time of donation including longer recovery, financial stressors, younger age, higher
BMI, lower education, smoking and higher expectations prior to donation, may identify donors more likely to
develop poor HRQoL, providing an opportunity for intervention [158, 159, 162]. It is paramount that a careful

18 RENAL TRANSPLANTATION - TEXT UPDATE 2021


risk–benefit assessment is done and that proper information is given to the prospective donor, this should also
include recommendations on health-promoting behaviour post-donation [163].

Summary of evidence LE
A systematic review and meta-analysis on complications in minimally invasive LDN concluded that the 1a
techniques used for minimally invasive LDN are safe and associated with low complication rates.
Survival rates and risk of end-stage renal disease are similar to those in the general population whilst 2b
donors HRQoL remains on average better than the general population.

Recommendations Strength rating


Restrict living donor nephrectomy to specialised centres. Strong
Offer long-term follow-up to all living kidney donors. Strong

3.1.7 Recipient complications


3.1.7.1 General complications
Surgical complications during and after kidney transplantation may expose the recipient to an increased risk
of morbidity and mortality. The incidence and management of such complications is therefore of primary
importance [138, 145, 164-176]. We herein describe in detail the most common surgical complications in renal
transplantation.

3.1.7.2 Haemorrhage
Haematomas are usually a minor complication in renal transplantation. Their incidence is reported to be
between 0.2-25% [177, 178]. Small and asymptomatic haematomas do not usually require any intervention. In
case of larger haematomas, clinical signs and symptoms due to external pressure with graft dysfunction and/or
thrombotic graft vessel complications can be present. These cases may be treated by percutaneous drainage
under computed tomography (CT) or ultrasound (US) guidance or may require surgical treatment [177].

3.1.7.3 Arterial thrombosis


Transplant renal artery thrombosis is a rare complication with a prevalence ranging from 0.5-3.5% [179].
Usually, it is a consequence of a technical error during the anastomosis although other causes may be
related to both the donor and recipient’s artery condition (i.e. atherosclerosis), intimal rupture during kidney
harvesting, acute rejection episodes, external compression by haematoma or lymphocele, hypercoagulative
state, severe hypotension, and toxicity of immunosuppressive agents (cyclosporine or sirolimus) [180]. The
clinical manifestations are acute reduction of urine output and the elevation of renal function tests, often
resulting in graft loss [177]. The diagnosis is obtained with eco-colour-Doppler [177]. Surgical exploration
is usually recommended to evaluate the status of the graft. In the rare event the graft appears salvageable,
a thrombectomy must be performed. In this situation, the iliac artery is clamped and an arteriotomy vs. a
dissection of the vascular anastomosis must be performed in order to remove the clot. The graft can be flushed
in-situ and re-vascularised [177]. Unfortunately, in the majority of the situations, the graft is not perfused
and therefore an allograft nephrectomy must be performed [177, 181]. Alternatively, thrombolytic agent
administration through a catheter directly into the transplant renal artery can be an efficient treatment, after the
first ten to fourteen post-transplantation days [177].

Summary of evidence LE
The diagnosis of renal artery thrombosis depends on eco-colour-Doppler followed by surgical 2b
exploration to assess the status of the graft.
Thrombectomy in the case of a viable graft and allograft nephrectomy in the case a non-viable graft 2b
are the treatment options for renal artery thrombosis.

Recommendations Strength rating


Perform ultrasound-colour-Doppler in case of suspected graft thrombosis. Strong
Perform surgical exploration in case of ultrasound finding of poor graft perfusion. Strong
Perform a surgical thrombectomy in case of a salvageable graft if arterial thrombosis is Weak
confirmed intra-operatively.
Perform an allograft nephrectomy in case of a non-viable graft. Strong

RENAL TRANSPLANTATION - TEXT UPDATE 2021 19


3.1.7.4 Venous thrombosis
Transplant renal vein thrombosis is an early complication (prevalence 0.5-4%) and one of the most important
causes of graft loss during the first post-operative month [182]. The aetiology includes technical errors and/or
difficulties during surgery [177] and the hypercoagulative state of the recipient [183, 184]. Colour-Doppler-flow-
ultrasonography shows absence of venous flow with an abnormal arterial signal (usually a plateau-like reversed
diastolic flow). Furthermore, it is common to see an enlargement of the graft due to venous congestion [185].
Surgical exploration is usually recommended despite the fact that the majority of the cases will result in graft
loss. In those cases where the venous thrombosis has not resulted in kidney loss at surgical exploration,
a venotomy with surgical thrombectomy after clamping the iliac vein can be performed. Alternatively, an
explantation and subsequent re-implantation can be considered [177]. Thrombolytic agents can also be used;
however, their results have not been satisfactory [177, 186, 187].

Summary of evidence LE
The diagnosis of renal vein thrombosis depends on colour-Doppler-flow-ultrasonography followed by 2b
surgical exploration to assess the status of the graft.
Thrombectomy in the case of a viable graft and allograft nephrectomy in the case a non-viable graft 2b
are the treatment options for renal vein thrombosis.

Recommendations Strength rating


Perform ultrasound-colour-Doppler in case of suspected graft thrombosis. Strong
Perform surgical exploration in case of ultrasound finding of poor graft perfusion. Weak
If venous thrombosis is confirmed intra-operatively, perform a surgical thrombectomy in Weak
case of a salvageable graft or an allograft nephrectomy in case of a non-viable graft.
Do not routinely use pharmacologic prophylaxis to prevent transplant renal vein thrombosis. Strong

3.1.7.5 Transplant renal artery stenosis.


The incidence of transplant renal artery stenosis is 1-25% [188, 189]. Risk factors include small calibre and
atherosclerosis of the donor artery, trauma to donor artery at procurement, absence of arterial patch, suturing
technique (interrupted vs. continuous), and damage to the iliac artery during transplantation [190, 191]. It is
more common at the site of the anastomosis [190, 191]. It can be suspected in case of arterial hypertension
refractory to medical treatment and/or an increase in serum creatinine without hydronephrosis or urinary
infection. The diagnosis is performed by US-colour-Doppler, showing a peak systolic velocity (PSV) of > 200
cm/s in the graft renal artery [190]. In cases of doubt a magnetic resonance angiogram or a CT angiogram can
be performed [192]. It is important to determine whether the stenosis is haemodynamically significant or not.
Usually, a stenosis of over 50% is considered a risk for kidney impairment [193]. In case of mild stenosis (<
50%) and absence of symptoms with no deterioration of the allograft, management is normally conservative;
although, a strict follow-up with US-colour-Doppler and clinical parameters has to be adopted due to the
possible risk of graft failure [190]. In cases of clinically significant stenosis and/or > 50% on US-colour-Doppler,
a confirmatory angiogram should be performed. If confirmed and a decision to treat is taken, treatments
include percutaneous transluminal angioplasty/stent or surgical intervention. Interventional radiology is typically
the first choice although patients considered unsuitable for radiological angioplasty due to recent transplant,
multiple, long and narrow stenosis, or after failure of angioplasty may benefit from surgical treatment [190, 191].

Summary of evidence LE
Suspect transplant renal artery stenosis in case of refractory arterial hypertension and/or increasing 3
serum creatinine without hydronephrosis/infection.
The diagnosis for transplant renal artery stenosis is by US-colour-Doppler, showing a peak systolic 2a
velocity of > 200 cm/s in the graft renal artery.
Interventional radiology is the first-line treatment option for transplant renal artery stenosis; however, in 3
patients considered unsuitable for radiological angioplasty surgical treatment may be considered.

Recommendations Strength rating


Perform ultrasound-colour-Doppler to diagnose an arterial stenosis, in case of Strong
undetermined results on ultrasound consider a magnetic resonance or computed
tomography angiogram.

20 RENAL TRANSPLANTATION - TEXT UPDATE 2021


Perform percutaneous transluminal angioplasty/stent, if feasible, as first-line treatment for Strong
an arterial stenosis.
Offer surgical treatment in case of recent transplant, multiple, long and narrow stenosis, or Strong
after failure of angioplasty.

3.1.7.6 Arteriovenous fistulae and pseudo-aneurysms after renal biopsy


Percutaneous biopsy may result in arteriovenous (AV) fistulae and/or intra-renal pseudo-aneurysms in 1-18%
of cases [194]. The aetiology of the AV fistula is related to the simultaneous injury of adjacent arterial and
venous branches. A pseudo-aneurysm occurs when only the arterial branch is damaged. Both conditions are
diagnosed with US-colour-Doppler [177]. The majority of AV fistulae are asymptomatic, resolving in one to
two years spontaneously, whilst approximately 30% of them persist and become symptomatic. Typically, the
symptoms are hypertension, haematuria, and graft dysfunction due to shunting between arterial and venous
vessels. There is an increased risk of spontaneous rupture in case of enlarging pseudo-aneurysms. For both AV
fistulae and pseudo-aneurysm, angiographic selective or super selective embolisation represents the treatment
of choice [195]. Partial or radical allograft nephrectomy is currently considered the last option [177].

Recommendations Strength rating


Perform a ultrasound-colour-Doppler if a arteriovenous fistulae or pseudo-aneurysm is Strong
suspected.
Perform angiographic embolisation as first-line treatment in symptomatic cases of Strong
arteriovenous fistulae or pseudo-aneurysm.

3.1.7.7 Lymphocele
Lymphocele is a relatively common (1-26%) complication [196]. There is a significant aetiological association
with diabetes, mammalian target of rapamycin (mTOR) inhibitors (i.e. sirolimus) therapy, and acute rejection
[197]. For large and symptomatic lymphocele, laparoscopic fenestration is associated with the lowest overall
recurrence (8%) and complication (14%) rate compared to open surgery and aspiration therapy [198].
Placement of a percutaneous drain (i.e. Pig-Tail) is an option with a success rate as high as 50% [163].
Percutaneous aspiration can be performed although the recurrence rate can be as high as 95% [198], with an
increased risk of local infection (6-17%) [198]. Furthermore, sclerosant agents such as ethanol, fibrin sealant,
gentamicin, or octreotide reduce the recurrence rate compared to simple aspiration [198, 199].

Recommendations Strength rating


Perform percutaneous drainage placement as first-line treatment for large and symptomatic Strong
lymphocele.
Perform fenestration when percutaneous treatments fail. Strong

3.1.7.8 Urinary leak


Urinary leakage occurs in 0-9.3% of cases [200]. Anastomotic urine leaks can be ureteral or vesical [201].
Ureteral necrosis and/or suture failure are the most important causes [202, 203]. Non-technical risk factors
include recipient age, number of renal arteries, site of arterial anastomosis, occurrence of acute rejection
episodes, bladder problems, and immunosuppressive regimen [204]. Urinary leak can be suspected by the
urine output and the creatinine level in the drain fluid [202]. In order to decrease the risk of ureteral necrosis,
it is important to preserve vascularisation of the distal ureter [202]. Furthermore, the routine use of a JJ-stent
is recommended [203, 205]. The management of urinary leak depends on the location (renal pelvis, proximal
or distal ureter, and bladder), the time of appearance and the volume of the leak. For early and low volume
urine leaks the treatment may be conservative (i.e. urethral catheter, percutaneous nephrostomy and JJ-stent)
[206]. In case of failure of the conservative management, or massive leak, surgical repair must be undertaken.
Ureteral re-implantation directly to the bladder or to the native ureter provide similar results [141, 206].

Summary of evidence LE
Suspect urinary leakage based on the urine output and the creatinine level in the drain fluid. 3
For early and low volume urine leaks conservative management may be considered. 3
Surgical repair should be undertaken when conservative management fails or massive urine leak 2b
occurs.

RENAL TRANSPLANTATION - TEXT UPDATE 2021 21


Recommendations Strength rating
Manage urine leak by JJ-stent and bladder catheter and/or percutaneous nephrostomy Strong
tube.
Perform surgical repair in cases of failure of conservative management. Strong

3.1.7.9 Ureteral stenosis


Ureteral stenosis is a common complication in recipients, with an incidence of 0.6-10.5% [207]. Early
stenosis (within three months of surgery) is usually caused by surgical technique or compromised ureteral
blood supply during surgery. Late stenosis (after > six months) is provoked by infection, fibrosis, progressive
vascular disease and/or rejection [202, 208]. Clinically significant ureteral stricture should be considered when
persistent hydronephrosis on US occurs in association with impaired renal function. The first approach in the
management of stricture is the placement of a percutaneous nephrostomy tube with an antegrade pyelogram
[207]. The following treatment options depend mainly on the timing, recoverable kidney function, anatomy
of the stricture, patient body habitus/comorbidities, and surgeon preference. Strictures < 3 cm in length may
be treated endoscopically either with percutaneous balloon dilation or antegrade flexible ureteroscopy and
holmium laser incision. In this scenario the success rate approaches 50%; although, maximum success is
obtained for strictures < 1 cm [209-211]. In case of a recurrence after a primary endourological approach
and/or stricture > 3 cm in length, surgical reconstruction should be performed [208] including direct ureteral
re-implantation, pyelo-vesical re-implantation (with or without psoas hitch and/or Boari Flap) or in cases with a
normal native ureter, uretero-ureterostomy [212, 213]. Long-term graft and patient survival are not significantly
affected [214].

Summary of evidence LE
Clinically significant ureteral stricture should be considered when persistent hydronephrosis on US 3
occurs in association with impaired renal function.
The first approach in the management of a stricture is the placement of a percutaneous nephrostomy 2b
tube with an antegrade pyelogram.
Strictures < 3 cm in length may be treated endoscopically. 3
For strictures > 3 cm in length or those which have reoccurred following a primary endourological 2b
approach surgical reconstruction should be performed.

Recommendations Strength rating


In case of ureteral stricture, place a nephrostomy tube for both kidney decompression and Strong
stricture diagnosis via an antegrade pyelogram.
Manage strictures < 3 cm in length either with surgical reconstruction or endoscopically Strong
(percutaneous balloon dilation or antegrade flexible ureteroscopy and holmium laser
incision).
Treat late stricture recurrence and/or stricture > 3 cm in length with surgical reconstruction Strong
in appropriate recipients.

3.1.7.10 Haematuria
The incidence of haematuria ranges from 1-34% [200]. According to the literature, the Lich-Gregoire technique
provides the lowest incidence of haematuria. Furthermore, meticulous haemostasis during re-implantation
results in minimal bleeding [138, 200, 201]. Bladder irrigation is the first-line treatment. Some cases require
cystoscopy with evacuation of clots and/or fulguration of bleeding sites [200].

3.1.7.11 Reflux and acute pyelonephritis


The frequency of vesicoureteral reflux is between 1-86% [200, 215]. Acute graft pyelonephritis occurs in 13%
of graft recipients. Patients with lower tract urinary infections and cytomegalovirus (CMV) infection present a
higher risk of acute graft pyelonephritis [216]. Endoscopic injection of dextranomer/hyaluronic acid copolymer
may be the first approach for treatment of vesicoureteral reflux associated with acute pyelonephritis, with a
success rate ranging from 57.9% after the first injection to 78.9% after the second injection [217]. Ureteral
re-implantation or pyelo-ureterostomy with the native ureter is a viable second treatment option [212].

Recommendation Strength rating


Use an endoscopic approach as first-line treatment for symptomatic reflux. Weak

22 RENAL TRANSPLANTATION - TEXT UPDATE 2021


3.1.7.12 Kidney stones
Urolithiasis occurs in 0.2-1.7% of recipients [218, 219]. The most frequent causes are hyper filtration, renal
tubular acidosis, recurrent UTIs, hypocitraturia, hyperuricaemia, hyperuricemia, excessive alkaline urine,
persistent tertiary hyperparathyroidism and ureteral strictures [220, 221]. Another risk factor can be urinary
anastomosis, with the lowest stone rate using Lich-Gregoir technique [219]. The most frequent clinical signs
are fever, increased serum creatinine level, decreased urine output, and haematuria. Pain is usually not referred
to due to impaired innervation. A US examination usually provides the diagnosis although a CT of the kidneys,
ureters and bladder may be needed to confirm the location and size of the stone [220]. The management
depends on the location and size of the stone, and the presence of obstruction. In case of obstructive
stones first-line treatment includes placement of a nephrostomy tube, or in some occasions a JJ-stent [222].
Extracorporeal shock wave lithotripsy (ESWL) is usually considered the first approach for stones < 15 mm with
stone-free rates varying between 40 and 80% depending on the location of the stone [222]. Ureteroscopy,
including antegrade and retrograde approaches, can be considered for stones < 20 mm, with a success rate
of up to 67% [140, 219, 223]. For larger stones (> 20 mm), percutaneous nephrolithotomy (PNL) can be offered
with high overall effective stone-free rates. In cases of large impacted stones, uretero-ureteral anastomosis,
pyelo-ureteral anastomosis, or uretero-vesical re-implantation may provide excellent results for both stone and
ureteral obstruction [219].

Summary of evidence LE
Extracorporeal shockwave lithotripsy should be considered as the first-line treatment option for stones 2b
< 15 mm.
Antegrade/retrograde ureteroscopy and PNL may be considered as treatment options as they provide 2b
high stone-free rates.
For larger stones (> 20 mm), PNL can be offered with a high overall effective stone-free rate. 2b

Recommendations Strength rating


Evaluate the causes of urolithiasis in the recipient. Strong
Treat ureteral obstruction due to a stone with a percutaneous nephrostomy tube or JJ-stent Strong
placement.
Perform shockwave lithotripsy or antegrade/retrograde ureteroscopy for stones < 15 mm. Strong
Perform percutaneous nephrolithotomy for stones > 20 mm. Weak

3.1.7.13 Wound infection


Wound infections occur in about 4% of cases. Risk factors include recipients > 60 years, high BMI, anaemia,
hypo-albuminemia, long surgical times (> 200 min) [224]. Bacteria commonly involved are Enterobacteriaceae,
Staphylococcus aureus and Pseudomonas [212]. Subcutaneous sutures, pre-dialysis transplantation, sealing
or ligation of lymphatic trunks, prophylactic fenestration, reducing corticosteroid load, and avoiding sirolimus/
everolimus therapy can decrease wound complication rates [224].

3.1.7.14 Incisional hernia


Incisional hernia occurs in approximately 4% of open kidney transplantations. Risk factors include age,
obesity, diabetes, haematoma, rejection, re-operation through the same transplant incision and use of m-TOR
inhibitors. Mesh infection is a risk factor for incisional hernia recurrence [225]. Open and laparoscopic repair
approaches are safe and effective [225].

3.1.8 Urological malignancy and renal transplantation


The following section is limited to a synopsis of three systematic reviews conducted by the EAU Renal
Transplantation Panel.

3.1.8.1 Malignancy prior to renal transplantation


3.1.8.1.1 In the recipient
Standard procedure for transplant candidates includes systematic screening for the presence of any active/
latent cancer or a past history of cancer. In candidates with a previous history of urological cancer, it can be
challenging to decide if patients are suitable for transplantation and, if so, how long the waiting period prior
to transplantation should be. To date, the waiting period has been primarily based on the Cincinnati Registry,
which takes into account the type of tumour and the time between its treatment and kidney transplantation.
However, the Cincinnati Registry has potential drawbacks as it does not consider the epidemiology of tumours

RENAL TRANSPLANTATION - TEXT UPDATE 2021 23


or that diagnostic and therapeutic procedures/tests have changed over time and that prognostic tools have
improved. Additionally, treatment and the staging of the disease are not defined.
According to a recent systematic review the risk of tumour recurrence was similar between
transplantation (n=786) and dialysis (n=1,733) populations for renal cell carcinoma (RCC) and prostate
cancer (PCa). This was especially true for low grade/stage PCa, for which the risk of recurrence was low
and consistent with nomograms [226]. For low stage/grade RCC the recurrence rate was significant for both
dialysis and renal transplantation; however, recurrences were actually contralateral RCC with no impact on
patient or graft survival [226].
Testicular cancer had a low risk of recurrence but case reports highlighted the possibility of late
recurrence even for stage I tumours [226].
For urothelial carcinoma, studies were mainly related to upper urinary tract carcinomas in the
context of aristolochic acid nephropathy for which the rate of synchronous bilateral tumour was 10-16% and
the rate of contralateral recurrence was 31-39% [226].
These findings imply that a kidney transplant candidate with a history of appropriately treated low
stage/grade PCa (PSA ≤ 10, Gleason score ≤ 6 and T1/T2a) or low grade T1 RCC could be listed for renal
transplantation without any additional delay compared to a cancer-free patient. However, as the level of
evidence was low, more studies are needed to standardise waiting periods before renal transplantation.

Summary of evidence LE
Renal Cell Carcinoma
The recurrence rates for transplanted vs. dialysed patients at <1, 1–5, and > 5 years were 0–8% vs. 2b
0%, 0–27% vs. 0–9% and 0–41% vs. 0–48%, respectively.
Overall five year survival rates for transplantation vs. dialysed patients were 80–100% vs. 76–100%,
respectively.
Prostate Cancer
The recurrence rates for transplantation patients at <1 and > 5 years were 0–9% and 4–20%, 2b
respectively.
Overall, 1–5 year survival rates for transplantation patients ranged from 62% to 100%.

Recommendation Strength rating


List for renal transplantation patients with a history of appropriately treated low stage/grade Weak
renal cell carcinoma or prostate cancer without additional delay.

3.1.8.1.2 In the potential donor kidney


In the general population, RCC constitutes 3% of all malignancies, with the incidence being highest in patients
aged > 60 years. The current increasing age of donors may lead to a higher number of incidental RCCs found
in donor kidneys and could theoretically decrease the number of kidneys suitable for transplantation. The main
surgical approach to these kidneys is ex vivo tumour excision on the back-table with an oncological margin,
frozen section biopsy, bench surgery renorraphy, and finally transplantation in the conventional fashion [227].
A recent systematic review assessed the effectiveness and harms of using kidneys with small renal
tumours, from deceased or living donors, as a source for renal transplantation and it reported that five year overall
and graft survival rates were 92% and 95.6%, respectively [227]. Tumour excision was performed ex vivo in all
cases except for two (107/109 patients), and the vast majority of excised tumours were RCCs (88/109 patients),
with clear-cell subtype the most common [227]. This systematic review, although with low-level evidence,
suggested that kidneys with small renal masses are an acceptable source for renal transplantation and do not
compromise oncological outcomes with similar functional outcomes to other donor kidneys.

Summary of evidence LE
Tumour excision was performed ex-vivo in all cases except for two (107/109 patients). 2b
Overall survival rates at one, three and five years were 97.7%, 95.4%, and 92%, respectively.
Mean graft survival rates at one, three and five years were 99.2%, 95%, and 95.6%, respectively.

Recommendation Strength rating


Do not discard a kidney for potential transplantation on the basis of a small renal mass Weak
alone.

24 RENAL TRANSPLANTATION - TEXT UPDATE 2021


3.1.8.2 Malignancy after renal transplantation
Cancer development after kidney transplant has become a major problem as it is one of the main causes of
death in this population. Urological cancers, have an increased incidence after kidney transplantation partly
due to the increasing age of recipients and their prolonged survival after transplantation.

Treatment of localised PCa following kidney transplantation is challenging due the presence of the kidney
graft in the pelvic cavity close to the prostate. Two systematic reviews reported that oncological outcomes
following PCa treatment in kidney transplant recipients are comparable to the non-transplanted population
[228, 229] and surgery (radical prostatectomy), carried out in tertiary high-volume referral centres was the
treatment choice in 75 to 85% of patients [228, 229]. Marra et al. reported cancer-specific survival rates of
96.8% for surgery, 88.2% for radiotherapy with androgen deprivation therapy and 100% for brachytherapy at
mean follow-up of 24 months [229]. Hevia et al. reported five year cancer-specific survival of 97.5% for surgery,
87.5% for external beam radiation and 94.4% for brachytherapy [228].

Summary of evidence LE
Surgery (radical prostatectomy) was the most frequently performed treatment for localised PCa after 2b
kidney transplant.
Overall oncological outcomes following PCa treatment in kidney transplant recipients were 2b
comparable to the non-transplanted population.

Recommendations Strength rating


Be aware of the presence of a kidney transplant in the pelvis and the possibility of Strong
subsequent transplants when planning treatment for prostate cancer.
Refer kidney transplant patients with prostate cancer to an integrated transplant urology Strong
centre.

3.1.9 Matching of donors and recipients


Histocompatibility antigens show remarkable polymorphism and human leukocyte antigen (HLA) matching
is still very important in kidney transplantation as transplant outcome correlates with the number of HLA
mismatches [230-233]. Human leukocyte antigen incompatibility can result in proliferation and activation of
the recipient’s CD4+ and CD8+ T-cells with concomitant activation of B-cell allo-antibody production. This
may lead to cellular and humoral graft rejection. Matching should concentrate on HLA antigens, which impact
outcome. Human leukocyte antigens A, B, C as well as DR must be determined in all potential recipients and
donors according to current guidelines and national allocation rules [230-235]. Additionally, it is recommended
to determine HLA-DQ antigens of donor and recipient. Furthermore, HLA-DP antigen characterisation may be
performed, especially for sensitised recipients [230-235].

All patients registered for renal transplantation must have their serum screened for anti-HLA antibodies,
which are particularly common after pregnancy, previous transplant, transplant rejection, and blood
transfusions [230-235]. Thorough pre-transplant testing for HLA antibodies must be performed according
to current recommendations [230-235]. Sera from potential organ recipients should be screened for
HLA-specific antibodies every three months or as stipulated by the national and/or international organ
exchange organisations [230-235]. In addition, screening for HLA-specific antibodies should be carried out
at two and four weeks after every immunising event, e.g. blood transfusion, transplantation, pregnancy, and
graft explantation [230-235]. Highly sensitised patients should have prioritised access to special allocation
programmes [232, 233, 235], such as the acceptable mismatch (AM) programme of Eurotransplant [236]. A
careful analysis of HLA antibody specificities must be carried out to avoid unacceptable HLA antigens and
to determine acceptable HLA antigens in potential donors, who are expected to give a negative cross-match
result. The definition of unacceptable HLA antigens should be implemented according to local allocation rules
and international recommendations [230-234, 237]. The information on unacceptable HLA antigens should be
highlighted with the patient’s details in the database of the national kidney-sharing programme, preventing the
unnecessary transport of kidneys to recipients with high antibody sensitivity.

To avoid hyper-acute rejection (HAR), adequate (e.g. CDC, virtual) cross-match tests must be performed before
each kidney and combined kidney/pancreas transplantation in accordance with national and international
recommendations [230-233, 235].

RENAL TRANSPLANTATION - TEXT UPDATE 2021 25


Laboratories which provide HLA-testing, HLA antibody testing and cross-matching for transplant centres must
have valid accreditation to ensure accuracy and reliability [224, 225, 230-232]. They must follow the standards
of national and international organisations, such as the European Federation for Immunogenetics [235].

Previously, compatibility for ABO blood group antigens and HLA antigens was of critical importance in kidney
transplantation. This may change in the future, e.g. in the new U.S. allocation system A2 and A2B donors
are transplanted into B recipients [233]. To avoid an increasing imbalance between demand and supply in
deceased donor kidney transplantation in O recipients, ABO identity is demanded by several organ allocation
organisations with a few exceptions, e.g. as in zero HLA-A+B+DR-mismatch kidneys [233, 234]. With the
introduction of antibody elimination methods, potent immunosuppression and novel agents (e.g. anti B-cell
drugs), successful ABO-incompatible living donor transplantations, with good long-term outcomes are possible
[238, 239]. However, higher costs and infection rates have been described.
Even the barrier of a positive cross-match due to preformed HLA antibodies is under discussion
with newer “desensitisation” techniques available in cases with available living donors [240, 241]. Success
rates are lower, antibody-mediated rejections are frequent, but survival may be better compared to
waiting list survival on dialysis. While this is a rapidly evolving field, further research is needed to define
standard protocols. Until then such “desensitisation” protocols are experimental and patients undergoing
“desensitisation” should be treated in specialised centres, where outcomes are documented. Patients should
be informed adequately of the risks and limitations and alternative strategies (e.g. acceptable mismatch
programmes, cross-over transplantation and donor chains) should be discussed.

Summary of evidence LE
Human leukocyte antigen matching is very important in kidney transplantation as transplant outcome 3
correlates with the number of HLA mismatches. Matching should concentrate on HLA antigens, which
impact outcome.
In accordance with national and international recommendations adequate (e.g. CDC, virtual) cross- 3
match tests must be performed before each kidney and combined kidney/pancreas transplantation to
avoid hyper-acute rejection.

Recommendations Strength rating


Determine the ABO blood group and the human leukocyte antigen A, B, C and DR Strong
phenotypes for all candidates awaiting kidney transplantation.
Test both the donor and recipient for human leukocyte antigen DQ. Human leukocyte Strong
antigen DP testing may be performed for sensitised patients.
Perform thorough testing for HLA antibodies before transplantation. Strong
Perform adequate cross-match tests to avoid hyper-acute rejection, before each kidney and Strong
combined kidney/pancreas transplantation.

3.1.10 Immunosuppression after kidney transplantation


The principle underlying successful immunosuppression is ‘the balance of survival’. Practitioners must
prescribe a dosage of drug high enough to suppress rejection without endangering the recipient’s health.
Increased understanding of immune rejection has led to the development of safe modern immune suppression
agents [242, 243], which suppress sensitised lymphocyte activity against a transplant. Immunosuppression
is particularly important during the initial post-transplant period when there is a high incidence of early post-
transplant rejection.
In later post-operative stages, ‘graft adaptation‘ occurs, resulting in the very low rejection rates seen
in maintenance patients. Rejection prophylaxis should therefore be reduced over time by steroid tapering and
gradual lowering of calcineurin inhibitor (CNI) [242-244].

Non-specific side effects of immunosuppression include a higher risk of malignancy and infection, particularly
opportunistic infections [242-244]. All immunosuppressants also have dose-dependent specific side effects.
Current immunosuppressive protocols aim to reduce drug-specific side effects using a synergistic regimen. A
truly synergistic regimen allows profound dose reductions of immunosuppressive drugs; therefore, reducing
side effects whilst still maintaining efficacy due to the synergistic effects of the immunosuppressants.

The currently recommended standard initial immunosuppression regime provides excellent efficacy with good
tolerability [242-245]. It is given to most patients and consists of:

26 RENAL TRANSPLANTATION - TEXT UPDATE 2021


• calcineurin inhibitors (preferably tacrolimus, alternatively cyclosporine);
• mycophenolate (MMF or enteric-coated mycophenolate sodium [EC-MPS]);
• steroids (prednisolone or methylprednisolone);
• induction therapy (preferably basiliximab in low and standard risk patients and anti-thymocyte globulin
[ATG] in high-risk patients).

This multidrug regimen reflects the current standard of care for the majority of transplant recipients worldwide
[242-244] and may be modified according to local needs and immunological risk. This standard regimen is
likely to change as new immunosuppressive drugs and new treatment regimens are developed [242-244]. In
addition, any initial drug regimen will need to be tailored to the individual needs of a patient as suggested by
the appearance of side effects, lack of efficacy or protocol-driven requirements.

Recommendation Strength rating


Perform initial rejection prophylaxis with a combination therapy of a calcineurin inhibitor Strong
(preferably tacrolimus), mycophenolate, steroids and an induction agent (either basiliximab
or anti-thymocyte globulin).

3.1.10.1 Calcineurin inhibitors


Both cyclosporine and tacrolimus have significant side effects that are hazardous to the graft and patient [242-
249]. Most importantly, both are nephrotoxic [250, 251], and long-term use is an important cause of chronic
allograft dysfunction [252], eventually leading to graft loss or severe chronic kidney disease in recipients of
non-renal organs. Both CNIs are considered to be ‘critical-dose’ drugs, so that any deviations from exposure
can lead to severe toxicity or failure of efficacy. Due to their narrow therapeutic window and the potential for
drug-to-drug interaction, CNIs should be monitored using trough levels, which provide a reasonable estimate
for exposure [249].

Meta-analysis of tacrolimus and cyclosporine has demonstrated similar outcomes with respect to overall
patient and graft survival [242-248, 253, 254]. Tacrolimus provided better rejection prophylaxis and was
associated with better graft survival, when censored for death in some analyses. Renal function was
favourable for tacrolimus treated patients, in a number of trials [254-259]. Therefore, both CNIs can be used
for the effective prevention of acute rejection, but due to higher efficacy tacrolimus is recommended by current
guidelines as first-line CNI [243].

For both CNIs several different formulations are available [249, 260-268]. Tacrolimus once-daily dosing seems
to be preferred by patients and is associated with better adherence and lower pharmacokinetic variability
[249, 269, 270]. Precautions (e.g. close surveillance and determination of drug levels) should be instituted
after conversion from one formulation to another [268, 271-275]. In case of specific side effects of a CNI (e.g.
hirsutism, alopecia, gingival hyperplasia, diabetes, polyoma nephropathy) conversion to another CNI can be a
successful strategy to reduce side effects [242-244, 276]. Due to differences in the efficacy and safety profile,
the choice of CNI should include the individual risks and benefits for each patient.

Despite their side effects, CNIs have been a cornerstone of modern immunosuppressive regimens for more
than thirty years as they have resulted in an exemplary improvement in kidney graft survival [242, 243]. Future
protocols aim to minimise or even eliminate CNIs [244, 247, 249, 277-280]. However, until such strategies
provide superior outcomes, CNIs remain the standard of care [242, 243, 281]. For severe CNI-related side
effects, CNI withdrawal, replacement, or profound reduction may be needed [242, 244, 247, 277, 278]. Special
attention should be paid to maintenance patients, who may need less CNIs than previously thought [242, 244,
249, 278, 279, 282].

Summary of evidence LE
Meta-analysis of tacrolimus and cyclosporine has demonstrated similar outcomes with respect to 1a
overall patient and graft survival; however, tacrolimus provided better rejection prophylaxis.
Due to differences in the efficacy and safety profile, the choice of CNI should take into account the 1
immunological risk, characteristics, concomitant immunosuppression, and socio-economic factors of
the recipient.

RENAL TRANSPLANTATION - TEXT UPDATE 2021 27


Recommendations Strength rating
Use calcineurin inhibitors for rejection prophylaxis as they represent current best practice Strong
pending publication of long-term results using newer agents.
Use tacrolimus as first-line calcineurin inhibitor due to its higher efficacy. Strong
Monitor blood-levels of both cyclosporine and tacrolimus to allow appropriate dose Strong
adjustment of calcineurin inhibitors.

3.1.10.2 Mycophenolates (MPA)


The mycophenolates, MMF and EC-MPS, are based on mycophenolic acid, which inhibits inosine
monophosphate dehydrogenase (IMPDH) [283-287]. This is the rate-limiting step for the synthesis of guanosine
monophosphate in the de novo purine pathway. As the function and proliferation of lymphocytes is more
dependent on de novo purine nucleotide synthesis compared to other cell types, IMPDH inhibitors may provide
more specific lymphocyte-targeted immunosuppression. The co-administration of MPA with prednisolone and
CNI has resulted in a profound reduction of biopsy-proven rejections [242, 245, 283-287]. Mycophenolic acid is
not nephrotoxic; however, it inhibits bone marrow function and may cause CMV infections and gastrointestinal
side effects, particularly diarrhoea [242, 245, 283-287]. There is also a higher incidence of polyoma
nephropathy, especially when mycophenolate is combined with tacrolimus [288].

Both MPA formulations are equally effective with an almost identical safety profile [240, 278, 281, 283-286],
though some prospective studies suggest a better gastrointestinal side-effect profile for EC-MPS in patients
who have suffered from MMF-related gastrointestinal complaints, although firm evidence from prospective
randomised studies is lacking [283-287, 289].

Mycophenolic acid is recommended by guidelines [243]. Standard doses in combination with cyclosporine are
MMF 1 g or EC-MPS 720 mg twice daily, although higher initial doses have been suggested [242, 243, 283-287].
Despite its frequent use with tacrolimus, there is insufficient evidence to support the optimal dosage for this
combination [242, 283, 285, 286, 290]. Tacrolimus has no influence on MPA exposure and leads to approximately
30% higher MPA exposure compared to cyclosporine. Most transplant centres use the same starting dose as in
cyclosporine-treated patients, however dose reductions are frequent, especially because of gastrointestinal side
effects. Weak evidence suggests that MPA dose reductions are associated with inferior outcomes, especially in
cyclosporine treated patients [284-286, 291, 292]. Due to the high incidence of side effects, some centres perform
a protocol-driven MPA dose reduction in tacrolimus treated patients [283, 285]. Regular monitoring for polyoma
(BK virus) is recommended in patients given MPA combined with tacrolimus [242, 288].

Due to a higher incidence of CMV disease with MPA [287], either CMV prophylaxis or a pre-emptive strategy
with regular screening for CMV viraemia should be instituted [242, 293]. Cytomegalovirus prophylaxis with
antiviral medications (e.g. valganciclovir) should be used routinely in CMV positive recipients and in CMV
negative recipients of CMV positive organ transplants, because prophylaxis has recently been shown to reduce
CMV disease, CMV-associated mortality in solid organ transplant recipients, and leads to better long-term graft
survival in kidney allograft recipients.

The benefit for MPA drug monitoring is uncertain and currently not recommended for the majority of patients
[283, 285, 286, 294].

In maintenance patients, the potency of MPA can be used for successful steroid withdrawal in most patients
[295] or for substantial dose reductions of nephrotoxic CNIs, which may lead to better renal function [242-245,
247, 278, 296]. Although there have been several studies of the potential for CNI-free protocols with MPA and
steroids, complete CNI avoidance or withdrawal over the first three years has been associated with a substantially
increased rejection risk and even worse outcomes in prospective randomised studies [242, 244, 278]. In contrast,
CNI withdrawal under MPA and steroids appeared to be safe in long-term maintenance patients beyond five years
post-transplant and resulted in improved renal function [242, 244, 247, 278, 296, 297].

Summary of evidence LE
The co-administration of MPA with prednisolone and CNI has resulted in a profound reduction of 1
biopsy-proven rejections.
Both MPA formulations, MMF and EC-MPS, are equally effective with an almost identical safety 1
profile.
Due to a higher incidence of CMV disease with MPA either CMV prophylaxis or a pre-emptive strategy 1
with regular screening for CMV viraemia should be instituted.

28 RENAL TRANSPLANTATION - TEXT UPDATE 2021


Recommendation Strength rating
Administer mycophenolate as part of the initial immunosuppressive regimen. Strong

3.1.10.3 Azathioprine
Mycophenolate is now routinely used as a primary therapy in place of azathioprine in most units worldwide. In
comparison to azathioprine, MPA reduced rejection rates significantly in prospective randomised trials [242,
243, 245, 283-287]. Although a large, prospective study found that azathioprine may give acceptable results
in a low-risk population [298], azathioprine is usually reserved for patients who cannot tolerate MPA [242,
243, 283, 284, 286]. When added to dual therapy with cyclosporine and steroids, a meta-analysis found no
significant benefit for azathioprine with respect to major outcome parameters [299].

Recommendation Strength rating


Azathioprine may be used in a low-risk population as an immunosuppressive drug, Weak
especially for those intolerant to mycophenolate formulations.

3.1.10.4 Steroids
Steroids have a large number of side effects [242-244, 295], especially with long-term use. Most practitioners
still consider steroids (either prednisolone or methylprednisolone) to be a fundamental adjunct to primary
immunosuppression, even though successful steroid withdrawal has been achieved in the vast majority of
patients in many prospective, randomised trials [242, 244, 245, 295, 300, 301]. The risk of steroid withdrawal
depends on the use of concomitant immunosuppressive medication, immunological risk, ethnicity, and
time after transplantation. Although the risk of rejection diminishes over time, potential benefits may be less
prominent after a prolonged steroid treatment period [242-245, 295]. Recent studies suggest similar efficacy
but less diabetes after early steroid withdrawal or steroid minimisation in low-risk patients treated with
tacrolimus, MPA and induction (either basiliximab or ATG) [302, 303].

Recommendations Strength rating


Initial steroid therapy should be part of immunosuppression in the peri-operative and early Strong
post-transplant period.
Consider steroid withdrawal in standard immunological risk patients on combination therapy Weak
with calcineurin inhibitors and mycophenolic acid after the early post-transplant period.

3.1.10.5 Inhibitors of the mammalian target of rapamycin (m-TOR)


The immunosuppressants, sirolimus and everolimus, inhibit the mammalian target of rapamycin and suppress
lymphocyte proliferation and differentiation [242, 277, 304-306]. They inhibit multiple intracellular pathways and
block cytokine signals for T-cell proliferation. Similar effects are seen on B-cells, endothelial cells, fibroblasts,
and tumour cells. Inhibitors of m-TOR are as effective as MPA when combined with CNIs in preventing rejection
[242, 245, 277, 304-307]. However, m-TOR inhibitors exhibit dose-dependent bone marrow toxicity [242, 277,
304-306]. Other potential side effects include hyperlipidaemia, oedema, development of lymphoceles, wound
healing problems, pneumonitis, proteinuria, and impaired fertility. The extensive side effect profile is responsible
for inferior tolerability compared to MPA and potential differences in outcome in early years, when higher doses
were used [308-313].

To date, no prospective comparative studies have been carried out on the m-TOR inhibitors sirolimus and
everolimus [314]. Both m-TOR inhibitors have an almost identical side effect profile and mainly differ in their
pharmacokinetic properties [242, 277, 304-306, 315]. Sirolimus has a half-life of about 60 hours, is given
once a day and is licensed for prophylaxis in kidney recipients only. Everolimus has a half-life of about 24
hours, is licensed for kidney, liver and heart recipients and is given twice a day. Everolimus is licenced for use
with cyclosporine and can be given simultaneously with cyclosporine, while sirolimus should be given four
hours after cyclosporine. The pharmacological drug-drug interaction with cyclosporine is far less relevant for
tacrolimus, resulting in the need for a higher starting dose of m-TOR inhibitors in combination with tacrolimus
[258, 316, 317]. Sirolimus is also licensed in combination therapy with steroids for cyclosporine withdrawal
from combination therapy with cyclosporine.

Therapeutic monitoring of trough levels is recommended because of the narrow therapeutic window and the
risk of drug-to-drug interactions [242, 277, 304-306, 315].

RENAL TRANSPLANTATION - TEXT UPDATE 2021 29


When combined with CNIs, antimicrobial prophylaxis for Pneumocystis jirovecii pneumonia should be
administered for one year following transplantation, e.g. low-dose cotrimoxazole [242, 304-306]. Most
importantly, combination therapy with CNIs aggravates CNI-induced nephrotoxicity, although m-TOR inhibitors
themselves are non-nephrotoxic [242]. Several studies suggest less favourable outcomes and increased drug
discontinuations due to adverse events for this combination, especially if CNIs are maintained at standard
dosages [242, 245, 247, 258, 307, 309, 310, 318-323]. Calcineurin inhibitor dosage should therefore be
substantially reduced in combination therapy with m-TOR inhibitors, which seems to have no impact on
efficacy, due to the highly synergistic potential of this combination therapy [277, 304-306, 312, 315].

Several studies suggest m-TOR inhibitors cannot replace CNIs in the initial phase after transplantation due to
lower efficacy and a less favourable side effect profile, particularly wound healing problems and lymphoceles
[240, 242, 243, 274, 298, 299, 304, 306, 314]. Other trials suggest that m-TOR inhibitors may replace CNI at
later stages, e.g. three months after transplantation, with improvements in renal function, predominately in
cyclosporine treated patients [242, 244, 245, 247, 256, 277, 304-306, 309, 310, 312, 324-326]. It is unclear
if there is a real benefit in comparison to patients on tacrolimus and MPA [256, 325]. However, there is an
increased risk of rejection and development of HLA antibodies [242, 244, 256, 277, 327], which may be offset
by the benefit of the non-nephrotoxic immunosuppression. Patients treated with m-TOR inhibitors develop less
leucopenia and opportunistic viral infections, especially less CMV infections compared to MPA [258, 309, 312,
322-324, 328].

Proteinuria and poor renal function at conversion are associated with inferior outcomes [242, 244, 277, 304-
306]. Conversion from CNIs is not advisable in patients with proteinuria > 800 mg/day, and a cautious and
individual approach should be followed in patients with GFR < 30 mL/min.

Due to an anti-proliferative effect and a lower incidence of malignancy in m-TOR inhibitor treated patients,
conversion from CNIs to m-TOR inhibitors may be beneficial for patients, who develop malignancy after
transplantation, or who are at a high risk for the development of post-transplant malignancy or skin cancer
[242, 244, 277, 304-306, 311-313, 329-332]. Several studies and case reports have suggested that patients
with Kaposi sarcoma under CNI therapy benefit from conversion to an m-TOR inhibitor [330].

In summary, m-TOR inhibitors are not recommended as initial immunosuppressive therapy due to their side
effect profile and higher discontinuation rates [243]. However, m-TOR inhibitors are a well-studied alternative
treatment option.

Summary of evidence LE
Combination therapy with CNIs aggravates CNI-induced nephrotoxicity. Therefore, CNI dosage should 1
be substantially reduced in combination therapy with m-TOR inhibitors, which seems to have no
impact on efficacy, due to the highly synergistic potential of this combination therapy.
Take into consideration impaired wound healing and prophylactic surgical measures when m-TOR 1
inhibitors are used as part of the initial immunosuppressive regimen or when patients treated with
m-TOR inhibitors undergo major surgery.
When combined with CNIs, antimicrobial prophylaxis for P. jirovecii pneumonia should be administered 1
for one year following transplantation.
Conversion from CNIs is not advisable in patients with proteinuria > 800 mg/day, and a cautious and 1
individual approach should be followed in patients with GFR < 30 mL/min.

Recommendations Strength rating


The m-TOR inhibitors may be used to prevent rejection in patients who are intolerant to Weak
standard therapy.
Significantly reduce calcineurin inhibitor dosage in a combination regimen with Strong
m-TOR inhibitors to prevent aggravated nephrotoxicity.
Do not convert patients with proteinuria and poor renal function to m-TOR inhibitors. Strong
Monitor blood-levels of both sirolimus and everolimus to allow for appropriate dose Strong
adjustment.

3.1.10.6 Induction with Interleukin-2 receptor antibodies


Basiliximab, a high-affinity anti-interleukin-2 (IL-2) receptor monoclonal antibody is approved for rejection
prophylaxis following organ transplantation [242, 243, 245, 333-337]. Basiliximab is given before

30 RENAL TRANSPLANTATION - TEXT UPDATE 2021


transplantation and on day four post-transplant. The drug is safe, and IL-2 receptor antibodies have been
shown in RCTs to reduce the prevalence of acute cellular rejection by approximately 40% [242, 243, 245,
333-335]. Meta-analyses [245, 333-335] have confirmed the efficacy, although no positive effect on patient
or graft survival could be demonstrated, large retrospective cohort studies and recent large prospective
studies suggest such a benefit [242, 243, 338, 339]. Several large controlled trials support the efficacy and
safety of quadruple therapy with tacrolimus, mycophenolate and steroids. Interleukin-2 receptor antibodies
may allow early steroid withdrawal [295], although higher rejection rates were described in some studies.
Most importantly, IL-2 receptor antibodies allow a substantial reduction in CNIs or steroids, while maintaining
excellent efficacy and renal function [242-245, 302, 333-335]. Therefore, this regimen is proposed as first-line
immunosuppression in patients with low to normal immunological risk [243, 339].

Recommendation Strength rating


Use interleukin-2 receptor antibodies for induction in patients with normal immunological Weak
risk in order to reduce incidence of acute rejection.

3.1.10.7 T-cell depleting induction therapy


Prophylactic immunosuppression regimens in many countries, particularly the U.S., use potent T-cell depleting
‘induction’ treatments [242, 243, 245, 333, 338, 340-343]. Most frequently, ATG is used for prevention of
rejection in immunological high-risk patients, as supported by meta-analysis [339], and recommended by
guidelines [243, 344]. In addition, these potent biological agents are used for the treatment of severe, steroid
resistant rejection episodes [340, 343].

Use of T-cell depleting antibodies in immunological low-risk patients has not been associated with improved
long-term outcomes but with an increased risk of severe opportunistic infections and malignancy, particularly
post-transplant lymphoproliferative disease [242, 243, 245, 333, 339-341]. Some centres use these agents to
provide effective rejection prophylaxis in order to facilitate steroid withdrawal [302, 338, 342].

Recommendation Strength rating


T-cell depleting antibodies may be used for induction therapy in immunologically high-risk Weak
patients.

3.1.10.8 Belatacept
Belatacept is a fusion protein, which effectively blocks the CD28 co-stimulatory pathway and thereby
prevents T-cell activation [277, 345, 346]. Belatacept is intravenously administered and indicated for use as
part of a CNI-free regimen together with basiliximab induction, MPA, and corticosteroids. Long-term data
from three randomised studies of de novo kidney transplant recipients demonstrated better renal function
vs. cyclosporine-based immunosuppression, although rates and grades of acute rejection were higher for
belatacept in the first year post-transplant [242, 245, 257, 277, 345-351]. In patients receiving a standard
deceased or living donor kidney, better graft survival was observed, while similar graft survival rates were
found with ECDs. Interestingly, belatacept-treated patients had better preserved histology and developed less
donor specific antibodies (DSA) compared to cyclosporine [352]. The long-term safety profile of belatacept
treated patients was similar to cyclosporine controls, less belatacept treated patients developed metabolic
complications or discontinued treatment due to adverse events [350, 351, 353, 354]. In addition, the option
of converting patients (either stable patients or due to CNI or m-TOR associated toxicity) was explored
with promising initial results [348, 355-357]. Specific safety signals include a higher rate of post-transplant
lymphoproliferative disorder (especially in Epstein-Barr virus (EBV) negative patients), more herpes infections,
and tuberculosis in patients from endemic areas [277, 345, 346]. Belatacept was approved in the U.S. and in
Europe for EBV positive patients, but is not yet available in many countries. Additional studies are ongoing to
fully explore the value of this compound.

Recommendation Strength rating


Belatacept may be used for immunosuppressive therapy in immunologically low-risk Weak
patients, who have a positive Epstein-Barr virus serology.

3.1.11 Immunological complications


Immunological rejection is a common cause of early and late transplant dysfunction [243, 358-362]. There is
great variation in the timing and severity of rejection episodes and how they respond to treatment. Today two

RENAL TRANSPLANTATION - TEXT UPDATE 2021 31


main types of immunological reactions are distinguished, T-cell mediated rejections (TCMR) and antibody-
mediated rejections (ABMR) [243, 358-360]. Antibody-mediated rejection and TCMR may be diagnosed
together, called mixed acute rejection. Antibody-mediated rejection may occur as hyperacute rejection (HAR),
active rejection or chronic rejection. Chronic ABMR is considered as one of the leading causes of late graft
loss.

The ultimate standard for the diagnosis of rejection is transplant biopsy [243], because it is impossible to
differentiate acute rejection solely on clinical indicators from other causes of renal dysfunction (e.g. acute
tubular necrosis, infection, disease recurrence or CNI nephrotoxicity). Therefore, all rejections should be
verified by renal biopsy and biopsies should be classified according to the most recent Banff criteria [363],
which are the basis for prognosis and treatment [241, 358, 361]. Renal transplant biopsy should be conducted
preferably under US control, using an automated needle biopsy system (e.g. Tru-Cut biopsy gun) [243, 358]
with a 16 G needle to assure specimen adequacy. The biopsy procedure is considered safe but complications
such as bleeding and AV fistulas may occur [243, 364, 365]. The reported risk of major complications
(including substantial bleeding, macroscopic haematuria with ureteric obstruction, peritonitis or graft loss) is
approximately 1%. Most important contraindications are anti-coagulant therapy including anti-platelet agents
and uncontrolled hypertension.

Summary of evidence LE
There must be routine access to US-guided biopsy of the transplant and sufficient expertise in the 2
hospital pathology department to allow a rapid and clear-cut diagnosis of rejection or other type of
allograft dysfunction.
Steroid treatment for rejection may start before the renal biopsy is performed. 2

Recommendations Strength rating


Monitor transplant recipients for signs of acute rejection, particularly during the first six Strong
months post-transplant.
Take regular blood samples in addition to regular monitoring of urine output and ultrasound Strong
examinations in order to detect graft dysfunction during hospitalisation.
Immediately rule out other potential causes of graft dysfunction in cases of suspected acute Strong
rejection. An ultrasound of the kidney transplant should be performed.
Perform a renal biopsy, graded according to the most recent Banff criteria, in patients with Strong
suspected acute rejection episodes.
Only if contraindications to renal biopsy are present, can ‘blind’ steroid bolus therapy be Strong
given.
Test patients who suffer acute rejection as soon as possible for anti-HLA antibodies against Strong
the graft.
Reassess the immunosuppressive therapy of all patients with rejection, including patient Strong
adherence to the medication, which is of particular importance in late rejections.

3.1.11.1 Hyper-acute rejection


Hyper-acute rejection is the most dramatic and destructive immunological attack on the graft [230, 243, 358,
359]. It results from circulating, complement-fixing IgG antibodies, specifically reactive against incompatible
donor antigen, which engages with and destroys the vascular endothelium within minutes or hours after
vascularisation. It occurs in ABO-incompatible grafts due to the presence of high titres of pre-existing iso-
antibodies against blood group antigens. In ABO-matched grafts, HAR is mediated by anti-donor HLA IgG
antibodies. With the development of the cross-match test before transplantation, HAR has become an
extremely uncommon complication [230]. Imaging and histology reveals generalised infarction of the graft,
which has to be treated by graft nephrectomy. Therefore, prevention is crucial, either by avoidance of high iso-
antibodies against incompatible blood group antigens in case of an ABO-incompatible renal transplant and/or
by performing a regular cross-match before transplantation (see section 3.1.9).

Recommendation Strength rating


Prevent hyper-acute rejection by adequate ABO blood group and HLA matching of donor Strong
and recipients.

32 RENAL TRANSPLANTATION - TEXT UPDATE 2021


3.1.11.2 Treatment of T-cell mediated acute rejection
As only a few randomised trials have investigated different treatment options for this clinical problem, therapy
is mainly based on empirical experience rather than on clinical evidence [243, 343, 358, 366]. Parenteral
methylprednisolone (500 mg to 1 g) should be given intravenously as one pulse per day for three days. Anuria
or a steep rise in the serum creatinine may indicate steroid-refractory rejection and the need for another
three day course of pulsed methylprednisolone therapy [243, 358]. In addition, baseline immunosuppression
should be optimised to ensure adequate drug exposure [243, 358, 366]. In severe rejection, a conversion from
cyclosporine to tacrolimus and/or from azathioprine to MPA is recommended [243, 358].

T-cell depleting biological agents, such as ATG may be given in severe steroid-refractory cases [243, 340,
343, 358, 366]. If biological agents are used, other immunological suppression should be adapted and
daily T-cell monitoring should be considered to minimise the dose of the biological agent [340]. Before
immunosuppression is intensified, especially before the use of T-cell depleting agents, the prognosis of the
graft should be critically assessed against the risks of the aggravated immunosuppression. The patient should
be counselled adequately.

Recommendations Strength rating


Use steroid bolus therapy as first-line treatment for T-cell mediated rejection in addition to Strong
ensuring adequate baseline immunosuppression.
In severe or steroid-resistant rejection, use intensified immunosuppression, high-dose Strong
steroid treatment, and eventually T-cell depleting agents.

3.1.11.3 Treatment of antibody mediated rejection (ABMR)


Treatment of ABMR relies mainly on retrospective studies and empirical treatment guidelines [367]. Consensus
is that it is important to classify the clinical and histological phenotype of the rejection in order to make
adequate treatment decisions [367]. Important clinical factors are time of rejection (early acute < 30 days
post-transplant vs. late), preformed vs. de novo donor-specific antibodies (DSA), and histology (active vs.
chronic rejection).

For active ABMR due to pre-existing DSA treatment with a steroid bolus (at least three days of 500 mg/day)
in combination with intravenous immunoglobulin (IVIG) and plasmapheresis or immune-adsorption is
recommended. Intravenous immunoglobulin (IVIG) [243, 358, 368-373] may modulate and/or suppress antibody
production. Intravenous immunoglobulin alone seems insufficient for effective treatment and IVIG is used today
in a multimodal regimen. Dosages vary widely from 0.2-2.0 g/kg bodyweight, and no comparative studies (e.g.
on the dose or optimal concomitant immunosuppression) have been published. Retrospective and prospective
case series clearly suggest efficacy of antibody removal using plasmapheresis or immune-adsorption columns
[243, 358, 368-373], although details of the procedures vary widely. Adjunctive therapies such as complement
inhibitors, rituximab or splenectomy might be considered in severe early acute cases. Despite controversial
data on the utility of anti-CD20 antibody [243, 343, 358, 368-373], rituximab may also be considered as
adjunctive therapy in late active ABMR according to expert consensus. Although T-cell depleting agents such
as ATG appear to have limited value they are frequently used during mixed acute rejection [241]. However,
retrospective series suggests aggravated toxicity, when rituximab is combined with ATG [374], or steroids [343].
Furthermore, many centres will optimise maintenance therapy with MPA and steroids and sufficient tacrolimus
trough levels should be achieved [243, 358, 368-370, 373].

Chronic acute mixed rejection due to pre-existent DSA has no specific treatment recommendations except
for optimisation of maintenance therapy and eventually IVIG as an adjunctive treatment with a low level of
evidence. In patients presenting with de novo DSA optimisation of maintenance, immunosuppression is
recommended and non-adherence should be addressed and managed accordingly. If histology shows active
ABMR plasmapheresis, rituximab and IVIG can be considered as potential adjunctive agents without good
evidence from clinical trials. If biopsy demonstrates pure chronic ABMR no special treatment is recommended
due to lack of convincing data, except for IVIG as potential treatment option without firm evidence. Treatment
for chronic ABMR appears to be less successful [358, 368, 370].

In summary, several regimens have proven some efficacy in ABMR. However, except for a beneficial effect
of early antibody removal, the lack of firm evidence does not permit evidence-based recommendations
for treatment. As a consequence, prevention of ABMR by adequate pre-transplant screening, regular DSA
monitoring, avoidance of suboptimal immunosuppression and reinforcement of adherence are crucial [230,
358, 372, 375].

RENAL TRANSPLANTATION - TEXT UPDATE 2021 33


Recommendation Strength rating
Treatment of antibody mediated rejection should include antibody elimination. Strong

3.1.12 Follow-up after transplantation


Long-term graft function is of critical importance for the success of a transplant [243, 244]. Therefore, regular
long-term follow-up by experienced transplant physicians is essential in order to detect complications
or graft dysfunction early and reassure adherence to the immunosuppressive regimen. Complications of
immunosuppression occur frequently including specific complications of the different drugs as well as over
immunosuppression (namely opportunistic infections and malignancy) [243, 244]. The risk of cancer and
cardiac disease is several-fold higher in transplanted patients than in the general population. Cancer is a cause
of significant morbidity and mortality in the transplanted population [243, 376, 377]. Cardiovascular disease
is the most frequent cause of death in renal allograft recipients [243, 378, 379]. Other important long-term
problems are non-adherence, the development of anti-HLA antibodies, recurrence of the original disease and
CNI associated nephrotoxicity [243, 244].

3.1.12.1 Chronic allograft dysfunction/interstitial fibrosis and tubular atrophy


Many patients lose their grafts due to chronic allograft dysfunction [243, 244, 380]. Histology will usually reveal
a chronic process of interstitial fibrosis and tubular atrophy (IF/TA) [381]. Some patients will have immunological
chronic ABMR [382], as discussed in section 3.1.11.3. Interstitial fibrosis and tubular atrophy takes months
or years to develop and is heralded by proteinuria and hypertension, with a simultaneous or delayed rise in
serum creatinine level over months [243, 380, 381]. It is likely that IF/TA is more common in patients who have
had early attacks of acute rejection or infection. The main differential diagnosis is chronic nephrotoxicity [383],
which is common in patients receiving CNIs, and pre-existing and/or aggravated chronic kidney damage from a
marginal donor kidney [243, 380, 381].
Diagnosis is by renal biopsy [243, 380]. In patients diagnosed early, particularly if there is evidence
for CNI toxicity, disease progression may be slowed by conversion to a CNI-free regimen [201-203, 263, 264].
Conversion to m-TOR inhibitors is an option for patients without significant proteinuria (< 800 mg/day), but
moderate renal function [242-244]. Alternatively, successful conversion to a mycophenolate based regimen
has been described, especially in patients beyond the first three years post-transplant [242, 244, 278]. If there
is intolerance to m-TOR inhibitors or MPA, conversion to belatacept or an azathioprine-based regimen may be
successful, though the higher risk of rejection warrants close surveillance [357]. If the risk of rejection seems
too high, another option is substantial reduction of CNI under the protection of MPA [244, 278].
In patients with proteinuria, intervention with an angiotensin converting enzyme inhibitor, or
angiotensin II receptor blocker [243, 380] together with tight blood pressure control may slow down renal
progression. Other supportive measures include the treatment of hypertension, hyperlipidaemia, diabetes,
anaemia, acidosis, and bone disease [243]. However, ultimately, the patient will require another transplant (if fit
enough to go on the transplant waiting list) or dialysis therapy.

Summary of evidence LE
Regular long-term follow-up by experienced transplant physicians is essential in order to detect 4
complications or graft dysfunction early and reassure adherence to the immunosuppressive regimen.
Annual screening should include a dermatological examination, cardiovascular history and exam, 4
tumour screening (including a nodal examination, faecal occult screening, chest x-ray, gynaecological
and urological examination), and an abdominal US, including US of the native and transplanted kidney.
If appropriate, further diagnostic tests should be prompted to treat or slow down the progression of
any identified complication.
In patients diagnosed early with IF/TA, particularly if there is evidence for CNI toxicity, disease 1
progression may be slowed by conversion to a CNI-free regimen. If the risk of rejection seems too
high, another option is substantial reduction of CNI under the protection of MPA.
Supportive measures should aim to adequately treat the consequences of chronic kidney disease (e.g. 4
anaemia, acidosis, bone disease).

Recommendations Strength rating


Provide lifelong regular post-transplant follow-up by an experienced and trained transplant Strong
specialist at least every six to twelve months.
Advise patients on appropriate lifestyle changes, potential complications, and the Strong
importance of adherence to their immunosuppressive regimen.

34 RENAL TRANSPLANTATION - TEXT UPDATE 2021


Regularly monitor (approximately every four to eight weeks) serum creatinine, estimated Strong
glomerular filtration rate, blood pressure, urinary protein excretion, immunosuppression and
complications after renal transplantation. Changes in these parameters over time should
trigger further diagnostic work-up including renal biopsy, a search for infectious causes and
anti-HLA antibodies.
Perform an ultrasound of the graft, in case of graft dysfunction, to rule out obstruction and Strong
renal artery stenosis.
In patients with interstitial fibrosis and tubular atrophy undergoing calcineurin inhibitor (CNI) Strong
therapy and/or with histological signs suggestive for CNI toxicity (e.g. arteriolar hyalinosis,
striped fibrosis) consider CNI reduction or withdrawal.
Initiate appropriate medical treatment, e.g. tight control of hypertension, diabetes, Strong
proteinuria, cardiac risk factors, infections, and other complications according to current
guidelines.

4. REFERENCES
1. Guyatt, G.H., et al. GRADE: an emerging consensus on rating quality of evidence and strength of
recommendations. BMJ, 2008. 336: 924.
https://pubmed.ncbi.nlm.nih.gov/18436948
2. Guyatt, G.H., et al. What is “quality of evidence” and why is it important to clinicians? BMJ, 2008.
336: 995.
https://pubmed.ncbi.nlm.nih.gov/18456631
3. Phillips B, et al. Oxford Centre for Evidence-based Medicine Levels of Evidence. Updated by
Jeremy Howick March 2009. 1998.
https://www.cebm.net/2009/06/oxford-centre-evidence-based-medicine-levels-evidence-march-2009/
4. Guyatt, G.H., et al. Going from evidence to recommendations. BMJ, 2008. 336: 1049.
https://pubmed.ncbi.nlm.nih.gov/18467413
5. Boissier, R., et al. Benefits and harms of benign prostatic obstruction treatments in renal
transplanted patients. PROSPERO, 2019. CRD42019136477.
https://www.crd.york.ac.uk/prospero/display_record.php?ID=CRD42019136477
6. Boissier, R., et al. Effectiveness of interventions on nephrolithiasis in transplanted kidney.
PROSPERO, 2019. CRD42019136474.
https://www.crd.york.ac.uk/prospero/display_record.php?ID=CRD42019136474
7. Lennerling, A., et al. Living organ donation practices in Europe - results from an online survey.
Transpl Int, 2013. 26: 145.
https://pubmed.ncbi.nlm.nih.gov/23198985
8. Antcliffe, D., et al. A meta-analysis of mini-open versus standard open and laparoscopic living donor
nephrectomy. Transpl Int, 2009. 22: 463.
https://pubmed.ncbi.nlm.nih.gov/19175543
9. Greco, F., et al. Laparoscopic living-donor nephrectomy: analysis of the existing literature. Eur Urol,
2010. 58: 498.
https://pubmed.ncbi.nlm.nih.gov/19175543
10. Wilson, C.H., et al. Laparoscopic versus open nephrectomy for live kidney donors. Cochrane
Database Syst Rev, 2011: CD006124.
https://pubmed.ncbi.nlm.nih.gov/22071829
11. Yuan, H., et al. The safety and efficacy of laparoscopic donor nephrectomy for renal transplantation:
an updated meta-analysis. Transplant Proc, 2013. 45: 65.
https://pubmed.ncbi.nlm.nih.gov/23375276
12. Serrano, O.K., et al. Evolution of Living Donor Nephrectomy at a Single Center: Long-Term
Outcomes with 4 Different Techniques in Greater Than 4000 Donors over 50 Years. Transplantation,
2016. 100: 1299.
https://pubmed.ncbi.nlm.nih.gov/27136265
13. Breda, A., et al. Mini-laparoscopic live donor nephrectomy with the use of 3-mm instruments and
laparoscope. World J Urol, 2015. 33: 707.
https://pubmed.ncbi.nlm.nih.gov/25182807

RENAL TRANSPLANTATION - TEXT UPDATE 2021 35


14. Elmaraezy, A., et al. Should hand-assisted retroperitoneoscopic nephrectomy replace the standard
laparoscopic technique for living donor nephrectomy? A meta-analysis. International Journal of
Surgery, 2017. 40: 83.
https://pubmed.ncbi.nlm.nih.gov/28216391
15. Creta, M., et al. Donor and Recipient Outcomes following Robotic-Assisted Laparoscopic Living
Donor Nephrectomy: A Systematic Review. Biomed Res Int, 2019. 2019: 1729138.
https://pubmed.ncbi.nlm.nih.gov/31143770/
16. Lentine, K.L., et al. Perioperative Complications After Living Kidney Donation: A National Study. Am
J Transplant, 2016. 16: 1848.
https://pubmed.ncbi.nlm.nih.gov/26700551
17. Wang, H., et al. Robot-assisted laparoscopic vs laparoscopic donor nephrectomy in renal
transplantation: A meta-analysis. Clin Transplant, 2019. 33: e13451.
https://pubmed.ncbi.nlm.nih.gov/30461073/
18. Autorino, R., et al. Laparoendoscopic single-site (LESS) vs laparoscopic living-donor nephrectomy:
a systematic review and meta-analysis. BJU Int, 2015. 115: 206.
https://pubmed.ncbi.nlm.nih.gov/24588876
19. Gupta, A., et al. Laparoendoscopic single-site donor nephrectomy (LESS-DN) versus standard
laparoscopic donor nephrectomy [Systematic Review]. Cochrane Database Syst Rev, 2016. 6: 6.
https://pubmed.ncbi.nlm.nih.gov/27230690
20. Alcaraz, A., et al. Feasibility of transvaginal natural orifice transluminal endoscopic surgery-assisted
living donor nephrectomy: is kidney vaginal delivery the approach of the future? Eur Urol, 2011.
59: 1019.
https://pubmed.ncbi.nlm.nih.gov/21458151
21. Liu, N., et al. Maximizing the donor pool: left versus right laparoscopic live donor nephrectomy--
systematic review and meta-analysis. Int Urol Nephrol, 2014. 46: 1511.
https://pubmed.ncbi.nlm.nih.gov/27230690/24595603
22. Khalil, A., et al. Trends and outcomes in right vs. left living donor nephrectomy: An analysis of the
OPTN/UNOS database of donor and recipient outcomes - should we be doing more right-sided
nephrectomies? Clinical Transplantation, 2016. 30: 145.
https://pubmed.ncbi.nlm.nih.gov/26589133
23. Hsi, R.S., et al. Analysis of techniques to secure the renal hilum during laparoscopic donor
nephrectomy: review of the FDA database. Urology, 2009. 74: 142.
24. Hsi, R.S., et al. Mechanisms of hemostatic failure during laparoscopic nephrectomy: review of Food
and Drug Administration database. Urology, 2007. 70: 888.
https://pubmed.ncbi.nlm.nih.gov/19406458
25. Ponsky, L., et al. The Hem-o-lok clip is safe for laparoscopic nephrectomy: a multi-institutional
review. Urology, 2008. 71: 593.
https://pubmed.ncbi.nlm.nih.gov/18295866
26. Allen, M.B., et al. Donor hemodynamics as a predictor of outcomes after kidney transplantation from
donors after cardiac death. Am J Transplant, 2016. 16: 181.
https://pubmed.ncbi.nlm.nih.gov/26361242
27. Heylen, L., et al. The duration of asystolic ischemia determines the risk of graft failure after
circulatory-dead donor kidney transplantation: A Eurotransplant cohort study. Am J Transplant,
2018. 18: 881.
https://pubmed.ncbi.nlm.nih.gov/28980391
28. Osband, A.J., et al. Extraction Time of Kidneys from Deceased Donors and Impact on Outcomes.
Am J of Transplant, 2016. 16: 700.
https://pubmed.ncbi.nlm.nih.gov/26414911
29. Redfield, R.R., et al. Predictors and outcomes of delayed graft function after living-donor kidney
transplantation. Transpl Int, 2016. 29: 81.
https://pubmed.ncbi.nlm.nih.gov/26432507
30. Irish, W.D., et al. A risk prediction model for delayed graft function in the current era of deceased
donor renal transplantation. Am J Transplant, 2010. 10: 2279.
https://pubmed.ncbi.nlm.nih.gov/20883559
31. de Boer, J., et al. Eurotransplant randomized multicenter kidney graft preservation study comparing
HTK with UW and Euro-Collins. Transpl Int, 1999. 12: 447.
https://pubmed.ncbi.nlm.nih.gov/10654357
32. Parsons, R.F., et al. Preservation solutions for static cold storage of abdominal allografts: which is
best? Curr Opin Organ Transplant, 2014. 19: 100.
https://pubmed.ncbi.nlm.nih.gov/24553501

36 RENAL TRANSPLANTATION - TEXT UPDATE 2021


33. Tillou, X., et al. Comparison of UW and Celsior: long-term results in kidney transplantation. Ann
Transplant, 2013. 18: 146.
https://pubmed.ncbi.nlm.nih.gov/23792514
34. Barnett, D., Black, D. W., Buckley, B., Campbell, D., Clarke, P.,. Machine perfusion systems and cold
static storage of kidneys from deceased donors. NICE Guidelines. Technology appraisal guidance
2009.
https://www.nice.org.uk/guidance/ta165
35. Kay, M.D., et al. Comparison of preservation solutions in an experimental model of organ cooling in
kidney transplantation. Br J Surg, 2009. 96: 1215.
https://pubmed.ncbi.nlm.nih.gov/19787767
36. Bond, M., et al. The effectiveness and cost-effectiveness of methods of storing donated kidneys from
deceased donors: a systematic review and economic model. Health Technol Assess, 2009. 13: iii.
https://pubmed.ncbi.nlm.nih.gov/19674537
37. Lledo-Garcia, E., et al. Spanish consensus document for acceptance and rejection of kidneys from
expanded criteria donors. Clin Transplant, 2014. 28: 1155.
https://pubmed.ncbi.nlm.nih.gov/25109314
38. Johnston, T.D., et al. Sensitivity of expanded-criteria donor kidneys to cold ischaemia time. Clin
Transplant, 2004. 18 Suppl 12: 28.
https://pubmed.ncbi.nlm.nih.gov/15217404
39. Peters-Sengers, H., et al. Impact of Cold Ischemia Time on Outcomes of Deceased Donor Kidney
Transplantation: An Analysis of a National Registry. Transplant Direct, 2019. 5: e448.
https://pubmed.ncbi.nlm.nih.gov/31165083/
40. Summers, D.M., et al. Analysis of factors that affect outcome after transplantation of kidneys
donated after cardiac death in the UK: a cohort study. Lancet, 2010. 376: 1303.
https://pubmed.ncbi.nlm.nih.gov/20727576
41. Aubert, O., et al. Long term outcomes of transplantation using kidneys from expanded criteria
donors: prospective, population based cohort study. BMJ, 2015. 351: h3557.
https://pubmed.ncbi.nlm.nih.gov/26232393
42. Kayler, L.K., et al. Impact of cold ischemia time on graft survival among ECD transplant recipients: a
paired kidney analysis. Am J Transplant, 2011. 11: 2647.
https://pubmed.ncbi.nlm.nih.gov/21906257
43. Chatauret, N., et al. Preservation strategies to reduce ischemic injury in kidney transplantation:
pharmacological and genetic approaches. Curr Opin Organ Transplant, 2011. 16: 180.
https://pubmed.ncbi.nlm.nih.gov/21415820
44. Jochmans, I., et al. Past, Present, and Future of Dynamic Kidney and Liver Preservation and
Resuscitation. Am J Transplant, 2016. 16: 2545.
https://pubmed.ncbi.nlm.nih.gov/26946212
45. O’Callaghan, J.M., et al. Systematic review and meta-analysis of hypothermic machine perfusion
versus static cold storage of kidney allografts on transplant outcomes. Br J Surg, 2013. 100: 991.
https://pubmed.ncbi.nlm.nih.gov/23754643
46. Martinez Arcos, L., et al. Functional Results of Renal Preservation in Hypothermic Pulsatile Machine
Perfusion Versus Cold Preservation: Systematic Review and Meta-Analysis of Clinical Trials.
Transplant Proc, 2018. 50: 24.
https://pubmed.ncbi.nlm.nih.gov/29407316/
47. Tingle, S.J., et al. Machine perfusion preservation versus static cold storage for deceased donor
kidney transplantation. Cochrane Database Syst Rev, 2019. 3: CD011671.
https://pubmed.ncbi.nlm.nih.gov/30875082
48. Jochmans, I., et al. Machine perfusion versus cold storage for the preservation of kidneys donated
after cardiac death: a multicenter, randomized, controlled trial. Ann Surg, 2010. 252: 756.
https://pubmed.ncbi.nlm.nih.gov/21037431
49. Reznik, O.N., et al. Machine perfusion as a tool to select kidneys recovered from uncontrolled
donors after cardiac death. Transplant Proc, 2008. 40: 1023.
https://pubmed.ncbi.nlm.nih.gov/18555105
50. Jochmans, I., et al. Hypothermic machine perfusion of kidneys retrieved from standard and high-risk
donors. Transpl Int, 2015. 28: 665.
https://pubmed.ncbi.nlm.nih.gov/25630347
51. Treckmann, J., et al. Machine perfusion versus cold storage for preservation of kidneys from
expanded criteria donors after brain death. Transpl Int, 2011. 24: 548.
https://pubmed.ncbi.nlm.nih.gov/21332580

RENAL TRANSPLANTATION - TEXT UPDATE 2021 37


52. Gill, J., et al. Pulsatile perfusion reduces the risk of delayed graft function in deceased donor kidney
transplants, irrespective of donor type and cold ischemic time. Transplantation, 2014. 97: 668.
https://pubmed.ncbi.nlm.nih.gov/24637865
53. Matsuno, N., et al. Machine perfusion preservation for kidney grafts with a high creatinine from
uncontrolled donation after cardiac death. Transplant Proc, 2010. 42: 155.
https://pubmed.ncbi.nlm.nih.gov/20172304
54. Jochmans, I., et al. Graft quality assessment in kidney transplantation: not an exact science yet!
Curr Opin Organ Transplant, 2011. 16: 174.
https://pubmed.ncbi.nlm.nih.gov/21383549
55. Jochmans, I., et al. Oxygenated Hypothermic Machine Perfusion of Kidneys Donated after Circulatory
Death: An International Randomised Controlled Trial [abstract]. Am J Transplant, 2019. 19.
https://atcmeetingabstracts.com/abstract/oxygenated-hypothermic-machine-perfusion-of-kidneys-
donated-after-circulatory-death-an-international-randomised-controlled-trial/
56. Hosgood, S.A., et al. Normothermic machine perfusion of the kidney: better conditioning and
repair? Transpl Int, 2015. 28: 657.
https://pubmed.ncbi.nlm.nih.gov/24629095
57. Reddy, S.P., et al. Normothermic perfusion: a mini-review. Transplantation, 2009. 87: 631.
https://pubmed.ncbi.nlm.nih.gov/19295304
58. Antoine, C., et al. Kidney Transplant From Uncontrolled Donation After Circulatory Death:
Contribution of Normothermic Regional Perfusion. Transplantation, 2020. 104: 130.
https://pubmed.ncbi.nlm.nih.gov/30985577
59. Reznik, O., et al. Kidney from uncontrolled donors after cardiac death with one hour warm ischemic
time: resuscitation by extracorporal normothermic abdominal perfusion “in situ” by leukocytes-free
oxygenated blood. Clin Transplant, 2011. 25: 511.
https://pubmed.ncbi.nlm.nih.gov/20973824
60. Hosgood, S.A., et al. Ex vivo normothermic perfusion for quality assessment of marginal donor
kidney transplants. Br J Surg, 2015. 102: 1433.
https://pubmed.ncbi.nlm.nih.gov/26313559
61. Hoyer, D.P., et al. Subnormothermic machine perfusion for preservation of porcine kidneys in a
donation after circulatory death model. Transpl Int, 2014. 27: 1097.
https://pubmed.ncbi.nlm.nih.gov/24963744
62. Naesens, M. Zero-Time Renal Transplant Biopsies: A Comprehensive Review. Transplantation, 2016.
100: 1425.
https://pubmed.ncbi.nlm.nih.gov/26599490
63. Kasiske, B.L., et al. The role of procurement biopsies in acceptance decisions for kidneys retrieved
for transplant. Clin J Am Soc Nephrol, 2014. 9: 562.
https://pubmed.ncbi.nlm.nih.gov/24558053
64. Marrero, W.J., et al. Predictors of Deceased Donor Kidney Discard in the United States.
Transplantation, 2016.
https://pubmed.ncbi.nlm.nih.gov/27163541
65. Sung, R.S., et al. Determinants of discard of expanded criteria donor kidneys: impact of biopsy and
machine perfusion. Am J Transplant, 2008. 8: 783.
https://pubmed.ncbi.nlm.nih.gov/18294347
66. Wang, C.J., et al. The Donor Kidney Biopsy and Its Implications in Predicting Graft Outcomes:
A Systematic Review. Am J Transplant, 2015. 15: 1903.
https://pubmed.ncbi.nlm.nih.gov/25772854
67. Hopfer, H., et al. Assessment of donor biopsies. Curr Opin Organ Transplant, 2013. 18: 306.
https://pubmed.ncbi.nlm.nih.gov/23492644
68. Gaber, L.W., et al. Glomerulosclerosis as a determinant of posttransplant function of older donor
renal allografts. Transplantation, 1995. 60: 334.
https://pubmed.ncbi.nlm.nih.gov/7652761
69. Solez, K., et al. Banff 07 classification of renal allograft pathology: updates and future directions. Am
J Transplant, 2008. 8: 753.
https://pubmed.ncbi.nlm.nih.gov/18294345
70. De Vusser, K., et al. The predictive value of kidney allograft baseline biopsies for long-term graft
survival. J Am Soc Nephrol, 2013. 24: 1913.
https://pubmed.ncbi.nlm.nih.gov/23949799
71. Anglicheau, D., et al. A simple clinico-histopathological composite scoring system is highly
predictive of graft outcomes in marginal donors. Am J Transplant, 2008. 8: 2325.
https://pubmed.ncbi.nlm.nih.gov/18785957

38 RENAL TRANSPLANTATION - TEXT UPDATE 2021


72. Balaz, P., et al. Identification of expanded-criteria donor kidney grafts at lower risk of delayed graft
function. Transplantation, 2013. 96: 633.
https://pubmed.ncbi.nlm.nih.gov/23912171
73. Lopes, J.A., et al. Evaluation of pre-implantation kidney biopsies: comparison of Banff criteria to a
morphometric approach. Kidney Int, 2005. 67: 1595.
https://pubmed.ncbi.nlm.nih.gov/15780116
74. Munivenkatappa, R.B., et al. The Maryland aggregate pathology index: a deceased donor kidney
biopsy scoring system for predicting graft failure. Am J Transplant, 2008. 8: 2316.
https://pubmed.ncbi.nlm.nih.gov/18801024
75. Liapis, H., et al. Banff Histopathological Consensus Criteria for Preimplantation Kidney Biopsies. Am
J Transplant, 2016.
https://pubmed.ncbi.nlm.nih.gov/27333454
76. Haas, M. Donor kidney biopsies: pathology matters, and so does the pathologist. Kidney Int, 2014.
85: 1016.
https://pubmed.ncbi.nlm.nih.gov/24786876
77. Azancot, M.A., et al. The reproducibility and predictive value on outcome of renal biopsies from
expanded criteria donors. Kidney Int, 2014. 85: 1161.
https://pubmed.ncbi.nlm.nih.gov/24284518
78. Peters, B., et al. Sixteen Gauge biopsy needles are better and safer than 18 Gauge in native and
transplant kidney biopsies. Acta Radiol, 2017. 58: 240.
https://pubmed.ncbi.nlm.nih.gov/27055922
79. Haas, M., et al. Arteriosclerosis in kidneys from healthy live donors: comparison of wedge and
needle core perioperative biopsies. Arch Pathol Lab Med, 2008. 132: 37.
https://pubmed.ncbi.nlm.nih.gov/18181671
80. Mazzucco, G., et al. The reliability of pre-transplant donor renal biopsies (PTDB) in predicting the
kidney state. A comparative single-centre study on 154 untransplanted kidneys. Nephrol Dial
Transplant, 2010. 25: 3401.
https://pubmed.ncbi.nlm.nih.gov/20356979
81. Wang, H.J., et al. On the influence of sample size on the prognostic accuracy and reproducibility of
renal transplant biopsy. Nephrol Dial Transplant, 1998. 13: 165.
https://pubmed.ncbi.nlm.nih.gov/9481734
82. Yushkov, Y., et al. Optimized technique in needle biopsy protocol shown to be of greater sensitivity
and accuracy compared to wedge biopsy. Transplant Proc, 2010. 42: 2493.
https://pubmed.ncbi.nlm.nih.gov/20832530
83. Muruve, N.A., et al. Are wedge biopsies of cadaveric kidneys obtained at procurement reliable?
Transplantation, 2000. 69: 2384.
https://pubmed.ncbi.nlm.nih.gov/10868645
84. Randhawa, P. Role of donor kidney biopsies in renal transplantation. Transplantation, 2001. 71: 1361.
https://pubmed.ncbi.nlm.nih.gov/11391219
85. Bago-Horvath, Z., et al. The cutting (w)edge--comparative evaluation of renal baseline biopsies
obtained by two different methods. Nephrol Dial Transplant, 2012. 27: 3241.
https://pubmed.ncbi.nlm.nih.gov/22492825
86. Jankovic, Z. Anaesthesia for living-donor renal transplant. Current Anaesthesia & Critical Care, 2008.
19: 175.
https://www.researchgate.net/publication/270283251_Jankovic_Z_Anaesthesia_for_living-donor_
renal_transplant_Curr_Anaesth_Crit_Care_2008_19_3_175-80
87. Karmarkar, S., et al. Kidney Transplantation. Anaesthesia And Intensive Care Medicine 2009. 10.5.
https://www.anaesthesiajournal.co.uk/article/S1472-0299(12)00070-7/abstract
88. Abramowicz, D., et al. European Renal Best Practice Guideline on kidney donor and recipient
evaluation and perioperative care. Nephrol Dial Transplant, 2015. 30: 1790.
https://pubmed.ncbi.nlm.nih.gov/25007790
89. Van Loo, A.A., et al. Pretransplantation hemodialysis strategy influences early renal graft function.
J Am Soc Nephrol, 1998. 9: 473.
https://pubmed.ncbi.nlm.nih.gov/9513911
90. Task Force for Preoperative Cardiac Risk, A., et al. Guidelines for pre-operative cardiac risk
assessment and perioperative cardiac management in non-cardiac surgery. Eur Heart J, 2009.
30: 2769.
https://pubmed.ncbi.nlm.nih.gov/24126879

RENAL TRANSPLANTATION - TEXT UPDATE 2021 39


91. Douketis, J.D., et al. Perioperative Management of Antithrombotic Therapy: Antithrombotic Therapy
and Prevention of Thrombosis, 9th ed: American College of Chest Physicians Evidence-Based
Clinical Practice Guidelines. Chest, 2012. 141.
https://pubmed.ncbi.nlm.nih.gov/22315266
92. Benahmed, A., et al. Ticlopidine and clopidogrel, sometimes combined with aspirin, only minimally
increase the surgical risk in renal transplantation: A case-control study. Nephrol Dial Transplant,
2014. 29: 463.
https://pubmed.ncbi.nlm.nih.gov/24275542
93. Osman, Y., et al. Necessity of Routine Postoperative Heparinization in Non-Risky Live-Donor Renal
Transplantation: Results of a Prospective Randomized Trial. Urology, 2007. 69: 647.
https://pubmed.ncbi.nlm.nih.gov/17445644
94. Orlando, G., et al. One-shot versus multidose perioperative antibiotic prophylaxis after kidney
transplantation: a randomized, controlled clinical trial. Surgery, 2015. 157: 104.
https://pubmed.ncbi.nlm.nih.gov/25304836
95. Choi, S.U., et al. Clinical significance of prophylactic antibiotics in renal transplantation. Transplant
Proc, 2013. 45: 1392.
https://pubmed.ncbi.nlm.nih.gov/23726580
96. O’Malley, C.M., et al. A randomized, double-blind comparison of lactated Ringer’s solution and
0.9% NaCl during renal transplantation. Anesth Analg, 2005. 100: 1518.
https://pubmed.ncbi.nlm.nih.gov/15845718
97. Othman, M.M., et al. The impact of timing of maximal crystalloid hydration on early graft function
during kidney transplantation. Anesth Analg, 2010. 110: 1440.
https://pubmed.ncbi.nlm.nih.gov/20418304
98. Dalton, R.S., et al. Physiologic impact of low-dose dopamine on renal function in the early post renal
transplant period. Transplantation, 2005. 79: 1561.
https://pubmed.ncbi.nlm.nih.gov/15940046
99. Ciapetti, M., et al. Low-dose dopamine in kidney transplantation. Transplant Proc, 2009. 41: 4165.
https://pubmed.ncbi.nlm.nih.gov/20005360
100. Hanif, F., et al. Outcome of renal transplantation with and without intra-operative diuretics. Int
J Surg, 2011. 9: 460.
https://pubmed.ncbi.nlm.nih.gov/21600319
101. Valeriani, G., et al. Bench surgery in right kidney transplantation. Transplant Proc, 2010. 42: 1120.
https://pubmed.ncbi.nlm.nih.gov/20534239
102. Wagenaar, S., et al. Minimally Invasive, Laparoscopic, and Robotic-assisted Techniques Versus
Open Techniques for Kidney Transplant Recipients: A Systematic Review. Eur Urol, 2017. 72: 205.
https://pubmed.ncbi.nlm.nih.gov/28262412
103. Chedid, M.F., et al. Living donor kidney transplantation using laparoscopically procured multiple
renal artery kidneys and right kidneys. J Am Coll Surg, 2013. 217: 144.
https://pubmed.ncbi.nlm.nih.gov/23791283
104. Kaminska, D., et al. The influence of warm ischemia elimination on kidney injury during
transplantation - clinical and molecular study. Sci Rep, 2016. 6: 36118.
https://pubmed.ncbi.nlm.nih.gov/27808277
105. Ozdemir-van Brunschot, D.M., et al. Is the Reluctance for the Implantation of Right Donor Kidneys
Justified? World J Surg, 2016. 40: 471.
https://pubmed.ncbi.nlm.nih.gov/26319261
106. Khalil, A., et al. Trends and outcomes in right vs. left living donor nephrectomy: an analysis of the
OPTN/UNOS database of donor and recipient outcomes--should we be doing more right-sided
nephrectomies? Clin Transplant, 2016. 30: 145.
https://pubmed.ncbi.nlm.nih.gov/26589133
107. Hsu, J.W., et al. Increased early graft failure in right-sided living donor nephrectomy. Transplantation,
2011. 91: 108.
https://pubmed.ncbi.nlm.nih.gov/21441855
108. Kulkarni, S., et al. Outcomes From Right Versus Left Deceased-Donor Kidney Transplants: A US
National Cohort Study. Am J Kidney Dis, 2020. 75: 725.
https://pubmed.ncbi.nlm.nih.gov/31812448/
109. Wang, K., et al. Right Versus Left Laparoscopic Living-Donor Nephrectomy: A Meta-Analysis. Exp
Clin Transplant, 2015. 13: 214.
https://pubmed.ncbi.nlm.nih.gov/26086831

40 RENAL TRANSPLANTATION - TEXT UPDATE 2021


110. Ciudin, A., et al. Transposition of iliac vessels in implantation of right living donor kidneys. Transplant
Proc, 2012. 44: 2945.
https://pubmed.ncbi.nlm.nih.gov/23195003
111. Phelan, P.J., et al. Left versus right deceased donor renal allograft outcome. Transpl Int, 2009.
22: 1159.
https://pubmed.ncbi.nlm.nih.gov/19891044
112. Feng, J.Y., et al. Renal vein lengthening using gonadal vein reduces surgical difficulty in living-donor
kidney transplantation. World J Surg, 2012. 36: 468.
https://pubmed.ncbi.nlm.nih.gov/21882021
113. Nghiem, D.D. Use of spiral vein graft in living donor renal transplantation. Clin Transplant, 2008. 22: 719.
https://pubmed.ncbi.nlm.nih.gov/18673376
114. Matheus, W.E., et al. Kidney transplant anastomosis: internal or external iliac artery? Urol J, 2009.
6: 260.
https://pubmed.ncbi.nlm.nih.gov/20027554
115. El-Sherbiny, M., et al. The use of the inferior epigastric artery for accessory lower polar artery
revascularization in live donor renal transplantation. Int Urol Nephrol, 2008. 40: 283.
https://pubmed.ncbi.nlm.nih.gov/17721826
116. Firmin, L.C., et al. The use of explanted internal iliac artery grafts in renal transplants with multiple
arteries. Transplantation, 2010. 89: 766.
https://pubmed.ncbi.nlm.nih.gov/20308866
117. Oertl, A.J., et al. Saphenous vein interposition as a salvage technique for complex vascular
situations during renal transplantation. Transplant Proc, 2007. 39: 140.
https://pubmed.ncbi.nlm.nih.gov/17275492
118. Tozzi, M., et al. Treatment of aortoiliac occlusive or dilatative disease concomitant with kidney
transplantation: how and when? Int J Surg, 2013. 11 Suppl 1: S115.
https://pubmed.ncbi.nlm.nih.gov/24380542
119. Franchin, M., et al. ePTFE suture is an effective tool for vascular anastomosis in kidney
transplantation. Ital J Vasc Endovasc Surg, 2015. 22: 61.
https://www.researchgate.net/publication/285219004_ePTFE_suture_is_an_effective_tool_for_
vascular_anastomosis_in_kidney_transplantation
120. Izquierdo, L., et al. Third and fourth kidney transplant: still a reasonable option. Transplant Proc,
2010. 42: 2498.
https://pubmed.ncbi.nlm.nih.gov/20832531
121. Blanco, M., et al. Third kidney transplantation: a permanent medical-surgical challenge. Transplant
Proc, 2009. 41: 2366.
https://pubmed.ncbi.nlm.nih.gov/19715921
122. Nourbala, M.H., et al. Our experience with third renal transplantation: results, surgical techniques
and complications. Int J Urol, 2007. 14: 1057.
https://pubmed.ncbi.nlm.nih.gov/18036037
123. Musquera, M., et al. Orthotopic kidney transplantation: an alternative surgical technique in selected
patients. Eur Urol, 2010. 58: 927.
https://pubmed.ncbi.nlm.nih.gov/20888120
124. Heylen, L., et al. The Impact of Anastomosis Time During Kidney Transplantation on Graft Loss: A
Eurotransplant Cohort Study. American journal of transplantation : official journal of the American
Society of Transplantation and the American Society of Transplant Surgeons, 2017. 17: 724.
https://pubmed.ncbi.nlm.nih.gov/27593738
125. Weissenbacher, A., et al. The faster the better: anastomosis time influences patient survival after
deceased donor kidney transplantation. Transpl Int, 2015. 28: 535.
https://pubmed.ncbi.nlm.nih.gov/25557890
126. McCulloch, P., et al. IDEAL framework for surgical innovation 1: the idea and development stages.
BMJ, 2013. 346: f3012.
https://pubmed.ncbi.nlm.nih.gov/23778427
127. Breda, A., et al. Robot-assisted Kidney Transplantation: The European Experience [Figure presented].
Eur Urol, 2018. 73: 273.
https://www.europeanurology.com/article/S0302-2838(17)30721-2/pdf
128. Gallioli, A., et al. Learning Curve in Robot-assisted Kidney Transplantation: Results from the
European Robotic Urological Society Working Group. Eur Urol, 2020.
https://www.europeanurology.com/article/S0302-2838(19)30947-9/fulltext

RENAL TRANSPLANTATION - TEXT UPDATE 2021 41


129. Tzvetanov, I.G., et al. Robotic kidney transplantation in the obese patient: 10-year experience from a
single center. Am J Transplant, 2020. 20: 430.
https://pubmed.ncbi.nlm.nih.gov/31571369
130. Basu, A., et al. Adult dual kidney transplantation. Curr Opin Organ Transplant, 2007. 12: 379.
https://journals.lww.com/co-transplantation/Abstract/2007/08000/Adult_dual_kidney_
transplantation.10.aspx
131. Haider, H.H., et al. Dual kidney transplantation using midline extraperitoneal approach: description
of a technique. Transplant Proc, 2007. 39: 1118.
https://pubmed.ncbi.nlm.nih.gov/17524907
132. Ekser, B., et al. Technical aspects of unilateral dual kidney transplantation from expanded criteria
donors: experience of 100 patients. Am J Transplant, 2010. 10: 2000.
https://pubmed.ncbi.nlm.nih.gov/20636454
133. Nghiem, D.D. Simultaneous double adult kidney transplantation using single arterial and venous
anastomoses. Urology, 2006. 67: 1076.
https://pubmed.ncbi.nlm.nih.gov/16581114
134. Veroux, P., et al. Two-as-one monolateral dual kidney transplantation. Urology, 2011. 77: 227.
https://pubmed.ncbi.nlm.nih.gov/20399490
135. Salehipour, M., et al. En-bloc Transplantation: an Eligible Technique for Unilateral Dual Kidney
Transplantation. Int J Organ Transplant Med, 2012. 3: 111.
https://pubmed.ncbi.nlm.nih.gov/25013633
136. Rigotti, P., et al. A single-center experience with 200 dual kidney transplantations. Clin Transplant,
2014. 28: 1433.
https://pubmed.ncbi.nlm.nih.gov/25297945
137. Al-Shraideh, Y., et al. Single vs dual (en bloc) kidney transplants from donors </= 5 years of age: A
single center experience. World J Transplant, 2016. 6: 239.
https://pubmed.ncbi.nlm.nih.gov/27011923
138. Alberts, V.P., et al. Ureterovesical anastomotic techniques for kidney transplantation: a systematic
review and meta-analysis. Transpl Int, 2014. 27: 593.
https://pubmed.ncbi.nlm.nih.gov/24606191
139. Slagt, I.K., et al. A randomized controlled trial comparing intravesical to extravesical
ureteroneocystostomy in living donor kidney transplantation recipients. Kidney Int, 2014. 85: 471.
https://pubmed.ncbi.nlm.nih.gov/24284515
140. Timsit, M.O., et al. Should routine pyeloureterostomy be advocated in adult kidney transplantation?
A prospective study of 283 recipients. J Urol, 2010. 184: 2043.
https://pubmed.ncbi.nlm.nih.gov/20850818
141. Suttle, T., et al. Comparison of urologic complications between ureteroneocystostomy and
ureteroureterostomy in renal transplant: A meta-analysis. Exp Clin Transplant, 2016. 14: 276.
https://pubmed.ncbi.nlm.nih.gov/26925612
142. Dadkhah, F., et al. Modified ureteroneocystostomy in kidney transplantation to facilitate endoscopic
management of subsequent urological complications. Int Urol Nephrol, 2010. 42: 285.
https://pubmed.ncbi.nlm.nih.gov/19760513
143. Kehinde, E.O., et al. Complications associated with using nonabsorbable sutures for
ureteroneocystostomy in renal transplant operations. Transplant Proc, 2000. 32: 1917.
https://pubmed.ncbi.nlm.nih.gov/11119999
144. Wilson, C.H., et al. Routine intraoperative ureteric stenting for kidney transplant recipients.
Cochrane Database Syst Rev, 2013: CD004925.
https://pubmed.ncbi.nlm.nih.gov/23771708
145. Tavakoli, A., et al. Impact of stents on urological complications and health care expenditure in renal
transplant recipients: results of a prospective, randomized clinical trial. J Urol, 2007. 177: 2260.
https://pubmed.ncbi.nlm.nih.gov/17509336
146. Cai, J.F., et al. Meta-analysis of Early Versus Late Ureteric Stent Removal After Kidney
Transplantation. Transplant Proc, 2018. 50: 3411.
https://pubmed.ncbi.nlm.nih.gov/30577214
147. Visser, I.J., et al. Timing of Ureteric Stent Removal and Occurrence of Urological Complications after
Kidney Transplantation: A Systematic Review and Meta-Analysis. J Clin Med, 2019. 8.
https://pubmed.ncbi.nlm.nih.gov/30577214
148. Patel, P., et al. Prophylactic Ureteric Stents in Renal Transplant Recipients: A Multicenter
Randomized Controlled Trial of Early Versus Late Removal. Am J Transplant, 2017. 17: 2129.
https://pubmed.ncbi.nlm.nih.gov/28188678

42 RENAL TRANSPLANTATION - TEXT UPDATE 2021


149. Heidari, M., et al. Transplantation of kidneys with duplicated ureters. Scand J Urol Nephrol, 2010.
44: 337.
https://pubmed.ncbi.nlm.nih.gov/20653492
150. Alberts, V.P., et al. Duplicated ureters and renal transplantation: a case-control study and review of
the literature. Transplant Proc, 2013. 45: 3239.
https://pubmed.ncbi.nlm.nih.gov/24182792
151. Surange, R.S., et al. Kidney transplantation into an ileal conduit: a single center experience of
59 cases. J Urol, 2003. 170: 1727.
https://pubmed.ncbi.nlm.nih.gov/14532763
152. Kortram, K., et al. Perioperative Events and Complications in Minimally Invasive Live Donor
Nephrectomy: A Systematic Review and Meta-Analysis. Transplantation, 2016.
https://pubmed.ncbi.nlm.nih.gov/27428715
153. Segev, D.L., et al. Perioperative mortality and long-term survival following live kidney donation.
JAMA, 2010. 303: 959.
https://pubmed.ncbi.nlm.nih.gov/20215610
154. Chu, K.H., et al. Long-term outcomes of living kidney donors: a single centre experience of
29 years. Nephrology (Carlton), 2012. 17: 85.
https://pubmed.ncbi.nlm.nih.gov/21919999
155. Fehrman-Ekholm, I., et al. Post-nephrectomy development of renal function in living kidney donors:
a cross-sectional retrospective study. Nephrol Dial Transplant, 2011. 26: 2377.
https://pubmed.ncbi.nlm.nih.gov/21459783
156. Li, S.S., et al. A meta-analysis of renal outcomes in living kidney donors. [Review]. Medicine, 2016. 95.
https://pubmed.ncbi.nlm.nih.gov/27310964
157. Thiel, G.T., et al. Investigating kidney donation as a risk factor for hypertension and
microalbuminuria: findings from the Swiss prospective follow-up of living kidney donors. BMJ Open,
2016. 6: 22.
https://pubmed.ncbi.nlm.nih.gov/27006347
158. Ibrahim, H.N., et al. Long-term consequences of kidney donation. N Engl J Med, 2009. 360: 459.
https://pubmed.ncbi.nlm.nih.gov/27006347
159. Li, S.S., et al. A meta-analysis of renal outcomes in living kidney donors. Medicine (Baltimore), 2016.
95: e3847.
https://pubmed.ncbi.nlm.nih.gov/27310964
160. Matas, A.J., et al. Causes and timing of end-stage renal disease after living kidney donation. Am
J Transplant, 2018. 18: 1140.
https://pubmed.ncbi.nlm.nih.gov/29369517
161. Locke, J.E., et al. Obesity increases the risk of end-stage renal disease among living kidney donors.
Kidney Int, 2017. 91: 699.
https://pubmed.ncbi.nlm.nih.gov/28041626
162. Gross, C.R., et al. Health-related quality of life in kidney donors from the last five decades: results
from the RELIVE study. Am J Transplant, 2013. 13: 2924.
https://pubmed.ncbi.nlm.nih.gov/24011252
163. Maggiore, U., et al. Long-term risks of kidney living donation: Review and position paper by the
ERA-EDTA DESCARTES working group. Nephrol Dial Transplant, 2017. 32: 216.
https://pubmed.ncbi.nlm.nih.gov/28186535
164. Lorenz, E.C., et al. The impact of urinary tract infections in renal transplant recipients. Kidney Int,
2010. 78: 719.
https://pubmed.ncbi.nlm.nih.gov/20877371
165. Ariza-Heredia, E.J., et al. Urinary tract infections in kidney transplant recipients: role of gender,
urologic abnormalities, and antimicrobial prophylaxis. Ann Transplant, 2013. 18: 195.
https://pubmed.ncbi.nlm.nih.gov/23792521
166. Chang, C.Y., et al. Urological manifestations of BK polyomavirus in renal transplant recipients. Can
J Urol, 2005. 12: 2829.
https://pubmed.ncbi.nlm.nih.gov/16274519
167. Hwang, J.K., et al. Comparative analysis of ABO-incompatible living donor kidney transplantation
with ABO-compatible grafts: a single-center experience in Korea. Transplant Proc, 2013. 45: 2931.
https://pubmed.ncbi.nlm.nih.gov/24157006
168. Habicht, A., et al. Increase of infectious complications in ABO-incompatible kidney transplant
recipients--a single centre experience. Nephrol Dial Transplant, 2011. 26: 4124.
https://pubmed.ncbi.nlm.nih.gov/21622990

RENAL TRANSPLANTATION - TEXT UPDATE 2021 43


169. Sorto, R., et al. Risk factors for urinary tract infections during the first year after kidney
transplantation. Transplant Proc, 2010. 42: 280.
https://pubmed.ncbi.nlm.nih.gov/20172330
170. Thrasher, J.B., et al. Extravesical versus Leadbetter-Politano ureteroneocystostomy: a comparison
of urological complications in 320 renal transplants. J Urol, 1990. 144: 1105.
https://pubmed.ncbi.nlm.nih.gov/2231880
171. Mangus, R.S., et al. Stented versus nonstented extravesical ureteroneocystostomy in renal
transplantation: a metaanalysis. Am J Transplant, 2004. 4: 1889.
https://pubmed.ncbi.nlm.nih.gov/15476491
172. Wilson, C.H., et al. Routine intraoperative ureteric stenting for kidney transplant recipients.
Cochrane Database Syst Rev, 2005: CD004925.
https://pubmed.ncbi.nlm.nih.gov/16235385
173. Osman, Y., et al. Routine insertion of ureteral stent in live-donor renal transplantation: is it
worthwhile? Urology, 2005. 65: 867.
https://pubmed.ncbi.nlm.nih.gov/15882713
174. Georgiev, P., et al. Routine stenting reduces urologic complications as compared with stenting “on
demand” in adult kidney transplantation. Urology, 2007. 70: 893.
https://pubmed.ncbi.nlm.nih.gov/17919691
175. Akoh, J.A., et al. Effect of ureteric stents on urological infection and graft function following renal
transplantation. World J Transplant, 2013. 3: 1.
https://pubmed.ncbi.nlm.nih.gov/24175202
176. Fayek, S.A., et al. Ureteral stents are associated with reduced risk of ureteral complications after
kidney transplantation: a large single center experience. Transplantation, 2012. 93: 304.
https://pubmed.ncbi.nlm.nih.gov/22179401
177. Dimitroulis, D., et al. Vascular complications in renal transplantation: a single-center experience in
1367 renal transplantations and review of the literature. Transplant Proc, 2009. 41: 1609.
https://pubmed.ncbi.nlm.nih.gov/19545690
178. Pawlicki, J., et al. Risk factors for early hemorrhagic and thrombotic complications after kidney
transplantation. Transplant Proc, 2011. 43: 3013.
https://pubmed.ncbi.nlm.nih.gov/
179. Rouviere, O., et al. Acute thrombosis of renal transplant artery: graft salvage by means of intra-
arterial fibrinolysis. Transplantation, 2002. 73: 403.
https://pubmed.ncbi.nlm.nih.gov/11884937
180. Domagala, P., et al. Complications of transplantation of kidneys from expanded-criteria donors.
Transplant Proc, 2009. 41: 2970.
https://pubmed.ncbi.nlm.nih.gov/19857652
181. Ammi, M., et al. Evaluation of the Vascular Surgical Complications of Renal Transplantation. Annals
of Vascular Surgery, 2016. 33: 23.
https://pubmed.ncbi.nlm.nih.gov/26995525
182. Giustacchini, P., et al. Renal vein thrombosis after renal transplantation: an important cause of graft
loss. Transplant Proc, 2002. 34: 2126.
https://pubmed.ncbi.nlm.nih.gov/12270338
183. Wuthrich, R.P. Factor V Leiden mutation: potential thrombogenic role in renal vein, dialysis graft and
transplant vascular thrombosis. Curr Opin Nephrol Hypertens, 2001. 10: 409.
https://pubmed.ncbi.nlm.nih.gov/11342806
184. Parajuli, S., et al. Hypercoagulability in Kidney Transplant Recipients. Transplantation, 2016. 100: 719.
https://pubmed.ncbi.nlm.nih.gov/26413991
185. Granata, A., et al. Renal transplant vascular complications: the role of Doppler ultrasound.
J Ultrasound, 2015. 18: 101.
https://pubmed.ncbi.nlm.nih.gov/26191097
186. Hogan, J.L., et al. Late-onset renal vein thrombosis: A case report and review of the literature. Int
J Surg Case Rep, 2015. 6C: 73.
https://pubmed.ncbi.nlm.nih.gov/25528029
187. Musso, D., et al. Symptomatic Venous Thromboembolism and Major Bleeding After Renal
Transplantation: Should We Use Pharmacologic Thromboprophylaxis? Transplantation Proceedings,
2016. 48: 2773.
https://pubmed.ncbi.nlm.nih.gov/27788816
188. Hurst, F.P., et al. Incidence, predictors and outcomes of transplant renal artery stenosis after kidney
transplantation: analysis of USRDS. Am J Nephrol, 2009. 30: 459.
https://pubmed.ncbi.nlm.nih.gov/19776559

44 RENAL TRANSPLANTATION - TEXT UPDATE 2021


189. Willicombe, M., et al. Postanastomotic transplant renal artery stenosis: association with de novo
class II donor-specific antibodies. Am J Transplant, 2014. 14: 133.
https://pubmed.ncbi.nlm.nih.gov/24354873
190. Ghazanfar, A., et al. Management of transplant renal artery stenosis and its impact on long-term
allograft survival: a single-centre experience. Nephrol Dial Transplant, 2011. 26: 336.
https://pubmed.ncbi.nlm.nih.gov/20601365
191. Seratnahaei, A., et al. Management of transplant renal artery stenosis. Angiology, 2011. 62: 219.
https://pubmed.ncbi.nlm.nih.gov/20682611
192. Rountas, C., et al. Imaging modalities for renal artery stenosis in suspected renovascular
hypertension: prospective intraindividual comparison of color Doppler US, CT angiography,
GD-enhanced MR angiography, and digital substraction angiography. Ren Fail, 2007. 29: 295.
https://pubmed.ncbi.nlm.nih.gov/17497443
193. Fervenza, F.C., et al. Renal artery stenosis in kidney transplants. Am J Kidney Dis, 1998. 31: 142.
https://pubmed.ncbi.nlm.nih.gov/9428466
194. Bach, D., et al. Percutaneous renal biopsy: three years of experience with the biopty gun in 761
cases--a survey of results and complications. Int Urol Nephrol, 1999. 31: 15.
https://pubmed.ncbi.nlm.nih.gov/10408297
195. Loffroy, R., et al. Management of post-biopsy renal allograft arteriovenous fistulas with selective
arterial embolization: immediate and long-term outcomes. Clin Radiol, 2008. 63: 657.
https://pubmed.ncbi.nlm.nih.gov/18455557
196. Atray, N.K., et al. Post transplant lymphocele: a single centre experience. Clin Transplant, 2004.
18 Suppl 12: 46.
https://pubmed.ncbi.nlm.nih.gov/15217407
197. Ulrich, F., et al. Symptomatic lymphoceles after kidney transplantation - multivariate analysis of risk
factors and outcome after laparoscopic fenestration. Clin Transplant, 2010. 24: 273.
https://pubmed.ncbi.nlm.nih.gov/19719727
198. Lucewicz, A., et al. Management of primary symptomatic lymphocele after kidney transplantation: a
systematic review. Transplantation, 2011. 92: 663.
https://pubmed.ncbi.nlm.nih.gov/21849931
199. Capocasale, E., et al. Octreotide in the treatment of lymphorrhea after renal transplantation: a
preliminary experience. Transplant Proc, 2006. 38: 1047.
https://pubmed.ncbi.nlm.nih.gov/16757259
200. Kayler, L., et al. Kidney transplant ureteroneocystostomy techniques and complications: review of
the literature. Transplant Proc, 2010. 42: 1413.
https://pubmed.ncbi.nlm.nih.gov/20620446
201. Secin, F.P., et al. Comparing Taguchi and Lich-Gregoir ureterovesical reimplantation techniques for
kidney transplants. J Urol, 2002. 168: 926.
https://pubmed.ncbi.nlm.nih.gov/12187192
202. Dinckan, A., et al. Early and late urological complications corrected surgically following renal
transplantation. Transpl Int, 2007. 20: 702.
https://pubmed.ncbi.nlm.nih.gov/17511829
203. Kumar, A., et al. Evaluation of the urological complications of living related renal transplantation at
a single center during the last 10 years: impact of the Double-J* stent. J Urol, 2000. 164: 657.
https://pubmed.ncbi.nlm.nih.gov/10953120
204. Mazzucchi, E., et al. Primary reconstruction is a good option in the treatment of urinary fistula after
kidney transplantation. Int Braz J Urol, 2006. 32: 398.
https://pubmed.ncbi.nlm.nih.gov/16953905
205. Davari, H.R., et al. Urological complications in 980 consecutive patients with renal transplantation.
Int J Urol, 2006. 13: 1271.
https://pubmed.ncbi.nlm.nih.gov/17010003
206. Sabnis, R.B., et al. The development and current status of minimally invasive surgery to manage
urological complications after renal transplantation. Indian J Urol, 2016. 32: 186.
https://pubmed.ncbi.nlm.nih.gov/27555675
207. Breda, A., et al. Incidence of ureteral strictures after laparoscopic donor nephrectomy. J Urol, 2006.
176: 1065.
https://pubmed.ncbi.nlm.nih.gov/16890691
208. Helfand, B.T., et al. Reconstruction of late-onset transplant ureteral stricture disease. BJU Int, 2011.
107: 982.
https://pubmed.ncbi.nlm.nih.gov/20825404

RENAL TRANSPLANTATION - TEXT UPDATE 2021 45


209. Kaskarelis, I., et al. Ureteral complications in renal transplant recipients successfully treated with
interventional radiology. Transplant Proc, 2008. 40: 3170.
https://pubmed.ncbi.nlm.nih.gov/19010224
210. Gabr, A.H., et al. Ureteral complications after hand-assisted laparoscopic living donor nephrectomy.
Transplantation, 2014. 97: 788.
https://pubmed.ncbi.nlm.nih.gov/24305639
211. Kristo, B., et al. Treatment of renal transplant ureterovesical anastomotic strictures using antegrade
balloon dilation with or without holmium:YAG laser endoureterotomy. Urology, 2003. 62: 831.
https://pubmed.ncbi.nlm.nih.gov/14624903
212. Nie, Z., et al. Comparison of urological complications with primary ureteroureterostomy versus
conventional ureteroneocystostomy. Clin Transplant, 2010. 24: 615.
https://pubmed.ncbi.nlm.nih.gov/19925475
213. Chaykovska, L., et al. Kidney transplantation into urinary conduits with ureteroureterostomy
between transplant and native ureter: single-center experience. Urology, 2009. 73: 380.
https://pubmed.ncbi.nlm.nih.gov/19022489
214. Kumar, S., et al. Long-term graft and patient survival after balloon dilation of ureteric stenosis after
renal transplant: A 23-year retrospective matched cohort study. Radiology, 2016. 281: 301.
https://pubmed.ncbi.nlm.nih.gov/27018575
215. Jung, G.O., et al. Clinical significance of posttransplantation vesicoureteral reflux during short-term
period after kidney transplantation. Transplant Proc, 2008. 40: 2339.
https://pubmed.ncbi.nlm.nih.gov/18790229
216. Giral, M., et al. Acute graft pyelonephritis and long-term kidney allograft outcome. Kidney Int, 2002.
61: 1880.
https://pubmed.ncbi.nlm.nih.gov/11967040
217. Pichler, R., et al. Endoscopic application of dextranomer/hyaluronic acid copolymer in the treatment
of vesico-ureteric reflux after renal transplantation. BJU Int, 2011. 107: 1967.
https://pubmed.ncbi.nlm.nih.gov/21059169
218. Abbott, K.C., et al. Hospitalized nephrolithiasis after renal transplantation in the United States. Am
J Transplant, 2003. 3: 465.
https://pubmed.ncbi.nlm.nih.gov/12694070
219. Verrier, C., et al. Decrease in and management of urolithiasis after kidney transplantation. J Urol,
2012. 187: 1651.
https://pubmed.ncbi.nlm.nih.gov/22425102
220. Oliveira, M., et al. Percutaneous nephrolithotomy in renal transplants: a safe approach with a high
stone-free rate. Int Urol Nephrol, 2011. 43: 329.
https://pubmed.ncbi.nlm.nih.gov/20848196
221. Silva, A., et al. Risk factors for urinary tract infection after renal transplantation and its impact on
graft function in children and young adults. J Urol, 2010. 184: 1462.
https://pubmed.ncbi.nlm.nih.gov/20727542
222. Challacombe, B., et al. Multimodal management of urolithiasis in renal transplantation. BJU Int,
2005. 96: 385.
https://pubmed.ncbi.nlm.nih.gov/16042735
223. Basiri, A., et al. Ureteroscopic management of urological complications after renal transplantation.
Scand J Urol Nephrol, 2006. 40: 53.
https://pubmed.ncbi.nlm.nih.gov/16452057
224. Roine, E., et al. Targeting risk factors for impaired wound healing and wound complications after
kidney transplantation. Transplant Proc, 2010. 42: 2542.
https://pubmed.ncbi.nlm.nih.gov/20832540
225. Yannam, G.R., et al. Experience of laparoscopic incisional hernia repair in kidney and/or pancreas
transplant recipients. Am J Transplant, 2011. 11: 279.
https://pubmed.ncbi.nlm.nih.gov/21272235
226. Boissier, R., et al. The Risk of Tumour Recurrence in Patients Undergoing Renal Transplantation for
End-stage Renal Disease after Previous Treatment for a Urological Cancer: A Systematic Review.
Eur Urol, 2018. 73: 94.
https://pubmed.ncbi.nlm.nih.gov/28803033
227. Hevia, V., et al. Effectiveness and Harms of Using Kidneys with Small Renal Tumors from Deceased
or Living Donors as a Source of Renal Transplantation: A Systematic Review. Eur Urol Focus, 2018.
https://pubmed.ncbi.nlm.nih.gov/29433988

46 RENAL TRANSPLANTATION - TEXT UPDATE 2021


228. Hevia, V., et al. Management of Localised Prostate Cancer in Kidney Transplant Patients: A
Systematic Review from the EAU Guidelines on Renal Transplantation Panel. Eur Urol Focus, 2018.
4: 153.
https://pubmed.ncbi.nlm.nih.gov/29921544
229. Marra, G., et al. Prostate cancer treatment in renal transplant recipients: a systematic review. BJU
Int, 2018. 121: 327.
https://pubmed.ncbi.nlm.nih.gov/28921938
230. Tait, B.D., et al. Consensus guidelines on the testing and clinical management issues associated
with HLA and non-HLA antibodies in transplantation. Transplantation, 2013. 95: 19.
https://pubmed.ncbi.nlm.nih.gov/23238534
231. European Renal Best Practice Transplantation Guideline Development Group. ERBP Guideline on
the Management and Evaluation of the Kidney Donor and Recipient. Nephrol Dial Transplant, 2013.
28 Suppl 2: ii1.
https://pubmed.ncbi.nlm.nih.gov/24026881
232. Poulton, K., et al. British Transplantation Society. Guidelines for the detection of clinically relevant
antibodies in allotransplantation. 2014.
https://bts.org.uk/wp-content/uploads/2016/09/06_BTS_BSHI_Antibodies-1.pdf
233. UNOS. Unitied Network For Organ Sharing Website:
https://www.unos.org/
234. Heidt, S., Eurotransplant Manual version 3.1 Chapter 10 Histocompatibility. 2015.
https://www.eurotransplant.org/wp-content/uploads/2020/01/H10-Histocompatibility.pdf
235. European Federation for Immunogenetics, EFI Standards for Histocompatibility and
Immunogenetics Testing Version 6.3. 2015.
https://efi-web.org/committees/standards-committee
236. De Meester, J., et al. Renal transplantation of highly sensitised patients via prioritised renal
allocation programs. Shorter waiting time and above-average graft survival. Nephron, 2002. 92: 111.
https://pubmed.ncbi.nlm.nih.gov/12187093
237. Susal, C., et al. Algorithms for the determination of unacceptable HLA antigen mismatches in kidney
transplant recipients. Tissue Antigens, 2013. 82: 83.
https://pubmed.ncbi.nlm.nih.gov/23718733
238. Bohmig, G.A., et al. Strategies to overcome the ABO barrier in kidney transplantation. Nat Rev
Nephrol, 2015. 11: 732.
https://pubmed.ncbi.nlm.nih.gov/26324199
239. Zschiedrich, S., et al. An update on ABO-incompatible kidney transplantation. Transpl Int, 2015.
28: 387.
https://pubmed.ncbi.nlm.nih.gov/25387763
240. Higgins, R.M., et al. Antibody-incompatible kidney transplantation in 2015 and beyond. Nephrol Dial
Transplant, 2015. 30: 1972.
https://pubmed.ncbi.nlm.nih.gov/25500804
241. Wongsaroj, P., et al. Modern approaches to incompatible kidney transplantation. World J Nephrol,
2015. 4: 354.
https://pubmed.ncbi.nlm.nih.gov/26167458
242. Bamoulid, J., et al. Immunosuppression and Results in Renal Transplantation. European Urology
Supplements, 2016. 15: 415.
https://www.eu-openscience.europeanurology.com/article/S1569-9056(16)30082-3/fulltext
243. Kidney Disease Improving Global Outcomes Transplant Work Group. KDIGO clinical practice
guideline for the care of kidney transplant recipients. Am J Transplant, 2009. 9 Suppl 3: S1.
https://pubmed.ncbi.nlm.nih.gov/19845597
244. Bamoulid, J., et al. The need for minimization strategies: current problems of immunosuppression.
Transpl Int, 2015. 28: 891.
https://pubmed.ncbi.nlm.nih.gov/25752992
245. Jones-Hughes, T., et al. Immunosuppressive therapy for kidney transplantation in adults: a
systematic review and economic model. Health Technol Assess, 2016. 20: 1.
https://pubmed.ncbi.nlm.nih.gov/27578428
246. Leas, B.F., et al., Calcineurin Inhibitors for Renal Transplant. 2016: Rockville (MD): AHRQ.
https://www.ncbi.nlm.nih.gov/books/NBK356377
247. Sawinski, D., et al. Calcineurin Inhibitor Minimization, Conversion, Withdrawal, and Avoidance
Strategies in Renal Transplantation: A Systematic Review and Meta-Analysis. Am J Transplant,
2016. 16: 2117.
https://pubmed.ncbi.nlm.nih.gov/26990455

RENAL TRANSPLANTATION - TEXT UPDATE 2021 47


248. Webster, A.C., et al. Tacrolimus versus ciclosporin as primary immunosuppression for kidney transplant
recipients: meta-analysis and meta-regression of randomised trial data. BMJ, 2005. 331: 810.
https://pubmed.ncbi.nlm.nih.gov/16157605
249. Brunet, M., et al. Therapeutic Drug Monitoring of Tacrolimus-Personalized Therapy: Second
Consensus Report. Therapeutic Drug Monitoring, 2019. 41: 261.
https://pubmed.ncbi.nlm.nih.gov/31045868/
250. Ekberg, H., et al. Relationship of tacrolimus exposure and mycophenolate mofetil dose with renal
function after renal transplantation. Transplantation, 2011. 92: 82.
https://pubmed.ncbi.nlm.nih.gov/21562449
251. Xia, T., et al. Risk factors for calcineurin inhibitor nephrotoxicity after renal transplantation: A
systematic review and meta-analysis. Drug Des Devel Ther, 2018. 12: 417.
https://pubmed.ncbi.nlm.nih.gov/29535503
252. Gallagher, M., et al. Cyclosporine withdrawal improves long-term graft survival in renal
transplantation. Transplantation, 2009. 87: 1877.
https://pubmed.ncbi.nlm.nih.gov/19543068
253. Liu, J.Y., et al. Tacrolimus versus cyclosporine as primary immunosuppressant after renal
transplantation: A meta-analysis and economics evaluation. Am J Ther, 2016. 23: e810.
https://pubmed.ncbi.nlm.nih.gov/25299636
254. Opelz, G., et al. Influence of immunosuppressive regimens on graft survival and secondary
outcomes after kidney transplantation. Transplantation, 2009. 87: 795.
https://pubmed.ncbi.nlm.nih.gov/19300179
255. Cheung, C.Y., et al. Long-term graft function with tacrolimus and cyclosporine in renal
transplantation: Paired kidney analysis. Nephrology, 2009. 14: 758.
https://pubmed.ncbi.nlm.nih.gov/20025685
256. de Fijter, J.W., et al. Early Conversion From Calcineurin Inhibitor- to Everolimus-Based Therapy
Following Kidney Transplantation: Results of the Randomized ELEVATE Trial. Am J Transplant, 2017.
17: 1853.
https://pubmed.ncbi.nlm.nih.gov/28027625
257. Goring, S.M., et al. A network meta-analysis of the efficacy of belatacept, cyclosporine and
tacrolimus for immunosuppression therapy in adult renal transplant recipients. Cur Med Res Opin,
2014. 30: 1473.
https://pubmed.ncbi.nlm.nih.gov/24628478
258. Pascual, J., et al. Everolimus with Reduced Calcineurin Inhibitor Exposure in Renal Transplantation.
J Am Soc Nephrol: JASN, 2018. 29: 1979.
https://pubmed.ncbi.nlm.nih.gov/29752413
259. Basso, G., et al. The effect of anti-thymocyte globulin and everolimus on the kinetics of
cytomegalovirus viral load in seropositive kidney transplant recipients without prophylaxis. Transpl
Infect Dis, 2018. 20: e12919.
https://onlinelibrary.wiley.com/doi/abs/10.1111/tid.12919
260. Bloom, R.D., et al. A randomized, crossover pharmacokinetic study comparing generic tacrolimus
vs. the reference formulation in subpopulations of kidney transplant patients. Clin Transplant, 2013.
27: E685.
https://pubmed.ncbi.nlm.nih.gov/24118450
261. Glander, P., et al. Bioavailability and costs of once-daily and twice-daily tacrolimus formulations in
de novo kidney transplantation. Clin Transplant, 2018: e13311.
https://pubmed.ncbi.nlm.nih.gov/29888809
262. Guirado, L., et al. Medium-Term Renal Function in a Large Cohort of Stable Kidney Transplant
Recipients Converted From Twice-Daily to Once-Daily Tacrolimus. Transplant Dir, 2015. 1: e24.
https://pubmed.ncbi.nlm.nih.gov/27500226
263. Melilli, E., et al. De novo use of a generic formulation of tacrolimus versus reference tacrolimus
in kidney transplantation: Evaluation of the clinical results, histology in protocol biopsies, and
immunological monitoring. Transplant Int, 2015. 28: 1283.
https://pubmed.ncbi.nlm.nih.gov/26088437
264. Robertsen, I., et al. Use of generic tacrolimus in elderly renal transplant recipients: Precaution is
needed. Transplantation, 2015. 99: 528.
https://pubmed.ncbi.nlm.nih.gov/25148382
265. Rostaing, L., et al. Novel once-daily extended-release tacrolimus versus twice-daily tacrolimus in de
novo kidney transplant recipients: Two-year results of phase 3, double-blind, randomized trial. Am
J Kidney Dis, 2016. 67: 648.
https://pubmed.ncbi.nlm.nih.gov/26717860

48 RENAL TRANSPLANTATION - TEXT UPDATE 2021


266. Saengram, W., et al. Extended release versus immediate release tacrolimus in kidney transplant
recipients: a systematic review and meta-analysis. Eur J Clin Pharmacol, 2018: 1.
https://pubmed.ncbi.nlm.nih.gov/29961086
267. Silva, H.T., et al. Long-term follow-up of a phase III clinical trial comparing tacrolimus extended-
release/MMF, tacrolimus/MMF, and cyclosporine/MMF in de novo kidney transplant recipients.
Transplantation, 2014. 97: 636.
https://pubmed.ncbi.nlm.nih.gov/24521771
268. Kahn, J., et al. Immunosuppression with generic tacrolimus in liver and kidney transplantation-
systematic review and meta-analysis on biopsy-proven acute rejection and bioequivalence.
Transplant Int, 2020. 33: 356.
https://www.ncbi.nlm.nih.gov/pmc/articles/PMC7154701/pdf/TRI-33-356.pdf
269. Lehner, L.J., et al. Evaluation of adherence and tolerability of prolonged-release tacrolimus
(AdvagrafTM) in kidney transplant patients in Germany: A multicenter, noninterventional study. Clin
Transplant, 2018. 32: e13142.
https://pubmed.ncbi.nlm.nih.gov/29052906
270. Paterson, T.S.E., et al. Impact of Once- Versus Twice-Daily Tacrolimus Dosing on Medication
Adherence in Stable Renal Transplant Recipients: A Canadian Single-Center Randomized Controlled
Trial. Can J Kidney Health, 2019. 6.
https://journals.sagepub.com/doi/pdf/10.1177/2054358119867993
271. Caillard, S., et al. Advagraf((R)) , a once-daily prolonged release tacrolimus formulation, in kidney
transplantation: literature review and guidelines from a panel of experts. Transpl Int, 2016. 29: 860.
https://pubmed.ncbi.nlm.nih.gov/26373896
272. McCormack, P.L. Extended-release tacrolimus: a review of its use in de novo kidney transplantation.
Drugs, 2014. 74: 2053.
https://pubmed.ncbi.nlm.nih.gov/25352392
273. Molnar, A.O., et al. Generic immunosuppression in solid organ transplantation: systematic review
and meta-analysis. BMJ, 2015. 350: h3163.
https://pubmed.ncbi.nlm.nih.gov/26101226
274. Staatz, C.E., et al. Clinical Pharmacokinetics of Once-Daily Tacrolimus in Solid-Organ Transplant
Patients. Clin Pharmacokinet, 2015. 54: 993.
https://pubmed.ncbi.nlm.nih.gov/26038096
275. van Gelder, T., et al. European Society for Organ Transplantation Advisory Committee
recommendations on generic substitution of immunosuppressive drugs. Transpl Int, 2011. 24: 1135.
https://pubmed.ncbi.nlm.nih.gov/22032583
276. Wissing, K.M., et al. Prospective randomized study of conversion from tacrolimus to cyclosporine
A to improve glucose metabolism in patients with posttransplant diabetes mellitus after renal
transplantation. Am J Transplant, 2018. 18: 1726.
https://pubmed.ncbi.nlm.nih.gov/29337426
277. Diekmann, F. Immunosuppressive minimization with mTOR inhibitors and belatacept. Transpl Int,
2015. 28: 921.
https://pubmed.ncbi.nlm.nih.gov/25959589
278. Kamar, N., et al. Calcineurin inhibitor-sparing regimens based on mycophenolic acid after kidney
transplantation. Transpl Int, 2015. 28: 928.
https://pubmed.ncbi.nlm.nih.gov/25557802
279. Park, S., et al. Reduced Tacrolimus Trough Level Is Reflected by Estimated Glomerular Filtration
Rate (eGFR) Changes in Stable Renal Transplantation Recipients: Results of the OPTIMUM Phase 3
Randomized Controlled Study. Ann Transplantn, 2018. 23: 401.
https://pubmed.ncbi.nlm.nih.gov/29891834
280. Sharif, A., et al. Meta-analysis of calcineurin-inhibitor-sparing regimens in kidney transplantation.,
J Am Soc Nephrol 2011. 22: 2107.
https://pubmed.ncbi.nlm.nih.gov/
281. Snanoudj, R., et al. Immunological risks of minimization strategies. Transpl Int, 2015. 28: 901.
https://pubmed.ncbi.nlm.nih.gov/25809144
282. Etienne, I., et al. A 50% reduction in cyclosporine exposure in stable renal transplant recipients:
Renal function benefits. Nephrol Dial Transplant, 2010. 25: 3096.
https://pubmed.ncbi.nlm.nih.gov/20299336
283. Budde, K., et al. Enteric-coated mycophenolate sodium. Expert Opin Drug Saf, 2010. 9: 981.
https://pubmed.ncbi.nlm.nih.gov/20795786

RENAL TRANSPLANTATION - TEXT UPDATE 2021 49


284. Cooper, M., et al. Enteric-coated mycophenolate sodium immunosuppression in renal transplant
patients: efficacy and dosing. Transplant Rev (Orlando), 2012. 26: 233.
https://pubmed.ncbi.nlm.nih.gov/22863029
285. Staatz, C.E., et al. Pharmacology and toxicology of mycophenolate in organ transplant recipients: an
update. Arch Toxicol, 2014. 88: 1351.
https://pubmed.ncbi.nlm.nih.gov/24792322
286. van Gelder, T., et al. Mycophenolate revisited. Transpl Int, 2015. 28: 508.
https://pubmed.ncbi.nlm.nih.gov/25758949
287. Wagner, M., et al. Mycophenolic acid versus azathioprine as primary immunosuppression for kidney
transplant recipients. Cochrane Database Syst Rev, 2015: CD007746.
https://pubmed.ncbi.nlm.nih.gov/26633102
288. Hirsch, H.H., et al. European perspective on human polyomavirus infection, replication and disease
in solid organ transplantation. Clin Microbiol Infect, 2014. 20 Suppl 7: 74.
https://pubmed.ncbi.nlm.nih.gov/24476010
289. Langone, A.J., et al. Enteric-coated mycophenolate sodium versus mycophenolate mofetil in
renal transplant recipients experiencing gastrointestinal intolerance: A multicenter, double-blind,
randomized study. Transplantation, 2011. 91: 470.
https://pubmed.ncbi.nlm.nih.gov/21245794
290. Doria, C., et al. Association of mycophenolic acid dose with efficacy and safety events in kidney
transplant patients receiving tacrolimus: An analysis of the Mycophenolic acid Observational REnal
transplant registry. Clin Transplant, 2012. 26: E602.
https://pubmed.ncbi.nlm.nih.gov/23121178
291. Langone, A., et al. Does reduction in mycophenolic acid dose compromise efficacy regardless of
tacrolimus exposure level? An analysis of prospective data from the Mycophenolic Renal Transplant
(MORE) Registry. Clin Transplant, 2013. 27: 15.
https://pubmed.ncbi.nlm.nih.gov/22861144
292. Su, V.C.H., et al. Impact of mycophenolate mofetil dose reduction on allograft outcomes in kidney
transplant recipients on tacrolimus-based regimens: A systematic review. Ann Pharmacother, 2011.
45: 248.
https://pubmed.ncbi.nlm.nih.gov/21304036
293. Kotton, C.N., et al. Updated international consensus guidelines on the management of
cytomegalovirus in solid-organ transplantation. Transplantation, 2013. 96: 333.
https://pubmed.ncbi.nlm.nih.gov/23896556
294. Le Meur, Y., et al. Therapeutic drug monitoring of mycophenolates in kidney transplantation: report
of The Transplantation Society consensus meeting. Transplant Rev (Orlando), 2011. 25: 58.
https://pubmed.ncbi.nlm.nih.gov/21454067
295. Haller, M.C., et al. Steroid avoidance or withdrawal for kidney transplant recipients. Cochrane
Database Syst Rev, 2016: CD005632.
https://pubmed.ncbi.nlm.nih.gov/27546100
296. Meier-Kriesche, H.U., et al. Mycophenolate mofetil initiation in renal transplant patients at different
times posttransplantation: The TranCept switch study. Transplantation, 2011. 91: 984.
https://pubmed.ncbi.nlm.nih.gov/21464796
297. Mathis, A.S., et al. Calcineurin inhibitor sparing strategies in renal transplantation, part one: Late
sparing strategies. World J Transplant, 2014. 4: 57.
https://pubmed.ncbi.nlm.nih.gov/25032096
298. Remuzzi, G., et al. Mycophenolate mofetil versus azathioprine for prevention of chronic allograft
dysfunction in renal transplantation: the MYSS follow-up randomized, controlled clinical trial. J Am
Soc Nephrol, 2007. 18: 1973.
https://pubmed.ncbi.nlm.nih.gov/17460145
299. Kunz, R., et al. Maintenance therapy with triple versus double immunosuppressive regimen in renal
transplantation: a meta-analysis. Transplantation, 1997. 63: 386.
https://pubmed.ncbi.nlm.nih.gov/9039928
300. Le Meur, Y., et al. Early steroid withdrawal and optimization of mycophenolic acid exposure in
kidney transplant recipients receiving mycophenolate mofetil. Transplantation, 2011. 92: 1244.
https://pubmed.ncbi.nlm.nih.gov/22067312
301. Suszynski, T.M., et al. Prospective randomized trial of maintenance immunosuppression with rapid
discontinuation of prednisone in adult kidney transplantation. Am J Transplant, 2013. 13: 961.
https://pubmed.ncbi.nlm.nih.gov/23432755

50 RENAL TRANSPLANTATION - TEXT UPDATE 2021


302. Thomusch, O., et al. Rabbit-ATG or basiliximab induction for rapid steroid withdrawal after renal
transplantation (Harmony): an open-label, multicentre, randomised controlled trial. Lancet, 2016.
388: 3006.
https://pubmed.ncbi.nlm.nih.gov/27871759
303. Torres, A., et al. Randomized Controlled Trial Assessing the Impact of Tacrolimus Versus Cyclosporine
on the Incidence of Posttransplant Diabetes Mellitus. Kidney Int Reports, 2018. 3: 1304.
https://www.ncbi.nlm.nih.gov/pmc/articles/PMC6224662/pdf/main.pdf
304. Halleck, F., et al. An evaluation of sirolimus in renal transplantation. Expert Opin Drug Metab Toxicol,
2012. 8: 1337.
https://pubmed.ncbi.nlm.nih.gov/22928953
305. Ventura-Aguiar, P., et al. Safety of mTOR inhibitors in adult solid organ transplantation. Expert Opin
Drug Saf, 2016. 15: 303.
https://pubmed.ncbi.nlm.nih.gov/26667069
306. Witzke, O., et al. Everolimus immunosuppression in kidney transplantation: What is the optimal
strategy? Transplant Rev (Orlando), 2016. 30: 3.
https://pubmed.ncbi.nlm.nih.gov/26603484
307. Montero, N., et al. Mammalian Target of Rapamycin Inhibitors Combined with Calcineurin Inhibitors as
Initial Immunosuppression in Renal Transplantation: A Meta-analysis. Transplantation, 2019. 103: 2031.
https://journals.lww.com/transplantjournal/Fulltext/2019/10000/Mammalian_Target_of_Rapamycin_
Inhibitors_Combined.16.aspx
308. Badve, S.V., et al. Mammalian target of rapamycin inhibitors and clinical outcomes in adult kidney
transplant recipients. Clin J Am Soc Nephrol, 2016. 11: 1845.
https://pubmed.ncbi.nlm.nih.gov/27445164
309. Lim, W.H., et al. A systematic review of conversion from calcineurin inhibitor to mammalian target
of rapamycin inhibitors for maintenance immunosuppression in kidney transplant recipients. Am
J Transplant, 2014. 14: 2106.
https://pubmed.ncbi.nlm.nih.gov/25088685
310. Liu, J., et al. Efficacy and safety of everolimus for maintenance immunosuppression of kidney
transplantation: A meta-analysis of randomized controlled trials. PLoS ONE, 2017. 12: e0170246.
https://pubmed.ncbi.nlm.nih.gov/28107397
311. Knoll, G.A., et al. Effect of sirolimus on malignancy and survival after kidney transplantation:
systematic review and meta-analysis of individual patient data. BMJ, 2014. 349: g6679.
https://pubmed.ncbi.nlm.nih.gov/25422259
312. Xie, X., et al. mTOR inhibitor versus mycophenolic acid as the primary immunosuppression regime
combined with calcineurin inhibitor for kidney transplant recipients: a meta-analysis. BMC Nephrol,
2015. 16: 91.
https://pubmed.ncbi.nlm.nih.gov/26126806
313. Wolf, S., et al. Effects of mTOR-Is on malignancy and survival following renal transplantation: A
systematic review and meta-analysis of randomized trials with a minimum follow-up of 24 months.
PLoS One, 2018. 13: e0194975.
https://pubmed.ncbi.nlm.nih.gov/29659588
314. Hahn, D., et al. Target of rapamycin inhibitors (TOR-I; sirolimus and everolimus) for primary
immunosuppression in kidney transplant recipients. Cochrane Database Syst Rev, 2019. 2019:
CD004290.
https://www.cochranelibrary.com/cdsr/doi/10.1002/14651858.CD004290.pub3/full
315. Shipkova, M., et al. Therapeutic Drug Monitoring of Everolimus: A Consensus Report. Ther Drug
Monit, 2016. 38: 143.
https://pubmed.ncbi.nlm.nih.gov/26982492
316. Rostaing, L., et al. The pharmacokinetics of everolimus in de novo kidney transplant patients
receiving tacrolimus: An analysis from the randomized ASSET study. Ann Transplant, 2014. 19: 337.
https://pubmed.ncbi.nlm.nih.gov/25017487
317. Shihab, F., et al. Association of Clinical Events With Everolimus Exposure in Kidney Transplant
Patients Receiving Low Doses of Tacrolimus. Am J Transplant, 2017. 17: 2363.
https://pubmed.ncbi.nlm.nih.gov/28141897
318. Kumar, J., et al. Systematic review on role of mammalian target of rapamycin inhibitors as an
alternative to calcineurin inhibitors in renal transplant: Challenges and window to excel. Exp Clin
Transplant, 2017. 15: 241.
https://pubmed.ncbi.nlm.nih.gov/27915965

RENAL TRANSPLANTATION - TEXT UPDATE 2021 51


319. Qazi, Y., et al. Efficacy and Safety of Everolimus Plus Low-Dose Tacrolimus Versus Mycophenolate
Mofetil Plus Standard-Dose Tacrolimus in De Novo Renal Transplant Recipients: 12-Month Data. Am
J Transplant, 2017. 17: 1358.
https://pubmed.ncbi.nlm.nih.gov/27775865
320. Rummo, O.O., et al. ADHERE: randomized controlled trial comparing renal function in de novo
kidney transplant recipients receiving prolonged-release tacrolimus plus mycophenolate mofetil or
sirolimus. Transplant Int, 2017. 30: 83.
https://pubmed.ncbi.nlm.nih.gov/27754567
321. Berger, S.P., et al. Two-year outcomes in de novo renal transplant recipients receiving everolimus-
facilitated calcineurin inhibitor reduction regimen from the TRANSFORM study. Am J Transplant,
2019. 19: 3018.
https://onlinelibrary.wiley.com/doi/full/10.1111/ajt.15480
322. Sommerer, C., et al. An open-label, randomized trial indicates that everolimus with tacrolimus or
cyclosporine is comparable to standard immunosuppression in de novo kidney transplant patients.
Kidney Int, 2019. 96: 231.
https://www.kidney-international.org/article/S0085-2538(19)30193-0/pdf
323. Tedesco-Silva, H., et al. Safety of Everolimus With Reduced Calcineurin Inhibitor Exposure in De
Novo Kidney Transplants: An Analysis From the Randomized TRANSFORM Study. Transplantation,
2019. 103: 1953.
https://journals.lww.com/transplantjournal/Fulltext/2019/09000/Safety_of_Everolimus_With_
Reduced_Calcineurin.36.aspx
324. He, L., et al. Efficacy and safety of everolimus plus lowdose calcineurin inhibitor vs. mycophenolate
mofetil plus standard-dose calcineurin inhibitor in renal transplant recipients: A systematic review
and meta-analysis. Clin Nephrol, 2018. 89: 336.
https://pubmed.ncbi.nlm.nih.gov/29292693
325. Liu, J.Y., et al. Sirolimus versus tacrolimus as primary immunosuppressant after renal
transplantation: A meta-analysis and economics evaluation. Am J Ther, 2016. 23: e1720.
https://pubmed.ncbi.nlm.nih.gov/25569597
326. Zhao, D.Q., et al. Sirolimus-based immunosuppressive regimens in renal transplantation: A systemic
review. Transplant Proc, 2016. 48: 3.
https://pubmed.ncbi.nlm.nih.gov/26915834
327. Liefeldt, L., et al. Donor-specific HLA antibodies in a cohort comparing everolimus with cyclosporine
after kidney transplantation. Am J Transplant, 2012. 12: 1192.
https://pubmed.ncbi.nlm.nih.gov/22300538
328. Wolf, S., et al. Infections after kidney transplantation: A comparison of mTOR-Is and CNIs as basic
immunosuppressants. A systematic review and meta-analysis. Transpl Infect Dis, 2020.
https://onlinelibrary.wiley.com/doi/full/10.1111/tid.13267
329. Halleck, F., et al. Transplantation: Sirolimus for secondary SCC prevention in renal transplantation.
Nat Rev Nephrol, 2012. 8: 687.
https://pubmed.ncbi.nlm.nih.gov/23026948
330. Ponticelli, C., et al. Skin cancer in kidney transplant recipients. J Nephrol, 2014. 27: 385.
https://pubmed.ncbi.nlm.nih.gov/24809813
331. Cheung, C.Y., et al. Conversion to mammalian target of rapamycin inhibitors in kidney transplant
recipients with de novo cancers. Oncotarget, 2017. 8: 44833.
https://pubmed.ncbi.nlm.nih.gov/28160552
332. Opelz, G., et al. Immunosuppression with mammalian target of rapamycin inhibitor and incidence of
post-transplant cancer in kidney transplant recipients. Nephrol Dial Transplant, 2016. 31: 1360.
https://pubmed.ncbi.nlm.nih.gov/27190384
333. Liu, Y., et al. Basiliximab or antithymocyte globulin for induction therapy in kidney transplantation: a
meta-analysis. Transplant Proc, 2010. 42: 1667.
https://pubmed.ncbi.nlm.nih.gov/20620496
334. Sun, Z.J., et al. Efficacy and Safety of Basiliximab Versus Daclizumab in Kidney Transplantation: A
Meta-Analysis. Transplant Proc, 2015. 47: 2439.
https://pubmed.ncbi.nlm.nih.gov/26518947
335. Webster, A.C., et al. Interleukin 2 receptor antagonists for kidney transplant recipients. Cochrane
Database Syst Rev, 2010: CD003897.
https://pubmed.ncbi.nlm.nih.gov/20091551

52 RENAL TRANSPLANTATION - TEXT UPDATE 2021


336. Lim, W., et al. Effect of interleukin-2 receptor antibody therapy on acute rejection risk and severity,
long-term renal function, infection and malignancy-related mortality in renal transplant recipients.
Transplant Int, 2010. 23: 1207.
https://pubmed.ncbi.nlm.nih.gov/20536789
337. McKeage, K., et al. Basiliximab: A review of its use as induction therapy in renal transplantation.
BioDrugs, 2010. 24: 55.
https://pubmed.ncbi.nlm.nih.gov/20055533
338. Hellemans, R., et al. Induction Therapy for Kidney Transplant Recipients: Do We Still Need Anti-IL2
Receptor Monoclonal Antibodies? Am J Transplant, 2017. 17: 22.
https://pubmed.ncbi.nlm.nih.gov/27223882
339. Ali, H., et al. Rabbit anti-thymocyte globulin (rATG) versus IL-2 receptor antagonist induction
therapies in tacrolimus-based immunosuppression era: a meta-analysis. Int Urol Nephrol, 2020.
https://link.springer.com/article/10.1007/s11255-020-02418-w
340. Bamoulid, J., et al. Anti-thymocyte globulins in kidney transplantation: focus on current indications
and long-term immunological side effects. Nephrol Dial Transplant, 2016.
https://pubmed.ncbi.nlm.nih.gov/27798202
341. Malvezzi, P., et al. Induction by anti-thymocyte globulins in kidney transplantation: a review of the
literature and current usage. J Nephropathol, 2015. 4: 110.
https://pubmed.ncbi.nlm.nih.gov/26457257
342. Hill, P., et al. Polyclonal and monoclonal antibodies for induction therapy in kidney transplant
recipients. Cochrane Database of Syst Rev, 2017. 2017: CD004759.
https://pubmed.ncbi.nlm.nih.gov/28073178
343. Webster, A.C., et al. Polyclonal and monoclonal antibodies for treating acute rejection episodes in
kidney transplant recipients. Cochrane Database of Syst Rev, 2017. 2017: CD004756.
https://pubmed.ncbi.nlm.nih.gov/28731207
344. Gill, J., et al. Induction immunosuppressive therapy in the elderly kidney transplant recipient in the
United States. Clinical Journal of the Am Soc Nephrol, 2011. 6: 1168.
https://pubmed.ncbi.nlm.nih.gov/21511836
345. Grinyo, J.M., et al. Belatacept utilization recommendations: an expert position. Expert Opin Drug
Saf, 2013. 12: 111.
https://pubmed.ncbi.nlm.nih.gov/23206310
346. Wojciechowski, D., et al. Current status of costimulatory blockade in renal transplantation. Curr Opin
Nephrol Hypertens, 2016. 25: 583.
https://pubmed.ncbi.nlm.nih.gov/27517137
347. Durrbach, A., et al. Long-Term Outcomes in Belatacept- Versus Cyclosporine-Treated Recipients of
Extended Criteria Donor Kidneys: Final Results From BENEFIT-EXT, a Phase III Randomized Study.
Am J Transplant, 2016. 16: 3192.
https://pubmed.ncbi.nlm.nih.gov/27130868
348. Vincenti, F., et al. Belatacept and Long-Term Outcomes in Kidney Transplantation. N Engl J Med,
2016. 374: 333.
https://pubmed.ncbi.nlm.nih.gov/27355541
349. De Graav, G.N., et al. A Randomized Controlled Clinical Trial Comparing Belatacept with Tacrolimus
after de Novo Kidney Transplantation. Transplantation, 2017. 101: 2571.
https://pubmed.ncbi.nlm.nih.gov/28403127
350. Masson, P., et al. Belatacept for kidney transplant recipients. The Cochrane database Syst Rev,
2014. 11: CD010699.
https://pubmed.ncbi.nlm.nih.gov/25416857
351. Talawila, N., et al. Does belatacept improve outcomes for kidney transplant recipients? A systematic
review. Transplant Int, 2015. 28: 1251.
https://pubmed.ncbi.nlm.nih.gov/25965549
352. Bray, R.A., et al. De novo donor-specific antibodies in belatacept-treated vs cyclosporine-treated
kidney-transplant recipients: Post hoc analyses of the randomized phase III BENEFIT and BENEFIT-
EXT studies. Am J Transplant, 2018.
https://pubmed.ncbi.nlm.nih.gov/29509295
353. Grannas, G., et al. Ten years experience with belatacept-based immunosuppression after kidney
transplantation. J Clin Med Res, 2014. 6: 98.
https://pubmed.ncbi.nlm.nih.gov/24578751
354. Schwarz, C., et al. Long-term outcome of belatacept therapy in de novo kidney transplant recipients
- A case-match analysis. Transplant Int, 2015. 28: 820.
https://pubmed.ncbi.nlm.nih.gov/25703346

RENAL TRANSPLANTATION - TEXT UPDATE 2021 53


355. Elhamahmi, D.A., et al. Early Conversion to Belatacept in Kidney Transplant Recipients with Low
Glomerular Filtration Rate. Transplantation, 2018. 102: 478.
https://pubmed.ncbi.nlm.nih.gov/29077658
356. Grinyo, J.M., et al. Safety and Efficacy Outcomes 3 Years After Switching to Belatacept From a
Calcineurin Inhibitor in Kidney Transplant Recipients: Results From a Phase 2 Randomized Trial. Am
J Kidney Dis, 2017. 69: 587.
https://pubmed.ncbi.nlm.nih.gov/27889299
357. Darres, A., et al. Conversion to Belatacept in Maintenance Kidney Transplant Patients: A
Retrospective Multicenter European Study. Transplantation, 2018. 102: 1545.
https://pubmed.ncbi.nlm.nih.gov/29570163
358. Bamoulid, J., et al. Advances in pharmacotherapy to treat kidney transplant rejection. Expert Opin
Pharmacother, 2015. 16: 1627.
https://pubmed.ncbi.nlm.nih.gov/26159444
359. Broecker, V., et al. The significance of histological diagnosis in renal allograft biopsies in 2014.
Transpl Int, 2015. 28: 136.
https://pubmed.ncbi.nlm.nih.gov/25205033
360. Halloran, P.F., et al. Molecular assessment of disease states in kidney transplant biopsy samples.
Nat Rev Nephrol, 2016. 12: 534.
https://pubmed.ncbi.nlm.nih.gov/27345248
361. Lentine, K.L., et al. The implications of acute rejection for allograft survival in contemporary U.S.
kidney transplantation. Transplantation, 2012. 94: 369.
https://pubmed.ncbi.nlm.nih.gov/22836133
362. Clayton, P.A., et al. Long-term outcomes after acute rejection in kidney transplant recipients: An
Anzdata analysis. J Am Soc Nephrol, 2019. 30: 1697.
https://jasn.asnjournals.org/content/30/9/1697.abstract
363. Loupy, A., et al. The Banff 2019 Kidney Meeting Report (I): Updates on and clarification of criteria for
T cell- and antibody-mediated rejection. Am J Transplant, 2020. 20: 2318.
https://pubmed.ncbi.nlm.nih.gov/32463180
364. Morgan, T.A., et al. Complications of Ultrasound-Guided Renal Transplant Biopsies. Am
J Transplant, 2016. 16: 1298.
https://pubmed.ncbi.nlm.nih.gov/26601796
365. Redfield, R.R., et al. Nature, timing, and severity of complications from ultrasound-guided
percutaneous renal transplant biopsy. Transpl Int, 2016. 29: 167.
https://pubmed.ncbi.nlm.nih.gov/26284692
366. Bouatou, Y., et al. Response to treatment and long-term outcomes in kidney transplant recipients
with acute T cell-mediated rejection. Am J Transplant, 2019. 19: 1972.
https://onlinelibrary.wiley.com/doi/full/10.1111/ajt.15299
367. Schinstock, C.A., et al. Recommended Treatment for Antibody-mediated Rejection After Kidney
Transplantation: the 2019 Expert Consensus From the Transplantion Society Working Group.
Transplantation, 2020.
https://pubmed.ncbi.nlm.nih.gov/31895348
368. Amore, A. Antibody-mediated rejection. Curr Opin Organ Transplant, 2015. 20: 536.
https://pubmed.ncbi.nlm.nih.gov/26284692
369. Burton, S.A., et al. Treatment of antibody-mediated rejection in renal transplant patients: a clinical
practice survey. Clin Transplant, 2015. 29: 118.
https://pubmed.ncbi.nlm.nih.gov/25430052
370. Haririan, A. Current status of the evaluation and management of antibody-mediated rejection in
kidney transplantation. Curr Opin Nephrol Hypertens, 2015. 24: 576.
https://pubmed.ncbi.nlm.nih.gov/26406806
371. Sautenet, B., et al. One-year Results of the Effects of Rituximab on Acute Antibody-Mediated
Rejection in Renal Transplantation: RITUX ERAH, a Multicenter Double-blind Randomized Placebo-
controlled Trial. Transplantation, 2016. 100: 391.
https://pubmed.ncbi.nlm.nih.gov/26555944
372. Loupy, A., et al. Antibody-Mediated Rejection of Solid-Organ Allografts. N Engl J Med, 2018.
379: 1150.
https://pubmed.ncbi.nlm.nih.gov/30231232
373. Wan, S.S., et al. The Treatment of Antibody-Mediated Rejection in Kidney Transplantation: An
Updated Systematic Review and Meta-Analysis. Transplantation, 2018. 102: 557.
https://pubmed.ncbi.nlm.nih.gov/29315141

54 RENAL TRANSPLANTATION - TEXT UPDATE 2021


374. Kamar, N., et al. Incidence and predictive factors for infectious disease after rituximab therapy in
kidney-transplant patients. Am J Transplant, 2010. 10: 89.
https://pubmed.ncbi.nlm.nih.gov/19656128
375. Velidedeoglu, E., et al. Summary of 2017 FDA Public Workshop: Antibody-mediated Rejection in
Kidney Transplantation. Transplantation, 2018. 102: e257.
https://pubmed.ncbi.nlm.nih.gov/29470345
376. Farrugia, D., et al. Malignancy-related mortality following kidney transplantation is common. Kidney
Int, 2014. 85: 1395.
https://pubmed.ncbi.nlm.nih.gov/24257690
377. Piselli, P., et al. Risk of de novo cancers after transplantation: results from a cohort of 7217 kidney
transplant recipients, Italy 1997-2009. Eur J Cancer, 2013. 49: 336.
https://pubmed.ncbi.nlm.nih.gov/23062667
378. Jardine, A.G., et al. Prevention of cardiovascular disease in adult recipients of kidney transplants.
Lancet, 2011. 378: 1419.
https://pubmed.ncbi.nlm.nih.gov/22000138
379. Liefeldt, L., et al. Risk factors for cardiovascular disease in renal transplant recipients and strategies
to minimize risk. Transpl Int, 2010. 23: 1191.
https://pubmed.ncbi.nlm.nih.gov/21059108
380. Nankivell, B.J., et al. Diagnosis and prevention of chronic kidney allograft loss. Lancet, 2011.
378: 1428.
https://pubmed.ncbi.nlm.nih.gov/22000139
381. Boor, P., et al. Renal allograft fibrosis: biology and therapeutic targets. Am J Transplant, 2015.
15: 863.
https://pubmed.ncbi.nlm.nih.gov/25691290
382. Westall, G.P., et al. Antibody-mediated rejection. Curr Opin Organ Transplant, 2015. 20: 492.
https://pubmed.ncbi.nlm.nih.gov/26262460
383. Chapman, J.R. Chronic calcineurin inhibitor nephrotoxicity-lest we forget. Am J Transplant, 2011.
11: 693.
https://pubmed.ncbi.nlm.nih.gov/21446974

5. CONFLICT OF INTEREST
All members of the EAU Renal Transplantation Guidelines Panel have provided disclosure statements on all
relationships that they have that might be perceived to be a potential source of a conflict of interest. This
information is publically accessible through the EAU website: http://www.uroweb.org/guidelines/. These
Guidelines were developed with the financial support of the EAU. No external sources of funding and support
have been involved. The EAU is a non-profit organisation, and funding is limited to administrative assistance,
travel and meeting expenses. No honoraria or other reimbursements have been provided.

6. CITATION INFORMATION
The format in which to cite the EAU Guidelines will vary depending on the style guide of the journal in which the
citation appears. Accordingly, the number of authors or whether, for instance, to include the publisher, location,
or an ISBN number may vary.

The compilation of the complete Guidelines should be referenced as:


EAU Guidelines. Edn. presented at the EAU Annual Congress Milan Italy 2021. ISBN 978-94-92671-13-4.

If a publisher and/or location is required, include:


EAU Guidelines Office, Arnhem, The Netherlands. http://uroweb.org/guidelines/compilations-of-all-guidelines/

References to individual guidelines should be structured in the following way:


Contributors’ names. Title of resource. Publication type. ISBN. Publisher and publisher location, year.

RENAL TRANSPLANTATION - TEXT UPDATE 2021 55


56 RENAL TRANSPLANTATION - TEXT UPDATE 2021
EAU Guidelines on
Bladder Stones
C. Türk (Chair), J.F. Donaldson, A. Neisius,
A. Petřík, C. Seitz, A. Skolarikos (Vice-chair), B. Somani,
K. Thomas, G. Gambaro (Consultant nephrologist)
Guidelines Associates: N. Davis, R. Lombardo, L. Tzelves

© European Association of Urology 2021


TABLE OF CONTENTS PAGE
1. INTRODUCTION 3
1.1 Aims and Scope 3
1.2 Panel Composition 3
1.3 Available Publications 3
1.4 Publication History and Summary of Changes 3
1.4.1 Publication History 3
1.4.2 Summary of Changes 3

2. METHODS 3
2.1 Data Identification 3
2.2 Review 4

3. GUIDELINES 4
3.1 Prevalence, aetiology and risk factors 4
3.2 Presentation 5
3.3 Diagnostic evaluation 5
3.3.1 Diagnostic investigations for bladder stones 5
3.3.2 Diagnosing the cause of bladder stones 6
3.4 Disease Management 6
3.4.1 Conservative treatment and Indications for active stone removal 6
3.4.2 Medical management of bladder stones 6
3.4.3 Bladder stone interventions 6
3.4.3.1 Suprapubic cystolithotomy 6
3.4.3.2 Transurethral cystolithotripsy 6
3.4.3.2.1 Transurethral cystolithotripsy in adults 6
3.4.3.2.1.1 Lithotripsy modalities used during
transurethral cystolithotripsy in adults 7
3.4.3.2.1.2 Transurethral cystolithotripsy in children 7
3.4.3.3 Percutaneous cystolithotripsy 7
3.4.3.3.1 Percutaneous cystolithotripsy in adults: 7
3.4.3.3.2 Percutaneous cystolithotripsy in children: 7
3.4.3.4 Extracorporeal shock wave lithotripsy 7
3.4.3.4.1 Shock wave lithotripsy in adults 7
3.4.3.4.2 Shock wave lithotripsy in children 8
3.4.3.5 Laparoscopic cystolithotomy 8
3.4.4 Treatment for bladder stones secondary to bladder outlet obstruction in adult men 8
3.4.5 Special situations 8
3.4.5.1 Neurogenic bladder and stone formation 8
3.4.5.2 Bladder Augmentation 9
3.4.5.3 Urinary diversion 9
3.4.5.4 Treatment of stones in patients with bladder augmentation or
urinary diversion 9

4. FOLLOW-UP 10

5. REFERENCES 13

6. CONFLICT OF INTEREST 19

7. CITATION INFORMATION 19

2 BLADDER STONES - LIMITED UPDATE MARCH 2021


1. INTRODUCTION
1.1 Aims and Scope
The European Association of Urology (EAU) Bladder Stones Guidelines Panel, a sub-panel of the EAU
Urolithiasis Guidelines Panel, has prepared these guidelines to help urologists assess evidence-based
management of calculi in native urinary bladders and urinary tract reconstructions and to incorporate
recommendations into clinical practice. The management of upper urinary tract stones is addressed in a
separate document: the EAU Guidelines on Urolithiasis.
It must be emphasised that clinical guidelines present the best evidence available to the experts
but following guideline recommendations will not necessarily result in the best outcome. Guidelines can never
replace clinical expertise when making treatment decisions for individual patients, but rather help to focus
decisions - also taking personal values and preferences/individual circumstances of patients into account.
Guidelines are not mandates and do not purport to be a legal standard of care.

1.2 Panel Composition


The EAU Bladder Stones Guidelines Panel consists of an international group of clinicians with particular expertise
in this area. All experts involved in the production of this document have submitted potential conflict of interest
statements which can be viewed on the EAU website Uroweb: http://uroweb.org/guideline/bladderstones/.

1.3 Available Publications


A quick reference document (Pocket Guidelines) is available, both in print and as an app for iOS and Android
devices. These are abridged versions, which may require consultation together with the full text versions. The
EAU Urolithiasis Panel has also published a number of scientific publications in the EAU journal European Urology
[1-3]. All documents can be accessed through the EAU website: http://uroweb.org/guideline/bladderstones/.

1.4 Publication History and Summary of Changes


1.4.1 Publication History
The EAU Bladder stones Guidelines were first published in 2019. This 2021 document presents a full update of
the 2020 text.

1.4.2 Summary of Changes


The literature throughout the entire document has been reassessed and updated (see Methods section below).

For 2021 the content of each chapter has been rephrased and re-assessed; in particular the discussion of
metabolic factors in section 3.1, Prevalence, aetiology and risk factors has seen significant revision. The
summary of evidence and recommendations tables have been completely revised and updated.

2. METHODS
2.1 Data Identification
For the 2021 Bladder Stones guideline, new and relevant evidence was identified, collated, and appraised
through a structured assessment of the literature. The search was limited to studies representing high levels
of evidence only (i.e., systematic reviews with meta-analysis (MA), randomised controlled trials (RCTs),
and prospective non-randomised comparative studies) published in the English language. The search was
restricted to articles published between February 2019 and April 2020. A detailed search strategy is available
online: http://uroweb.org/guideline/bladder-stones/?type=appendices-publications.

The chapters on the treatment of bladder stones in adults and children are based on a systematic review [4].

All methodological information can be found in the general Methodology section of this print, and online at the
EAU website; http://www.uroweb.org/guideline/. A list of associations endorsing the EAU Guidelines can also
be viewed online at the above address.

BLADDER STONES - LIMITED UPDATE MARCH 2021 3


For each recommendation within the guidelines there is an accompanying online strength rating form, the
basis of which is a modified GRADE methodology [5-7]. Each strength-rating form addresses a number of key
elements, namely:
1. the overall quality of the evidence which exists for the recommendation, references used in
this text are graded according to a classification system modified from the Oxford Centre for
Evidence-Based Medicine Levels of Evidence [8];
2. the magnitude of the effect (individual or combined effects);
3. the certainty of the results (precision, consistency, heterogeneity and other statistical or
study related factors);
4. the balance between desirable and undesirable outcomes;
5. the impact of patient values and preferences on the intervention;
6. the certainty of those patient values and preferences.

These key elements are the basis which panels use to define the strength rating of each recommendation.
The strength of each recommendation is represented by the words ‘strong’ or ‘weak’ [9]. The strength of each
recommendation is determined by the balance between desirable and undesirable consequences of alternative
management strategies, the quality of the evidence (including certainty of estimates), and nature and variability
of patient values and preferences.

2.2 Review
This document was peer reviewed prior to publication in 2019.

3. GUIDELINES
3.1 Prevalence, aetiology and risk factors
Bladder stones constitute only approximately 5% of all urinary tract stones [10], yet are responsible for 8% of
urolithiasis-related mortalities in developed nations [11]. The incidence is higher in developing countries [12].
The prevalence of bladder stones is higher in males, with a reported male:female ratio between 10:1 and 4:1
[13, 14]. The age distribution is bimodal: incidence peaks at three years in children in developing countries [13,
15], and 60 years in adulthood [14].

The aetiology of bladder stones is typically multi-factorial [14]. Bladder stones can be classified as primary,
secondary or migratory [16].

Primary or endemic bladder stones occur in the absence of other urinary tract pathology, typically seen in
children in areas with poor hydration, recurrent diarrhoea and a diet deficient in animal protein [17].

Secondary bladder stones occur in the presence of other urinary tract abnormalities, which include: bladder
outlet obstruction (BOO), neurogenic bladder dysfunction, chronic bacteriuria, foreign bodies (including
catheters), bladder diverticula and bladder augmentation or urinary diversion. In adults, BOO is the most
common predisposing factor for bladder stone formation and accounts for 45-79% of vesical calculi
[14, 18-21].

Migratory bladder stones are those which have passed from the upper urinary tract where they formed and
may then serve as a nidus for bladder stone growth. Patients with bladder calculi are more likely to have a
history of upper tract stones and risk factors for their formation [22].

A wide range of metabolic urinary abnormalities can pre-dispose to calculi anywhere in the urinary tract, which
is covered in more detail in the EAU Urolithiasis Guideline [23]. There is a paucity of studies on the specific
metabolic abnormalities which predispose to bladder stones.

Bladder stones will form in 3-4.7% of men undergoing surgery for benign prostatic obstruction (BPO) [24, 25],
15-36% of spinal cord injury patients [26-28], and 2.2% of patients with long-term catheters [29]. Of 57 men
with chronic urinary retention secondary to BPO, the urine of the 30 men with bladder stones had a higher uric
acid concentration (2.2 vs. 0.6 mmol/L, p < 0.01), lower magnesium (106 vs. 167 mmol/L, p = 0.01) and lower
pH (5.9 vs. 6.4, p = 0.02) than the 27 men without bladder stones [22]. It is therefore likely that patients with
these conditions who form bladder stones also have an abnormal urine composition which pre-disposes them
to bladder stone formation.

4 BLADDER STONES - LIMITED UPDATE MARCH 2021


The metabolic abnormalities which pre-dispose patients to form secondary bladder stones are poorly
understood. Stone analysis of 86 men with a BPO-related bladder stone demonstrated 42% had calcium-
based stones (oxalate, phosphate), 33% had magnesium ammonium phosphate, 10% had mixed stones and
14% had urate stones [14]. Similar findings were reported in more recent studies [30-32] and it is therefore
likely that multiple metabolic factors pre-dispose patients to secondary bladder stone formation.

The exact metabolic basis for primary bladder stones is poorly understood and likely multi-factorial. Low urine
volume (poor hydration) is the most consistently demonstrable abnormality [33-35]. Twenty-four-hour urine
analysis in children with endemic bladder stones is reported in two studies. Of 57 children in Pakistan, 89.5%
had hypocitraturia, 49% had a low urine volume, 44% had hyperoxaluria and 42% had hypocaluria [33]. Of 61
children in India, stone formers had higher urine calcium and uromucoid concentrations than controls [34]. One
study from Thailand compared 24 hour urine analyses from children from a rural area with a high prevalence
of bladder stones with those from an urban area: rural children had lower urine volumes and, despite equal
calcium, oxalate and uric acid concentrations, crystalluria with uric acid and calcium oxalate crystals was more
prevalent in rural children [35].

Table 3.1 Bladder stones classified by aetiology


Type of Primary Secondary Migratory
bladder
stone
Cause / Occur in the absence of BOO (e.g., BPO, Urethral Form in the upper urinary
Associations other urinary tract pathology, stricture) tract, then passed into the
typically in children in areas Neurogenic bladder bladder where they may be a
with poor hydration, recurrent dysfunction nidus for stone growth
diarrhoea, and a diet deficient Chronic bacteriuria
in animal protein Foreign bodies (including
catheters)
Bladder diverticula
Bladder augmentation
Urinary diversion
BOO = Bladder Outlet Obstruction; BPO = Benign Prostatic Obstruction.

3.2 Presentation
The symptoms most commonly associated with bladder stones are urinary frequency, haematuria (which is
typically terminal) and dysuria or suprapubic pain, which are worst towards the end of micturition. Sudden
movement and exercise may exacerbate these symptoms. Detrusor over-activity is found in over two thirds of
adult male patients with vesical calculi and is significantly more common in patients with larger stones (> 4 cm).
However, recurrent urinary tract infections (UTIs) may be the only symptom [19, 20].

In children, symptoms may also include pulling of the penis, difficulties in micturition, urinary retention, enuresis
and rectal prolapse (resulting from straining due to bladder spasms). Bladder stones may also be an incidental
finding in 10% of cases [17, 36].

3.3 Diagnostic evaluation


3.3.1 Diagnostic investigations for bladder stones
Plain X-ray of kidney ureter bladder (KUB) has a reported sensitivity of 21%-78% for cystoscopically detected
bladder stones in adults [19, 37]. Larger (> 2.0 cm) stones are more likely to be radiopaque [37]. However, plain
X-Ray provides information on radio-opacity which may guide treatment and follow-up [23].

Ultrasound (US) has a reported sensitivity and specificity of 20-83% and 98-100%, respectively for the
detection of bladder stones in adults [38, 39]. Computer tomography (CT) and cystoscopy have a higher
sensitivity for detecting bladder stones than US or X-Ray in adults [38, 39]. No study compares cystoscopy and
CT for the diagnosis of bladder stones. Cystoscopy has the advantage of detecting other potential causes for a
patient’s symptoms (e.g., bladder cancer), whilst CT can also assess upper tract urolithiasis [23, 40].

There is a paucity of evidence for the investigation of bladder stones, particularly in children [41, 42]. See also
EAU Guidelines on Urolithiasis, Section 3.3, for further information on diagnostic imaging for urolithiasis [23].
The principle of ALARA (As Low As Reasonably Achievable) should be applied, especially in children [43].

BLADDER STONES - LIMITED UPDATE MARCH 2021 5


3.3.2 Diagnosing the cause of bladder stones
The cause of the bladder stone should be considered prior to bladder stone treatment as eliminating the
underlying cause will reduce recurrence rates [44]. The following should be performed where possible prior to
(or at the time of) bladder stone treatment:
• physical examination of external genitalia, peripheral nervous system (including digital rectal examination,
peri-anal tone and sensation in men);
• uroflowmetry and post-void residual urine assessment;
• urine dipstick to include pH ± culture;
• metabolic assessment (see also EAU Guideline on Urolithiasis section 3.3.2.3) including: serum
(creatinine, (ionised) calcium, uric acid, sodium, potassium, blood cell count);
• urine pH;
• stone analysis: in first-time formers using a valid procedure (X-ray diffraction or infrared spectroscopy).

The following investigations should also be considered for selected patients:


• upper tract imaging (in patients with a history of urolithiasis or loin pain);
• cysto-urethroscopy or urethrogram.

3.4 Disease Management


3.4.1 Conservative treatment and Indications for active stone removal
Migratory bladder stones in adults may typically be left untreated, especially asymptomatic small stones. Rates
of spontaneous stone passage are unknown, but data on ureteric stones suggest stones < 1 cm are likely to
pass in the absence of BOO, bladder dysfunction or long-term catheterisation [23].
Primary and secondary bladder stones are usually symptomatic and are unlikely to pass
spontaneously: active treatment of such stones is usually indicated.

3.4.2 Medical management of bladder stones


There is a paucity of evidence on chemolitholysis of bladder stones. However, guidance on the medical
management of urinary tract stones in Chapter 3.4.9 of the EAU Urolithiasis Guidelines [23] can be applied
to urinary stones in all locations. Uric acid stones can be dissolved by oral urinary alkalinisation when a
PH > 6.5 is consistently achieved, typically using an alkaline citrate or sodium bicarbonate. Regular monitoring
is required during therapy [23]. Irrigation chemolysis is also possible using a catheter; however, this is time
consuming and may cause chemical cystitis and is therefore not commonly employed [45, 46].

3.4.3 Bladder stone interventions


Minimally invasive techniques for the removal of bladder stones have been widely adopted to reduce the
risk of complications and shorten hospital stay and convalescence. Bladder stones can be treated with
open, laparoscopic, robotic assisted laparoscopic, endoscopic (transurethral or percutaneous) surgery or
extracorporeal shock wave lithotripsy (SWL) [4].

3.4.3.1 Suprapubic cystolithotomy


Open suprapubic cystolithotomy is very effective, but is associated with a need for catheterisation and longer
hospital stay in both adults and children compared to all other stone removal modalities [4]. In children, a
non-randomised study found that, if the bladder was closed meticulously in two layers, “tubeless” (drain-less
and catheter-less) cystolithotomy was associated with a significantly shorter length of hospital stay compared
with traditional cystolithotomy, without significant differences regarding late or intra-operative complications
provided that children with prior UTI, recurrent stones, or with previous surgery for anorectal malformation (or
other relevant surgery) were excluded [47].

3.4.3.2 Transurethral cystolithotripsy


In both adults and children, transurethral cystolithotripsy provides high stone-free rates (SFR) and appears to
be safe, with a very low-risk of unplanned procedures and major post-operative and late complications [4].

3.4.3.2.1 Transurethral cystolithotripsy in adults


In adults, meta-analysis of four RCTs including 409 patients demonstrated that transurethral cystolithotripsy
has a shorter hospital stay and convalescence with less pain, but equivalent SFR and complications compared
to percutaneous cystolithotripsy [4]. Transurethral cystolithotripsy with a nephroscope was quicker than
percutaneous cystolithotripsy in three RCTs, although transurethral cystolithotripsy with a cystoscope was
slower than percutaneous cystolithotripsy [4].

6 BLADDER STONES - LIMITED UPDATE MARCH 2021


Rates of urethral strictures following transurethral procedures were not robustly reported: studies report rates
between 2.9% and 19.6% during a follow up of 12 – 24 months [4, 30, 48].

One small RCT demonstrated a shorter duration of catheterisation, hospital stay and procedure with
transurethral cystolithotripsy than open cystolithotomy with similar SFR [4]. Meta-analysis of four RCTs
found shorter procedure duration for transurethral cystolithotripsy using a nephroscope vs. cystoscope
with similar SFRs, hospital stay, convalescence, pain, and complications [4, 30, 49-51]. Two retrospective
studies (n=188) reported that using a resectoscope or nephroscope was associated with a shorter procedure
duration (p < 0.05) than a cystoscope for transurethral cystolithotripsy [52, 53]. This suggests that transurethral
cystolithotripsy is quicker when using a continuous flow instrument.

3.4.3.2.1.1 Lithotripsy modalities used during transurethral cystolithotripsy in adults


When considering lithotripsy modalities for transurethral cystolithotripsy, the Panel’s systematic review found
very low-quality evidence from five non-randomised studies (n=385) which found no difference in SFR between
modalities (mechanical, laser, pneumatic, ultrasonic, electrohydraulic lithotripsy [EHL] or washout alone) [4].
Unplanned procedures and major post-operative complications were low-rate events and were not significantly
different between lithotripsy modalities, although one non-randomised study (NRS) suggested these might be
higher with EHL or mechanical lithotripsy than pneumatic or ultrasonic lithotripsy [54]. All outcomes had very
low-quality of evidence (GRADE) [4]. While the laser power setting (30W vs. 100W) does not seem to influence
lithotripsy time significantly [31], laser lithotripsy was faster than pneumatic lithotripsy (MD 16.6 minutes; CI:
23.51-9.69, p < 0.0001) in one NRS (n=62); however, a laser was used with a resectoscope and the pneumatic
device with a cystoscope [55]. Continuous vs. intermittent irrigating instrument may affect the operation time
more significantly than the choice of lithotripsy device [4].

3.4.3.2.1.2 Transurethral cystolithotripsy in children


In children, three NRS suggest that transurethral cystolithotripsy has a shorter hospital stay and catheterisation
time than open cystolithotomy, but similar stone-free and complication rates [4, 56]. One small quasi RCT
found a shorter procedure time using laser vs. pneumatic lithotripsy for < 1.5 cm bladder stones with no
difference in SFR or other outcomes [4, 57].

3.4.3.3 Percutaneous cystolithotripsy


3.4.3.3.1 Percutaneous cystolithotripsy in adults:
One NRS found a shorter duration of procedure and catheterisation and less blood loss for percutaneous,
compared with open surgery in adult male patients with urethral strictures; all patients in both groups were
rendered stone-free [32].

Meta-analysis of four RCTs comparing transurethral and percutaneous cystolithotripsy found a shorter hospital
stay for transurethral cystolithotripsy over percutaneous surgery. Transurethral cystolithotripsy was quicker
when using a nephroscope. There were no significant differences in SFRs, major post-operative complications
or re-treatment [4].

3.4.3.3.2 Percutaneous cystolithotripsy in children:


In children, three NRS suggest that percutaneous cystolithotripsy has a shorter hospital stay and
catheterisation time, but a longer procedure duration and more peri-operative complications than open
cystolithotripsy; SFRs were similar [4, 36, 56].

Two small NRS compared percutaneous and transurethral cystolithotripsy and both found similar SFRs, but
that transurethral surgery offers a shorter duration of catheterisation and hospital stay [36, 56]. One small NRS
found a non-significant increased risk of unplanned procedures (within 30 days of primary procedure) and
major post-operative complications for percutaneous operations compared with transurethral procedures;
however, age and stone size determined which intervention children underwent and all patients were rendered
stone-free [36]. Urethral stricture rates were not robustly compared in either study.

3.4.3.4 Extracorporeal shock wave lithotripsy


Extracorporeal SWL is the least invasive therapeutic procedure [4].

3.4.3.4.1 Shock wave lithotripsy in adults


In adults, one RCT compared SWL with transurethral cystolithotripsy in 100 patients with ≤ 2 cm bladder
stones presenting with acute urinary retention. Stone free rate after one SWL session favored transurethral

BLADDER STONES - LIMITED UPDATE MARCH 2021 7


cystolithotripsy (86% vs. 98%, p=0.03); however, following up to three sessions of SWL, there was no
significant difference in SFR (94% vs. 98%, p=0.3) [4, 58].

Two NRS compared transurethral cystolithotripsy vs. SWL and found no significant difference in SFR (97.0%
vs. 93.9%, p=0.99, 97.7% vs. 89.7% p=0.07) despite larger stones in transurethral cystolithotripsy patients
(4.2 vs. 2.5 cm, p=0.014; and 3.6 vs. 2.6 cm [p value not reported]) [59, 60].

Length of hospital stay appeared to favour SWL in all three studies (0 vs. 1 day, 4.8 vs. 0 days, p=0.02,
0.8 vs. 2.4 days, respectively) [58-60]. No significant differences in major post-operative or intra-operative
complications were reported in any study [58-60].

One NRS compared SWL vs. open cystolithotomy in just 43 patients. Stone sizes were not comparable (2.5
vs. 7.4 cm, p < 0.001). SFR were not significantly different (93.9% vs. 100%, p=0.50). Length of stay favoured
SWL. There was no significant difference in intra-operative or major post-operative complications [59].

3.4.3.4.2 Shock wave lithotripsy in children


One large NRS found lower SFR for SWL than both transurethral cystolithotripsy and open cystolithotomy,
despite treating smaller stones with SWL. However, the length of hospital stay favoured SWL over open
cystolithotomy, although this appeared to be comparable between SWL and transurethral cystolithotripsy [61].

3.4.3.5 Laparoscopic cystolithotomy


Laparoscopic cystolithotomy has been described in adults and is typically performed in combination with
simple prostatectomy using either traditional laparoscopy or with robotic assistance [62, 63]. A SR found no
studies comparing laparoscopic surgery with other procedures [4].

3.4.4 Treatment for bladder stones secondary to bladder outlet obstruction in adult men
Bladder stones in men aged over 40 years may be caused by BPO, the management of which should also
be considered. Bladder stones were traditionally an indication for a surgical intervention for BPO: a doctrine
which has been questioned by recent studies. One prospective study reports urodynamics (cystometrogram)
findings in 46 men aged > 60 years before and after bladder stone treatment [20]. Only 51% of men had BOO
while 10% had detrusor under-activity. Eighteen percent of men had a completely normal urodynamic study
and 68% had detrusor over-activity. There was no significant difference between pre- and post-bladder stone
removal urodynamic findings [20].

One NRS compared 64 men undergoing transurethral cystolithotripsy with either transurethral resection of
prostate (TURP) or medical management for BPO (α-blocker with or without 5-alpha reductase inhibitor). After
28 months follow-up, no men on medication had had a recurrence, but 34% underwent TURP: a high post-
void residual urine volume predicted the need for subsequent TURP [64]. Another observational study of 23
men undergoing cystolithotripsy and commencing medical management for BPO found 22% developed a BPO
related complication, including 17% who had recurrent stones [44].

Large studies support the safety of performing BPO and bladder stone procedures during the same operation
with no difference in major complications compared to a BPO procedure alone [65-67]. An observational
study on 2,271 patients undergoing TURP found no difference in complications except UTIs, which occurred
slightly more frequently in patients with simultaneously treated bladder stones: 0% vs. 0.6%, p=0.044 [65]. An
observational study of 321 men undergoing Holmium laser enucleation of the prostate (HoLEP) found a higher
rate of early post-operative incontinence (26.8% vs. 12.5%, p=0.03) in men having concomitant transurethral
cystolithotripsy, but no difference in long-term continence rates [66].

3.4.5 Special situations


3.4.5.1 Neurogenic bladder and stone formation
Patients with a neurogenic bladder secondary to spinal cord injury or myelomeningocele are at increased risk
of forming bladder stones. Within eight to ten years, 15-36% of patients with spinal cord injury will develop
a bladder stone [26-28]. The absolute annual risk of stone formation in spinal cord injury patients with an
indwelling catheter is 4% compared with 0.2% for those voiding with clean intermittent self-catheterisation
(CISC) [68].

A study of 2,825 spinal cord injury patients over eight years found a 3.3% incidence of bladder stones: 2% with
CISC, 6.6% with indwelling urethral catheter, 11% with a suprapubic catheter and 1.1% in patients voiding
using reflex micturition [69]. However, another study of 457 spinal cord injury patients for six months found no

8 BLADDER STONES - LIMITED UPDATE MARCH 2021


difference in bladder stones between urethral and suprapubic catheterisation [68]. Spinal cord injury patients
with an indwelling urethral catheter are six times more likely to develop bladder stones than patients with
normal micturition [28, 69].

The risk of stone recurrence after complete removal in spinal cord injury patients is 16% per year [68]. A RCT of
78 spinal cord injury patients who perform CISC found a significant reduction in bladder stone formation when
twice weekly manual bladder irrigations were performed for six months (49% vs. 0%, p=< 0.0001), as well as
less symptomatic UTIs (41% vs. 8%; p=0.001) [70]. However, this study excluded patients who developed
autonomic dysreflexia during bladder irrigations. The irrigation volume used was not reported.

3.4.5.2 Bladder Augmentation


The incidence of vesical calculus formation after bladder augmentation is 2-44% in adults [71-79], and 4-53%
in children [80-94]. Following cystoplasty, stones form after 24-31 months in adults [72, 74, 79], and after 25-68
months in children [84, 86, 88, 92, 95-97].The reported cumulative incidence of bladder stone formation after
ten years is 28-36% and after twenty years is 41% [93, 98].

Risk factors for bladder stone formation after augmentation include excess mucus production, incomplete
bladder emptying, non-compliance with CIC or bladder irrigations, bacteriuria or urinary tract infections (due
to urease-producing bacteria), foreign bodies (including staples, mesh, non-absorbable sutures), drainage
by vesico-entero-cystostomy (Mitrofanoff or Monti) [72, 75, 77, 78, 83, 84, 86, 98, 99] and voiding by CISC
compared with those voiding spontaneously [76]. Gastric segment augmentation confers a lower risk of
bladder stones than ileal or colononic segment cystoplasty [80, 83, 84, 86].

In previous stone formers, the rate of recurrence is 15-44% in adults [72-74, 76, 79], and 19-56% in children
[80, 83, 84, 86, 88-90, 93, 97, 99]. The risk of recurrence is greatest during the first two years, at about 12% per
patient per year, with the risk decreasing with time [97].

Daily, or three-times-weekly bladder irrigations reduce the incidence of bladder stones following bladder
augmentation or continent urinary diversion [75, 99]. A randomised study found that daily bladder irrigation
with 240 mL of saline reduced stone recurrences (p< 0.0002, p=0.0152) and symptomatic UTIs (p< 0.0001,
p< 0.0001) compared to 60mL or 120mL [75]. The frequency of bladder irrigations required is unclear.

3.4.5.3 Urinary diversion


The incidence of stone formation after urinary diversion with an ileal or colon conduit is 0-3% [100, 101]. The
incidence of stone formation is 0-34% in orthotopic ileal neobladders (Hautmann, hemi-Kock, Studer, T-pouch
or w-neobladder) [76, 100, 102-110], and 4-6% in orthotopic sigmoid neobladders (Reddy) [106, 111]. The risk
of pouch stone formation is 4-43% in adults with an ileocaecal continent cutaneous urinary diversion (Indiana,
modified Indiana, Kock or Mainz I) [76, 100, 101, 109, 112, 113]. The average interval from construction of
the urinary diversion to stone detection is 71-99 months [105, 114]. In children, the incidence of neobladder
stone formation is 30% after Mainz II diversion (rectosigmoid reservoir) [81], and 27% after Kock ileal reservoir
construction [91].

3.4.5.4 Treatment of stones in patients with bladder augmentation or urinary diversion


Stones may be removed by open or endoscopic surgery in patients with bladder augmentation or diversion
[90]. However, often access cannot be obtained through a continent vesico-entero-cystostomy without
damaging the continence apparatus; hence a percutaneous or open approach is typically preferred [90].

No studies comparing outcomes following procedures for stones in reconstructed or augmented bladders
were found. Two observational studies indicate that percutaneous lithotomy can be safely performed with
US or CT guidance in patients with reconstructed or augmented bladders [115, 116] and is proposed to offer
similar advantages over open surgery to those for percutaneous native bladder surgery. Stone recurrence after
successful removal has been reported to be 10-42% [115, 116], but appears to be unrelated to the modality
used for stone removal [79, 83, 84, 86, 89, 97].

BLADDER STONES - LIMITED UPDATE MARCH 2021 9


Figure 3.4 Management of Bladder stones

* Lithotripsy modality at surgeon’s discretion (e.g., mechanical, laser, pneumatic, ultrasonic).


† Prefer “tubeless” procedure (without placing a catheter or drain) for children with primary bladder stones and

no prior infection, surgery, or bladder dysfunction where open cystolithotomy is indicated.


** Stone analysis should be sent for all first-time stone formers and in patients who develop a recurrence under
pharmacological prevention, early recurrence after interventional therapy with complete stone clearance or
late recurrence after a prolonged stone-free period (see main Urolithiasis guideline).
†† Use an alkaline citrate or sodium bicarbonate with frequent urine pH monitoring and dose titration to achieve

a consistent pH > 6.5.

BOO = Bladder Outlet Obstruction, TUCL = Trans-urethral cystolithotripsy, PCCL = Percutaneous


Cystolithotripsy, SWL = Shock-wave Lithotripsy.

4. FOLLOW-UP
There are no studies examining the merits of differing follow-up modalities or frequencies following
conservative, medical, or operative treatment of bladder stones in adults or children. Identification and
prevention of the cause of bladder stone formation will be crucial to prevent recurrence (see section 3.3.2).

In adults, there is a paucity of evidence on dietary modification or medical treatment for the prevention of
bladder stone recurrence. Recommendations in the EAU Guideline on Urolithiasis, based on evidence from
upper tract stones, constitutes the best available recommendations, especially for migratory bladder stones
(see chapter 4 in the main EAU Urolithiasis guideline) [23].

10 BLADDER STONES - LIMITED UPDATE MARCH 2021


Where it is possible to address the cause of secondary bladder stones (e.g., treatment of BPO), it is unclear
whether metabolic intervention would offer any significant additional benefit in preventing stone recurrence.
However, especially where the secondary cause cannot be addressed (e.g., indwelling catheter, neuropathic
bladder, bladder augmentation or urinary diversion); metabolic interventions are likely to reduce bladder stone
recurrence rates.

Regular bladder irrigation reduces the chances of bladder stone recurrence in adults and children with bladder
augmentation or continent cutaneous urinary diversion and adults with spinal cord injury who perform CISC
(see section 3.3.5) [70, 75, 100].

In children with primary (endemic) bladder stones maintenance of hydration, avoidance of diarrhoea and a
mixed cereal diet with milk and Vitamins A and B supplements, with the addition of eggs, meat and boiled
cows’ milk after one year of age are recommended to prevent recurrence [33].

Finally, there are contradictory reports on a possible association between bladder calculi and future
development of bladder cancer [117-119]. The need for follow-up with regular cystoscopy therefore remains
controversial.

Summary of evidence LE
The incidence of bladder stones peaks at three years in children (endemic/primary stones in 2c
developing countries) and 60 years in adults.
The aetiology of bladder stones is typically multi-factorial. Bladder stones can be classified as primary 4
(endemic), secondary (associated with lower urinary tract abnormalities e.g., BPO, neuropathic
bladder, foreign body, chronic bacteriuria) or migratory (having formed in the upper tract).
In adults, BOO is the most common pre-disposing factor for bladder stone formation. 2c
Of men undergoing surgery for BPO, 3-4.7% form bladder stones. 2b
Metabolic abnormalities are also likely to contribute to bladder stone formation in patients with 2b
secondary bladder stones.
Primary (endemic) bladder stones typically occur in children in areas with poor hydration, recurrent 5
diarrhoea, and a diet deficient in animal protein. The following measures are proposed to reduce their
incidence: maintenance of hydration, avoidance of diarrhoea, and a mixed cereal diet with milk and
Vitamins A and B supplements; with the addition of eggs, meat, and boiled cows’ milk after one year
of age.
In adults, US has a sensitivity of 20-83% for diagnosing bladder stones. 2b
In adults, XR-KUB has a sensitivity of 21-78%; sensitivity increases with stone size. 2b
Computer tomography has a higher sensitivity than US for the detection of bladder stones. 2b
Cystoscopy has a higher sensitivity than XR-KUB or US for the detection of bladder stones. 2b
Endoscopic bladder stone treatments (trans-urethral or percutaneous) are associated with comparable 1a
SFRs, but a shorter length of hospital stay, duration of procedure and duration of catheterisation
compared to open cystolithotomy in adults.
Stone-free rates are lower in patients treated with SWL than those treated with open or endoscopic 2a
procedures in both adults and children.
Transurethral cystolithotripsy is associated with a shorter length of hospital stay, less pain and a 1b
shorter convalescence period than percutaneous cystolithotripsy in adults.
Transurethral cystolithotripsy with a nephroscope is quicker than when using a cystoscope with no 1a
difference in SFR in adults.
Transurethral cystolithotripsy with a resectoscope is quicker than when using a cystoscope with no 2a
difference in SFR in adults.
Mechanical, pneumatic and laser appear equivalent lithotripsy modalities for use in endoscopic 2a
bladder stone treatments in adults and children.
Open cystolithotomy without a retropubic drain or urethral catheter (“tubeless”) is associated with a 2b
shorter length of hospital stay than traditional cystolithotomy and can be performed safely in children
with primary stones and no prior bladder surgery or infections.
Bladder stone removal with concomitant treatment for BOO is associated with no significant 2b
difference in major post-operative complications when compared to BOO treatment alone in adults.
However, concomitant bladder stone treatment does increase the rates of short-term post-operative
incontinence and UTI.

BLADDER STONES - LIMITED UPDATE MARCH 2021 11


The incidence of bladder stone formation in spinal cord injury patients is 15-36% after eight to ten 2b
years. The absolute annual risk of stone formation in spinal cord injury patients is significantly higher
with an indwelling catheter compared to those voiding with CISC or spontaneously.
The incidence of bladder stone formation after bladder augmentation or vesico-entero-cystostomy is 2b
between 2-53% in adults and children.
Urinary diversion including orthotopic ileal neobladders, ileocaecal continent cutaneous urinary 2b
diversion and rectosigmoid reservoirs is associated with urinary reservoir stone formation in 0-43%.
The risk of bladder stone formation in spinal cord injury, bladder augmentation or continent urinary 2b
diversion patients is reduced by performing regular bladder irrigation.

Recommendations Strength rating


Use ultrasound (US) as first-line imaging with symptoms suggestive of a bladder stone. Strong
Use cystoscopy or computer tomography (CT), or kidney-ureter-bladder X-Ray (KUB) to Strong
investigate adults with persistent symptoms suggestive of a bladder stone if US is negative.
Use X-Ray KUB for adults with confirmed bladder stones to guide treatment options and Weak
follow-up.
All patients with bladder stones should be examined and investigated for the cause of Weak
bladder stone formation, including:
• uroflowmetry and post-void residual;
• urine dipstick, pH, ± culture;
• metabolic assessment and stone analysis (see sections 3.3.2.3 and 4.1 of the
Urolithiasis guideline for further details).
In selected patients, consider:
• upper tract imaging (in patients with a history of urolithiasis or loin pain);
• cysto-urethroscopy or urethrogram.
Offer oral chemolitholysis for radiolucent or known uric acid bladder stones in adults. Weak
Offer adults with bladder stones transurethral cystolithotripsy where possible. Strong
Perform transurethral cystolithotripsy with a continuous flow instrument in adults (e.g., Weak
nephroscope or resectoscope) where possible.
Offer adults percutaneous cystolithotripsy where transurethral cystolithotripsy is not Strong
possible or advisable.
Suggest open cystolithotomy as an option for very large bladder stones in adults and Weak
children.
Offer children with bladder stones transurethral cystolithotripsy where possible. Weak
Offer children percutaneous cystolithotripsy where transurethral cystolithotripsy is not Weak
possible or is associated with a high risk of urethral stricture (e.g., young children, previous
urethral reconstruction, and spinal cord injury).
Open, laparoscopic, and extracorporeal shock wave lithotripsy are alternative treatments Weak
where endoscopic treatment is not advisable in adults and children.
Prefer “tubeless” procedure (without placing a catheter or drain) for children with primary Weak
bladder stones and no prior infection, surgery, or bladder dysfunction where open
cystolithotomy is indicated.
Perform procedures for the stone and underlying bladder outlet obstruction (BOO) Strong
simultaneously in adults with bladder stones secondary to BOO, where possible.
Individualise imaging follow up for each patient as there is a paucity of evidence. Weak
Factors affecting follow up will include:
• whether the underlying functional predisposition to stone formation can be treated (e.g.,
TURP);
• metabolic risk.
Recommend regular irrigation therapy with saline solution to adults and children with Weak
bladder augmentation, continent cutaneous urinary reservoir or neuropathic bladder
dysfunction, and no history of autonomic dysreflexia, to reduce the risk of stone recurrence.

12 BLADDER STONES - LIMITED UPDATE MARCH 2021


5. REFERENCES
1. Skolarikos, A., et al. Metabolic evaluation and recurrence prevention for urinary stone patients: EAU
guidelines. Eur Urol, 2015. 67: 750.
https://pubmed.ncbi.nlm.nih.gov/25454613
2. Turk, C., et al. EAU Guidelines on Diagnosis and Conservative Management of Urolithiasis. Eur Urol,
2016. 69: 468.
https://pubmed.ncbi.nlm.nih.gov/26318710
3. Turk, C., et al. EAU Guidelines on Interventional Treatment for Urolithiasis. Eur Urol, 2016. 69: 475.
https://pubmed.ncbi.nlm.nih.gov/26344917
4. Donaldson, J.F., et al. Treatment of Bladder Stones in Adults and Children: A Systematic Review
and Meta-analysis on Behalf of the European Association of Urology Urolithiasis Guideline Panel.
Eur Urol, 2019. 76: 352.
https://pubmed.ncbi.nlm.nih.gov/31311676
5. Balshem, H., et al. GRADE guidelines: 3. Rating the quality of evidence. J Clin Epidemiol, 2011. 64: 401.
https://pubmed.ncbi.nlm.nih.gov/21208779
6. Guyatt, G.H., et al. What is “quality of evidence” and why is it important to clinicians? BMJ, 2008.
336: 995.
https://pubmed.ncbi.nlm.nih.gov/18456631
7. Guyatt, G.H., et al. GRADE: an emerging consensus on rating quality of evidence and strength of
recommendations. BMJ, 2008. 336: 924.
https://pubmed.ncbi.nlm.nih.gov/18436948
8. Phillips B, et al. Oxford Centre for Evidence-based Medicine Levels of Evidence. Updated by
Jeremy Howick March 2009. 1998.
https://www.cebm.net/2009/06/oxford-centre-evidence-based-medicine-levels-evidence-march-2009/
9. Guyatt, G.H., et al. Going from evidence to recommendations. BMJ, 2008. 336: 1049.
https://pubmed.ncbi.nlm.nih.gov/18467413
10. Schwartz, B.F., et al. The vesical calculus. Urol Clin North Am, 2000. 27: 333.
https://pubmed.ncbi.nlm.nih.gov/10778475
11. Kum, F., et al. Do stones still kill? An analysis of death from stone disease 1999-2013 in England
and Wales. BJU Int, 2016. 118: 140.
https://pubmed.ncbi.nlm.nih.gov/26765522
12. Ramello, A., et al. Epidemiology of nephrolithiasis. J Nephrol, 2000. 13 Suppl 3: S45.
https://pubmed.ncbi.nlm.nih.gov/11132032
13. Halstead, S.B. Epidemiology of bladder stone of children: precipitating events. Urolithiasis, 2016.
44: 101.
https://pubmed.ncbi.nlm.nih.gov/26559057
14. Takasaki, E., et al. Chemical compositions of 300 lower urinary tract calculi and associated
disorders in the urinary tract. Urol Int, 1995. 54: 89.
https://pubmed.ncbi.nlm.nih.gov/7538235
15. Naqvi, S.A., et al. Bladder stone disease in children: clinical studies. J Pak Med Assoc, 1984. 34: 94.
https://pubmed.ncbi.nlm.nih.gov/6429380
16. Philippou, P., et al. The management of bladder lithiasis in the modern era of endourology. Urology,
2012. 79: 980.
https://pubmed.ncbi.nlm.nih.gov/22119259
17. Lal, B., et al. Childhood bladder stones-an endemic disease of developing countries. J Ayub Med
Coll Abbottabad, 2015. 27: 17.
https://pubmed.ncbi.nlm.nih.gov/26182729
18. Douenias, R., et al. Predisposing factors in bladder calculi: Review of 100 cases. Urology, 1991.
37: 240.
https://pubmed.ncbi.nlm.nih.gov/2000681
19. Smith, J.M., et al. Vesical stone: the clinical features of 652 cases. Irish Med J, 1975. 68: 85.
https://pubmed.ncbi.nlm.nih.gov/1112692
20. Millan-Rodriguez, F., et al. Urodynamic findings before and after noninvasive management of
bladder calculi. BJU Int, 2004. 93: 1267.
https://pubmed.ncbi.nlm.nih.gov/15180620
21. Yang, X., et al. The value of respective urodynamic parameters for evaluating the occurrence of
complications linked to benign prostatic enlargement. Int Urol Nephrol, 2014. 46: 1761.
https://pubmed.ncbi.nlm.nih.gov/24811567

BLADDER STONES - LIMITED UPDATE MARCH 2021 13


22. Childs, M.A., et al. Pathogenesis of bladder calculi in the presence of urinary stasis. J Urol, 2013.
189: 1347.
https://pubmed.ncbi.nlm.nih.gov/23159588
23. Türk, C., et al., EAU Guidelines on Urolithiasis, in European Association of Urology Guidelines. 2021,
EAU Guidelines Office: Arnhem, The Netherlands.
https://uroweb.org/guideline/urolithiasis/
24. Krambeck, A.E., et al. Experience with more than 1,000 holmium laser prostate enucleations for
benign prostatic hyperplasia. J Urol, 2010. 183: 1105.
https://pubmed.ncbi.nlm.nih.gov/20092844
25. Mebust, W.K., et al. Transurethral prostatectomy: immediate and postoperative complications. a
cooperative study of 13 participating institutions evaluating 3,885 patients. 1989. J Urol, 2002.
167: 999.
https://pubmed.ncbi.nlm.nih.gov/11908420
26. Chen, Y., et al. Bladder stone incidence in persons with spinal cord injury: Determinants and trends,
1973-1996. Urology, 2001. 58: 665.
https://pubmed.ncbi.nlm.nih.gov/
27. Hall, M.K., et al. Renal calculi in spinal cord-injured patient: association with reflux, bladder stones,
and foley catheter drainage. Urology, 1989. 34: 126.
https://pubmed.ncbi.nlm.nih.gov/2789449
28. DeVivo, M.J., et al. The risk of bladder calculi in patients with spinal cord injuries. Arch Int Med,
1985. 145: 428.
https://pubmed.ncbi.nlm.nih.gov/3977510
29. Kohler-Ockmore, J., et al. Long-term catheterization of the bladder: prevalence and morbidity. Br
J Urol, 1996. 77: 347.
https://pubmed.ncbi.nlm.nih.gov/8814836
30. Bansal, A., et al. Prospective randomized comparison of three endoscopic modalities used in
treatment of bladder stones. Urologia, 2016. 83: 87.
https://pubmed.ncbi.nlm.nih.gov/27103095
31. Kawahara, T., et al. Correlation between the operation time using two different power settings of a
Ho: YAG laser: laser power doesn’t influence lithotripsy time. BMC Res Notes, 2013. 6: 80.
https://pubmed.ncbi.nlm.nih.gov/23510531
32. Liu, G., et al. Minimally invasive percutaneous suprapubic cystolithotripsy: An effective treatment for
bladder stones with urethral strictures. Int J Clin Exp Med, 2016. 9: 19907.
http://www.ijcem.com/files/ijcem0023634
33. Soliman, N.A., et al. Endemic bladder calculi in children. Pediatr Nephrol, 2017. 32: 1489.
https://pubmed.ncbi.nlm.nih.gov/27848095
34. Aurora, A.L., et al. Bladder stone disease of childhood. II. A clinico-pathological study. Acta Paediatr
Scand, 1970. 59: 385.
https://pubmed.ncbi.nlm.nih.gov/5447682
35. Valyasevi, A., et al. Studies of bladder stone disease in Thailand. VI. Urinary studies in children, 2-10
years old, resident in a hypo- and hyperendemic area. Am J Clin Nutr, 1967. 20: 1362.
https://pubmed.ncbi.nlm.nih.gov/6074673
36. Al-Marhoon, M.S., et al. Comparison of Endourological and Open Cystolithotomy in the
Management of Bladder Stones in Children. J Urol, 2009. 181: 2684.
https://pubmed.ncbi.nlm.nih.gov/19375100
37. Linsenmeyer, M.A., et al. Accuracy of bladder stone detection using abdominal x-ray after spinal
cord injury. J Spinal Cord Med, 2004. 27: 438.
https://pubmed.ncbi.nlm.nih.gov/15648797
38. Bakin, S., et al. Accuracy of ultrasound versus computed tomography urogram in detecting urinary
tract calculi. Med J Malaysia, 2015. 70: 238.
http://www.e-mjm.org/2015/v70n4/urinary-tract-calculi.pdf
39. Ahmed, F.O., et al. A comparison between transabdominal ultrasonographic and cystourethroscopy
findings in adult Sudanese patients presenting with haematuria. Int Urol Nephrol, 2014. 47: 223.
https://pubmed.ncbi.nlm.nih.gov/25374263
40. Babjuk, M., et al., EAU Guidelines on Non-musle-invasive Bladder Cancer (TaT1 and CIS), in
European Association of Urology Guidelines 2021 edition. 2021, The European Association of
Urology: Arnhem, The Netherlands.
https://uroweb.org/guideline/non-muscle-invasive-bladder-cancer/

14 BLADDER STONES - LIMITED UPDATE MARCH 2021


41. Johnson, E.K., et al. Are stone protocol computed tomography scans mandatory for children with
suspected urinary calculi? Urology, 2011. 78: 662.
https://pubmed.ncbi.nlm.nih.gov/21722946
42. Passerotti, C., et al. Ultrasound versus computerized tomography for evaluating urolithiasis. J Urol,
2009. 182: 1829.
https://pubmed.ncbi.nlm.nih.gov/19692054
43. ICRP. The 2007 Recommendations of the International Commission on Radiological Protection.
ICRP Publication 103. Ann. ICRP 37: 2.
https://www.icrp.org/publication.asp?id=ICRP%20Publication%20103
44. O’Connor, R.C., et al. Nonsurgical management of benign prostatic hyperplasia in men with bladder
calculi. Urology, 2002. 60: 288.
https://pubmed.ncbi.nlm.nih.gov/12137828
45. Lopez, J.R., et al. Irrigating solutions in bladder stone dissolution. Drug Intell Clin Pharm, 1987. 21: 872.
https://pubmed.ncbi.nlm.nih.gov/3678056
46. Rodman, J.S., et al. Dissolution of uric acid calculi. J Urol, 1984. 131: 1039.
https://pubmed.ncbi.nlm.nih.gov/6726897
47. Rattan, K.N., et al. Catheterless and drainless open suprapubic cystolithotomy in children: A safe
procedure. Pediatr Surg Int, 2006. 22: 255.
https://pubmed.ncbi.nlm.nih.gov/16416282
48. Ullah, S., et al. Comparison of open vesicolithotomy and cystolitholapaxy. Pakistan J Med Sci, 2007.
23: 47.
https://www.pjms.com.pk/issues/janmar07/article/article7.html
49. Singh, K.J., et al. Comparison of three different endoscopic techniques in management of bladder
calculi. Indian J Urol, 2011. 27: 10.
https://pubmed.ncbi.nlm.nih.gov/
50. Ozdemir A.T., et al. Randomized comparison of the transurethral use of nephroscope via amplatz
sheath with cystoscope in transurethral cystolithotripsy of bladder stones in male patients.
J Endourol, 2012. 26: A142. [No abstract available].
51. Ener, K., et al. The randomized comparison of two different endoscopic techniques in the
management of large bladder stones: Transurethral use of nephroscope or cystoscope? J Endourol,
2009. 23: 1151.
https://pubmed.ncbi.nlm.nih.gov/19530944
52. Wu, J.H., et al. Combined usage of Ho:YAG laser with monopolar resectoscope in the treatment of
bladder stone and bladder outlet obstruction. Pak J Med Sci, 2014. 30: 908.
https://pubmed.ncbi.nlm.nih.gov/25097543
53. Halis, F., et al. The comparison of percutaneous and transurethral cystolithotripsy methods
simultaneously performed with Transurethral Resection of Prostate in patients with BPH and bladder
stone. Kuwait Med J, 2019. 51: 189.
http://www.kmj.org.kw/previous-issues
54. Razvi, H.A., et al. Management of Vesical Calculi: Comparison of Lithotripsy Devices. J Endourol,
1996. 10: 559.
https://pubmed.ncbi.nlm.nih.gov/8972793
55. Ercil, H., et al. Comparison of Ho:Yag laser and pneumatic lithotripsy combined with transurethral
prostatectomy in high burden bladder stones with benign prostatic hyperplasia. Asian J Surg, 2016.
39: 238.
https://pubmed.ncbi.nlm.nih.gov/25937584
56. Javanmard, B., et al. Surgical Management of Vesical Stones in Children: A Comparison Between
Open Cystolithotomy, Percutaneous Cystolithotomy and Transurethral Cystolithotripsy With
Holmium-YAG Laser. J Lasers Med Sci, 2018. 9: 183.
https://pubmed.ncbi.nlm.nih.gov/30809329
57. Gangkak, G., et al. Pneumatic cystolithotripsy versus holmium:yag laser cystolithotripsy in the
treatment of pediatric bladder stones: a prospective randomized study. Pediatr Surg Int, 2016. 32: 609.
https://pubmed.ncbi.nlm.nih.gov/26879752
58. Ali, M., et al. Shock wave lithotripsy versus endoscopic cystolitholapaxy in the management of
patients presenting with calcular acute urinary retention: a randomised controlled trial. World J Urol,
2019. 37: 879.
https://pubmed.ncbi.nlm.nih.gov/30105456
59. Deswanto, I.A., et al. Management of bladder stones: The move towards non-invasive treatment.
Med J Indonesia, 2017. 26: 128.
https://mji.ui.ac.id/journal/index.php/mji/article/view/1602

BLADDER STONES - LIMITED UPDATE MARCH 2021 15


60. Bhatia, V., et al. A comparative study of cystolithotripsy and extracorporeal shock wave therapy for
bladder stones. Int Urol Nephrol, 1994. 26: 27.
https://pubmed.ncbi.nlm.nih.gov/8026920
61. Rizvi, S.A., et al. Management of pediatric urolithiasis in Pakistan: experience with 1,440 children.
J Urol, 2003. 169: 634.
https://pubmed.ncbi.nlm.nih.gov/12544331
62. Autorino, R., et al. Perioperative Outcomes of Robotic and Laparoscopic Simple Prostatectomy:
A European-American Multi-institutional Analysis. Eur Urol, 2015. 68: 86.
https://pubmed.ncbi.nlm.nih.gov/25484140
63. Matei, D.V., et al. Robot-assisted simple prostatectomy (RASP): does it make sense? BJU Int, 2012.
110: E972.
https://pubmed.ncbi.nlm.nih.gov/22607242
64. Philippou, P., et al. Prospective comparative study of endoscopic management of bladder lithiasis:
Is prostate surgery a necessary adjunct? Urology, 2011. 78: 43.
https://pubmed.ncbi.nlm.nih.gov/21296391
65. Guo, R.Q., et al. Correlation of benign prostatic obstruction-related complications with clinical
outcomes in patients after transurethral resection of the prostate. Kaohsiung J Med Sci, 2017. 33: 144.
https://pubmed.ncbi.nlm.nih.gov/28254117
66. Tangpaitoon, T., et al. Does Cystolitholapaxy at the Time of Holmium Laser Enucleation of the
Prostate Affect Outcomes? Urology, 2017. 99: 192.
https://pubmed.ncbi.nlm.nih.gov/27637344
67. Romero-Otero, J., et al. Analysis of Holmium Laser Enucleation of the Prostate in a High-Volume
Center: The Impact of Concomitant Holmium Laser Cystolitholapaxy. J Endourol, 2019. 33: 564.
https://pubmed.ncbi.nlm.nih.gov/30773913
68. Ord, J., et al. Bladder management and risk of bladder stone formation in spinal cord injured
patients. J Urol, 2003. 170: 1734.
https://pubmed.ncbi.nlm.nih.gov/14532765
69. Bartel, P., et al. Bladder stones in patients with spinal cord injury: a long-term study. Spinal Cord,
2014. 52: 295.
https://pubmed.ncbi.nlm.nih.gov/24469146
70. Chen, H., et al. Can bladder irrigation reduce the morbidity of bladder stones in patients with spinal
cord injury? Open J Urol, 2015. 5: 42.
https://www.ncbi.nlm.nih.gov/pmc/articles/PMC4842518/
71. Awad, S.A., et al. Long-term results and complications of augmentation ileocystoplasty for
idiopathic urge incontinence in women. British J Urol, 1998. 81: 569.
https://pubmed.ncbi.nlm.nih.gov/9598629
72. Blyth, B., et al. Lithogenic properties of enterocystoplasty. J Urol, 1992. 148: 575.
https://pubmed.ncbi.nlm.nih.gov/1640525
73. Flood, H.D., et al. Long-term results and complications using augmentation cystoplasty in
reconstructive urology. Neurourol Urodyn, 1995. 14: 297.
https://pubmed.ncbi.nlm.nih.gov/7581466
74. Hayashi, Y., et al. Review of 86 Patients With Myelodysplasia and Neurogenic Bladder Who
Underwent Sigmoidocolocystoplasty and Were Followed More Than 10 Years. J Urol, 2006. 176: 1806.
https://pubmed.ncbi.nlm.nih.gov/16945655
75. Husmann, D.A. Long-term complications following bladder augmentations in patients with spina
bifida: Bladder calculi, perforation of the augmented bladder and upper tract deterioration. Transl
Androl Urol, 2016. 5: 3.
https://pubmed.ncbi.nlm.nih.gov/26904407
76. Nurse, D.E., et al. Stones in enterocystoplasties. British J Urol, 1996. 77: 684.
https://pubmed.ncbi.nlm.nih.gov/8689111
77. Shekarriz, B., et al. Surgical complications of bladder augmentation: Comparison between various
enterocystoplasties in 133 patients. Urology, 2000. 55: 123.
https://pubmed.ncbi.nlm.nih.gov/10654908
78. Welk, B., et al. Population based assessment of enterocystoplasty complications in adults. J Urol,
2012. 188: 464.
https://pubmed.ncbi.nlm.nih.gov/22704106
79. Zhang, H., et al. Bladder stone formation after sigmoidocolocystoplasty: Statistical analysis of risk
factors. J Pediatr Surg, 2005. 40: 407.
https://pubmed.ncbi.nlm.nih.gov/15750938

16 BLADDER STONES - LIMITED UPDATE MARCH 2021


80. DeFoor, W., et al. Bladder calculi after augmentation cystoplasty: Risk factors and prevention
strategies. J Urol, 2004. 172: 1964.
https://pubmed.ncbi.nlm.nih.gov/15540766
81. Hanna, M.K., et al. Challenges in salvaging urinary continence following failed bladder exstrophy
repair in a developing country. J Pediatr Urol, 2017. 13: 270.
https://pubmed.ncbi.nlm.nih.gov/28262536
82. Inouye, B.M., et al. Urologic complications of major genitourinary reconstruction in the exstrophy-
epispadias complex. J Pediatr Urol, 2014. 10: 680.
https://pubmed.ncbi.nlm.nih.gov/25082713
83. Kaefer, M., et al. Reservoir calculi: a comparison of reservoirs constructed from stomach and other
enteric segments. J Urol, 1998. 160: 2187.
https://pubmed.ncbi.nlm.nih.gov/9817364
84. Kronner, K.M., et al. Bladder calculi in the pediatric augmented bladder. J Urol, 1998. 160: 1096.
https://pubmed.ncbi.nlm.nih.gov/9719284
85. Lima, S.V.C., et al. Nonsecretory Intestinocystoplasty: A 15-Year Prospective Study of 183 Patients.
J Urol, 2008. 179: 1113.
https://pubmed.ncbi.nlm.nih.gov/18206934
86. Metcalfe, P.D., et al. What is the Need for Additional Bladder Surgery After Bladder Augmentation in
Childhood? J Urol, 2006. 176: 1801.
https://pubmed.ncbi.nlm.nih.gov/16945653
87. Novak, T.E., et al. Complications of Complex Lower Urinary Tract Reconstruction in Patients With
Neurogenic Versus Nonneurogenic Bladder-Is There a Difference? J Urol, 2008. 180: 2629.
https://pubmed.ncbi.nlm.nih.gov/18951557
88. Palmer, L.S., et al. Urolithiasis in children following augmentation cystoplasty. J Urol, 1993. 150: 726.
https://pubmed.ncbi.nlm.nih.gov/8326634
89. Silver, R.I., et al. Urolithiasis in the exstrophy-epispadias complex. J Urol, 1997. 158: 1322.
https://pubmed.ncbi.nlm.nih.gov/9258206
90. Surer, I., et al. Continent urinary diversion and the exstrophy-epispadias complex. J Urol, 2003.
169: 1102.
https://pubmed.ncbi.nlm.nih.gov/12576862
91. Wagstaff, K.E., et al. Blood and urine analysis in patients with intestinal bladders. British J Urol,
1991. 68: 311.
https://pubmed.ncbi.nlm.nih.gov/1913074
92. Wang, K., et al. Complications after sigmoidocolocystoplasty: Review of 100 cases at one
institution. J Pediatr Surg, 1999. 34: 1672.
https://pubmed.ncbi.nlm.nih.gov/10591568
93. Szymanski, K.M., et al. Additional Surgeries after Bladder Augmentation in Patients with Spina
Bifida in the 21st Century. J Urol, 2020. 203: 1207.
https://pubmed.ncbi.nlm.nih.gov/31951496
94. Ross, J.P.J., et al. Pediatric bladder augmentation - Panacea or Pandora’s box? Can Urol Assoc J,
2020. 14: E251.
https://pubmed.ncbi.nlm.nih.gov/31977304
95. Breda, A., et al. Percutaneous Cystolithotomy for Calculi in Reconstructed Bladders: Initial UCLA
Experience. J Urol, 2010. 183: 1989.
https://pubmed.ncbi.nlm.nih.gov/20303534
96. Kisku, S., et al. Bladder calculi in the augmented bladder: A follow-up study of 160 children and
adolescents. J Pediatr Urol, 2015. 11: 66.
https://pubmed.ncbi.nlm.nih.gov/25819600
97. Szymanski, K.M., et al. Cutting for stone in augmented bladders - What is the risk of recurrence and
is it impacted by treatment modality? J Urol, 2014. 191: 1375.
https://pubmed.ncbi.nlm.nih.gov/24316089
98. Schlomer, B.J., et al. Cumulative incidence of outcomes and urologic procedures after
augmentation cystoplasty. J Pediatr Urol, 2014. 10: 1043.
https://pubmed.ncbi.nlm.nih.gov/24766857
99. Hensle, T.W., et al. Preventing reservoir calculi after augmentation cystoplasty and continent urinary
diversion: the influence of an irrigation protocol. BJU Int, 2004. 93: 585.
https://pubmed.ncbi.nlm.nih.gov/15008735
100. Knap, M.M., et al. Early and late treatment-related morbidity following radical cystectomy. Scan
J Urol Nephrol, 2004. 38: 153.
https://pubmed.ncbi.nlm.nih.gov/15204405

BLADDER STONES - LIMITED UPDATE MARCH 2021 17


101. Turk, T.M., et al. Incidence of urolithiasis in cystectomy patients after intestinal conduit or continent
urinary diversion. World J Urol, 1999. 17: 305.
https://pubmed.ncbi.nlm.nih.gov/10552149
102. Arai, Y., et al. Orthotopic ileal neobladder in male patients: Functional outcomes of 66 cases. Int
J Urol, 1999. 6: 388.
https://pubmed.ncbi.nlm.nih.gov/10466450
103. Badawy, A.A., et al. Orthotopic diversion after cystectomy in women: A single-centre experience
with a 10-year follow-up. Arab J Urol, 2011. 9: 267.
https://pubmed.ncbi.nlm.nih.gov/26579310
104. Ji, H., et al. Identification and management of emptying failure in male patients with orthotopic
neobladders after radical cystectomy for bladder cancer. Urology, 2010. 76: 644.
https://pubmed.ncbi.nlm.nih.gov/20573379
105. Madbouly, K. Large orthotopic reservoir stone burden: Role of open surgery. Urol Ann, 2010. 2: 96.
https://pubmed.ncbi.nlm.nih.gov/20981195
106. Miyake, H., et al. Experience with various types of orthotopic neobladder in Japanese men: Long-
term follow-up. Urol Int, 2010. 84: 34.
https://pubmed.ncbi.nlm.nih.gov/20173366
107. Moeen, A.M., et al. Management of neobladder complications: endoscopy comes first. Scan J Urol,
2017. 51: 146.
https://pubmed.ncbi.nlm.nih.gov/28635567
108. Simon, J., et al. Neobladder emptying failure in males: incidence, etiology and therapeutic options.
J Urol, 2006. 176: 1468.
https://pubmed.ncbi.nlm.nih.gov/16952662
109. Stein, J.P., et al. The orthotopic T pouch ileal neobladder: Experience with 209 patients. J Urol,
2004. 172: 584.
https://pubmed.ncbi.nlm.nih.gov/15247737
110. Stein, J.P., et al. Complications of the afferent antireflux valve mechanism in the Kock ileal reservoir.
J Urol, 1996. 155: 1579.
https://pubmed.ncbi.nlm.nih.gov/8627827
111. Miyake, H., et al. Orthotopic sigmoid neobladder after radical cystectomy: Assessment of
complications, functional outcomes and quality of life in 82 Japanese patients. BJU Int, 2010. 106:
412.
https://pubmed.ncbi.nlm.nih.gov/19888974
112. Holmes, D.G., et al. Long-term complications related to the modified Indiana pouch. Urology, 2002.
60: 603.
https://pubmed.ncbi.nlm.nih.gov/12385916
113. Khalil, F., et al. Long-term follow-up after ileocaecal continent cutaneous urinary diversion (Mainz
i pouch): A retrospective study of a monocentric experience. Arab J Urol, 2015. 13: 245.
https://pubmed.ncbi.nlm.nih.gov/26609442
114. Marien, T., et al. Characterization of Urolithiasis in Patients Following Lower Urinary Tract
Reconstruction with Intestinal Segments. J Endourol, 2017. 31: 217.
https://pubmed.ncbi.nlm.nih.gov/27936931
115. Davis, W.B., et al. Percutaneous imaging-guided access for the treatment of calculi in continent
urinary reservoirs. CardioVasc Intervent Radiol, 2002. 25: 119.
https://pubmed.ncbi.nlm.nih.gov/11901429
116. Paez, E., et al. Percutaneous treatment of calculi in reconstructed bladder. J Endourol, 2007. 21: 334.
https://pubmed.ncbi.nlm.nih.gov/17444782
117. Chung, S.-D., et al. A case-control study on the association between bladder cancer and prior
bladder calculus. BMC Cancer, 2013. 13: 117.
https://pubmed.ncbi.nlm.nih.gov/23497224
118. Jhamb, M., et al. Urinary tract diseases and bladder cancer risk: a case-control study. Cancer
Causes Control, 2007. 18: 839.
https://pubmed.ncbi.nlm.nih.gov/17593531
119. La Vecchia, C., et al. Genital and urinary tract diseases and bladder cancer. Cancer Res, 1991. 51: 629.
https://pubmed.ncbi.nlm.nih.gov/1985779

18 BLADDER STONES - LIMITED UPDATE MARCH 2021


6. CONFLICT OF INTEREST
All members of the Bladder Stones Guidelines Panel have provided disclosure statements of all relationships
that they have that might be perceived as a potential source of a conflict of interest. This information is publicly
accessible through the European Association of Urology website: http://www.uroweb.org/guidelines/. This
guidelines document was developed with the financial support of the European Association of Urology. No
external sources of funding and support have been involved. The EAU is a non-profit organisation and funding
is limited to administrative assistance and travel and meeting expenses. No honoraria or other reimbursements
have been provided.

7. CITATION INFORMATION
The format in which to cite the EAU Guidelines will vary depending on the style guide of the journal in which the
citation appears. Accordingly, the number of authors or whether, for instance, to include the publisher, location,
or an ISBN number may vary.

The compilation of the complete Guidelines should be referenced as:


EAU Guidelines. Edn. presented at the EAU Annual Congress Milan 2021. ISBN 978-94-92671-13-4.

If a publisher and/or location is required, include:


EAU Guidelines Office, Arnhem, the Netherlands. http://uroweb.org/guidelines/compilations-of-all-guidelines/

References to individual guidelines should be structured in the following way:


Contributors’ names. Title of resource. Publication type. ISBN. Publisher and publisher location, year.

BLADDER STONES - LIMITED UPDATE MARCH 2021 19


20 BLADDER STONES - LIMITED UPDATE MARCH 2021
EAU Guidelines on
Chronic
Pelvic Pain
D. Engeler (Chair), A.P. Baranowski, B. Berghmans,
J. Borovicka, A.M. Cottrell, P. Dinis-Oliveira, S. Elneil,
J. Hughes, E.J. Messelink (Vice-chair), A.C. de C Williams
Guidelines Associates: B. Parsons, S. Goonewardene,
P. Abreu-Mendes, V. Zumstein

© European Association of Urology 2021


TABLE OF CONTENTS PAGE
1. INTRODUCTION 5
1.1 Aim 5
1.2 Publication history 5
1.3 Available Publications 5
1.4 Panel composition 5
1.5 Terminology 6

2. METHODOLOGY 14
2.1 Methods 14
2.2 Review 15

3. EPIDEMIOLOGY, AETIOLOGY AND PATHOPHYSIOLOGY 15


3.1 Chronic visceral pain 15
3.1.1 Incidence 16
3.1.2 Prevalence 16
3.1.3 Influence on Quality of Life 16
3.1.4 Costs 16
3.1.5 Risk Factors and underlying causes 16
3.1.5.1 Risk factors 16
3.1.5.2 Underlying causes 17
3.1.5.3 Clinical paradigms in visceral pain 20
3.2 Pelvic Pain 21
3.2.1 Incidence 21
3.2.2 Prevalence 21
3.2.2.1 Primary prostate pain syndrome 21
3.2.2.2 Primary bladder pain syndrome 21
3.2.2.3 Sexual pain syndrome 21
3.2.2.4 Myofascial pain syndromes 22
3.2.3 Influence on QoL 22
3.2.4 Costs 22
3.2.5 Risk factors and underlying causes 22
3.2.5.1 Primary Prostate Pain Syndrome 22
3.2.5.2 Primary Bladder Pain Syndrome 22
3.2.5.3 Primary Scrotal Pain Syndrome 22
3.2.5.4 Primary Urethral Pain Syndrome 23
3.2.5.5 Primary Vaginal and Vulvar Pain Syndromes 23
3.2.5.6 Chronic Pelvic Pain and Prolapse/Incontinence Mesh 23
3.2.5.7 Chronic post-surgical pain 24
3.2.5.8 Associated conditions in pelvic pain syndromes 24
3.3 Abdominal aspects of pelvic pain 27
3.3.1 Incidence 27
3.3.2 Prevalence 27
3.3.3 Influence on QOL 27
3.3.4 Costs 27
3.3.5 Risk factors & underlying causes 27
3.4 Summary of evidence and recommendations: CPPPS and mechanisms 27

4. DIAGNOSTIC EVALUATION 28
4.1 General Evaluation 28
4.1.1 History 28
4.1.1.1 Anxiety, depression, and overall function 28
4.1.1.2 Urological aspects 28
4.1.1.3 Gynaecological aspects 29
4.1.1.4 Gastrointestinal aspects 29
4.1.1.5 Peripheral nerve aspects 29
4.1.1.6 Myofascial aspects 30
4.1.2 Physical Evaluation 30
4.2 Supplemental evaluation 31

2 CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021


4.2.1 Assessing pelvic pain and related symptoms 31
4.2.2 Focused myofascial evaluation 32
4.2.3 Neurological 32
4.2.4 Imaging 33
4.2.5 Laboratory Tests 33
4.2.6 Invasive tests 34
4.3 Diagnostic algorithm 35
4.4 Other painful conditions without a urological cause 36
4.5 Summary of evidence and recommendations: diagnostic evaluation 37
4.5.1 Diagnostic evaluation - general 37
4.5.2 Diagnostic evaluation of PPS 37
4.5.3 Diagnostic evaluation of primary bladder pain syndrome 38
4.5.4 Diagnostic evaluation of scrotal pain syndrome 38
4.5.5 Diagnostic evaluation of urethral pain syndrome 38
4.5.6 Diagnostic evaluation of gynaecological aspects chronic pelvic pain 38
4.5.7 Diagnostic evaluation of anorectal pain syndrome 39
4.5.8 Diagnostic evaluation of nerves to the pelvis 39
4.5.9 Diagnostic evaluation of sexological aspects in chronic pelvic pain 39
4.5.10 Diagnostic evaluation of psychological aspects of chronic pelvic pain 40
4.5.11 Diagnostic evaluation of pelvic floor function 40

5. MANAGEMENT 40
5.1 Conservative management 40
5.1.1 Pain education 40
5.1.2 Physical therapy 41
5.1.3 Psychological therapy 42
5.1.4 Dietary treatment 43
5.2 Pharmacological management 43
5.2.1 Drugs for chronic primary pelvic pain syndrome 43
5.2.1.1 Mechanisms of action 43
5.2.1.2 Comparisons of agents used in pelvic pain syndromes 43
5.2.2 Analgesics 47
5.2.2.1 Mechanisms of action 48
5.2.2.2 Comparisons within and between groups in terms of efficacy and
safety 48
5.3 Further management 50
5.3.1 Nerve blocks 50
5.3.2 Neuromodulation 50
5.3.3 Surgery 51
5.4 Summary of evidence and recommendations: management 54
5.4.1 Management of primary prostate pain syndrome 54
5.4.2 Management of primary bladder pain syndrome 54
5.4.3 Management of scrotal pain syndrome 55
5.4.4 Management of primary urethral pain syndrome 55
5.4.5 Management of gynaecological aspects of chronic pelvic pain 56
5.4.6 Management of primary anorectal pain syndrome 56
5.4.7 Management of pudendal neuralgia 56
5.4.8 Management of sexological aspects in chronic pelvic pain 56
5.4.9 Management of psychological aspects in chronic pelvic pain 57
5.4.10 Management of pelvic floor dysfunction 57
5.4.11 Management of chronic/non-acute urogenital pain by opioids 57

6. EVALUATION OF TREATMENT RESULTS 57


6.1 Evaluation of treatment 57
6.1.1 Treatment has not been effective 57
6.1.1.1 Alternative treatment 57
6.1.1.2 Referral to next envelope of care 57
6.1.1.3 Self-management and shared care 58
6.1.2 Treatment has been effective 58

CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021 3


7. REFERENCES 58

8. CONFLICT OF INTEREST 85

9. CITATION INFORMATION 85

4 CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021


1. INTRODUCTION
1.1 Aim
This guideline plays an important role in the process of consolidation and improvement of care for patients with
pelvic pain and associated lower abdominal pain. From both literature and daily practice it has become clear
that lower abdominal and pelvic pain are areas still under development. This guideline has been recognised as
a cornerstone for important developments that have taken place in the past ten years.

This guideline aims to expand the awareness of caregivers in the field of abdominal and pelvic pain and to
assist those who treat patients with abdominal and pelvic pain in their daily practice. The guideline is a useful
instrument not only for urologists, but also for gynaecologists, surgeons, physiotherapists, psychologists and
pain doctors.

It must be emphasised that guidelines present the best evidence available to the experts. However following
guideline recommendations will not necessarily result in the best outcome. Guidelines can never replace
clinical expertise when making treatment decisions for individual patients, but rather help to focus decisions -
also taking personal values and preferences/individual circumstances of patients into account. Guidelines are
not mandates and do not purport to be a legal standard of care.

Structure and scope


The panel wishes to take advantage of modern methods of delivering guideline information to clinicians
dealing with these patients. In 2016, a stepped information structure was made, in alignment with stepped
care protocols, using new digital information sources like websites and apps to aid this process. Furthermore,
the guideline was changed according to the template used in all other non-oncology guidelines of the EAU. It
was recognised that structuring a guideline on chronic pain is quite different from structuring one on another
subject. A multi-disciplinary approach is of utmost importance and demands a broad view. In 2016, the
guideline was rewritten to be centred around pain instead of being organ-centred. It is partly theoretical to
show the importance of using this pain-centred approach. The biggest part, however, deals with the practical
approach to diagnostics, treatment and management of patients with abdominal and pelvic pain.

1.2 Publication history


The EAU Guidelines on Chronic Pelvic Pain were first published in 2003 [1] which formed the basis of a
scientific publication in European Urology in 2004 [2]. Also, in the 2003 edition the concept of Chronic Pelvic
Pain Syndromes (CPPS) was introduced, which is now referred to as “pain as a disease process”. Partial
updates of the Chronic Pelvic Pain Guidelines were published in 2008 and formed the basis for another
scientific publication in European Urology in the year 2010 [3, 4]. Two chapters were added at that time:
Chapter 5 ‘Gastrointestinal aspects of chronic pelvic pain’ and Chapter 7 ‘Sexological aspects of chronic pelvic
pain’. In the 2014 edition minor revisions were made in Chapter 5 ‘Gastrointestinal aspects of chronic pelvic
pain’ and Chapter 8 ‘Psychological aspects of chronic pelvic pain’.

For the 2015 edition the panel critically reviewed the sub-chapter on chronic primary bladder pain syndrome
which is now a comprehensive part of the guideline. The fact that this part was so extensive shows that the
roots of talking about abdominal pain and pelvic pain lies in the bladder, where Interstitial Cystitis was one of
the first subjects addressed talking about pain in urology. The panel has illustrated this in the publication in
European Urology in 2013 [5].

1.3 Available Publications


Alongside the full text version, a quick reference document (Pocket Guidelines) is available in print and as an
app for iOS and Android devices. These are abridged versions which may require consultation together with
the full text version. This reference document follows the updating cycle of the underlying large texts.

All available material can be viewed at the EAU website. The EAU website also includes a selection of EAU
Guideline articles as well as translations produced by national urological associations: uroweb.org/guideline/
chronicpelvicpain/.

1.4 Panel composition


The panel of experts responsible for this document include four urologists, (one of which has a sub-
specialisation in neuro-urology and one is a sexologist), two consultants in pain medicine, a uro-gynaecologist,
a psychologist, a gastroenterologist and a pelvic physiotherapist, health scientist and (clinical) epidemiologist.

CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021 5


The panel is also grateful to Ms. J. Birch for her expertise, time and diligence in undertaking a review of these
guidelines from a patient perspective.

1.5 Terminology

Definitions of chronic pelvic pain terminology


Classification
Much debate over the classification of chronic pelvic pain has occurred, is ongoing and will continue in the
future. Classification involves three aspects of defining a condition: phenotyping, terminology and taxonomy.

Phenotyping
Phenotyping is describing the condition. For example, chronic bladder pain may be associated with the
presence of Hunner’s lesions and glomerulation on cystoscopy, whereas other bladder pain conditions may
have a normal appearance on cystoscopy. These are two different phenotypes. The same is true for irritable
bowel syndrome (IBS), which may be sub-divided into that associated primarily with diarrhoea or that with
constipation. Phenotyping is based upon mechanisms when they are known (e.g., infection, ischaemic,
auto-immune, or neuropathic). In the absence of well-defined mechanisms, describing the condition by its
symptoms, signs and, where possible, by investigations, has been demonstrated to have clinical and research
validity in many situations. When pain is the main symptom and pain as a disease process is considered the
cause, the condition is often referred to as a pain syndrome - a well-defined collection of symptoms, signs
and investigation results associated with pain mechanisms and pain perception as the primary complaint. The
World Health Organization (WHO) International Classification of Diseases 11th Revision (ICD-11) uses the term
Chronic Primary Pain to distinguish these conditions from pain associated with another diagnosis that they
refer to as Chronic Secondary Pain (see below).

Terminology
Terminology is the word(s) that are used within classification, both to name the phenotype and within the
definition of the phenotype. Examples of names for phenotypes associated with the bladder include interstitial
cystitis, painful bladder syndrome or bladder pain syndrome (BPS). The EAU, the International Society for the
study of BPS (known as ESSIC), the International Association for the Study of Pain (IASP) and several other
groups have preferred the term bladder pain syndrome. In the pain syndromes, the role of the nervous system
in generating the sensations is thought to be pivotal, but the term syndrome is also comprehensive and takes
into account the emotional, cognitive, behavioural, sexual and functional consequences of the chronic pain.

When defining the phenotype, the terminology used in that definition must also be clear and if necessary
defined. One of the most important guiding principles is that spurious terminology should be avoided. Terms
that end in “itis” in particular should be avoided unless infection and or inflammation is proven and considered
to be the cause of the pain [6]. It must be appreciated that end-organ inflammation may be secondary and
neurogenic in origin and not a primary cause of the pain.

Taxonomy
Taxonomy places the phenotypes into a relationship hierarchy. The EAU approach sub-divides chronic pelvic
pain into conditions that are pain syndromes with no obvious diagnosis, chronic primary pelvic pain syndromes
(CPPPS) (consistent with ICD-11 Chronic Primary Pain) and those that are non-pain syndromes. The latter
are conditions that have well-recognised pathology (e.g., infection, neuropathy or inflammation), whereas
the former syndromes do not, and pain as a disease process is the mechanism. Other terms for the non-
pain syndromes include “classical conditions”, “well-defined conditions” and “confusable diseases” and the
ICD-11 Chronic Secondary Pain. Although the EAU approach deals primarily with urological conditions, this
approach to classification can be applied to all conditions associated with pain perception within the pelvis; the
classification has been developed to include non-urological pain and was accepted by the IASP for publication
in January 2012.

Classification of Chronic Pelvic Pain


Importance of classification
It should be obvious to all that a condition cannot be treated unless it is defined. However, the reasons for
classifying chronic pelvic pain go far beyond that.

Clues to the mechanism


As a result of systematic phenotypic and taxonomic classifications, similarities and differences between
conditions become clear. Drawing comparisons between the phenotypes of different disorders allows comparison
between disorders such as bladder and bowel pain syndromes, thus facilitating research and treatment.

6 CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021


Guidelines for best treatment options
As conditions become better defined, more specific treatment approaches can be adopted. In particular, there
will be a move away from treatments based upon spurious terms (e.g., antibiotics and non-steroidal anti-
inflammatory drugs for the “-itis” conditions). Generic treatments aimed at groups of conditions will be more
commonplace and based upon research evidence.

Research platform
Only by clearly defining the phenotype being investigated can research be valued or applied to the clinical
situation.

Patient needs
A diagnosis, or name, for a set of symptoms can provide patients with a sense of being understood, as well
as hope for relief. It may therefore help in acceptance of the problem as chronic, resolution of unfounded
fears about its implications (if not life-threatening), and engagement in therapeutic endeavours, as well as in
self-management. However, it may also lead to accessing information of variable quality associated with the
diagnosis or name, and the possibility of generating new concerns about long-term consequences or about
appropriateness of treatment.

IASP definitions
Sub-dividing pain syndromes
There is much debate on the sub-divisions of the pain syndromes within the hierarchical taxonomy. The EAU
has led the way in this regard and the guiding principles are as follows [2]:

1. The pain syndromes are defined by a process of exclusion. In particular, there should be no evidence of
infection or inflammation. Investigations by end-organ specialists should therefore be aimed at obtaining
a differential diagnosis; repeated, unnecessary investigations are detrimental in the management of
chronic pain syndromes.
2. A sub-division phenotype should only be used if there is adequate evidence to support its use. For
instance, in non-specific, poorly localised pelvic pain without obvious pathology, only the term chronic
primary pelvic pain syndrome (CPPPS) should be used. If the pain can be localised to an organ, then a
more specific term, such as rectal pain syndrome, may be used, also potentially with the term primary
added. If the pain is localised to multiple organs, then the syndrome is a regional pain syndrome and the
term CPPPS should once again be considered. As well as defining the patient by a specific end-organ
phenotype, there are several other more general descriptors that need to be considered. These are
primarily psychological (e.g., cognitive or emotional), sexual, behavioural and functional. Psychological
and behavioural factors are well-established factors which relate to quality of life (QoL) issues and
prognosis. In North America a research programme, the MAPP program (Multi-disciplinary Approach
to the study of Chronic Pelvic Pain research) has been devised to investigate the importance of these
factors and looks at all types of pelvic pain irrespective of the end-organ where the pain is perceived. It
also looks at systemic disorder associations, such as the co-occurrence of fibromyalgia, facial pain, or
auto-immune disorders.
3. In 2004 the panel introduced the concept of managing the polysymptomatic nature of CPPPS, since then
others have developed their own schemes, such as Nickel’s UPOINT [7], modified by Magri et al. [8]. In
light of these and other publications, the symptom classification table has been updated (Table 1).

The debate in relation to sub-dividing the pain syndromes remains ongoing. As more information is collected
suggesting that the central nervous system (CNS) is involved, and indeed may be the main cause of many
CPPPS conditions (e.g., bladder, genitalia, colorectal or myofascial), there is a general tendency to move away
from end-organ nomenclature. Only time and good research will determine whether this is appropriate. To
enable such research, it is essential to have a framework of classification within which to work. Any hierarchical
taxonomy must be flexible to allow change.

ICD classification: purpose and uses

The International Classification of Diseases is the foundation for the identification of health trends and statistics
globally, and the international standard for reporting diseases and health conditions. It is the diagnostic
classification standard for all clinical and research purposes. It defines the universe of diseases, disorders,
injuries and other related health conditions, listed in a comprehensive, hierarchical fashion [9]. The latest
version ICD-11 is available for member states to report with from January 2022.

CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021 7


The ICD-11 classification for the first time included chronic pain (“chronic pain is pain that persists or
recurs for longer than 3 months”) and divided the coding into Chronic Primary Pain (“chronic primary pain is
multifactorial: biological, psychological and social factors contribute to the pain syndrome”) and a number of
Chronic Secondary Pain conditions (related to cancer, post surgical, musculoskeletal, visceral, neuropathic,
headache/orofacial, other).

The significance of the inclusion of Chronic Pain as a condition within the ICD-11 should not be
underestimated. There are, however, unresolved issues regarding this classification, such as when a condition
ends and pain persists, does that term Chronic Secondary Pain become Chronic Primary Pain? [10, 11].
Similarly, the contents of recent draft NICE guidelines [12] (https://www.nice.org.uk/guidance/GID-NG10069/
documents/draft-guideline), were found to be contentious as the guidelines considered all Chronic Primary
Pain as being essentially the same and the ‘biological’ nature of the pain appeared to have been missed.
Whereas in the final guidelines this may be corrected, it does illustrate the risk behind the term Chronic Primary
Pain.

The panel will change the EAU terminology previously used in the Guidelines to show conformity with ICD 11
definitions. This will include changing terminology used in originally cited works.

8 CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021


Axis I Axis II Axis III Axis IV Axis V Axis VI Axis VII Axis VIII
Region System End-organ as pain syndrome Referral Temporal Character Associated Psychological
as identified from Hx, Ex characteristics characteristics symptoms symptoms
and Ix
Chronic Chronic Urological Prostate Suprapubic ONSET Aching UROLOGICAL ANXIETY
pelvic secondary Inguinal Acute Burning Frequency About pain
Bladder
pain pelvic pain Urethral Chronic Stabbing Nocturia or putative
syndrome, Scrotal Penile/clitoral Electric Hesitancy cause of pain
formally known Testicular Perineal ONGOING Dysfunctional flow
as specific Epididymal Rectal Sporadic Urgency Catastrophic thinking about
disease Back Cyclical Incontinence Pain
associated Penile Buttocks Continuous
pelvic pain Urethral Thighs GYNAECOLOGICAL DEPRESSION
Post-vasectomy TIME Menstrual Attributed to
OR Filling Menopause pain or impact
Gynaecological Vulvar Emptying of pain
Chronic primary Vestibular Immediate post GASTROINTESTINAL
pelvic pain Clitoral Late post Constipation Attributed to
syndrome, Diarrhoea other causes

CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021


Endometriosis associated
formally known TRIGGER Bloatedness
as pelvic pain CPPPS with cyclical exacerbations Provoked Urgency Unattributed
syndrome Spontaneous Incontinence
Dysmenorrhoea
PTSD
Gastrointestinal Irritable bowel NEUROLOGICAL SYMPTOMS
Dysaesthesia Re-experiencing
Chronic anal
Hyperaesthesia
Table 1: EAU classification of chronic pelvic pain syndromes

Avoidance
Intermittent chronic anal Allodynia
Hyperalgesia
Peripheral nerves Pudendal pain syndrome

Sexological Dyspareunia SEXUOLOGICAL


Satisfaction
The classification has been set up according to the axis system used by IASP.

Pelvic pain with sexual dysfunction Female dyspareunia


Sexual avoidance
Psychological Any pelvic organ
Erectile dysfunction
Medication
Musculo-skeletal Pelvic floor muscle
Abdominal muscle

Hx = History; Ex = Examination; Ix = Investigation; PTSD = post-traumatic stress disorder.


MUSCLE
Spinal
Function impairment
Coccyx Fasciculation

CUTANEOUS
Trophic changes
Sensory changes

9
Pain syndromes
The original EAU classification [2] was inspired by the IASP classification [13] and much work around what has
become known as “pain as a disease” and its associated psychological, behavioural, sexual, social and organ
function aspects. After ten years of work developing the initial ideas, an updated version was accepted by the
IASP Council for publication in January 2012.

EAU Definition of chronic pelvic pain


Chronic pelvic pain is chronic or persistent pain perceived* in structures related to the pelvis of either men or
women. It is often associated with negative cognitive, behavioural, sexual and emotional consequences as well
as with symptoms suggestive of lower urinary tract, sexual, bowel, pelvic floor or gynaecological dysfunction.
[*Perceived indicates that the patient and clinician, to the best of their ability from the history, examination and
investigations (where appropriate) have localised the pain as being discerned in a specified anatomical pelvic
area.]

In the case of documented nociceptive pain that becomes chronic/persistent through time, pain must have
been continuous or recurrent for at least three months (in accordance with ICD-11). For cyclical pain, a longer
period of more than six months may be appropriate. Cyclical pain is included in the classification, particularly
if there is evidence of central sensitisation and hence dysmenorrhoea (hormonally dependent) needs to be
considered as a chronic pain syndrome, if it is persistent and associated with negative cognitive, behavioural,
sexual, or emotional consequences.

Chronic pelvic pain may be sub-divided into conditions with well-defined classical pathology (such as infection
or cancer) and those with no obvious pathology but still including biological mechanisms. For the purpose of
the EAU’s classification, the term “specific disease-associated pelvic pain” has been accepted for the former,
and “chronic pelvic pain syndrome” for the latter. In the new ICD-11 these conditions have new names: the
former will be called Chronic Secondary Pelvic Pain and the latter Chronic Primary Pelvic Pain.

The following classification only deals with Chronic Primary Pelvic Pain Syndromes.

EAU Definition of chronic primary pelvic pain syndrome


Chronic primary pelvic pain syndrome (CPPPS) is the occurrence of chronic pain when there is no proven
infection or other obvious local pathology that may account for the pain. It is often associated with negative
cognitive, behavioural, sexual or emotional consequences, as well as with symptoms suggestive of lower
urinary tract, sexual, bowel or gynaecological dysfunction. Chronic Primary Pelvic Pain Syndrome is a sub-
division of chronic pelvic pain. Throughout the text below in the 2021 update, CPPS is replaced with CPPPS if
it is appropriate.

Further subdivision of CPPPS


Pain perception in CPPPS may be focused within a single organ, more than one pelvic organ and even
associated with systemic symptoms such as chronic fatigue syndrome (CFS), fibromyalgia (FM) or Sjögren’s
syndrome. When the pain is localised to a single organ, some specialists may wish to consider using an
end organ term such as Bladder Pain Syndrome (Table 2), also using the term primary. The use of such a
phrase with the terminology “syndrome” indicates that, although peripheral mechanisms may exist, CNS
neuromodulation may be more important and systemic associations may occur. When the pain is localised
to more than one organ site, the generic term CPPPS should be used. Many, including some of the panel
members never sub-divide by anatomy and prefer to refer to patients with pain perceived within the pelvis, and
no specific disease process, as suffering from CPPPS, sub-divided by psychological and functional symptoms.

Psychological considerations for classification


Many CPPPSs are associated with a range of concurrent negative psychological, behavioural and sexual
consequences that must be described and assessed. Examples that need to be considered are depression,
anxiety, fears about pain or its implications, unhelpful coping strategies, and distress in relationships. Both
anxiety and depression can be significant important concomitant symptoms that are relevant to pain, disability
and poor QoL. Catastrophic interpretation of pain has been shown to be a particularly salient variable,
predicting patients’ report of pain, disability, and poor QoL, over and above psychosocial variables such
as depression or behavioural factors such as self-reported sexual dysfunction. It is suggested that CPPPS
sometimes creates a sense of helplessness that can be reported as overwhelming, and may be associated
with the refractory nature of the patients’ symptoms. It is important to note that many of these biopsychosocial
consequences are common to other persistent pain problems but may show varying degrees of importance
for any one individual suffering from CPPPS. In all patients with CPPPS, these consequences must be

10 CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021


clearly described as part of the phenotype (where the term phenotype is used to indicate the observable
characteristics of the syndrome).

Functional considerations for classification


Functional disorders, for the purpose of this document, are pathologies that have arisen secondary to changes
in the control mechanisms of an organ or system. That is, they are disorders characterised by disturbance of
function. As an example, slow colonic transit is a functional disorder of the bowel - the normal function of the
bowel is not occurring as a result of changes in the mechanisms that produce defecation, and therefore bowel
control is altered. The term is not used in the sense of a psychiatric functional disorder. Many CPPPSs are
associated with functional abnormalities at a local and even systemic level. These also need to be defined as a
part of the phenotype.

Functional pain disorders may not include significant pathology in the organs that appear responsible for the
primary symptoms, but they are associated with substantial neurobiological, physiological and sometimes
anatomical changes in the CNS.

Multi-system sub-division
It is recognised that the end-organ where the pain is perceived may not be the centre of pain generation.
This classification is based upon the most effective and accepted method of classifying and identifying
different pain syndromes, that is, by site of presentation. It is argued that keeping the end-organ name in the
classification is inappropriate because, in most cases, there are multi-systemic causes and effects, with the
result that symptoms are perceived in several areas. This is an area in which discussions are ongoing, and
despite there being strong arguments for both keeping and dispensing with end-organ classification, the
panel have not taken the umbrella approach of referring to all pain perceived in the pelvis as CPPS, primary or
secondary.

Dyspareunia
Dyspareunia is defined as pain perceived within the pelvis associated with penetrative sex. It tells us nothing
about the mechanism and may be applied to women and men. It is usually applied to penile penetration, but is
often associated with pain during insertion of any object. It may apply to anal as well as vaginal intercourse. It
is classically sub-divided into superficial and deep.

Primary perineal pain syndrome


Perineal pain syndrome is a neuropathic-type pain that is perceived in the distribution area of the pudendal
nerve, and may be associated with symptoms and signs of rectal, urinary tract or sexual dysfunction. There is
no proven obvious pathology. It is often associated with negative cognitive, behavioural, sexual and emotional
consequences, as well as with symptoms suggestive of lower urinary tract, sexual, bowel or gynaecological
dysfunction. Primary perineal pain syndrome should be distinguished from pudendal neuralgia which is a
specific disease associated with perineal pain that is caused by nerve damage.

Table 2: C
 hronic Primary Pelvic Pain Syndromes (the term primary can be included in any of the
following)

Urological Pain Syndromes


Primary prostate Primary prostate pain syndrome (PPPS) is the occurrence of persistent or recurrent
pain syndrome episodic pain (which is convincingly reproduced by prostate palpation). There is no
proven infection or other obvious local pathology. Primary prostate pain syndrome
is often associated with negative cognitive, behavioural, sexual or emotional
consequences, as well as with symptoms suggestive of lower urinary tract and sexual
dysfunction. The term “chronic prostatitis” continues to be equated with that of
PPPS. In the authors’ and others’ opinion, this is an inappropriate term, although it
is recognised that it has a long history of use. The National Institutes of Health (NIH)
consensus [14] includes infection (types I and II), which the authors feel should not
be considered under PPPS, but as specific disease-associated pelvic pain. The term
prostadynia has also been used in the past but is no longer recommended by the
expert panel. Please note that some of the authors of the IASP document disagree with
this term and suggest that CPPPS of the male is used instead of PPPS, which has been
agreed by the majority.

CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021 11


Primary bladder Primary bladder pain syndrome (PBPS) is the occurrence of persistent or recurrent pain
pain syndrome perceived in the urinary bladder region, accompanied by at least one other symptom,
such as pain worsening with bladder filling and day-time and/or night-time urinary
frequency. There is no proven infection or other obvious local pathology. Primary
bladder pain syndrome is often associated with negative cognitive, behavioural, sexual
or emotional consequences, as well as with symptoms suggestive of lower urinary
tract and sexual dysfunction. PBPS is believed to represent a heterogeneous spectrum
of disorders. There may be specific types of inflammation as a feature in subsets of
patients. Localisation of the pain can be difficult by examination, and consequently,
another localising symptom is required. Cystoscopy with hydrodistension and biopsy
may be indicated to define phenotypes. Recently, ESSIC has suggested a standardised
scheme of sub-classifications [15] to acknowledge differences and make it easier to
compare various studies. Other terms that have been used include “interstitial cystitis”,
“painful bladder syndrome”, and “PBS/IC” or “BPS/IC”. These terms are no longer
recommended.
Primary scrotal Primary scrotal pain syndrome is the occurrence of persistent or recurrent episodic
pain syndrome pain localised to the scrotum or the structure within it and may be associated with
symptoms suggestive of lower urinary tract or sexual dysfunction. There is no proven
infection or other obvious local pathology. Primary scrotal pain syndrome is often
associated with negative cognitive, behavioural, sexual or emotional consequences.
Primary scrotal pain syndrome is a generic term and is used when the site of the pain is
not clearly testicular or epididymal. The pain is not in the skin of the scrotum as such,
but perceived within its contents, in a similar way to idiopathic chest pain.
Primary testicular Primary testicular pain syndrome is the occurrence of persistent or recurrent episodic
pain syndrome pain perceived in the testes, and may be associated with symptoms suggestive of
lower urinary tract or sexual dysfunction. There is no proven infection or other obvious
local pathology. Primary testicular pain syndrome is often associated with negative
cognitive, behavioural, sexual or emotional consequences. Previous terms have
included orchitis, orchialgia and orchiodynia. These terms are no longer recommended.
Primary epididymal Primary epididymal pain syndrome is the occurrence of persistent or recurrent episodic
pain syndrome pain perceived in the epididymis, and may be associated with symptoms suggestive of
lower urinary tract or sexual dysfunction. There is no proven infection or other obvious
local pathology. Primary epididymal pain syndrome is often associated with negative
cognitive, behavioural, sexual or emotional consequences.
Primary penile Primary penile pain syndrome is the occurrence of pain within the penis that is not
pain syndrome primarily in the urethra, in the absence of proven infection or other obvious local
pathology. Primary penile pain syndrome is often associated with negative cognitive,
behavioural, sexual or emotional consequences, as well as with symptoms suggestive
of lower urinary tract and sexual dysfunction.
Primary urethral Primary urethral pain syndrome is the occurrence of chronic or recurrent episodic pain
pain syndrome perceived in the urethra, in the absence of proven infection or other obvious local
pathology. Primary urethral pain syndrome is often associated with negative cognitive,
behavioural, sexual or emotional consequences, as well as with symptoms suggestive
of lower urinary tract, sexual, bowel or gynaecological dysfunction. Primary urethral
pain syndrome may occur in men and women.
Post-vasectomy Post-vasectomy scrotal pain syndrome is a scrotal pain syndrome that follows
scrotal pain vasectomy. Post-vasectomy scrotal pain syndrome is often associated with negative
syndrome cognitive, behavioural, sexual or emotional consequences, as well as with symptoms
suggestive of lower urinary tract and sexual dysfunction. Post-vasectomy pain may be
as frequent as 1% following vasectomy, possibly more frequent. The mechanisms are
poorly understood and for that reason it is considered by some a special form of
primary scrotal pain syndrome.

12 CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021


Primary Gynaecological Pain Syndromes: external genitalia
Primary vulvar Primary vulvar pain syndrome is the occurrence of persistent or recurrent episodic
pain syndrome vulvar pain. There is no proven infection or other local obvious pathology. It is often
associated with negative cognitive, behavioural, sexual or emotional consequences,
as well as with symptoms suggestive of lower urinary tract, sexual, bowel or
gynaecological dysfunction. Although pain perceived in the vulva was included under
sexual disorders in the DSM-IV-R manual for classifying psychiatric disorders, there
is no scientific basis for this classification, and pain perceived in the vulva is best
understood as a pain problem that usually has psychological consequences. There is
no evidence for its classification as a psychiatric disorder. The International Society for
the Study of Vulvovaginal Disease (ISSVD) has used the term vulvodynia, where the
panel use the term primary vulvar pain syndrome. According to the ISSVD, vulvodynia
is vulvar pain that is not accounted for by any physical findings. The ISSVD has defined
vulvodynia as “vulvar discomfort, most often described as burning pain, occurring in
the absence of relevant visible findings or a specific, clinically identifiable, neurologic
disorder”. If physical findings are present, the patient is said to have vulvar pain due
to a specified cause. The ISSVD has sub-divided vulvodynia based on pain location
and temporal characteristics of the pain (e.g., provoked or unprovoked). The following
definitions are based on that approach.
Primary Primary generalised vulvar pain syndrome refers to a vulvar pain syndrome in which the
generalised vulvar pain/burning cannot be consistently and precisely localised by point-pressure mapping
pain syndrome via probing with a cotton-tipped applicator or similar instrument. Rather, the pain is
diffuse and affects all parts of the vulva. The vulvar vestibule (the part that lies between
the labia minora into which the urethral meatus and vaginal introitus open) may be
involved but the discomfort is not limited to the vestibule. This pain syndrome is often
associated with negative cognitive, behavioural, sexual or emotional consequences.
Previous terms have included “dysesthetic vulvodynia” and “essential vulvodynia”, but
these are no longer recommended.
Primary localised Primary localised vulvar pain syndrome refers to pain that can be consistently and
vulvar pain precisely localised by point-pressure mapping to one or more portions of the vulva.
syndrome Clinically, the pain usually occurs as a result of provocation (touch, pressure or friction).
Primary localised vulvar pain syndrome can be sub-divided into primary vestibular pain
syndrome and primary clitoral pain syndrome.
Primary vestibular Primary vestibular pain syndrome refers to pain that can be localised by point-pressure
pain syndrome mapping to the vestibule or is well perceived in the area of the vestibule.
Primary clitoral Primary clitoral pain syndrome refers to pain that can be localised by point-pressure
pain syndrome mapping to the clitoris or is well-perceived in the area of the clitoris.
Gynaecological system: internal pelvic pain syndromes
Endometriosis Endometriosis-associated pain syndrome is chronic or recurrent pelvic pain in patients
associated pain with laparoscopically confirmed endometriosis, and the term is used when the
syndrome symptoms persist despite adequate endometriosis treatment. It is often associated
with negative cognitive, behavioural, sexual or emotional consequences, as well as
with symptoms suggestive of lower urinary tract, sexual, bowel or gynaecological
dysfunction. Many patients have pain above and beyond the endometriotic lesions;
this term is used to cover that group of patients. Endometriosis may be an incidental
finding, is not always painful, and the degree of disease seen laparoscopically does not
correlate with severity of symptoms. As with other patients, they often have more than
one end-organ involved. It has been suggested that this phenotype should be removed
from the classification because the endometriosis may be irrelevant.
Chronic primary Chronic primary pelvic pain syndrome with cyclical exacerbations covers the non-
pelvic pain gynaecological organ pain that frequently shows cyclical exacerbations (e.g., IBS or
syndrome BPS) as well as pain similar to that associated with endometriosis/adenomyosis but
with cyclical where no pathology is identified. This condition is different from dysmenorrhoea, in
exacerbations which pain is only present with menstruation.
Primary Primary dysmenorrhoea is pain with menstruation that is not associated with well-
dysmenorrhoea defined pathology. Dysmenorrhoea needs to be considered as a chronic primary pain
syndrome if it is persistent and associated with negative cognitive, behavioural, sexual
or emotional consequences.

CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021 13


Gastrointestinal Pelvic Pain Syndromes
Irritable bowel Irritable bowel syndrome is the occurrence of chronic or recurrent episodic pain
syndrome perceived in the bowel, in the absence of proven infection or other obvious local
pathology. Bowel dysfunction is frequent. Irritable bowel syndrome is often associated
with worry and pre-occupation about bowel function, and negative cognitive,
behavioural, sexual or emotional consequences, as well as with symptoms suggestive
of lower urinary tract or gynaecological dysfunction. The above classification is based
upon the Rome III Criteria [16]: three months of continuous or recurring symptoms
of abdominal pain or irritation that may be relieved with a bowel movement, may be
coupled with a change in frequency, or may be related to a change in stool consistency.
Two or more of the following are present at least 25% of the time: change in stool
frequency (> three bowel movements per day or < three per week); noticeable
difference in stool form (hard, loose, watery or poorly formed stools); passage of mucus
in stools; bloating or feeling of abdominal distension; or altered stool passage (e.g.,
sensation of incomplete evacuation, straining, or urgency). Extra-intestinal symptoms
include: nausea, fatigue, full sensation after even a small meal, and vomiting.
Chronic primary Chronic primary anal pain syndrome is the occurrence of chronic or recurrent episodic
anal pain pain perceived in the anus, in the absence of proven infection or other obvious local
syndrome pathology. Chronic primary anal pain syndrome is often associated with negative
cognitive, behavioural, sexual or emotional consequences, as well as with symptoms
suggestive of lower urinary tract, sexual, bowel or gynaecological dysfunction.
Intermittent Intermittent chronic primary anal pain syndrome refers to severe, brief, episodic
chronic primary pain that seems to arise in the rectum or anal canal and occurs at irregular intervals.
anal pain This is unrelated to the need to or the process of defecation. It may be considered a
syndrome sub-group of the chronic primary anal pain syndromes. It was previously known as
“proctalgia fugax” but this term is no longer recommended.
Musculoskeletal System
Primary pelvic Primary pelvic floor muscle pain syndrome is the occurrence of persistent or recurrent
floor muscle pain episodic pelvic floor pain. There is no proven well-defined local pathology. It is often
syndrome associated with negative cognitive, behavioural, sexual or emotional consequences,
as well as with symptoms suggestive of lower urinary tract, sexual, bowel or
gynaecological dysfunction. This syndrome may be associated with over-activity of,
or trigger points within, the pelvic floor muscles. Trigger points may also be found in
several muscles, such as the abdominal, thigh and paraspinal muscles and even those
not directly related to the pelvis.
Primary coccyx Primary coccyx pain syndrome is the occurrence of chronic or recurrent episodic pain
pain syndrome perceived in the region of the coccyx, in the absence of proven infection or other
obvious local pathology. Primary coccyx pain syndrome is often associated with
negative cognitive, behavioural, sexual or emotional consequences, as well as
with symptoms suggestive of lower urinary tract, sexual, bowel or gynaecological
dysfunction. The term “coccydynia” was used but is no longer recommended.
Chronic Pain Post-Surgery
Chronic post- The definition of chronic post-surgical pain is chronic pain that develops or increases
surgical pain in intensity after a surgical procedure and persists beyond the healing process, i.e., at
syndrome least three months after the surgery. There is a separate category for this in the ICD-11
classification.

2. METHODOLOGY
2.1 Methods
For each recommendation within the guidelines there is an accompanying strength rating form, the basis
of which is a modified GRADE methodology [17, 18]. Each strength rating form addresses a number of key
elements namely:
1. the overall quality of the evidence which exists for the recommendation, references used in this text
are graded according to a classification system modified from the Oxford Centre for Evidence-Based
Medicine Levels of Evidence [19];

14 CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021


2. the magnitude of the effect (individual or combined effects);
3. the certainty of the results (precision, consistency, heterogeneity and other statistical or study related
factors);
4. the balance between desirable and undesirable outcomes;
5. the impact of patient values and preferences on the intervention;
6. the certainty of those patient values and preferences.

These key elements are the basis which panels use to define the strength rating of each recommendation. The
strength of each recommendation is represented by the words ‘strong’ or ‘weak’ [17]. The strength of each
recommendation is determined by the balance between desirable and undesirable consequences of alternative
management strategies, the quality of the evidence (including certainty of estimates), and nature and variability
of patient values and preferences.

Additional information can be found in the general Methodology section of this print, and online at the EAU
website: http://www.uroweb.org/guideline/. A list of associations endorsing the EAU Guidelines can also be
viewed online at the above address.

The 2012 full text update was based on a systematic review of literature using the Embase and Medline
databases, the Cochrane Central Register of controlled trials and the PsycINFO and Bandolier databases to
identify the best evidence from randomised controlled trials (RCTs) (Level of Evidence 1 [LE: 1]) according to
the rating schedule adapted from the Oxford Centre for Evidence-based Medicine Levels of Evidence. Where
no LE: 1 literature could be identified the search was moved down to the next lower level on the rating scale.
Extensive use of free text ensured the sensitivity of the searches, resulting in a substantial body of literature
to scan. Searches covered the period January 1995 to July 2011 and were restricted to English language
publications. In 2017, a scoping search for the previous five years was performed and the guideline was
updated accordingly.

For the 2021 print, a new section was included on Post-Surgical Pain Syndrome. In addition, the classifications
in the Guideline have been amended to reflect ICD-11 released by WHO. The latest version of ICD-11 will be
available for member states to report with as from January 2022.

2.2 Review
This document was subject to peer review prior to publication in 2021.

3. EPIDEMIOLOGY, AETIOLOGY AND


PATHOPHYSIOLOGY
3.1 Chronic visceral pain
Definition of pain
Pain is defined as an unpleasant sensory and emotional experience associated with actual or potential tissue
damage, or described in terms of such damage (IASP Taxonomy).

Introduction to chronic pelvic primary pain syndromes


Over the years much of the focus for CPPPS has been on peripheral-end-organ mechanisms, such as
inflammatory or infective conditions. However, both animal and clinical research have indicated that many of
the mechanisms for the CPPPSs are based within the CNS. Although a peripheral stimulus such as infection
may initiate the start of a CPPPS condition, the condition may become self-perpetuating as a result of CNS
modulation. As well as pain, these central mechanisms are associated with several other sensory, functional,
behavioural and psychological phenomena. It is this collection of phenomena that forms the basis of the pain
syndrome diagnosis and each individual phenomenon needs to be addressed in its own right through multi-
specialty and multi-disciplinary care. Although ongoing peripheral organ pathology can produce persistent and
chronic pain, the main focus of these guidelines is on CPPPSs in which no peripheral ongoing pathology (such
as infection or neoplastic disease) is detected. The main exception is when pain is due to peripheral nerve
damage.

CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021 15


3.1.1 Incidence
No adequate data on incidence were found.

3.1.2 Prevalence
Across the world [20] chronic pain is prevalent, seriously affecting the quality of people’s social, family, and
working lives, with differences between countries attributable to multiple causes, including study methodology.
A more recent study in the UK found a prevalence of chronic pelvic pain of 14.8% in women over 25 years [21].

3.1.3 Influence on Quality of Life


Assessing QoL in pelvic pain patients is challenging due to the complex pathology, the multi-faceted nature of
the complaints and the overlap between the different pelvic pain syndromes [22, 23].

Pelvic pain syndromes have an impact in terms of QoL [24, 25], depression, anxiety, impaired emotional
functioning, insomnia and fatigue [24, 26]. If these aspects are identified and targeted early in the diagnostic
process, the associated pain symptoms may also improve. Addressing comorbidities will help in further
improving QoL [27]. Quality of life assessment is therefore important in patients with pelvic pain and should
include physical, psychosocial and emotional tools, using standardised and validated instruments [25].

Chronic pain is, in many countries, the leading cause of years lost to disability [20], although these figures are
dominated by musculoskeletal pain and headache. Chronic pain is often associated with depression and other
psychological problems; with loss or reduction of work and of ability to carry out domestic tasks; and, with
substantial use of healthcare, often with disappointing outcomes.

3.1.4 Costs
No adequate data on costs were found.

3.1.5 Risk Factors and underlying causes


3.1.5.1 Risk factors
Risk factors include many different factors from various areas, including genetic, psychological state, recurrent
physical trauma and endocrine factors.

The endocrine system is involved in visceral function. Significant life events, and in particular, early life events
may alter the development of the hypothalamic-pituitary-adrenal axis and the chemicals released. Increased
vulnerability to stress is thought to be partly due to increased corticotrophin-releasing hormone (CRH) gene
expression. Up-regulation of CRH has been implicated in several pain states such as rectal hypersensitivity to
rectal distension. This model suggests an action of CRH on mast cells. A range of stress-related illnesses have
been suggested, e.g., IBS and BPS. There is evidence accumulating to suggest that the sex hormones also
modulate both nociception and pain perception. Stress can also produce long-term biological changes which
may form the relation between chronic pain syndromes and significant early life and adverse life events [28].
Asking the patient about these events is important as they have an effect on a patient’s psychological wellbeing
[29, 30].

Genetics also play a role in assessing the risk of developing chronic pain. An individual who has one chronic
pain syndrome is more likely to develop another. Family clusters of pain conditions are also observed and
animals can be bred to be more prone to apparent chronic pain state. A range of genetic variations have
been described that may explain the pain in certain cases; many of these are to do with subtle changes
in transmitters and their receptors. However, the picture is more complicated in that developmental,
environmental and social factors also influence the situation. Evidence that PBPS may have a genetic
component has been presented in several identical twin studies, but genetics may contribute to less than one
third of total variation in susceptibility to PBPS [31, 32].

Studies about integrating the psychological factors of CPPPSs are few but the quality is high. Psychological
factors are consistently found to be relevant in the maintenance of persistent pelvic and urogenital pain
[33]. Beliefs about pain contribute to the experience of pain [34] and symptom-related anxiety and central
pain amplification may be measurably linked, as in IBS [35], and catastrophic thinking about pain and
perceived stress predict worsening of urological chronic pain over a year [33, 36]. Central sensitisation has
been demonstrated in symptomatic endometriosis [37] and central changes are evident in association with
dysmenorrhoea and increasingly recognised as a risk for female pelvic pain [38]. The various mechanisms
of CNS facilitation, amplification and failure of inhibition, mean that there is no simple relationship between
physical findings, pain experienced and resulting distress and restriction of activities. Division of aetiology into

16 CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021


organic vs. psychogenic is unscientific. Diagnoses that assign women’s pain to psychological origins due to
scepticism about the reality or severity of their pain [39, 40] undermines any therapeutic relationship [41]. Pelvic
pain and distress may be related [42] in both men and women [43]; as are painful bladder and distress [36].
In a large population based study of men, CPPPS was associated with prior anxiety disorder [44]. The only
systematic review [45] of risk factors for chronic non-cyclical pelvic pain in women included, as well as medical
variables: sexual or physical abuse (Odds Ratio (OR): 1.51 to 3.49); psychological problems such as anxiety
(OR: 2.28, 95% Confidence Interval (CI): 1.41- 3.70) and depression (OR: 2.69, 95% CI: 1.86-3.88); multiple
somatic problems (OR: 4.83, 95% CI: 2.50-9.33 and OR: 8.01, 95% CI: 5.16-12.44).

Many studies have reported high rates of childhood sexual abuse in adults with persistent pain, particularly
in women with pelvic pain [46]. It is hard to establish a causal role for sexual abuse or trauma history, anxiety
or depression in women with CPPPS [47-49], the attribution of current pain to past sexual or physical
abuse is associated both with current depression [50] and with current overall physical health [51]. There is
some evidence for a specific relationship between rape and CPPPS (and with fibromyalgia and functional
gastrointestinal disorders) [52]; and, recent sexual assault may prompt presentation of pelvic pain [46, 53].
Few studies have been found of sexual or physical abuse in childhood and pelvic pain in men, although it has
known adverse effects on health [52], but men who reported having experienced sexual, physical or emotional
abuse had increased odds (3.3 compared to 1.7) for symptoms suggestive of CPPPS [54]. Both sexes should
be screened for sexual abuse when presenting with symptoms suggestive of CPPPS, and clinicians should
inquire about pelvic pain in patients who have experienced abuse [54].

3.1.5.2 Underlying causes


The mechanisms that serve as an underlying cause for chronic pelvic pain are:

1. Ongoing acute pain mechanisms [55] (such as those associated with inflammation or infection), which
may involve somatic or visceral tissue.
2. Chronic pain mechanisms, which especially involve the CNS [6].
3. Emotional, cognitive, behavioural and sexual responses and mechanisms [56-58].

Symptoms and signs of neuropathic pain appear to be common in CPPPS patients and assessment of
neuropathic pain should be considered in that group of patients. The presence or absence of endometriosis
does not seem to change this [59].

Chronic pain mechanisms may include altered resting state neuromotor connectivity, for instance in men with
chronic prostatitis/CPPPS [60].

Table 3 illustrates some of the differences between the somatic and visceral pain mechanisms. These underlie
some of the mechanisms that may produce the classical features of visceral pain; in particular, referred pain
and hyperalgesia.

CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021 17


Table 3: Comparison between visceral and somatic pain

Visceral pain Somatic pain


Effective painful stimuli Stretching and distension, Mechanical, thermal, chemical and
producing poorly localised pain. electrical stimuli, producing well
localised pain.
Summation Widespread stimulation produces Widespread stimulation produces a
significantly magnified pain. modest increase in pain.
Autonomic involvement Autonomic features (e.g., nausea Autonomic features less frequent.
and sweating) frequently present.
Referred pain Pain perceived at a site distant to Pain is relatively well localised and
the cause of the pain is common. well recognised.
Referred hyperalgesia Referred cutaneous and muscle Hyperalgesia tends to be localised.
hyperalgesia is common, as is
involvement of other visceral organs.
Innervation Low density, unmyelinated C fibres Dense innervation with a wide
and thinly myelinated A∂ fibres. range of nerve fibres.
Primary afferent physiology Intensity coding. As stimulation Two fibre coding. Separate fibres
increases, afferent firing increases for pain and normal sensation.
with an increase in sensation and
ultimately pain.
Silent afferents 50-90% of visceral afferents These fibres are very important in
are silent until the time they are the central sensitisation process.
switched on. Silent afferents present, but form a
lower percentage.
Central mechanisms Play an important part in the Sensations not normally perceived
hyperalgesia, viscero-visceral, become perceived and non-
viscero-muscular and musculo- noxious sensations become
visceral hyperalgesia. painful. Responsible for the
allodynia and hyperalgesia of
chronic somatic pain.
Abnormalities of function Central mechanisms associated Somatic pain associated with
with visceral pain may be somatic dysfunction, e.g., muscle
responsible for organ dysfunction. spasm.
Central pathways and As well as classical pathways, Classical pain pathways.
representation there is evidence for a separate
dorsal horn pathway and central
representation.

Ongoing peripheral pain mechanisms in visceral pain


In most cases of chronic pelvic pain, ongoing tissue trauma, inflammation or infection is absent [61, 62].
However, conditions that produce recurrent trauma, infection or ongoing inflammation may result in chronic
pelvic pain in a small proportion of cases. For example, out of a large cohort with acute bacterial prostatitis,
10.5% ended up with a state of CPPPS [63]. It is for this reason that the early stages of assessment include
looking for these pathologies [15]. Once excluded, ongoing investigations for these causes are rarely helpful
and indeed may be detrimental.

When acute pain mechanisms are activated by a nociceptive event, as well as direct activation of the peripheral
nociceptor transducers, sensitisation of those transducers may also occur; therefore, magnifying the afferent
signalling. Afferents that are not normally active may also become activated by the change, that is, there may
be activation of the so-called silent afferents. Although these are mechanisms of acute pain, the increased
afferent signalling is often a trigger for the chronic pain mechanisms that maintain the perception of pain in the
absence of ongoing peripheral pathology (see below) [64].

There are a number of mechanisms by which the peripheral transducers may exhibit an increase in sensibility:
1. Modification of the peripheral tissue, which may result in the transducers being more exposed to
peripheral stimulation.
2. There may be an increase in the chemicals that stimulate the receptors of the transducers [65].
3. There are many modifications in the receptors that result in them being more sensitive.

18 CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021


In general, the effect of 1 and 2 is to lower the threshold and the effect of 3 is to increase responsiveness to
external stimuli. Some of the chemicals responsible for the above changes may be released from those cells
associated with inflammation, but the peripheral nervous system may also release chemicals in the positive
and inhibitory loops [66, 67].

Central sensitisation as a mechanism in visceral pain


It is important to appreciate that nociception is the process of transmitting information to centres involved in
perception of a stimulus that has the potential to cause tissue damage. Pain is far more complex and involves
activation of the nociceptive pathways but also the emotional response. The brain may affect the modulation of
pain pathways at the spinal cord level.

Neuronal sensitisation is responsible for a decrease in threshold and an increase in response duration and
magnitude of dorsal horn neurons. It is associated with an expansion of the receptive field. As a result,
sensitisation increases signalling to the CNS and amplifies what we perceive from a peripheral stimulus.
For example, for cutaneous stimuli, light touch would not normally produce pain, however, when central
sensitisation is present, light touch may be perceived as painful (allodynia). In visceral hyperalgesia (so called
because the afferents are primarily small fibres), visceral stimuli that are normally sub-threshold and not usually
perceived, may be perceived. For instance, with central sensitisation, stimuli that are normally sub-threshold
may result in a sensation of fullness and a need to void or to defecate. Non-noxious stimuli may be interpreted
as pain and stimuli that are normally noxious may be magnified (true hyperalgesia) with an increased
perception of pain. As a consequence, one can see that many of the symptoms of PBPS and IBS may be
explained by central sensitisation. A similar explanation exists for the muscle pain in FM.

It is now well accepted that there are both descending pain-inhibitory and descending pain-facilitatory
pathways that originate from the brain [68]. Several neurotransmitters and neuromodulators are involved in
descending pain-inhibitory pathways. The main ones are the opioids, 5-hydroxytryptamine and noradrenaline.

The autonomic nervous system also plays a role in sensitisation. There is good evidence that damaged afferent
fibres may develop a sensitivity to sympathetic stimulation, both at the site of injury and more centrally,
particularly in the dorsal horns. In visceral pain, the efferent output of the CNS may be influenced by central
changes (again, those changes may be throughout the neuraxis), and such modification of the efferent
message may produce significant end-organ dysfunction. These functional abnormalities can have a significant
effect on QoL and must be managed as appropriate.

Psychological mechanisms in visceral pain


Psychological processes of emotions, thought and behaviour involve networks rather than distinct centres.
Some of these processes are sophisticated and others fundamental in evolutionary terms, and their interaction
with pain processing is complex.

Various psychological processes affect pain neuromodulation at a higher level. Inhibiting or facilitating both
the strength of the nociceptive signal reaching the consciousness and appraisal and interpretation of that
signal, will also modulate the response to the nociceptive message and hence the pain experience. Further,
descending pathways represent cognitive, emotional and behavioural states at spinal and peripheral levels.
Functional magnetic resonance imaging (fMRI) has indicated that the psychological modulation of visceral pain
probably involves multiple pathways. For instance, mood and attentional focus probably act through different
areas of the brain when involved in reducing pain [69].

This psychological modulation may act to reduce nociception within a rapid time frame but may also result
in long-term vulnerability to chronic visceral pain, through long-term potentiation. This involvement of
higher centre learning may be at both a conscious and subconscious level, and is clearly significant in the
supratentorial neuroprocessing of nociception and pain. Long-term potentiation [70] may occur at any level
within the nervous system, so that pathways for specific or combinations of stimuli may become established,
resulting in an individual being vulnerable to perceiving sensations that would not normally be experienced as
painful.

An important review [28] of chronic pelvic pain in women dismantles the notion that women without relevant
physical findings differ in psychological characteristics from women with relevant physical findings. It argues
for better methodology, and for greater use of idiographic methods. Women with pelvic pain often have other
non-pain somatic symptoms and current or lifetime anxiety and depression disorder [21]; they may have a
history of physical or sexual abuse in childhood of unclear significance. Studies that describe these non-pain

CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021 19


somatic symptoms as ‘medically unexplained’ or ‘psychosomatic’ or ‘somatoform’ disorders are unhelpful,
misinterpreting absence of physical findings to indicate psychological origins of the complaint. Pain studies
describe multiple processes by which pain may spread from one site to another, or in time, including central
sensitisation (see previous section), viscero-visceral cross sensitisation in relation to multiple pain sites [71],
activation of the hypothalamic-pituitary axis and dysregulation of serotonergic pathways [72] that can render
pain levels sensitive to stress. Some pain problems which affect sexual activity are diagnosed as sexual
problems (e.g., ‘dyspareunia’) when pain is the central problem and not contingent on sexual activity alone [73].
Better integration of sexology and mainstream psychology for pelvic pain in both men and women is needed,
building on a biopsychosocial formulation [74, 75].

The term psychosomatic symptoms can best be understood as multiple somatic symptoms not associated
with or indicative of any serious disease process. Medical and surgical history may also be important [76].
There have been a few studies of maintenance of, or recovery from, pelvic pain in relation to psychological
factors of importance in pain. Those that described pelvic pain as medically unexplained or psychosomatic,
due to the lack of physical findings, have been discarded, because such a distinction is inconsistent with
known pain mechanisms.

Understanding the psychological components of pain


Psychological processes of emotions, thought and behaviour involve distributed networks, whose interactions
with pain processing are complex, producing inhibition and facilitation of signal processing, appraisal, and
response. Models that integrate psychological factors involved in maintaining persistent pelvic and urogenital
pain with current neurobiological understanding of pain are few, but the quality is high (see Section 3.1.5.1).

There is no evidence that women with CPPPS without physical findings are primarily presenting a
psychological problem [28]. Anxiety and post-traumatic stress symptoms are common in some women with
CPPPS [40, 77] and with vulvar pain [78], and may account for substantial variance in health status, treatment
use and treatment outcome; for instance, women’s expectations about vulvar pain on penetration predicted
pain, sexual function and sexual satisfaction [79]. Negative investigative findings do not necessarily resolve
women’s anxieties about the cause of pain [80, 81] and anxiety often focuses on what might be ‘wrong’.
Depression may be related to pain in various ways, as described above. Until measures are available that are
adequately standardised in patients with pain, assessment of anxiety and distress requires questions about
the patient’s beliefs about the cause of pain, the hope that diagnosis will validate pain, the struggle with
unpredictability, and the implications of pain for everyday life [82, 83]. Reference to the studies of the IMMPACT
group [84] is recommended for guidance on outcome measures suitable for pain trials.

Stress can modify the nervous system to produce long-term biological changes. These structural changes
may be responsible for significant early life and adverse life events which are associated with chronic pain
syndromes [30]. The patient should be asked about adverse life events that may produce these biological
responses and affect a patient’s general psychological well-being [30, 85].

3.1.5.3 Clinical paradigms in visceral pain


Referred pain
Referred pain is frequently observed and its identification is important for diagnosis and treatment. Referral
is usually somatic to somatic, or visceral to somatic. However, there is no reason why pain cannot also be
perceived within the area of an organ with the nociceptive signal having arisen from a somatic area. Referred
pain may occur as a result of several mechanisms but the main theory is one of convergence-projection. In the
convergence-projection theory, afferent fibres from the viscera and the somatic site of referred pain converge
onto the same second order projection neurons. The higher centres receiving messages from these projection
neurons are unable to separate the two possible sites from the origin of the nociceptive signal [64].

Hyperalgesia refers to an increased sensitivity to normally painful stimuli. In patients that have passed a renal
stone, somatic muscle hyperalgesia is frequently present, even a year after expulsion of the stone. Pain to
non-painful stimuli (allodynia) may also be present in certain individuals. Somatic tissue hyperaesthesia is
associated with urinary and biliary colic, IBS, endometriosis, dysmenorrhoea, and recurrent bladder infections.
Primary vulvar pain syndromes are examples of cutaneous allodynia that, in certain cases, may be associated
with visceral pain syndromes, such as BPS. Referred pain with hyperalgesia is thought to be due to central
sensitisation of the converging viscero-somatic neurons. Central sensitisation also stimulates efferent activity
that could explain the trophic changes that are often found in the somatic tissues.

20 CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021


Musculo-skeletal system and pelvic pain
In the urogenital pain syndromes, muscle tenderness and trigger points may be implicated as a source of pain.
Central mechanisms are of great importance in the pathogenesis of this muscle hyperalgesia. The muscles
involved may be a part of the spinal, abdominal or pelvic complex of muscles. It is not unknown for adjacent
muscles of the lower limbs and the thorax to become involved. Pain may be localised to the trigger points but it
is more often associated with classical referral patterns. As well as trigger points, inflammation of the ligaments
and tendons to the bones (enthesitis) and of the bursa (bursitis) may be found [86]. Certain postures affect the
different muscles in different ways, and as a consequence, may exacerbate or reduce the pain. Stress has
been implicated as both an initiator of pelvic myalgia and as a maintenance factor. As a result, negative sexual
encounters may also have a precipitating effect [28].

Visceral hyperalgesia
The increased perception of stimuli in the viscera is known as visceral hyperalgesia, and the underlying
mechanisms are thought to be responsible for IBS, PBPS and dysmenorrhoea. The mechanisms involved are
often acute afferent input (e.g., due to infection) followed by long-term central sensitisation. Viscero-visceral
hyperalgesia is thought to be due to two or more organs with converging sensory projections and central
sensitisation. For instance, overlap of bladder and uterine afferents or uterine and colon afferents.

3.2 Pelvic Pain


3.2.1 Incidence
No adequate data on incidence were found.

3.2.2 Prevalence
3.2.2.1 Primary prostate pain syndrome
There is only limited information on the true prevalence of PPPS in the population. As a result of significant
overlap of symptoms with other conditions (e.g., benign prostatic enlargement and PBPS), purely symptom
based case definitions may not reflect the true prevalence of PPPS [87, 88]. In the literature, population-based
prevalence of prostatitis symptoms ranges from 1 to 14.2% [89, 90]. The risk of prostatitis increases with age
(men aged 50-59 years have a 3.1-fold greater risk than those aged 20-39 years).

3.2.2.2 Primary bladder pain syndrome


Reports of PBPS prevalence have varied greatly, along with the diagnostic criteria and populations studied.
Recent reports range from 0.06 to 30% [91-100]. There is a female predominance of about 10:1 [97] but
possibly no difference in race or ethnicity [87, 101, 102]. The relative proportions of Hunner’s lesion and non-
lesion disease are unclear. Incidence in studies has ranged from 5 to 50% [103-106]. There is increasing
evidence that children under eighteen may also be affected, although prevalence figures are low; therefore,
PBPS cannot be excluded on the basis of age [107].

3.2.2.3 Sexual pain syndrome


In the 1980s an association between chronic pelvic pain and sexual dysfunction was postulated. In a review the
relationship between Primary Prostate Pain Syndrome and health status, with influence on sexual activity,
was addressed [108]. In a Chinese study of men with chronic pelvic pain, 1,768 males completed the
questionnaires. The overall prevalence of sexual dysfunction was 49%. Erectile dysfunction (ED) is the most
investigated sexual dysfunction in PPPS patients. The reported prevalence of ED ranges from 15.1 to 48%,
varying with evaluation tools and populations [109, 110]. Erectile dysfunction was prevalent in 27.4% of Italian
men aged 25-50 [111], 15.2% among Turkish men (significantly higher than in the control group) [112] and
43% among Finnish men with PPPS [113]. The prevalence of ED was found to be higher in young men with
PPPS than in the general population. According to other studies men with pelvic pain had a higher chance of
suffering from ED [114]. Recently, a significant correlation between “chronic prostatitis”, chronic pelvic pain
symptoms (measured by NIH-CPSI) and ED (measured by International Index of Erectile Function [IIEF]) was
confirmed [115], while other studies using the same questionnaires were not able to confirm such a correlation
[75, 116]. Some studies also report ejaculatory dysfunction, mainly premature ejaculation [109, 110, 117, 118].

In community-based studies in the UK [119], New Zealand [120] and Australia [121], a substantially larger
proportion of the women with chronic pelvic pain reported dyspareunia (varying between 29 and 42%) than
women without chronic pelvic pain (varying between 11 and 14%). Only a few studies have investigated
sexual problems within clinical populations [122]. Another study showed that all of the sexual function domains
(desire, arousal, lubrication, orgasm, satisfaction, and pain) were significantly lower in women with chronic
pelvic pain than in women without chronic pelvic pain [122]. In line with the results of community based studies,
patients with chronic pelvic pain reported more sexual problems such as dyspareunia, problems with desire or

CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021 21


arousal and lubrication than women without chronic pelvic pain [122, 123]. One study of patients enrolled in
chronic pain treatment programs in England has reported that 73% had pain-related sexual problems [124].

3.2.2.4 Myofascial pain syndromes


The relationship between muscular dysfunction (especially over-activity) and pelvic pain has been found in
several studies [125]. Rectal pain treated with pelvic floor muscle therapy is only relieved when patients learn to
relax their pelvic floor muscles [126, 127]. The vast majority (92.2%) of men visiting a tertiary centre for pelvic
pain had dysfunction of the pelvic floor muscles. This finding was true regardless of evidence of inflammation
(prostatitis or cystitis) [128]. This relationship has been found in chronic prostatitis [129], PBPS [130] and vulvar
pain [131]. Dysfunction of the pelvic floor directly affects function of the pelvic viscera and vice versa. Both
systems can act as the primary signal to the spinal cord, with a cascade of reactions ascending to the CNS as
a result. The muscle itself ends up shortened, leading to restrictions even in a relaxed state.

3.2.3 Influence on QoL


Data on the influence on QoL will be included in a future version of the guidelines.

3.2.4 Costs
No adequate data on costs were found.

3.2.5 Risk factors and underlying causes


The risk factors are unspecific for most of the pain syndromes in the pelvic area. They are described in
Section 3.1.5.1. The underlying causes, including the mechanisms for the different clinical pain syndromes are
described here.

3.2.5.1 Primary Prostate Pain Syndrome


Pain is the main symptom in PPPS. As a common feature of primary chronic pain syndromes, no single aetiological
explanation has been found. One explanation is that the condition probably occurs in susceptible men exposed
to one or more initiating factors, which may be single, repetitive or continuous. Several of these potential
initiating factors have been proposed, including infectious, genetic, anatomical, neuromuscular, endocrine,
immune (including autoimmune), or psychological mechanisms. These factors may then lead to a peripheral self-
perpetuating immunological, inflammatory state and/or neurogenic injury, creating acute and then chronic pain.
One recent study showed that chronic but not acute histological inflammation of the prostate was significantly
associated with symptomatic progression [132]. Based on the peripheral and the central nervous system,
sensitisation involving neuroplasticity may lead to a centralised neuropathic pain state [133]. This could also
explain why tissue damage is not usually found in PPPS. There is growing evidence for a neuropathic origin and
association with CNS changes of pain in PPPS, and anxiety appears to be a risk factor for its development [44].

3.2.5.2 Primary Bladder Pain Syndrome


An initial unidentified insult to the bladder, leading to urothelial damage, neurogenic inflammation and pain is
thought to be a trigger of PBPS. However, PBPS might be a local manifestation of a systemic disorder. No
infection has as yet been implicated. Nevertheless, urinary infections are significantly more frequent during
childhood and adolescence, in patients with PBPS in adulthood [134]. Experimental induction of chronic
pelvic pain by O-antigen deficient bacterial strains supports the bacterial hypothesis [135]. Pancystitis, with
associated perineural inflammatory infiltrates, and mast cell count increase is an essential part of PBPS type
3 C [136], but is rare in non-lesion PBPS [30, 69, 137, 138]. Cystoscopic and biopsy findings in both lesion
and non-lesion PBPS are consistent with defects in the urothelial glycosaminoglycan (GAG) layer, which might
expose submucosal structures to noxious urine components [139-145] and a consequent cytotoxic effect [146,
147]. Basic and clinical studies indicate that autonomic dysfunction with sympathetic predominance may be
implicated in PBPS [148, 149].
An association has been reported between PBPS and non-bladder syndromes such as FM, CFS,
IBS, vulvodynia, depression, panic disorders, migraine, sicca syndrome, temporomandibular joint disorder,
allergy, asthma and systemic lupus erythematosus [150-154].
Risk of PBPS correlates with a number of non-bladder syndromes in each patient [155]. Recent
work showing non-lesion PBPS to have significantly more FM, migraine, temporomandibular joint disorder and
depression than PBPS type 3 C patients, which emphasises the need for subtyping [156].

3.2.5.3 Primary Scrotal Pain Syndrome


Often scrotal pain is not associated with any specific pathology. Pain is perceived in the testes, epididymis, or the
vas deferens. The ilioinguinal, genitofemoral and the pudendal nerves innervate the scrotum [157]. Any pathology
or intervention at the origin or along the course of these nerves may result in pain perceived in the scrotum [158].

22 CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021


Two special forms of scrotal pain syndrome can be described. The first is post-vasectomy scrotal pain
syndrome which occurs following vasectomy. The mechanisms are poorly understood, and for that reason it
is considered by some a special form of primary scrotal pain syndrome. Incidence of post-vasectomy pain is
2-20% among all men who have undergone a vasectomy [159]. In men with post-vasectomy pain, 2-6% have
a Visual Analogue Scale (VAS) score > 5 [160]. In a large cohort study of 625 men, the likelihood of scrotal pain
after six months was 14.7%. The mean pain severity on a VAS score was 3.4/10. In the pain group, 0.9% had
quite severe pain, noticeably affecting their daily life. In this cohort, different techniques were used to perform
the vasectomy. The risk of post-vasectomy pain was significantly lower in the no-scalpel vasectomy group
(11.7% vs. the scalpel group 18.8%) [161].

The second special form of scrotal pain is post-inguinal hernia repair pain. It is seen as a complication of hernia
repair, but in trials it is seldom reported, or it is put under the term chronic pain (not specified). In studies that
have explicitly mentioned scrotal pain, there was a difference in incidence between laparoscopic and open
hernia repair. In almost all studies, the frequency of scrotal pain was significantly higher in the laparoscopic
than in the open group [158, 162]. In one particular study, there was no difference at one year but after five
years, the open group had far fewer patients with scrotal pain [163]. Inguinal hernia repair can lead to chronic
post-surgical pain (CPSP) in up to 10% of patients at 6 months [164] and may present with groin and/or scrotal
pain. Testicular injury is uncommon (< 1%) but if associated with pain, orchidectomy can lead to symptomatic
relief in 2/3 of patients [165]. Careful identification and preservation of nerves has been found to be associated
with a reduced risk of chronic pain.

3.2.5.4 Primary Urethral Pain Syndrome


Several mechanisms for the development of primary urethral pain syndrome have been proposed. The intimate
relationship of the urethra with the bladder (both covered with urothelium) suggests that primary urethral pain
syndrome may be a form of PBPS. Mechanisms thought to be basic for PBPS may also apply to the urethra.
This means that the specific testing with potassium has been used to support the theory of epithelial leakage
[166, 167]. Another possible mechanism is neuropathic hypersensitivity following urinary tract infection [168].
The relationship with gynaecological and obstetric aspects is unclear. In a small group of patients with urethral
pain, it has been found that grand multi-parity and delivery without episiotomy were more often seen in patients
with urethral syndrome, using univariate analysis [169].

3.2.5.5 Primary Vaginal and Vulvar Pain Syndromes


Pain in the vagina or the female external genital organs is often due to infection or trauma, as a consequence
of childbirth or surgery. Pain is usually a precedent to dyspareunia. When the pain persists for more than three
months, it can be diagnosed as primary vulvar pain syndrome previously known as “vulvodynia” or “chronic
vaginal pain” with no known cause. It is still a poorly understood condition, and therefore difficult to treat.

There are two main sub-types of primary vulvar pain syndrome: generalised, where the pain occurs in different
areas of the vulva at different times; and focal, where the pain is at the entrance of the vagina. In primary
generalised vulvar pain syndrome, the pain may be constant or occur occasionally, but touch or pressure does
not initiate it, although it may make the pain worse. In primary focal vulvar pain syndrome, the pain is described
as a burning sensation that comes on only after touch or pressure, such as during intercourse.

The possible causes of primary vulvar pain syndrome are many and include:
• history of sexual abuse;
• history of chronic antibiotic use;
• hypersensitivity to yeast infections, allergies to chemicals or other substances;
• abnormal inflammatory response (genetic and non-genetic) to infection and trauma;
• nerve or muscle injury or irritation;
• hormonal changes.

3.2.5.6 Chronic Pelvic Pain and Prolapse/Incontinence Mesh


Continence and prolapse mesh implants were developed as simple flexible polypropylene plastic acting as a
scaffold to treat urinary stress incontinence (USI) and uterovaginal prolapse, respectively. They were deemed
easy to insert, but no credence was given as to how safe they were, whether they could be removed should
they cause complications, or what to do should they not be effective [170, 171]. Most meshes took less than
an hour to implant surgically and most patients were treated as day cases, allowing women to leave hospital
quickly and get on with their lives. Therefore, rather than undergo complex traditional surgery, women were
offered permanent mesh implants, particularly in the treatment of USI where they were considered to be the
gold standard [172, 173]. However, over the last few years the insertion of mesh has come with significant
‘health and safety warnings’ [174, 175].

CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021 23


For many, mesh was initially seen not just as an effective treatment but as a permanent one. Complications
were thought not be a significant issue and the figure of 1-3% was often quoted. However, we now know the
complication rate was closer to 10% [176]. They included chronic pain [177, 178], as well as chronic infections
[179], erosion into the surrounding organs including the vagina, urethra and bladder, as well as nerve and musculo-
skeletal damage affecting mobility [177, 178, 180, 181]. All had a significant impact on the patients’ QoL.

It is as a result of severely debilitating complications following mesh implantation [177], that the field of mesh
removal medicine and surgery has emerged [182].

Early recognition of possible mesh complications is very important. It is normal to wake up in some degree
of discomfort after any surgery. However, if the pain after the operation is very severe and much more than
expected after this type of surgery, it can be a sign that there was added trauma to the surrounding organs
during the procedure. Most pain is often managed with analgesia, but some women might not fully respond to
therapy. If the pain is difficult to treat and does not improve over time, it may become necessary to remove the
mesh. Leaving a painful mesh in the pelvis, can lead to chronic pelvic pain. The precise mechanism is unknown
but it is thought to be a ‘neuro-inflammatory’ process [183], as has been proposed in hernia mesh neuralgia.
The impact of the mesh, regardless of site, appears to be similar.

3.2.5.7 Chronic post-surgical pain


Chronic pain may develop following surgical procedures and has a significant impact on the individual. The
ICD-11 has recently classified chronic post-surgical pain (CPSP) as a chronic pain condition. The definition of
CPSP is chronic pain that develops or increases in intensity after a surgical procedure and persists beyond the
healing process, i.e., at least 3 months after the surgery [184].

Chronic post-surgical pain (as start of sentence) may occur in a significant number of patients, and is more
prevalent following some operations rather than others. Procedures with a higher risk of CPSP include limb
amputation (30-85%), thoracotomy (5-65%) and mastectomy (11-57%) [185].

Risk factors for CPSP include a number of pre-, peri- and post-operative factors. Younger age, female gender,
chronic pain pre-operatively elsewhere, higher number of previous operations, use of opioids and a higher
post-operative pain score have been found to be associated with a higher risk of CPSP in a prospective cohort
of patients undergoing laparoscopy and laparotomies. Older age, malignant indication for surgery, a higher
pre-operative mental health score and the use of epidural analgesia in addition to general anaesthesia were
protective [186, 187].

There are a number of procedures specific to the abdomen and pelvis that are associated with an increased
risk of chronic pain post-surgery including bariatric procedures, inguinal hernia repair, vasectomy, hysterectomy
and caesarean section.

The estimated prevalence of CPSP following bariatric surgery is 30% [188]. In affected individuals careful
assessment that may include laparoscopy could identify a treatable cause (such as adhesions, mesenteric
defect or cholecystitis) and lead to a significant reduction in post-operative pain [189].

Inguinal hernia repair can lead to CPSP in up to 10% of patients at 6 months [164] and may present with groin
and/or scrotal pain.

The incidence of post vasectomy pain ranges from 2-20% [159, 160]. The risk is significantly lower following
the no scalpel technique [161].

The incidence of post-surgical pain following hysterectomy is difficult to determine as pain is a common
indication for the operation. When defined as CPSP, rates are estimated at 28-30% [190, 191]. Careful case
selection and management of patient expectation is therefore important.

The frequency of caesarean section has increased over time. A meta-analysis has shown a significant
incidence of CPSP both at three months and at more than 12 months (15% and 11% respectively) [192],
therefore careful counselling is needed in non-emergency cases.

3.2.5.8 Associated conditions in pelvic pain syndromes


Nerve damage
Spinal pathology and any pathology along the course of the nerve involved may result in neuropathic pain in

24 CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021


the distribution of these nerves. Neoplastic disease, infection, trauma, surgical incisions and post-operative
scarring may result in nerve injury [193].

Pudendal neuralgia is the most often mentioned form of nerve damage in the literature. Anatomical variations
may pre-dispose the patient to developing pudendal neuralgia over time or with repeated low-grade trauma
(such as sitting for prolonged periods of time or cycling) [194, 195].

The pudendal nerve may be damaged at the level of:


1. The piriformis muscle. For example, as part of a piriformis syndrome: in some cases, the nerve may pass
through the muscle and hence be trapped; or in other cases, muscle hypertrophy or spasm is implicated.
2. The sacrospinal/sacrotuberous ligaments, possibly accounting for 42% of cases.
3. Within Alcock’s canal (medial to the obturator internus muscle, within the fascia of the muscle), possibly
accounting for 26% of cases.
4. Multiple levels in 17% of cases.
The site of injury determines the location of perceived pain and the nature of associated symptoms (e.g., the
more distal the damage, the less likely the anal region will be involved).

The clinical presentation depends on different factors. There is a wide age range, as one would expect, with
a condition that has so many potential causes. It is suggested that, the younger the patient, the better the
prognosis. Essentially, the sooner the diagnosis is made, as with any compression nerve injury, the better
the prognosis, and older patients may have a more protracted problem [196, 198]. Six out of ten cases are
observed in women. Some special situations can be listed:
• In orthopaedic hip surgery, pressure from the positioning of the patient, where the perineum is placed
hard against the brace, can result in pudendal nerve damage [199, 200]. The surgery itself may also
directly damage the nerve. Pelvic surgery such as sacrospinous fixation is clearly associated with
pudendal nerve damage in some cases [201, 202]. In many types of surgery, including colorectal,
urological and gynaecological, pudendal nerve injury may be implicated.
• Fractures of the sacrum or pelvis may result in pudendal nerve/root damage and pain. Falls and trauma to
the gluteal region may also produce pudendal nerve damage if associated with significant tissue injury or
prolonged pressure.
• Tumours in the pre-sacral space must be considered. Tumours invading the pudendal nerve may occur
and there may also be damage from surgery for pelvic cancer [203].
• The pudendal neuralgia of birth trauma is thought to resolve in most cases over a period of months.
However, rarely, it appears to continue as painful neuropathy. Multiple pregnancies and births may
predispose to stretch neuropathy in later life. This is more difficult to be certain about [204].
• Child birth and repeated abdominal straining associated with chronic constipation [205] are thought to
pre-dispose elderly women to post-menopausal pelvic floor descent and stretching of the pudendal
nerve with associated pain. Changes in the hormone status may also be a factor. In Urogenital Pain
Management Centres, the commonest associations with pudendal neuralgia appear to be: history of
pelvic surgery; prolonged sitting (especially young men working with computer technology); and post-
menopausal older women.

Sexual dysfunction
Chronic pelvic pain is a clinical condition that results from complex interactions of physiological and
psychological factors and has a direct impact on the social, personal and professional lives of men and women.

Men
Chronic pain as well as its treatment can impair our ability to express sexuality. In a study in England, 73%
of patients with any chronic pain had some degree of sexual problems as a result of the pain [124]. These
problems can occur because of several factors. Psychological factors like decrease in self-esteem, depression
and anxiety can contribute to loss of libido. Physiological factors like fatigue, nausea and pain itself can cause
sexual dysfunction. Pain medications (opioids, and the selective serotonin re-uptake inhibitors [SSRIs]) can also
decrease libido [206] and delay ejaculation. The number of studies on the effects of CPPPS on sexual function
is limited. Sexual dysfunction is often ignored because of a lack of standardised measurements. At present, the
most commonly used tool is the IIEF questionnaire [146].

The presence of pelvic pain may increase the risk for ED independent of age [207]. On the other hand, cross-
sectional data suggest no improvement of lower urinary tract symptoms (LUTS) by an increased frequency
of ejaculation [208]. Although mental distress and impaired QoL related to illness could contribute to sexual
dysfunction observed in patients with PPPS, the presence of erectile and ejaculatory disorders is more

CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021 25


frequently related to symptoms suggestive of a more severe inflammatory condition [118]. These arguments
are important for the understanding of the close relationship between CPPPS symptoms, disturbed sexuality,
impact on QoL, and psychological implications including depression and more failure anticipation thoughts,
[108-110, 208-210]. Sexual dysfunction heightens anger, frustration and depression, all of which place a
strain on the patients’ relationships. The female partners of men with sexual dysfunction and depression
often present with similar symptoms including pain upon intercourse and depressive symptoms. Men with
CPPPS have reported a high frequency of sexual relationship dissolution and psychological symptoms, such
as depression and suicidal thinking [108, 211]. Primary Prostate Pain Syndrome patients reported substantial
sexual and relationship problems [108, 211]. On the other hand, it was found that men with PPPS did not report
significantly decreased sexual satisfaction compared to controls [212]. There is consensus that therapeutic
strategies reducing symptoms of pelvic pain are of relevance in relation to changes in sexual function. Also
intimacy and having sex can yield positive experiences that will reduce the pain. The CNS plays an important
role in this mechanism.

Women
Chronic pelvic pain leads to substantial impairment in QoL and several sexual dysfunctions [120, 213-215]. It
seems reasonable to expect that pain, extreme fatigue, depressive mood and pain drugs will affect women’s
sexuality. Women with CPPPS reported significantly more pain, depression, and anxiety symptoms and were
physically more impaired than women in the control group. In comparison with controls, women with CPPPS
reported significantly more sexual avoidance behaviour, non-sensuality, and complaints of “vaginismus”
[216]. Patients with CPPPS reported more sexual problems than women with any other type of chronic pain
problem [217]. The quality of intimate relationships is closely connected with sexual function [218]. Satisfaction
with sexual relationships appears to be associated with higher marital functioning [219]. In addition sexual
dissatisfaction is related to sexual dysfunction. When one partner suffers from chronic pain, the ability of both
partners to cope with the pain and the extent to which partners are supportive of the chronic pain sufferer have
been found to be a predictor of sexual functioning [219].

Approximately two-thirds of patients in another study reported reduced frequency in their sexual relations as
a result of CPPPS [220]. One study demonstrated that CPPPS patients reported worse sexual function with
regard to desire, arousal, lubrication, orgasm, satisfaction, and more frequent and severe pain with vaginal
penetration than women without CPPPS [221]. In an interview with 50 chronic pain sufferers and their spouses,
78% of the pain sufferers and 84% of partners described deterioration, including cessation of their sex life
[222]. In a study in patients with back pain, half reported decreased frequency of sex since the onset of
chronic pain [124]. The Female Sexual Function Index (FSFI) has been developed as a brief, multi-dimensional
self-report instrument for assessing the key dimensions of sexual function in women, which includes desire,
subjective arousal, lubrication, orgasm, satisfaction, and pain. Using the FSFI, women with CPPPS reported
worse sexual function in all subscales and total score than women without CPPPS. The largest differences
between women with CPPPS and without CPPPS were seen for the domains of pain and arousal. The total
score and the subscales of the FSFI had high levels of internal consistency and test-retest reliability when
assessed in a sample of women with CPPPS. The FSFI also showed good ability to discriminate between
women with and without CPPPS [221].

Myofascial pain
Myalgia is too often overlooked as a form of chronic pelvic pain. The pelvic floor and adjacent muscles
are used in an abnormal way. Studies in the field of chronic prostatitis support the idea that patients with
CPPPS have more muscle spasm and increased muscle tone and report pain when the pelvic floor muscles
are palpated [223]. Learning pelvic floor muscle relaxation can diminish spasm and pain [224]. Repeated or
chronic muscular overload can activate trigger points in the pelvic floor muscles. A report from the Chronic
Prostatitis Cohort Study showed that 51% of patients with prostatitis and only 7% of controls had any muscle
tenderness. Tenderness in the pelvic floor muscles was only found in the CPPPS group [129].

The first ideas about the neurological aspects of the pelvic floor muscles in relation to chronic pelvic pain were
published in 1999. The possibility of CNS changes in the regulation of pelvic floor function was suggested as
a mechanism for development of CPPPS. Of the patients presenting with pelvic pain, 88% had poor to absent
pelvic floor muscle function [128]. Animal studies on the role of neurogenic inflammation have also elucidated
some important phenomena. Irritation of the prostate, bladder and pelvic floor muscles results in expression
of C-fos-positive cells in the CNS. There appears to be convergence of afferent information onto central
pathways. Once the central changes have become established, they become independent of the peripheral
input that initiated them [225].

26 CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021


Repeated or chronic muscular overload can activate trigger points in the muscle. Trigger points are defined as
hyper-irritable spots within a taut band. Other criteria for trigger points are recognition of the pain as ‘familiar’,
and pain on stretching the muscle. Apart from pain, trigger points prevent full lengthening of the muscle,
thereby restricting the range of movement. Pain as a result of these trigger points is aggravated by specific
movements and alleviated by certain positions. Positions and movements in which the shortened muscle is
stretched are painful. Patients know which activities and postures influence pain. Trigger points can be located
within the pelvic floor muscles and in adjacent muscles such as the abdominal, gluteal and iliopsoas muscles.
Pain is aggravated by pressure on the trigger point (e.g., pain related to sexual intercourse). Pain also worsens
after sustained or repeated contractions of pelvic floor muscles (e.g., pain related to voiding or defecation).

3.3 Abdominal aspects of pelvic pain


3.3.1 Incidence
Epidemiological data on IBS and CPPPS are scarce [226]. Chronic Pelvic Pain has been shown to be one of
the most common functional disorders in women of reproductive age. The monthly incidence rate of CPPPS
published by Zondervan et al. was 1.58/1000 [227].

3.3.2 Prevalence
Using a vague definition of continuous or episodic pain situated below the umbilicus over six months, one
study reported that CPPPS was one of the most common diagnoses in primary care units in Great Britain [227].
The monthly prevalence rate of CPPPS in this study was 21.5/1,000, with an annual prevalence of 38.3/1,000.
The prevalence rates increase significantly with older age and vary significantly between regions in the UK.
The overall prevalence of anorectal pain in a sample of USA householders was 6.6% and was more common
in women [228]. Irritable bowel syndrome is associated with common gynaecologic problems (endometriosis,
dyspareunia, and dysmenorrhoea) [229]. Fifty per cent of women who presented with abdominal pain to
the gynaecologic clinic or were scheduled for laparoscopy due to CPPPS had symptoms of IBS [230]. In a
survey from Olmsted county 20% of women reported CPPPS and 40% of those met criteria for IBS [22]. This
overlap of CPPPS and IBS was associated with an increased incidence of somatisation. Not gynaecological
surgical procedures but only psychosocial variables predict pain development without a different incidence
of IBS in a prospective and controlled study [231]. Clinical features of pelvic floor dysfunction, gynaecological
and psychological features are related to disordered anorectal function in IBS patients but do not predict
physiological anorectal testing.

3.3.3 Influence on QOL


There is little known on health related quality of life (HRQoL) in patients with CPPPS. There is a need to develop
validated disease specific HRQoL instruments for CPPPS in addition to sound measurement properties. More
data are available in patients with IBS treated at referral centres who have comparable HRQoL scores as
patients with other common disorders such as diabetes, end-stage renal disease, and inflammatory bowel
disease [232]. Sub-groups of IBS with predominance of diarrhoea or constipation show no difference in
HRQoL. Multi-variate analysis shows that HRQoL in patients with IBS is affected by sex and psychological
conditions.

3.3.4 Costs
Costs combine direct health-care costs and societal costs (productivity loss) such as under-performance
and absenteeism from work. The annual costs to society can be calculated by using the average population
earnings. In Germany direct care costs are estimated at €791 and societal costs €995 per patient with IBS per
year which may be comparable to patients with CPPPS [233].

3.3.5 Risk factors & underlying causes


Risk factors are covered in Section 3.1.5.

3.4 Summary of evidence and recommendations: CPPPS and mechanisms

Summary of evidence LE
CPPPS mechanisms are well defined and involve mechanisms of neuroplasticity and neuropathic pain. 2
The mechanisms of neuroplasticity and neuropathic pain result in increased perception of afferent 1
stimuli which may produce abnormal sensations as well as pain.
End-organ function can also be altered by the mechanisms of neuroplasticity so that symptoms of 1
function can also occur.
The diagnosis of a CPPPS as a pain syndrome is essential as it encourages a holistic approach to 2
management with multi-specialty and multi-disciplinary care.

CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021 27


Recommendations Strength rating
All of those involved in the management of chronic pelvic pain should have knowledge of Strong
peripheral and central pain mechanisms.
The early assessment of patients with chronic pelvic pain should involve investigations Strong
aimed at excluding disease-associated pelvic pain.
Assess functional, emotional, behavioural, sexual and other quality of life issues, such as Strong
effect on work and socialisation, early in patients with chronic pelvic pain and address these
issues as well as the pain.
Build up relations with colleagues so as to be able to manage chronic pelvic primary pain Strong
syndrome comprehensively in a multi-specialty and multi-disciplinary environment with
consideration of all their symptoms.

4. DIAGNOSTIC EVALUATION
4.1 General Evaluation
4.1.1 History
History is very important for the evaluation of patients with chronic pelvic pain. Pain syndromes are
symptomatic diagnoses, which are derived from a history of pain perceived in the region of the pelvis, and
absence of other pathology, for a minimum of three months. This implies that specific disease-associated
pelvic pain caused by bacterial infection, cancer, drug-induced pathology (e.g., ketamine use) [234], primary
anatomical or functional disease of the pelvic organs, and neurogenic disease must be ruled out.

4.1.1.1 Anxiety, depression, and overall function


Distress is best understood in the context of pain and of the meaning of pain to the individual and is
best assessed ideographically rather than normatively. Almost all diagnostic measures and standardised
instruments of anxiety and depression are designed for people without significant physical problems, so are
difficult to interpret in chronic pelvic pain [235].

Anxiety about pain often refers to fears of missed pathology (particularly cancer) as the cause of pain [34], or
to uncertainties about treatment and prognosis. These can drive healthcare seeking behaviour. The question:
“What do you believe or fear is the cause of your pain?” has been suggested [236]. Anxiety may also concern
urinary urgency and frequency that are problematic in social settings.

Depression or depressed mood are common in chronic pain [237] e.g., often related to losses consequent
to chronic pain (work, leisure activities, social relationships, etc.). Due to the lack of suitable assessment
instruments, it is better to ask a simple question such as “How does the pain affect you emotionally?” If the
answer gives cause for concern about the patient’s emotional state, further assessment should be undertaken
by an appropriately qualified colleague.

Most measures of restricted function are designed primarily for musculoskeletal pain and may emphasise mobility
problems rather than the difficulties of the individual with pelvic or urogenital pain. A promising specific measure,
UPOINT, was introduced and in a later version the sexological aspects were added [238]. However, it may under-
assess relevant psychological variables [43]. Generic QoL measures are helpful. If such an instrument is not already
used in the clinic, the Brief Pain Inventory [239] provides a broad and economical assessment of interference of
pain with various aspects of life in multiple languages. (For further suggested instruments see [240]). In a study,
more pain, pain-contingent rest, and urinary symptoms were associated with poorer function [58].

4.1.1.2 Urological aspects


Pain may be associated with urological symptoms. A detailed history of lower urinary tract functions should
be taken. Dysfunctions of the lower urinary tract may exacerbate symptoms, as pain may interfere with the
function of the lower urinary tract. Micturition in all its aspects should be addressed. Special attention should
be paid to the influence of micturition on the experience of pain.

Primary prostate pain syndrome


Primary prostate pain syndrome is diagnosed from a history of pain perceived in the region of the prostate
(convincingly reproduced by prostate palpation), and absence of other lower urinary tract pathology, for
a minimum of three months. As mentioned above, specific disease-associated pelvic pain must be ruled

28 CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021


out. A thorough history is an important first step in the evaluation of PPPS. It should include type of pain
and localisation. Pain is often reported in other pelvic areas outside the prostate such as perineum, rectum,
penis, testicles and abdomen [50]. In addition, associated lower urinary tract symptoms, sexual function,
psychological, social and economic factors should be addressed. Determination of the severity of disease,
its progression and treatment response can be assessed only by means of a validated symptom-scoring
instrument (see Section 4.2.3). These subjective outcome measures are recommended for the basic evaluation
and therapeutic monitoring of patients in urological practice.

Primary bladder pain syndrome


Primary bladder pain syndrome should be diagnosed on the basis of pain, pressure or discomfort associated
with the urinary bladder, accompanied by at least one other symptom, such as daytime and/or night-time
increased urinary frequency, the exclusion of confusable diseases as the cause of symptoms, and if indicated,
cystoscopy with hydrodistension and biopsy (Table 4) [15].

The nature of pain is key to disease definition:


1. pain, pressure or discomfort perceived to be related to the bladder, increasing with increasing bladder
content;
2. located suprapubically, sometimes radiating to the groins, vagina, rectum or sacrum;
3. relieved by voiding but soon returns [241, 242];
4. aggravated by food or drink [242].
Primary bladder pain syndrome type 3 can lead to a small capacity fibrotic bladder with or without upper
urinary tract outflow obstruction.

4.1.1.3 Gynaecological aspects


A detailed medical history outlining the nature, frequency and site of pain; its relationship to precipitating
factors and the menstrual cycle, may help define the aetiology. A menstrual and sexual history, including a
history of sexually transmitted diseases, vaginal discharge, as well as previous sexual trauma is mandatory
as well as up to date cervical cancer screening. A history of obstetric and/or gynaecological surgery is also
warranted, particularly if devices such as synthetic mesh were used.

4.1.1.4 Gastrointestinal aspects


The predominant symptoms that patients are interviewed about are discomfort or pain in relation to their
bowel habits, daily activities, and eating. A precise history of dysfunctional voiding or defecation should be
asked, ideally applying symptom questionnaires for urinary and anorectal symptoms (e.g., Rome III criteria
for anorectal pain). Excessive straining at most defecations, anal digitations in dyssynergic defecation, and a
sensation of anal blockage may be found in patients with chronic anal pain. History of anxiety and depression
with impaired QoL is often encountered in anorectal functional disorders and should be evaluated.

Diagnostic criteria for primary chronic anal pain syndrome (chronic proctalgia) according to the Rome III criteria
are as follows and must include all of the following: chronic or recurrent rectal pain or aching, episodes last at
least 20 minutes and exclusion of other causes of rectal pain such as ischaemia, inflammatory bowel disease,
cryptitis, intramuscular abscess and fissure, haemorrhoids, prostatitis, and Coccyx Pain Syndrome. These criteria
should be fulfilled for the past three months with symptom onset at least six months before diagnosis [243, 244].

The primary chronic anal pain syndrome includes the above diagnostic criteria and exhibits exquisite
tenderness during posterior traction on the puborectalis muscle (previously called “Levator Ani Syndrome”).
Pathophysiology of pain is thought to be due to over-activity of the pelvic floor muscles.

Primary intermittent chronic anal pain syndrome (proctalgia fugax) consists of all the following diagnostic
criteria, which should be fulfilled for three months: recurrent episodes of pain localised to the anus or lower
rectum, episodes last from several seconds to minutes and there is no anorectal pain between episodes.
Stressful life events or anxiety may precede the onset of the intermittent chronic anal pain syndrome. The
attacks may last from a few seconds to as long as 30 minutes. The pain may be cramping, aching or stabbing
and may become unbearable. However, most patients do not report it to their physicians and pain attacks
occur less than five times a year in 51% of patients.

4.1.1.5 Peripheral nerve aspects


A proportion of patients will be able to relate the onset of pain to an acute event such as surgery, sepsis or
trauma, and occasionally, cycling for a prolonged period. Chronic injury is more frequent, such as associated
with sitting for prolonged periods over time. Many will be idiopathic.

CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021 29


The pain is classically perceived in the perineum from anus to clitoris/penis. However, less-specific pain
distribution may occur, and this may be due to anatomical variation, involvement of branches of the nerve
rather that the main nerve, CNS central sensitisation, and consequently, the involvement of other organs
and systems in a regional pain syndrome. Other nerves in the vicinity may also be involved, for example,
inferior cluneal nerve and perineal branches of the posterior femoral cutaneous nerve. The musculoskeletal
system may become involved, confusing the pain picture as aches and pain developing in the muscles due to
immobility and disability, possibly magnified by the CNS changes.

Burning is the most predominant adjective used to describe the pain. Crushing and electric may also be used,
indicating the two components - a constant pain often associated with acute sharp episodes. Many patients
may have the feeling of a swelling or foreign body in the rectum or perineum, often described as a golf or
tennis ball. The term pain has different meanings to patients and some would rather use the term discomfort or
numbness.

Aggravating factors include any type of pressure being applied, either directly to the nerve or indirectly to
other tissue, resulting in pudendal traction. Allodynia is pain on light touch due to involvement of the CNS, and
may make sexual contact and the wearing of clothes difficult. These patients often remain standing, and as a
consequence, develop a wide range of other aches and pains. Soft seats are often less well-tolerated, whereas
sitting on a toilet seat is said to be much better tolerated. If unilateral, sitting on one buttock is common. The
pain may be exacerbated by bowel or bladder evacuation.

Pudendal nerve damage may be associated with a range of sensory phenomena. In the distribution of the
nerve itself, as well as unprovoked pain; the patient may have paraesthesia (pins and needles); dysaesthesia
(unpleasant sensory perceptions usually but not necessarily secondary to provocation, such as the sensation
of running cold water); allodynia (pain on light touch); or hyperalgesia (increased pain perception following a
painful stimulus, including hot and cold stimuli). Similar sensory abnormalities may be found outside of the area
innervated by the damaged nerve, particularly for visceral and striated muscle hyperalgesia.

The cutaneous sensory dysfunction may be associated with superficial dyspareunia, but also irritation and
pain associated with clothes brushing the skin. There may also be a lack of sensation and pain may occur in
the presence of numbness. Visceral hypersensitivity may result in an urge to defecate or urinate. This is usually
associated with voiding frequency, with small amounts of urine being passed. Pain on visceral filling may occur.
Anal pain and loss of motor control may result in poor bowel activity, with constipation and/or incontinence.
Ejaculation and orgasm may also be painful or reduced.

Many of those suffering from pudendal neuralgia complain of fatigue and generalised muscle cramps,
weakness and pain. Being unable to sit is a major disability, and over time, patients struggle to stand and they
often become bedbound. The immobility produces generalised muscle wasting, and minimal activity hurts. As
a consequence of the widespread pain and disability, patients often have emotional problems, and in particular,
depression. Patients with chronic pelvic pain are also often anxious and have the tendency to catastrophise.
Depression, catastrophising and disability are all poor prognostic markers. Cutaneous colour may change
due to changes in innervation but also because of neurogenic oedema. The patient may describe the area as
swollen due to this oedema, but also due to the lack of afferent perception.

4.1.1.6 Myofascial aspects


When taking a history from a patient with pelvic pain, it is important to address the function of all the organs
in the pelvic area. The following items certainly should be addressed: lower urinary tract function, anorectal
function, sexual function, gynaecological items, presence of pain and psychosocial aspects. One cannot state
that there is a pelvic floor dysfunction based only on the history. But there is a suspicion of pelvic floor muscle
dysfunction when two or more pelvic organs show dysfunction, for instance a combination of micturition and
defecation problems.

4.1.2 Physical Evaluation


The clinical examination often serves to confirm or refute the initial impressions gained from a good history.
The examination should be aimed at specific questions where the outcome of the examination may change
management. Prior to an examination, best practice requires the medical practitioner to explain what will
happen and what the aims of the examination are to the patient. Consent to the examination should occur
during that discussion and should cover an explanation around the aim to maintain modesty as appropriate
and, if necessary, why there is a need for rectal and/or vaginal examination. Finally, the risk of exacerbating
the pain should form a part of that request. A record of the discussion should be noted. The possibility of

30 CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021


the presence of a chaperone should be discussed with the patient. As well as a local examination, a general
musculoskeletal and neurological examination should be considered an integral part of the assessment and
undertaken. Following the examination, it is good practice to ask the patient if they had any concerns relating
to the conduct of the examination and that discussion should be noted.

There is no specific diagnostic test for CPPPS, therefore, procedures are on the one hand directed towards
identification and exclusion of specific diseases associated with pelvic pain, and on the other hand may be used
for phenotypic description. Abdominal and pelvic examination to exclude gross pelvic pathology, as well as to
demonstrate the site of tenderness is essential. It is important to look for abnormalities in muscle function.

Examination of the external genitalia is a part of the evaluation. In patients with scrotal pain, gentle palpation
of each component of the scrotum is performed to search for masses and painful spots. The penis and urethra
may be palpated in a similar way. Many authors recommend that one should assess cutaneous allodynia
along the dermatomes of the abdomen (T11-L1) and the perineum (S3), and the degree of tenderness should
be recorded. The bulbocavernosus reflex in the male may also provide useful information concerning the
intactness of the pudendal nerves. Clinical pelvic examination should be a single digit examination if possible.
The usual bi-manual examination can generate severe pain so the examiner must proceed with caution. A
rectal examination is done to look for prostate abnormalities in male patients including pain on palpation and to
examine the rectum and the pelvic floor muscles regarding muscle tenderness and trigger points as well as the
ability to contract and relax these muscles.

At clinical examination, perianal dermatitis may be found as a sign of faecal incontinence or diarrhoea. Fissures
may be easily overlooked and should be searched for thoroughly in patients with anal pain. A rectal digital
examination findings may show high or low anal sphincter resting pressure, a tender puborectalis muscle in
patients with the Levator Ani Syndrome, and occasionally increased perineal descent. The tenderness during
posterior traction on the puborectalis muscle differentiates between Levator Ani Syndrome and unspecified
Functional Anorectal Pain and is used in most studies as the main inclusion criterion. Dyssynergic (paradoxical)
contraction of the pelvic muscles when instructed to strain during defecation is a frequent finding in patients
with pelvic pain. Attention should be paid to anal or rectal prolapse at straining, and ideally during combined
rectal and vaginal examination to diagnose pelvic organ prolapse.

A full clinical examination of the musculo-skeletal, nervous and urogenital systems is necessary to aid in
diagnosis of pudendal neuralgia, especially to detect signs indicating another pathology. Often, there is little
to find in pudendal neuralgia and frequently findings are non-specific. The main pathognomonic features are
the signs of nerve injury in the appropriate neurological distribution, for example, allodynia or numbness.
Tenderness in response to pressure over the pudendal nerve may aid the clinical diagnosis. This may be
elicited by per rectal or per vaginal examination and palpation in the region of the ischial spine and/or Alcock’s
canal. Muscle tenderness and the presence of trigger points in the muscles may confuse the picture. Trigger
points may be present in a range of muscles, both within the pelvis (levator ani and obturator internus muscles)
or externally (e.g., the piriformis, adductors, rectus abdominis or paraspinal muscles).

4.2 Supplemental evaluation


If history is suggestive of lower urinary tract, gynaecological, anorectal or other disease of known aetiology,
diagnostic work-up should follow respective guidelines.

4.2.1 Assessing pelvic pain and related symptoms


Determination of the severity of pain and associated symptoms, its progression and treatment response can
be assessed only by means of a reliable and validated symptom-scoring instrument. These subjective outcome
measures are recommended for the basic evaluation and therapeutic monitoring of patients. Pain should
always be assessed at presentation and (see below) to identify progression and treatment response. As well as
doing this in the clinic, the patient can keep a daily record (pain diary). This may need to include other relevant
variables such as voiding, sexual activity, activity levels, or analgesic use.

Increased attention to patient reported outcomes gives prominence to patients’ views on their disease and pain
diaries, in patients’ own environments, improve data quality.

Quality of life should also be measured because it can be very poor compared to other chronic diseases [245,
246]. In a study more pain, pain-contingent rest, and urinary symptoms were associated with greater disability
(also measured by self-report), and pain was predicted by depression and by catastrophising (helplessness
subscale) [58].

CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021 31


Where the primary outcome of treatment is pain relief, it is useful before starting treatment to agree a clinically
useful level of relief [247]. The most reliable methods are:
• a five point verbal scale: none, mild, moderate, severe, very severe pain;
• a VAS score from one to ten;
• an eleven point numerical scale (as below).

An 11 point numerical scale

Pain assessment ratings are not independent of cognitive and emotional variables [57]. Target outcomes of
pain severity, distress and disability co-vary only partly, and improvement in one does not necessarily imply
improvement in the others. When the primary outcome is pain its meaning should be anchored in discussion of
clinically important difference [247].

Primary prostate pain syndrome


Reliable, valid indices of symptoms and QoL are the NIH-CPSI [248] and the International Prostate Symptom
Score (I-PSS) [249].

Primary bladder pain syndrome


Symptom scores may help to assess the patient and act as outcome measures. The O’Leary-Sant Symptom
Index, also known as the Interstitial Cystitis Symptom Index (ICSI) was validated in a large study [250].

Gastrointestinal questionnaire
Functional anorectal pain disorders (anorectal pelvic pain) are defined and characterised by duration,
frequency, and quality of pain. More complex questionnaires are used in the setting of IBS. The validated IBS-
Symptom Severity Scale (IBS-SSS) includes the broadest measurement of pain-related aspects [251, 252].
However, as different instruments measure different endpoints of chronic abdominal pain in IBS, a comparison
of published studies is often impossible.

Sexual function assessment


In males the most frequent effects on sexual function are ED and premature ejaculation. These can be
evaluated by proper questionnaires namely IIEF and PEDT (Premature Ejaculation Diagnostic Tool). In
comparison with controls, women with chronic pelvic pain reported significantly more sexual avoidance
behaviour, non-sensuality, and complaints of “vaginismus” [205]. The Female Sexual Function Index (FSFI) has
been developed as a brief, multi-dimensional self-report instrument for assessing the key dimensions of sexual
function in women, which includes desire, subjective arousal, lubrication, orgasm, satisfaction, and pain. The
corresponding evidence in men is lacking.

4.2.2 Focused myofascial evaluation


Pelvic floor muscle testing can be done by the medical doctor but a consultation of the pelvic floor by a
physiotherapist is a good alternative, but either should have had appropriate training in pelvic assessment. A
vaginal or rectal examination is performed to assess the function of the pelvic floor muscles, according to the
International Continence Society (ICS) report. This assessment has been tested and shows satisfactory validity
and intra-observer reliability. It can therefore be considered suitable for use in clinical practice [253]. Rectal
examination is a good way to test the pelvic floor function in men [254]. There is a growing number of reports
on the use of ultrasound (US) in establishing the function of the pelvic floor muscles. The exact place in the
diagnostic setting needs to be addressed in the future [255]. In a cohort study of 72 men with chronic pelvic
pain, the relationship between the locations of the trigger point and the referred pain was examined. Ninety
percent of the patients showed tenderness in the puborectalis muscle and 55% in the abdominal wall muscles.
Of the patients in whom trigger points were found in the puborectalis, 93% reported pain in the penis and
57% in the suprapubic region. Patients with trigger points in the abdominal muscles reported pain in the penis
(74%), perineum (65%) and rectum (46%) [256]. In addition, a broad musculoskeletal (tender point) evaluation,
including muscles outside the pelvis, helps to diagnose the myofascial pain aspects of the pelvic pain in
phenotyping pelvic pain patients [257, 258].

4.2.3 Neurological

Injections
An injection of local anaesthetic and steroid at the site of nerve injury may be diagnostic. Differential block

32 CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021


of the pudendal nerve helps to provide information in relation to the site where the nerve may be trapped
[259, 260]. Infiltration at the ischial spine requires the use of a nerve stimulator/locator. Both motor (anal
contraction) and sensory endpoints may be noted. The anatomical target may be localised by fluoroscopy,
computed tomography (CT) guidance, or the use of US. Ultrasound avoids any form of radiation, whereas CT
guidance involves a significant amount of radiation. Currently, fluoroscopy is probably the imaging technique
most frequently used because it is readily available to most anaesthetists that perform the block. Currently,
infiltration of the pudendal nerve within Alcock’s canal is primarily undertaken with the use of CT. As well as
injecting around the pudendal nerve, specific blocks of other nerves arising from the pelvis may be performed.

Electrophysiological studies
These may reveal signs of perineal denervation, increased pudendal nerve latency, or impaired
bulbocavernosus reflex [196, 199, 261-263]. However, for an abnormality to be detected, significant nerve
damage is probably necessary. Pain may be associated with limited nerve damage, therefore, these
investigations are often normal.

4.2.4 Imaging
Ancillary studies should be performed according to appropriate guidelines for exclusion of diseases with known
aetiology presenting with symptoms identical to those of CPPS. Once the latter diagnosis is established,
studies can be useful to assess functional abnormalities and phenotype conditions such as PBPS, and primary
chronic anal pain syndrome.

Ultrasound
Ultrasound has limited value but may reassure patients. However, over-investigating may be detrimental.

MRI
Magnetic resonance neurography has been increasingly used in specialised centres for the diagnosis of the
location (proximal vs. peripheral) and degree (total vs. partial) of nerve injury in the peripheral nervous system,
earlier and with higher specificity than conduction studies. This may show benefits for CPPPS in the coming
years.

MR defecating proctogram
Magnetic resonance imaging in conjunction with MR defecography has become the most valuable imaging
technique to assess anorectal function dynamically. Magnetic resonance imaging studies simultaneously
outline the anatomy of the pelvic floor and visualise different structural and functional pathologies, by applying
dynamic sequences after filling of the rectum with a viscous contrast medium (e.g., US gel). The following
pathologies can be visualised: pelvic floor descent, an abnormal anorectal angle while squeezing and straining,
rectal intussusception, rectocele, enterocele and cystocele. However, limitations of MR defecography are the
left lateral position and the limited space for the patient, which may reduce the ability to strain and thereby
reduce the sensitivity of the method, underestimating the size of entero- and rectoceles as well as the amount
of intussusception.

Functional neuroimaging
Functional neuroimaging, functional magnetic resonance imaging (fMRI) is currently being re-evaluated as a
research tool and some groups have raised issues around over interpretation [264]. With regards to pain, fMRI
findings may represent a pain matrix or may represent non-specific threat processing [265]. Currently this panel
cannot recommend fMRI as a clinical tool.

4.2.5 Laboratory Tests

Microbiology tests
Primary prostate pain syndrome
Laboratory diagnosis of prostatitis has been classically based on the four-glass test for bacterial localisation
[266]. Besides sterile pre-massage urine (voided bladder urine-2), PPPS shows < 103 cfu/mL of uropathogenic
bacteria in expressed prostatic secretions and insignificant numbers of leukocytes or bacterial growth in
ejaculates. However, this test is too complex for use by practising urologists. Diagnostic efficiency may be
enhanced cost-effectively by a simple screening procedure, that is, the two-glass test or pre-post-massage
test (PPMT) [267, 268]. Overall, these tests help only a little in the diagnosis of PPPS, because 8% of patients
with suggested PPPS have been found to have positive prostatic localisation cultures, similar to the percentage
of asymptomatic men [269].

CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021 33


Primary bladder pain syndrome
Urine dipstick and urine culture (including culture for Tuberculosis if sterile pyuria) are recommended in all
patients suspected of having PBPS. Urine cytology is also recommended in risk groups.

Gynaecological aspects of chronic pelvic pain


Vaginal and endocervical swabs to exclude infection are recommended. In specific cases, imaging may be
required to help rule out a defined pathology such as sacral neuropathy in endometriosis [270].

4.2.6 Invasive tests

Anorectal pain
Anorectal manometry with sensory testing (pressure volume measurement: barostat) may be useful to diagnose
dyssynergic defecation and hypersensitivity of the rectum which are typical for patients with CPPPS and IBS.
Flexible rectosigmoidoscopy or colonoscopy should be considered in patients with anorectal pain to rule out
coincidental colorectal pathology.

Laparoscopy for females


Laparoscopy is perhaps the most useful invasive investigation to exclude gynaecological pathology [271, 272]
and to assist in the differential diagnosis of CPPPS in women [273]. Often, it is combined with cystoscopy [274,
275] and/or proctoscopy to help identify the site of multi-compartment pain.

Psychological considerations around laparoscopy


Three very different studies of laparoscopy suggest that it can improve pain through resolving concerns about
serious disease [276], although showing women the photograph of their pelvic contents did not improve pain
on explanation alone [277]. Integrating somatic and psychological assessment from the start rather than
dealing with psychological concerns only after excluding organic causes of pelvic pain is helpful [278].

Cystoscopy and bladder biopsy


Despite controversy on the diagnostic and follow-up value of cystoscopy in PBPS [279-283], the panel
believes that objective findings are important for diagnosis, prognosis and ruling out other treatable conditions
(a standardised scheme of diagnostic criteria will also contribute to uniformity and comparability of different
studies) [284]. Endoscopically, PBPS type 3 displays reddened mucosal areas often associated with small
vessels radiating towards a central scar, sometimes covered by a small clot or fibrin deposit - the Hunner's
lesion [241]. The scar ruptures with increasing bladder distension, producing a characteristic waterfall type of
bleeding. There is a strong association between PBPS type 3 and reduced bladder capacity under anaesthesia
[285]. Non-lesion disease displays a normal bladder mucosa at initial cystoscopy. The development of
glomerulations after hydrodistension is considered to be a positive diagnostic sign although they can be
observed without PBPS [286]. Biopsies are helpful in establishing or supporting the clinical diagnosis of both
classic and non-lesion types of the disease [140, 166, 284, 287, 288]. Important differential diagnoses to
exclude, by histological examination, are carcinoma in situ and tuberculous cystitis.

Table 4: E
 SSIC classification of PBPS types according to results of cystoscopy with hydrodistension and
biopsies [15]

Cystoscopy with hydrodistension


Not done Normal Glomerulationsa Hunner's lesionb
Biopsy
Not done XX 1X 2X 3X
Normal XA 1A 2A 3A
Inconclusive XB 1B 2B 3B
Positivec XC 1C 2C 3C
aCystoscopy: glomerulations grade 2-3.
bLesion per Fall’s definition with/without glomerulations.
cHistology showing inflammatory infiltrates and/or detrusor mastocytosis and/or granulation tissue and/or

intrafascicular fibrosis.

34 CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021


4.3 Diagnostic algorithm

Figure 1: Diagnosing chronic pelvic pain

Chronic Pelvic Pain

Physical
History
examinaon

Chronic secondary
yes pelvic pain
Symptom of a well
known disease
no
Chronic primary
pelvic pain syndrome

Organ specific
symptoms present

yes

Urology Gynaecology Gastro- Neurology Sexology Pelvic


enterology floor

Phenotype and proceed according to Chronic Pelvic Pain Guideline.

CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021 35


Figure 2: Phenotyping of pelvic pain - UPOINT classification

Phenotyping Assessment

Urology Urinary flow, micturi on diary, cystoscopy, ultrasound, uroflowmetry.

Psychology Anxiety about pain, depression and loss of func on, history of nega ve sexual experiences.

Ask for gynaecological, gastro-intes nal, ano-rectal, sexological complaints.


Organ specific
Gynaecological examina on, rectal examina on.

Infec on Semen culture and urine culture, vaginal swab, stool culture.

Ask for neurological complaints (sensory loss, dysaesthesia).


Neurological
Neurological tes ng during physical examina on: sensory problems, sacral reflexes and muscular func on.

Tender muscle Palpa on of the pelvic floor muscles, the abdominal muscles and the gluteal muscles.

Sexological Erec le func on, ejaculatory func on, post-orgasmic pain.

4.4 Other painful conditions without a urological cause

Dysmenorrhoea
Menstrual pain or ‘dysmenorrhoea’ may be primary or secondary. Primary dysmenorrhoea classically begins
at the onset of ovulatory menstrual cycles and tends to decrease following childbirth [273]. Secondary
dysmenorrhoea suggests the development of a pathological process, such as endometriosis [272],
adenomyosis or pelvic infection, which need to be excluded.

Infection
In pre-menopausal women, a history of Pelvic Inflammatory Disease (PID) must be excluded. A patient’s sexual
history should be taken along with swabs to exclude chlamydia and gonorrhoea infection. Bacterial and viral
genital tract pathogens should also be excluded [289], as they can cause severe pelvic/vaginal/vulvar pain
[290] and are associated with ulcerating lesions and inflammation, which may lead to urinary retention [291].
If there is any doubt about the diagnosis, laparoscopy may be helpful, as one of the differential diagnoses is
endometriosis.

Endometriosis and adenomyosis


The incidence of endometriosis is rising in the developed world. It has widespread impact on women’s lives
[292], with pain more important than physical findings in determining QoL [293]. The precise aetiology is
unknown, but an association with infertility is recognised [294]. A diagnosis is usually made when a history of
secondary dysmenorrhoea and/or dyspareunia exists. On examination, there is often tenderness in the lateral
vaginal fornices, reduced uterine mobility, tenderness in the recto-vaginal septum, and on occasion, adnexal
masses. Laparoscopy is the most useful diagnostic tool [295-298]. Adenomyosis is associated with augmented
pain during menses [299]. It is diagnosed by an US scan of the uterus, which often shows cystic dilatation of
the myometrium [300].

Gynaecological malignancy
The spread of gynaecological malignancy of the cervix, uterine body or ovary will cause pelvic pain depending
on the site of spread.

Injuries related to childbirth


Trauma occurring at the time of childbirth may lead to chronic pelvic pain related to the site of injury [298]. Female
sexual dysfunction is perhaps the commonest presenting problem [301], though increasingly women are reporting
other symptoms such as pelvic girdle pain and other genito-pelvic pain of different aetiology [302]. There is often
a transient problem with oestrogen deficiency in the post-partum period and during breastfeeding, which can
compound this situation. Denervation of the pelvic floor can similarly compound the situation [303].

36 CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021


Pain associated with pelvic organ prolapse and prolapse surgery
Pelvic organ prolapse is often asymptomatic, unless it is so marked that it causes back pain, vaginal pain
and skin excoriation [304]. Prolapse is often a disease of older women, and it is often associated with post-
menopausal oestrogen deficiency, which may lead to pain associated with intercourse. Prolapse surgery has
entailed the use of non-absorbable mesh (usually in the form of “mesh kits”). Although they may have a role in
supporting the vagina, they are also associated with several complications including bladder, bowel and vaginal
trauma [305] and neuropathy [306]. Patients need to be fully evaluated and may need specialised imaging,
using contrast mediums if necessary, to make a diagnosis of the possible cause of the pain [307-310].

Haemorrhoids
Chronic pelvic pain is rare in haemorrhoidal disease because endoscopic and surgical treatment is mostly
effective in acute disease. The most frequent aetiology of pain without significant bleeding is thrombosed
external haemorrhoids or an anal fissure. Haemorrhoidal pain on defecation associated with bleeding is usually
due to prolapse or ulceration of internal haemorrhoids. Anaemia from haemorrhoidal bleeding is rare but may
arise in patients on anti-coagulation therapy, or those with clotting disorders.

Anal fissure
Anal fissures are tears in the distal anal canal and induce pain during and after defecation. The pain can last
for several minutes to hours. Persistence of symptoms beyond six weeks or visible transversal anal sphincter
fibres define chronicity. Fissures located off the midline are often associated with specific diseases such as
Crohn’s disease or anal cancer. Internal anal sphincter spasms and ischaemia are associated with chronic
fissures.

Proctitis
Abdominal and pelvic pain in patients with inflammatory bowel disease and proctitis are often difficult to
interpret. Faecal calprotectin may help to differentiate between inflammation and functional pain, to spare
steroids.

Irritable bowel syndrome


Although IBS can be associated with pelvic pain, the panel consider a full discussion of this topic beyond the
scope of these guidelines. A number of high quality clinical guidelines address this topic [243, 311].

4.5 Summary of evidence and recommendations: diagnostic evaluation

4.5.1 Diagnostic evaluation - general

Summary of evidence LE
Clinical history and examination are mandatory when making a diagnosis. 2a

Recommendation Strength rating


Take a full history and evaluate to rule out a treatable cause in all patients with chronic Strong
pelvic pain.

4.5.2 Diagnostic evaluation of PPPS

Summary of evidence LE
Primary prostate pain syndrome (PPPS) is associated with negative cognitive, behavioural, sexual, 2b
or emotional consequences, as well as with symptoms suggestive of lower urinary tract and sexual
dysfunction.
Primary prostate pain syndrome has no known single aetiology. 3
Pain in PPPS involves mechanisms of neuroplasticity and neuropathic pain. 2a
Primary prostate pain syndrome has a high impact on QoL. 2b
Depression and catastrophic thinking are associated with more pain and poorer adjustment. 3
The prevalence of PPPS-like symptoms is high in population-based studies (> 2%). 2b
Reliable instruments assessing symptom severity as well as phenotypic differences exist. 2b

CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021 37


Recommendations Strength rating
Adapt diagnostic procedures to the patient. Exclude specific diseases with similar Strong
symptoms.
Use a validated symptom and quality of life scoring instrument, such as the National Strong
Institutes of Health Chronic Prostatitis Symptom Index, for initial assessment and follow-up.
Assess primary prostate pain syndrome associated negative cognitive, behavioural, Strong
sexual, or emotional consequences, as well as symptoms of lower urinary tract and sexual
dysfunctions.

4.5.3 Diagnostic evaluation of primary bladder pain syndrome

Summary of evidence LE
Primary bladder pain syndrome has no known single aetiology. 3
Pain in PBPS does not correlate with bladder cystoscopic or histologic findings. 2a
Primary bladder pain syndrome Type 3 C can only be confirmed by cystoscopy and histology. 2a
Lesion/non-lesion disease ratios of PBPS are highly variable between studies. 2a
The prevalence of PBPS-like symptoms is high in population-based studies. 2a
Primary bladder pain syndrome occurs at a level higher than chance with other pain syndromes. 2a
Primary bladder pain syndrome has an adverse impact on QoL. 2a
Reliable instruments assessing symptom severity as well as phenotypical differences exist. 2a

Recommendations Strength rating


Perform general anaesthetic rigid cystoscopy in patients with bladder pain to subtype and Strong
rule out confusable disease.
Diagnose patients with symptoms according to the EAU definition, after primary exclusion of Strong
specific diseases, with primary bladder pain syndrome (PBPS) by subtype and phenotype.
Assess PBPS associated non-bladder diseases systematically. Strong
Assess PBPS associated negative cognitive, behavioural, sexual, or emotional Strong
consequences.
Use a validated symptom and quality of life scoring instrument for initial assessment and Strong
follow-up.

4.5.4 Diagnostic evaluation of scrotal pain syndrome

Summary of evidence LE
The nerves in the spermatic cord play an important role in scrotal pain. 2b
Ultrasound of the scrotal contents does not aid in diagnosis or treatment of scrotal pain. 2b
Post-vasectomy pain is seen in a substantial number of men undergoing vasectomy. 2b
Scrotal pain is more often noticed after laparoscopic than after open inguinal hernia repair. 1b

4.5.5 Diagnostic evaluation of urethral pain syndrome

Summary of evidence LE
Primary urethral pain syndrome may be a part of BPS. 2a
Urethral pain involves mechanisms of neuroplasticity and neuropathic pain. 2b

4.5.6 Diagnostic evaluation of gynaecological aspects chronic pelvic pain

Summary of evidence LE
Laparoscopy is well-tolerated and does not appear to have negative psychological effects. 1b

38 CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021


Recommendations Strength rating
Take a full uro-gynaecological history in those who have had a continence or prolapse non- Strong
absorbable mesh inserted and consider specialised imaging of the mesh.
Refer to a gynaecologist if clinical suspicion of a gynaecological cause for pain following Strong
complete urological evaluation. Laparoscopy should be undertaken in accordance with
gynaecological guidelines.

4.5.7 Diagnostic evaluation of anorectal pain syndrome

Summary of evidence LE
Tenderness on traction is the main criterion of the chronic anal pain syndrome. 1a

Recommendation Strength rating


Anorectal function tests are recommended in patients with anorectal pain. Strong

4.5.8 Diagnostic evaluation of nerves to the pelvis

Summary of evidence LE
Multiple sensory and functional disorders within the region of the pelvis/urogenital system may occur 2
as a result of injury to one or more of many nerves. The anatomy is complex.
There is no single aetiology for the nerve damage and the symptoms and signs may be few or multiple. 1
Investigations are often normal. 2
The peripheral nerve pain syndromes are frequently associated with negative cognitive, behavioural, 1
sexual, or emotional consequences.

Recommendations Strength rating


Rule out confusable diseases, such as neoplastic disease, infection, trauma and spinal Strong
pathology.
If a peripheral nerve pain syndrome is suspected, refer early to an expert in the field, Weak
working within a multidisciplinary team environment.
Imaging and neurophysiology help diagnosis but image and nerve locator guided local Weak
anaesthetic injection is preferable.

4.5.9 Diagnostic evaluation of sexological aspects in chronic pelvic pain

Summary of evidence LE
Chronic pain can lead to decline in sexual activity and satisfaction and may reduce relationship 2a
satisfaction.
Men who reported having sexual, physical or emotional abuse show a higher rate of reporting 2b
symptoms of CPPPS.
Sexual dysfunctions are prevalent in men with PPPS. 2b
In men with PPPS the most prevalent sexual complaints are ED and ejaculatory dysfunction. 3
In females with CPPPS all sexual function domains are lower. The most reported dysfunctions are 2a
sexual avoidance, dyspareunia and “vaginismus”.
Vulvar pain syndrome is associated with PBPS. 2a
Women with PBPS suffer significantly more from fear of pain, dyspareunia and decreased desire. 2a
Pelvic floor muscle function is involved in the excitement and orgasm phases of sexual response. 3
Chronic pain can cause disturbances in each of the sexual response cycle phases. 2b

Recommendation Strength rating


Screen patients presenting with symptoms suggestive for chronic primary pelvic pain Weak
syndrome for abuse, without suggesting a causal relation with the pain.

CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021 39


4.5.10 Diagnostic evaluation of psychological aspects of chronic pelvic pain

Summary of evidence LE
There is no evidence that distress generates complaints of pelvic pain, or that multiple symptoms 2b
suggest unreality of pain.
Current or recent sexual abuse are possible contributory factors in pelvic pain. 2a

Recommendations Strength rating


Assess patient psychological distress in relation to their pain. Strong
Ask patients what they think is the cause of their pain and other symptoms to allow the Strong
opportunity to inform and reassure.

4.5.11 Diagnostic evaluation of pelvic floor function

Summary of evidence LE
The ICS classification is suitable for clinical practice. 2a
Over-activity of the pelvic floor muscles is related to chronic pelvic pain, prostate, bladder and vulvar 2a
pain.
Over-activity of the pelvic floor muscles is an input to the CNS causing central sensitisation. 2b
There is no accepted standard for diagnosing myofascial trigger points. 2a
There is a relation between the location of trigger point and the region where the pain is perceived. 3

Recommendations Strength rating


Use the International Continence Society classification on pelvic floor muscle function and Strong
dysfunction.
In patients with chronic primary pelvic pain syndrome, it is recommended to actively look for Weak
the presence of myofascial trigger points.

5. MANAGEMENT
The philosophy for the management of chronic pelvic pain is based on a bio-psychosocial model. This is a
holistic approach with the patients’ active involvement. Single interventions rarely work in isolation and need to
be considered within a broader personalised management strategy.

The management strategy may well have elements of self-management. Pharmacological and non-
pharmacological interventions should be considered with a clear understanding of the potential outcomes and
end points. These may well include: psychology, physiotherapy, drugs and more invasive interventions.

Treatment philosophy
Providing information that is personalised and responsive to the patient’s problems, conveying belief and
concern, is a powerful way to allay anxiety [312]. Additional written information or direction to reliable sources
of information is useful; practitioners tend to rely on locally produced material or pharmaceutical products of
variable quality while endorsing the need for independent materials for patients [313].

5.1 Conservative management

5.1.1 Pain education


It is always valuable to include education about the causes of pain, including eliciting from patients their
anxieties about undiscovered pathology and addressing them. Information improves adherence to treatment
and underpins self-management, as shown in many other painful and non-painful disorders but not specifically
in pelvic and abdominal pain except by a small qualitative study [314].

40 CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021


5.1.2 Physical therapy
The physiotherapist is part of the pain management team; (including doctors, psychologists and nurses). The
therapeutic options for physiotherapists may not be the same in every country. Physiotherapists can either
specifically treat the pathology of the pelvic floor muscles, or more generally treat myofascial pain if it is part of
the pelvic pain syndrome. In most studies that have been done looking at the effect of physiotherapy in pelvic
pain, the treatment of the pelvic floor is only part of the pain management. In a review about physiotherapy
in women with pelvic pain, it was concluded that recommendations for physiotherapy should be given with
caution [315]. The review found six RCTs, of which three showed level 1b evidence with low-risk of bias. One
of these three found that Mensendieck somatocognitive therapy showed a pain reduction after one year follow-
up of 64%. This approach consists of myofascial relaxation and tension, improving posture and movement in
combination with cognitive behaviour therapy (CBT) [316].

Pelvic floor muscle pain


Treating pelvic floor over-activity and myofascial trigger points should be considered in the management of
chronic pelvic pain. Treatment should be done by specialised physiotherapists who are trained not only in the
musculo-skeletal aspects of pain, but also in the psychological mechanisms and the role of the CNS in chronic
pain.

For patients with chronic pelvic pain and dysfunction of the pelvic floor muscles, it is very helpful to learn how
to relax the muscles when the pain starts. By doing this, the circle of pain-spasm-pain can be interrupted.
In the case of shortened muscles, relaxation alone is not enough. Stretching of the muscle is mandatory to
regain length and function. Studies on physical therapy for pelvic floor pain syndrome have been sparse. A
single blinded RCT with myofascial physical therapy and general body massage was carried out in patients
with prostate or bladder pain. The global response rate to treatment with massage was significantly better in
the prostate than in the bladder pain group (57% vs. 21%). In the prostate pain group, there was no difference
between the two treatment arms. In the bladder pain group, myofascial treatment did significantly better than
massage. Massage only improved complaints in the prostate pain group. The fact that gender distribution was
different in each group is mentioned as a possible confounding factor [317].

Myofascial trigger point release


Treatment of myofascial trigger points can be done by manual therapy, dry needling and wet needling. The
evidence for all the different treatments is weak, with most studies showing no significant difference between
these techniques, though most studies were small and heterogeneous with regards to the patients and
methods. There is no evidence that manual techniques are more effective than no treatment [318]. Most studies
of dry needling have compared with wet needling. Different systematic reviews have come to the conclusion
that, although there is an effect of needling on pain, it is neither supported nor refuted that this effect is better
than placebo [319].

Physiotherapy in PBPS
Transvaginal manual therapy of the pelvic floor musculature (Thiele massage) in PBPS patients with high-tone
dysfunction of the pelvic floor significantly improved several assessment scales [320]. The role of specific levator
ani trigger point injections in women with chronic pelvic pain has been studied [321]. Each trigger point was
identified by intravaginal palpation and injected with bupivacaine, lidocaine and triamcinolone. Seventy-two
percent of women improved with the first trigger point injection, with 33% being completely pain-free. Efficacy
and safety of pelvic floor myofascial physical therapy has been compared with global therapeutic massage in
women with PBPS; global response assessment (GRA) rate was 59% and 26%, respectively. Pain, urgency and
frequency ratings, and symptoms decreased in both groups during follow-up, and did not differ significantly
between the groups. This suggests that myofascial physical therapy is beneficial in women with PBPS [322].

Primary Anal Pain Syndrome


An RCT demonstrated that biofeedback treatment was superior to electrogalvanic stimulation and massage
of the Levator muscle for treating chronic primary anal pain syndrome [126]. One hundred and fifty-seven
patients who had at least weekly rectal pain were investigated, but only patients with tenderness on traction
of the pelvic floor showed a significant treatment benefit. In patients with tenderness of the puborectalis
muscle (Rome II: “Highly likely Levator Ani Syndrome”), 87% reported adequate relief after one month of
biofeedback vs. 45% for electrogalvanic stimulation, and 22% for massage. These results were maintained
at twelve months with adequate relief after nine sessions of biofeedback in 58% of the whole group (Rome II:
“Highly likely” and “Possible Levator Ani Syndrome”), after galvanic stimulation in 27% and massage in 21% of
patients. As previously described in dyssynergic defecation, the ability to expel a 50 mL water filled balloon and
to relax pelvic floor muscles after biofeedback treatment were predictive of a favourable therapeutic outcome

CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021 41


[126]. The pathophysiology of the chronic primary anal pain syndrome is therefore similar to that of dyssynergic
defecation, and this favours the role of the pelvic floor muscles in the pathophysiology of both conditions.
Other treatment modalities have been less successful.

Treatment of sexual dysfunctions and chronic pelvic pain


Couples often benefit from early referral for relationship and sexual counselling during their treatment course
[323]. It needs to be remembered that sexual difficulties will arise as a result of pelvic pain syndromes as well
as those disorders potentially being primary. Specific behavioural strategies for women who have urogenital
complaints and female sexual dysfunction often include exploring alternatives to sexual intercourse (manual
or oral pleasuring), different coital positions (female superior or side lying), and pacing, such as limiting the
activity to less than that which causes pain. Planning for the time of intercourse is important, and scheduling
a clinic visit after intercourse might be useful to identify specific sites and causes of post-coital flares. The
corresponding evidence in men is lacking, but similar principles would apply. Other behavioural changes
involve pre- and post-coital voiding, application of ice packs to the genital or suprapubic area [323, 324], and
increased use of vaginal dilators, fingers or sex toys. Lubricants can also be used and women with signs of
vulvovaginal atrophy may benefit from oestrogen cream [325]. Optimising the pelvic floor muscle is indicated
when dysfunction is present and will relief the pain [326-328].

Other physical therapy interventions


Electromagnetic therapy. A small, sham-controlled, double-blind study of four weeks showed a significant,
sustained effect over a one-year period for CPPPS [329].
Microwave thermotherapy. In uncontrolled studies significant symptomatic improvement has been reported
from heat therapy, for example, transrectal and transurethral thermotherapy [330, 331].
Extracorporeal shockwave therapy. A small sham-controlled double-blind study of four times weekly perineal
extracorporeal shockwave therapy (n=30) in men with CPPPS showed significant improvement in pain, QoL,
and voiding compared to the control group (n=30) over twelve weeks [332]. Two other randomised sham-
controlled studies, have been published more recently, one comparing ten treatment sessions over two weeks
(n=40 vs. n=40) [333], another with four times weekly treatments (n=20 vs. n=20) [334]. Both concluded there
was a significant effect in terms of total NIH-CPSI score and pain at twelve weeks. Unfortunately, no long term
effects at 24 weeks could be shown in a published follow-up study of the second [335]. A recent Cochrane
review of non-pharmacological interventions for chronic pelvic pain reported a reduction in symptoms
following treatment compared with control and concluded that extracorporeal shockwave therapy may improve
symptoms without an increase in adverse events [336].
Acupuncture. In a small three-arm randomised trial of CPPPS in men, electro-acupuncture was superior
to sham treatment and advice and exercise alone [337]. Another more recent randomised study comparing
acupuncture (n=50) vs. sham-controlled (n=50) once weekly treatment for six weeks showed significant long
lasting improvement at 24 weeks in terms of response rate and overall symptom scores [338]. Another RCT
showed a significant effect for a follow-up of 32 weeks [339]. Two systematic reviews and meta-analyses
were published in 2016 analysing seven randomised-controlled studies on a total of 471 participants
comparing acupuncture to sham control or oral medical treatment [340, 341]. Both came to the conclusion that
acupuncture was effective and safe, significantly reducing total NIH-CPSI scores compared to sham or medical
treatment, and should be considered as a treatment option. This is in line with the conclusion of a recent
Cochrane systematic review [336] on non-pharmacological treatment options. However, the durability of this
effect is not known.
Posterior tibial nerve stimulation. See section 5.3.2, Neuromodulation.
Transcutaneous electrical nerve stimulation. See section 5.3.2, Neuromodulation.

5.1.3 Psychological therapy


Psychological interventions may be directed at pain itself or at adjustment to pain in terms of function
and mood and reduced health-care use, with or without pain reduction. Ideally, treatment follows general
principles and practice in the field of chronic pain [342, 343] but these have been neglected in pelvic pain.
Two systematic reviews and meta-analyses of the few heterogeneous trials of psychologically based treatment
for pelvic pain [344, 345] found benefits over a few months for pain, of around 50%, comparable to that from
pharmacotherapy, but this was not sustained at follow-up. Exposure to pain-related fears in women with
chronic pelvic pain proved superior to manual therapy in reducing those fears and overall pain disability, albeit
assessed only by self-report [346]. More standard multi-component psychologically-based programmes are in
the pilot stages [347]. One that combined mixed psychological therapies with acupuncture for endometriosis-
related pain, reported significant pain reduction at two year follow-up [348]. Indeed acupuncture is the only
complementary treatment to have alleviated pain in this group [349]. Three more standard multicomponent
(including psychological) treatments for pain [278, 316, 350] did not provide pain or symptom relief. Another

42 CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021


RCT of multi-component treatment showed no effect on pain but benefits for distress [351], as did an RCT of
mindfulness meditation for women with bladder pain [352]. The importance of multi-disciplinary treatment is
emphasised by several reviews [43, 353, 354]. For less disabled and distressed patients, this can be delivered
in part over the internet [355].

5.1.4 Dietary treatment


Scientific data are limited and dietary restriction alone does not produce significant symptomatic relief; however,
consider the involvement of a dietician.

5.2 Pharmacological management

5.2.1 Drugs for chronic primary pelvic pain syndrome


In this section the evidence available for specific CPPPSs is presented. Where there is no evidence the reader
is directed to the section on analgesics below (Section 5.2.2) where more generic use is discussed. There
is a large discrepancy in the treatment effects reported in case series and controlled trials that results from
a large placebo effect or publication bias. As a result of the multifactorial origin of for example PPPS, one
reason for treatment failure in some large randomised placebo-controlled trials may be the heterogeneity of the
patient population. One strategy for improving treatment effects may be stratification of patient phenotypes.
A prospective series of phenotypically directed treatment for PPPS has shown significant improvement of
symptoms and QoL [356]. Monotherapeutic strategies for the treatment of PPPS may fail [357], therefore, most
patients require multimodal treatment aimed at the main symptoms, and taking comorbidity into account. In the
past ten years, results from RCTs have led to advances in standard and novel treatment options.

5.2.1.1 Mechanisms of action


Mechanisms of action are discussed as appropriate under the drugs headings below.

5.2.1.2 Comparisons of agents used in pelvic pain syndromes

Primary Prostate Pain Syndrome (PPPS)

Anti-inflammatory drugs
For non-steroidal anti-inflammatory agents (NSAIDs), a trial with celecoxib reported that the pain sub-score,
QoL sub-score, and total NIH-CPSI score were in favour of the treatment arm vs. placebo, but effects
were limited to the duration of therapy [358]. In a meta-analysis, two studies of NSAIDs [269, 358] and one
with prednisolone [359] were pooled. Anti-inflammatory drugs were 80% more likely to have a favourable
response than placebo. In an updated network meta-analysis with more restrictive inclusion criteria regarding
documented outcome measures but a wider spectrum of drugs (including glycosaminoglycans, phytotherapy
and tanezumab), a significant effect on total NIH-CPSI scores and treatment response rates could be
demonstrated. Overall, a moderate treatment effect has been shown for anti-inflammatory drugs, but larger
studies are needed for confirmation, and long-term side-effects have to be taken into account.

α-blockers
Positive results from RCTs of α-blockers, i.e. terazosin [360, 361], alfuzosin [362], doxazosin [363, 364],
tamsulosin [365, 366], and silodosin [367] have led to widespread use of α-antagonists in the treatment of
PPPS in recent years. Whereas one systematic review and meta-analysis has not reported a relevant effect of
α-blockers due to study heterogeneity [368], another network meta-analysis of α-blockers [367] has shown
significant improvement in total symptoms, pain, voiding, and QoL scores. In addition, they had a higher rate
of favourable response compared to placebo [relative risk (RR): 1.4, 95% CI: 1.1-1.8, p=0.013]. However,
treatment responsiveness, i.e., clinically perceptive or significant improvement, may be lower than expected
from the change in mean symptom scores. Overall, α-blockers seem to have moderate but significant
beneficial effects. This probably is not the case for long-standing PPPS patients [369]. Future studies should
show if longer duration of therapy or some sort of phenotypically directed (e.g., patients with PPPS and
relevant voiding dysfunction) treatment strategies will improve treatment outcomes.

Antibiotic therapy
Empirical antibiotic therapy is widely used because some patients have improved with antimicrobial therapy.
Patients responding to antibiotics should be maintained on medication for four to six weeks or even longer.
Unfortunately, culture, leukocyte and antibody status of prostate-specific specimens do not predict antibiotic
response in patients with PPS [370], and prostate biopsy culture findings do not differ from those of healthy
controls [371]. The only randomised placebo-controlled trials of sufficient quality have been done for oral

CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021 43


antibiotic treatment with ciprofloxacin (six weeks) [372], levofloxacin (six weeks) [373], and tetracycline
hydrochloride (twelve weeks) [374]. The studies have been analysed in meta-analyses [367, 375]. Although
direct meta-analysis has not shown significant differences in outcome measures, network meta-analysis has
suggested significant effects in decreasing total symptom, pain, voiding, and QoL scores compared with
placebo. Combination therapy of antibiotics with α-blockers has shown even better outcomes in network
meta-analysis. Despite significant improvement in symptom scores, antibiotic therapy did not lead to
statistically significant higher response rates [375]. In addition, the sample sizes of the studies were relatively
small and treatment effects only modest and most of the time below clinical significance. It may be speculated
that patients profiting from treatment have had some unrecognised uropathogens. If antibiotics are used, other
therapeutic options should be offered after one unsuccessful course of a quinolone or tetracycline antibiotic
over six weeks.

5-α-reductase inhibitors
Although a few small pilot studies with 5-α-reductase inhibitors supported the view that finasteride may improve
voiding and pain, the first RCT published in a peer-reviewed journal did not support this, although the study
lacked power [376]. In another RCT, finasteride provided better amelioration of symptoms compared to saw
palmetto over a one-year period, but lacked a placebo-control arm [377]. A six-month placebo-controlled study
showed a non-significant tendency towards better outcome in favour of finasteride, possibly because of a lack of
statistical power [366]. The NIH-CPSI scores decreased significantly in a subgroup of men enrolled in a prostate
cancer risk reduction study treated with dutasteride compared to placebo [367]. Patients (n=427, age 50 to 75,
with elevated prostate-specific antigen [PSA]) were included if they had significant “prostatitis-like” symptoms at
baseline. Based on the evidence, 5-α-reductase inhibitors cannot be recommended for use in PPPS in general,
but symptom scores may be reduced in a restricted group of older men with an elevated PSA [367].

Phytotherapy
Phytotherapy applies scientific research to the practice of herbal medicine. An adequately powered placebo-
controlled RCT of a pollen extract (Cernilton) showed clinically significant symptom improvement over a twelve-
week period in inflammatory PPPS patients (NIH Cat. IIIA) [378]. The effect was mainly based on a significant
effect on pain. Another pollen extract (DEPROX 500) has been shown to significantly improve total symptoms,
pain and QoL compared to ibuprofen [379]. A systematic review and meta-analysis of pollen extract for the
treatment of PPPS showed significant improvement in overall QoL [380]. Quercetin, a polyphenolic bioflavonoid
with documented antioxidant and anti-inflammatory properties, improved NIH-CPSI scores significantly in
a small RCT [381]. In contrast, treatment with saw palmetto, most commonly used for “benign prostatic
hyperplasia”, did not improve symptoms over a one-year period [377]. In a SR and meta-analysis, patients
treated with phytotherapy were found to have significantly lower pain scores than those treated with placebo
[367]. In addition, overall response rate in network meta-analysis was in favour of phytotherapy (RR: 1.6; 95%
CI: 1.1-1.6).

Pregabalin is an anti-epileptic drug that has been approved for use in neuropathic pain. In an adequately
powered randomised placebo-controlled study, which was the only report included in a published Cochrane
review [382], a six-week course of pregabalin (n=218) compared to placebo (n=106) did not result in a
significant reduction of NIH-CPSI total score [383].

Pentosane polysulphate is a semi-synthetic drug manufactured from beech-wood hemicellulose. One study
using oral high-dose (3 x 300 mg/day) demonstrated a significant improvement in clinical global assessment
and QoL over placebo in men with PPPS, suggesting a possible common aetiology [384].

Muscle relaxants (diazepam, baclofen) are claimed to be helpful in sphincter dysfunction or pelvic floor/
perineal muscle spasm, but there have been few prospective clinical trials to support these claims. In one
RCT, a triple combination of a muscle relaxant (thiocolchicoside), an anti-inflammatory drug (ibuprofen) and an
α-blocker (doxazosin) was effective in treatment-naïve patients, but not superior to an α-blocker alone [364].

Botulinum toxin type A (BTX-A) showed some effect in the global response assessment and the NIH-
CPSI pain subdomain score in a small randomised placebo-controlled study of perineal skeletal muscle
injection (100 U). However, patient numbers were low (thirteen in the BTX-A group and sixteen in the placebo
group), and follow-up was too short to draw definitive conclusions. Side-effects are unclear [385]. In another
randomised-controlled study of intraprostatic injection of BTX-A (100 or 200 U depending on prostate volume)
vs. placebo (n=30 in both groups) a significant improvement of total NIH-CPSI and subdomain scores could
be shown at six months [386]. However, no real placebo effect could be demonstrated, which suggests
unblinding. No definitive conclusion can be drawn.

44 CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021


Zafirlukast, a leukotriene antagonist, and prednisone in two low-power placebo-controlled studies failed to
show a benefit [359, 387]. More recently, a placebo-controlled phase IIa study of tanezumab, a humanised
monoclonal antibody that specifically inhibits the pain mediating neurotrophin, nerve growth factor, failed to
demonstrate significant effect [388] and should only be used in clinical trials.

Allopurinol
There is insufficient evidence for the use of allopurinol in PPPS [389, 390].

Primary Bladder Pain Syndrome (PBPS)

Treatments of significant value for BPS

Anti-histamines
Mast cells may play a role in PBPS. Histamine is one of the substances released by mast cells. Histamine
receptor antagonists have been used to block the H1 [391] and H2 [392] receptor subtypes, with variable
results. A prospective placebo-controlled RCT of hydroxyzine or oral pentosane polysulphate did not show a
significant effect [393].

Amitriptyline
Amitriptyline is a tricyclic antidepressant. Several reports have indicated improvement of PBPS symptoms after
oral amitriptyline [394]. Amitriptyline has been shown to be beneficial when compared with placebo plus
behavioural modification [395]. Drowsiness is a limiting factor with amitriptyline, nortriptyline is sometimes
considered instead.

Pentosane polysulphate
Is a semi-synthetic drug manufactured from beech-wood hemicellulose. Subjective improvement of pain,
urgency, frequency, but not nocturia, has been reported [396, 397]. Pentosane polysulphate had a more
favourable effect in PBPS type 3 C than in non-lesion disease [398]. Response was not dose dependent but
related more to treatment duration. At 32 weeks, about half the patients responded. Combination therapy
showed a response rate of 40% compared to 13% with placebo. For patients with an initial minor response to
pentosane polysulphate, additional subcutaneous heparin was helpful [399, 400].

Immunosuppressants
Azathioprine treatment has resulted in disappearance of pain and urinary frequency [401]. Initial evaluation of
cyclosporin A (CyA) [402] and methotrexate [403] showed good analgesic effect but limited efficacy for urgency
and frequency. Corticosteroids are not recommended in the management of patients with PBPS because of a
lack of evidence.

Intravesical Treatments
Intravesical drugs are administered due to poor oral bio-availability establishing high drug concentrations within
the bladder, with few systemic side-effects. Disadvantages include the need for intermittent catheterisation
which can be painful in PBPS patients, cost and risk of infection [404].

• Local anaesthetics
There are sporadic reports of successful treatment of PBPS with intravesical lidocaine [405, 406]. Alkalisation
of lidocaine improves its pharmacokinetics [407]. Combination of heparin, lidocaine and sodium bicarbonate
gave immediate symptom relief in 94% of patients and sustained relief after two weeks in 80% [408].
Intravesical instillation of alkalised lidocaine or placebo for five consecutive days resulted in significantly
sustained symptom relief for up to one month [409].

• Hyaluronic acid and chondroitin sulphate


These are described to repair defects in the GAG layer. Despite the fact that intravesical GAG replenishment
has been in use for about twenty years for PBPS/IC, most of the studies are uncontrolled and with a small
number of patients. Based on the studies available there are differences by virtue of substance classes,
whether they are natural GAG layer components, dosage formulations, or concentrations. A recent RCT seems
to reinforce the case for GAG layer replenishment, however it lacks a placebo arm [410]. A meta-analysis
confirms usefulness of GAG layer replenishment. However most retrieved studies are non-randomised and with
scarce numbers [411].

CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021 45


• Intravesical heparin
Primary bladder pain syndrome patients were treated with heparin for three months, and over half had
control of symptoms, with continued improvement after one year of therapy [412]. Kuo reported another
trial of intravesical heparin for three months in women with frequency-urgency syndrome and a positive
potassium test. Symptomatic improvement was reported in 80% of PBPS patients [413]. Intravesical heparin
plus peripheral neuromodulation in patients with refractory PBPS was studied and it was shown that voiding
frequency, pain score and maximum cystometric capacity were significantly better after two and twelve months
[414].

• Hyperbaric oxygen
This has a moderate effect on a small subgroup of PBPS patients. Disadvantages include high cost, limited
availability of treatment sites, and time-consuming treatment [399].

Treatments of limited value for BPS

Cimetidine
There are limited data to suggest that cimetidine improves symptoms of PBPS in the short-term. Compared
with placebo for three months, cimetidine significantly improved symptom scores, pain and nocturia, although
the bladder mucosa showed no histological changes in either group [415].

Prostaglandins
Misoprostol is a prostaglandin that regulates various immunological cascades. After three months of treatment
with misoprostol, 14 out of 25 patients had significantly improved, with twelve showing a sustained response
after a further six months [416]. The incidence of adverse drug effects was 64%.

L-Arginine
Oral treatment with the nitric oxide (NO) synthase substrate L-arginine decreases PBPS-related symptoms
[417-419]. Nitric oxide is elevated in patients with PBPS [420]. However, others have not demonstrated
symptomatic relief or changes in NO production after treatment [421, 422].

Oxybutynin is an anti-cholinergic drug used in overactive detrusor dysfunction. Intravesical oxybutynin


combined with bladder training improves functional bladder capacity, volume at first sensation, and
cystometric bladder capacity [423]. However, an effect on pain has not been reported.

Duloxetine (a serotonin-noradrenaline re-uptake inhibitor antidepressant with a licence for the management
of neuropathic pain) did not significantly improve symptoms of PBPS [424]. Administration was safe, but
tolerability was poor due to nausea. Based on these preliminary data, duloxetine cannot be recommended for
treatment of PBPS.

Primary Scrotal Pain Syndrome (PSPS)


Treatment of primary scrotal pain syndrome is based on the principles of treating chronic pain syndromes, as
described throughout these guidelines.

In men with pain post inguinal hernia repair, there is limited evidence from case series showing that neurectomy
of the damaged nerves can lead to symptomatic benefit [192, 425].

For scrotal pain post vasectomy, affected men may find that reversal of vasectomy can cure symptoms
especially in those in whom patency is achieved [426]. In a prospective RCT, pulsed radio-frequency to the
ilioinguinal and genitofemoral nerves is associated with high rates of symptomatic improvement (80%) but
follow up was limited to 3 months [427]. The evidence for epididymectomy is poor but if considered, is less
likely to provide benefit if the epididymis has a normal sonographic appearance [428].

Chronic gynaecological pain


It is difficult to compare the wide variation of drugs from an efficacy and safety perspective as they have such
diverse uses/indications.

In those gynaecological patients where chronic pelvic pain is unrelated to any of the well-defined conditions,
it is often difficult to determine a therapeutic pathway other than a multi-disciplinary chronic abdomino-pelvic
pain management plan. A Cochrane review suggests there may be some evidence (moderate) supporting the
use of progestogens [344]. Though efficacious, physicians need to be knowledgeable of progestogenic side

46 CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021


effects (e.g., weight gain, bloatedness - the most common adverse effects) which can stop some patients from
accepting such medication. Gonadotrophins, such as goserelin, are also thought to help such pain. However,
when compared with progestogens, their efficacy remains limited. The quality of evidence is generally low and
is drawn from single studies [344].

Gonadotropin-releasing hormone (GnRH) on the other hand binds to specific receptors on pituitary gonadotrophs,
leading to desensitisation and consequently to suppressed gonadotropin secretion. By contrast, GnRH
antagonists compete with GnRH for receptors thus gonadotrophin secretion, which may be beneficial in certain
clinical applications, such as reducing the size of fibroids, endometrial bleeding and endometriosis [429].

Pelvic Floor, Abdominal and Chronic Anal Pain


Botulinum toxin type A (pelvic floor)
Botulinum toxin type A has been injected into trigger points. It is more expensive than lidocaine and has not
been proven to be more effective [430]. Reviews do not support the injection of BTX-A into trigger points
[431]. Pelvic floor muscle over-activity plays a role in CPPS. Botulinum toxin type A, as a muscle relaxant,
can be used to reduce the resting pressure in the pelvic floor muscles. In women with high resting pressure
in the pelvic floor muscles, it has been found that BTX-A lowers this pressure significantly. The magnitude
of reduction was significantly higher than that in the placebo group. On the VAS pain score, no intergroup
differences were found in this relatively small randomised study [432]. Botulinum toxin type A can also be
injected at the sphincter level to improve urination or defecation. Relaxation of the urethral sphincter alleviates
bladder problems and secondarily the spasm. In a cohort study of thirteen patients with CPPPS, BTX-A was
injected into the external urethral sphincter. Subjectively, eleven patients reported a substantial change in pain
symptoms, from a score of 7.2 to 1.6 on a VAS [433].

Botulinum toxin type A (chronic primary anal pain syndrome)


In CPPPS associated with spasm of the levator ani muscles, treatment of the puborectalis and pubococcygeus
muscle by BTX-A appears to be promising in some women, as shown in a pilot study (n=12). The inclusion
criteria were dependent only on vaginal manometry with over-activity of the pelvic floor muscles, defined as
a vaginal resting pressure > 40 cm H2O. Although dyspareunia and dysmenorrhoea improved, non-menstrual
pelvic pain scores were not significantly altered [434]. In the following double-blinded, randomised, placebo-
controlled trial, the same group defined pelvic floor myalgia according to the two criteria of tenderness on
contraction and raised muscle tension (> 40 cm H2O) and included 60 women. In this larger study, non-
menstrual pelvic pain was significantly improved compared to those treated with placebo (VAS score 51 vs.
22; p=0.009). It was concluded therefore that BTX-A is effective at reducing pelvic floor-muscle associated
pain with acceptable adverse effects such as occasional urinary and faecal stress incontinence [420]. However,
recently, a small RCT failed to show any benefit of BTX-A [435].

Intermittent chronic primary anal pain syndrome


Due to the short duration of the episodes, medical treatment and prevention is often not feasible. Inhaled β-2
adrenergic agonist salbutamol was effective in an RCT in patients with frequent symptoms and shortened
pain duration [436]. Other treatment options are topic diltiazem and BTX-A [437]. However, there is still some
controversy regarding the duration of pain of intermittent chronic and chronic primary anal pain syndrome.
Randomised controlled trials often use different definitions, extending the pain duration (with a shift to chronic
pain) in order to include more patients and to better evaluate the study-drug action.

Abdominal pain associated with Irritable Bowel Syndrome


Linaclotide, a minimally absorbed peptide guanylate cyclase-C agonist at a dose of 290 μg once daily
significantly improved abdominal pain (48.9% vs. 34.5% placebo-treated) and bowel symptoms associated
with IBS with constipation over 26 weeks of treatment [438]. Diarrhoea was the most common adverse event in
patients treated with linaclotide (4.5%). Although it is known to overlap with IBS pelvic pain, effect on the latter
was not assessed in this study.

Delta-9-tetrahydrocannabinol (THC) shows only equivocal evidence of analgesic effects in chronic primary
abdominal pain. In a recently published phase II trial, no difference was found between THC tablet and a
placebo tablet in reducing pain outcome in patients with chronic abdominal pain [439].

5.2.2 Analgesics
If the use of simple analgesics fails to provide adequate benefit, then consider using neuropathic agents, and
if there is no improvement, consider involving a specialist pain management centre with an interest in pelvic
pain. Chronic pelvic pain is well defined and involves multiple mechanisms as described in previous sections.

CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021 47


The management requires a holistic approach with biological, psychological and social components. Few
studies have specifically looked at medications used in CPPPS [440], therefore, a wider look at the literature
has been undertaken and further specific research is required. The agents concerned are divided for ease of
description. Combinations often provide a greater benefit than individual agents. They may also allow lower
individual dosages and thus minimise side-effects. The aim of using these drugs is to allow patients to improve
their QoL. This is best measured by assessing their function as well as pain severity. If the use of these agents
does not allow this, then they should be withdrawn. Unfortunately, the failure of one agent does not exclude
potential benefit of an alternative. If the benefit is limited by side-effects, then the lowest effective dose should
be found (by dose titration). Sometimes, patients will prefer a higher level of pain and have fewer side-effects.

5.2.2.1 Mechanisms of action


Mechanisms of action are discussed as appropriate under the drug headings below.

5.2.2.2 Comparisons within and between groups in terms of efficacy and safety

Paracetamol (acetaminophen)
Paracetamol is a well-tolerated analgesic in a class of its own. This is an antipyretic analgesic with a central
mechanism of action [441]. It is often available over the counter without prescription. A review questions its
routine use as a first-line analgesic based on inadequate evidence of efficacy in many pain conditions including
dysmenorrhoea [442]. It will not be effective for all patients and individual responses should be reviewed when
deciding on longer term use.

Non-steroidal anti-inflammatory agents (NSAIDs)


These agents are anti-inflammatory, antipyretic analgesics that act by inhibiting the enzyme cyclooxygenase
(COX). They have a peripheral effect, hence their use in conditions involving peripheral or inflammatory
mechanisms. They are commonly used for pelvic pain; many are available over the counter and are usually
well-tolerated. There is insufficient evidence to suggest one NSAID over another for pelvic pain. Guidelines
for use of NSAIDs and COX-2 selective agents have been developed. They have more side-effects than
paracetamol, including indigestion, headaches and drowsiness.

The evidence for their benefit in chronic pelvic pain is weak or non-existent and is often limited by side-effects.
For pelvic pain in which inflammatory processes are considered important, such as dysmenorrhoea [443],
NSAIDs are more effective than placebo and paracetamol, but with a higher incidence of side-effects. For
pelvic pain in which central mechanisms may be incriminated, such as endometriosis [444], then the evidence
is lacking for NSAIDs despite their common use.

At a practical level, if NSAIDs are considered for use, they should be tried (having regard for the cautions and
contraindications) and the patient reviewed for improvement in function as well as analgesia. If this is not
achieved, or side-effects are limiting, then they should be withdrawn.

Neuromodulators
These are agents that are not simple analgesics but used to modulate neuropathic or centrally mediated pain.
There are several classes commonly used with recognised benefits in pain medicine. They are taken on a
regular basis and all have side-effects that may limit their use. In the UK, the National Institute for Health and
Clinical Excellence (NICE) has reviewed the pharmacological management of neuropathic pain [445]. Not all
the agents are licensed for use in pain management but there is a history and evidence to demonstrate their
benefit. The evidence for treatment of CPPPS is lacking but is present for other painful conditions. For this
chapter, most of the evidence is from non-pelvic pain sources. The general method for using these agents
is by titrating the dose against benefit and side-effects. The aim is for patients to have an improvement in
their QoL, which is often best assessed by alterations in their function. It is common to use these agents in
combination but studies comparing different agents against each other, or in combination, are lacking. Some
of these agents are also used for specific conditions. Early identification of neuropathic pain with a simple
questionnaire could facilitate targeted therapy with neuromodulators [59].

Antidepressants
Tricyclic antidepressants
The tricyclic antidepressants (TCAs) have multiple mechanisms of action including, blockade of acetylcholine
receptors, inhibition of serotonin and noradrenaline re-uptake, and blockade of histamine H1 receptors. They
also have anxiolytic affects [446] and are frequently limited by their side-effects. Tricyclic antidepressants have
a long history of use in pain medicine and have been subjected to a Cochrane review [447], suggesting that

48 CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021


they are effective for neuropathic pain. Amitriptyline is the most commonly used at doses from 10 to 75 mg/day
(sometimes rising to 150 mg/day). This is titrated against benefit or side-effects and can be taken at night [445].
Nortriptyline and imipramine are used as alternatives.

Other Antidepressants
Duloxetine is a serotonin-noradrenaline re-uptake inhibitor (SNRI) antidepressant licensed for use in
depression, urinary stress incontinence and neuropathic pain. There is moderately strong evidence of benefit in
diabetic neuropathy and fibromyalgia at a dose of 60 mg/day [448]. Side-effects are common and may result in
its discontinuation.

Selective serotonin re-uptake inhibitors are antidepressants with fewer side-effects. They are effective for
depression, but there have been insufficient studies to demonstrate their benefit in pelvic or neuropathic pain
[447-449].

Anticonvulsants
Anticonvulsants are commonly used in the management of neuropathic pain. There are general studies and
some looking more particularly at pelvic pain. Individual agents have been systematically reviewed. Their use is
suggested in the NICE Neuropathic Pain Guidelines [445].

There is a growing awareness and evidence of the risk for dependence and misuse of gabapentinoids [450]. A
formal assessment of efficacy against benefit and side-effects (both pain and QoL) is required with the patient
in order to determine the lowest effective dose and if longer-term treatment is to be used.

Carbamazepine has a long history of use in neuropathic pain. Evidence exists for its benefit [451]. Trials trend to
be of short duration, showing only moderate benefit. There are side-effects; some of which may be serious. It is
no longer a first choice agent. Other anticonvulsant agents are available with fewer serious side-effects.

Gabapentin is commonly used for neuropathic pain and has been systematically reviewed [452]. It provides
good quality relief with number needed to treat (NNT) of approximately six. Side-effects are common, notably
drowsiness, dizziness and peripheral oedema. For higher dose levels, reference should be made to local
formularies, and many clinicians do not routinely exceed 2.4 g/day in divided doses (most commonly three
times daily). One study of women with CPPPS has suggested that gabapentin alone or in combination with
amitriptyline provides better analgesia than amitriptyline alone [453]. A more recent pilot study suggests that
gabapentin is beneficial and tolerable; a larger study is required to provide a definitive result [454].

Pregabalin is a commonly used neuromodulator with good evidence of efficacy in some neuropathic conditions
but the NNT varies depending on the condition [455]. The dose for benefit is in the range of 300 to 600 mg/
day. The same SR found that doses less than 150 mg/day are unlikely to provide benefit. A review for CPPS
(prostate) only found a single reviewable study that does not show overall symptom improvement but suggests
individual symptoms may improve (e.g., pain, QoL) and side-effects were common demonstrating the need
for further robust studies [382]. As with gabapentin, side-effects are common and may not be tolerated by
patients. A formal assessment of efficacy against side-effects is required with the patient in order to determine
longer-term treatment. Other anticonvulsants are available but not commonly used for managing pain. Other
agents can be used in the management of neuropathic pain but they are best administered only by specialists
in the management of pain and familiar with their use. They tend to be considered after the standard options
have been exhausted. As with all good pain management, they are used as part of a comprehensive multi-
dimensional management plan.

Opioids
Over recent years opioids have been used extensively for managing chronic non-cancer pain. There is
increasing evidence that their role is limited in this population but may be beneficial for a small number of
patients at a low dose in a managed setting [456]. There is clear evidence of harm and significant professional,
public and political interest. Their use is beneficial for both acute pain and for cancer pain management
particularly towards the end of life.

Often patients will stop taking oral opioids due to side effects or insufficient analgesic effect [457]. There
is clear evidence of harm including effects on the endocrine and immune systems as well as a growing
understanding of opioid-induced hyperalgesia [458]. There is limited guidance on the best method for tapering
the dose of opioids with the aim of stopping or finding the lowest effective dose [459].

CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021 49


Opioids should only be used in conjunction with a management plan with consultation between clinicians
experienced in their use. It is suggested that a pain management unit should be involved along with the patient
and their primary care physician. Ensure there are arrangements for formal monitoring, follow-up and review.
If opioids are used and the pain remains, then they are not working and should be stopped even if there is no
alternative [458].

The risk of harm increases substantially at doses above 120 mg/day morphine equivalence [458] and guidance
suggests regular (at least annual) review for patients with over 50 mg/day morphine equivalence and pain
specialist involvement above 90 mg/day morphine equivalence [460].

There are well-established guidelines for the use of opioids in pain management as well as considering the
potential risks [458, 460]. Opioid reduction and optimisation should be undertaken where opioids are not
providing clear measurable benefit. There is also information available online for patients [458]. Opioids
Aware is a web-based resource for patients and healthcare professionals, jointly produced by the Faculty of
Pain Medicine of Royal College of Anaesthetists and Public Health England, to support prescribing of opioid
medicines for pain. https://fpm.ac.uk/opioids-aware.

Cannabinoids
There has been increasing interest and changes in national regulations regarding the use of cannabinoids for
medicinal use. Regarding pain the evidence base for the use of cannabinoids is weak [461-463] and further
well conducted clinical trials are necessary. This is an area where further guidance and research is likely in the
coming years.

5.3 Further management


5.3.1 Nerve blocks
Nerve blocks for pain management are usually carried out by specialists in pain medicine as part of a broader
management plan [464]. They may have a diagnostic or therapeutic role. Textbooks have been written on the
subject and practitioners using them should be trained in appropriate patient selection, indications, risks and
benefits. Many such interventions also require understanding and expertise in using imaging techniques to
perform the blocks accurately. Diagnostic blocks can be difficult to interpret due to the complex mechanisms
underlying the painful condition or syndrome. Sustained but limited benefit may lead to more permanent
procedures (e.g., radiofrequency procedures). There is a weak evidence base for these interventions for chronic
non-malignant pain [465].

Pudendal Neuralgia
The role of injections may be divided into two. First, an injection of local anaesthetic and steroid at the sight
of nerve injury may produce a therapeutic action. The possible reasons for this are related to the fact that
steroids may reduce any inflammation and swelling at the site of nerve irritation, but also because steroids may
block sodium channels and reduce irritable firing from the nerve [466]. However, a recent paper by Labat et al.
challenges this [467]. The second possible benefit is diagnostic. It has already been indicated that when the
pudendal nerve is injured there are several sites where this may occur. Differential block of the pudendal nerve
helps to provide information in relation to the site where the nerve may be trapped [260, 468-475].

Infiltration at the ischial spine requires the use of a nerve stimulator/locator. Both motor (anal contraction) and
sensory endpoints may be noted. The anatomical endpoint may be localised by fluoroscopy, CT guidance, or
the use of US, the latter avoids any radiation, whereas CT guidance involves a significant amount of radiation.
Currently, fluoroscopy is probably the imaging technique most frequently used because it is readily available
to most anaesthetists that perform the block. Currently, infiltration of the pudendal nerve within Alcock’s canal
is primarily undertaken with the use of CT. As well as injecting around the pudendal nerve, specific blocks
of other nerves arising from the pelvis may be performed. Pulsed radio frequency stimulation has also been
suggested as a treatment [476]. Pulsed radio frequency lesioning for pudendal neuralgia is being developed
with a paper demonstrating potential benefit. Follow-up is short term and further research is required to better
elucidate its place in management [477].

5.3.2 Neuromodulation
The role of neuromodulation in the management of pelvic pain should only be considered by specialists in
pelvic pain management. These techniques are used as part of a broader management plan and require regular
follow-up. The research base is developing and the techniques broadening (e.g., spinal cord stimulation [SCS],
sacral root stimulation, dorsal root ganglion stimulation or peripheral nerve stimulation). These are expensive
interventional techniques for patients refractory to other therapies. Neuromodulation is still finding its role in

50 CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021


pelvic pain management. There has been growing evidence but more detailed, high quality research is required
[478]. Its role in overactive bladder and faecal incontinence is more robust but is limited for pain. Two recent
systematic reviews have evaluated neuromodulation techniques for CPPS [479, 480]. Both studies concluded
that neuromodulation may be effective in reducing pain and improving QoL in patients with CPPS; however,
studies were of a low quality and long-term results were needed.

Transcutaneous Electrical Nerve Stimulation (TENS)


Transcutaneous electrical nerve stimulation is a non-invasive technique used in many pain conditions. A
SR identified twelve studies of TENS in chronic pelvic pain conditions including four RCTs [479]. All RCTs
demonstrated a significant reduction in pain following twelve weeks of treatment for pain conditions including
dysmenorrhoea and CPPPS. Pain was also found to improve following TENS for provoked vestibular pain.
There was conflicting data with regard to improvement of QoL following TENS; where validated questionnaires
were used, no significant improvement was found, whereas in trialist-defined studies, an improvement was
seen in TENS for dysmenorrhea and CPPPS. The beneficial effects of a course of TENS may be sustained; one
study demonstrating a persistent benefit at 43 months in 73% of men with CPPPS and another demonstrating
a prolonged significant improvement in women with provoked vestibular pain at ten months post-treatment.
Where reported there were no adverse events recorded. Transcutaneous electrical nerve stimulation could offer
an effective non-invasive treatment option for patients with CPPPS.

Percutaneous Tibial Nerve Stimulation (PTNS)


Percutaneous tibial nerve stimulation is a minimally invasive technique that can be use in an outpatient setting.
Two SRs have shown that PTNS is effective in reducing pain in patients with CPPPS [479, 480]. Three RCTs
identified showed a significant improvement in pain scores and QoL as measured by validated questionnaires.
Where recorded, adverse events were rare and minor including temporary slight pain at application site and
haematoma.

Sacral Nerve Stimulation (SNS)


Sacral nerve stimulation is an invasive technique requiring sedation or general anaesthesia for implantation of a
device following trial stimulation. A SR review identified ten studies of SNS in CPPPS, either retrospective case
series or prospective cohort studies and no RCTs. Where reported, a mean of 69% of participants progressed
to implantation of device following test stimulation (range 52-91%). All studies reported an improvement in
pain, statistically significant in five studies. Quality of Life was measured in three studies and a significant
improvement demonstrated in two of three studies. There was a large variation in adverse events reported
ranging from 0-50%. Complications not requiring surgical intervention included pain, failure of device, wound
infection and seroma. Re-operation rate ranged between 11-50% for complications including lead migration,
systemic infection, intrathecal implantation, loss of efficacy and erosion. In clinical practice, a patient should
be appropriately counselled regarding the need for a period of trial stimulation and whilst there may be an
improvement in symptoms, this should be weighed against a notable complication rate.

Other neuromodulation techniques


A variety of other techniques of neuromodulation for patients with CPPPS were identified by a recent
systematic review [479]. These techniques include intravaginal electrical stimulation for women with CPPPS,
pudendal nerve stimulation for CPPPS, spinal cord stimulation for pudendal neuralgia, transcutaneous
interferential electrical stimulation for IBS, electrical acupuncture for dysmenorrhoea and electrical stimulation/
biofeedback and electromagnetic stimulation for men with CPPPS. Whilst an improvement in pain has been
reported in these studies, it is noted that they are largely of low quality and further work is needed in this area
to enable robust clinical recommendations to be made.

5.3.3 Surgery

Primary Bladder Pain Syndrome (PBPS)

Bladder distension
Although bladder hydrodistension is a common treatment for PBPS, the scientific justification is scarce. It can
be part of the diagnostic evaluation, but has limited therapeutic role.

Hydrodistension and Botulinum toxin type A


Botulinum toxin type A may have an anti-nociceptive effect on bladder afferent pathways, producing
symptomatic and urodynamic improvements [481]. Treatment with hydrodistension and hydrodistension plus
intravesical BTX-A has been compared [482]. There was symptomatic improvement in all patients. However,

CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021 51


in the hydrodistension-only group, 70% returned to their previous symptoms after one month, while in the
BTX-A-treated patients, VAS score and functional and cystometric bladder capacity improved at three months.
Botulinum toxin type A trigonal-only injection seems effective and long-lasting as 87% of patients reported
improvement after three months follow-up [483]. Over 50% reported continued benefit nine months after the
first treatment. When re-treatment was needed, similar results were obtained. The authors concluded that
this treatment is safe, effective and can be repeated. Adverse effects of BTX-A administration for IC/PBPS
were significantly less than for overactive bladder syndrome, namely in increased post-void residual volumes
and decreased voiding efficiency [484]. Recent RCTs have confirmed benefits and long efficacy of BTX-A
administration [485-489]. The American Urological Association (AUA) guidelines panel has upgraded BTX-A
treatment from fifth to a fourth line treatment. There is conflicting data on results, hindering issuance of a clear
guideline for the use of botox in PBPS phenotypes [490].

Transurethral resection (TUR), coagulation and laser


Endourological destruction of bladder tissue aims to eliminate urothelial Hunner lesions. Since the 1970s,
resection and fulguration have been reported to achieve symptom relief, often for more than three years [491,
492]. Prolonged amelioration of pain and urgency has been described for transurethral laser ablation as well
[493].

Open Surgery for PBPS


Primary bladder pain syndrome is benign and does not shorten life, thus operative procedures rank last in
the therapeutic algorithm. There is no evidence that it relieves pain. Surgery for PBPS is only appropriate
as a last resort for patients with refractory disease. Major surgery should be preceded by thorough pre-
operative evaluation, with an emphasis on determining the relevant disease location and subtype. If surgery
is considered, the panel’s advice is to refer the patient to a specialist centre experienced in managing CPPPS
with a multi-disciplinary team approach.

Four major techniques are common:


1. Urinary diversion without cystectomy is performed to minimise the duration and complexity of surgery,
but complications related to the retained bladder commonly occur. Reports that un-resected PBPS
bladders cease to induce symptoms after loss of contact with urine are scarce [104, 494].

2. Supratrigonal cystectomy with bladder augmentation represents the most favoured continence-
preserving surgical technique particularly in younger patients [495]. Various intestinal segments have
been used [496-498]. After orthotopic bladder augmentation, bladder emptying may be incomplete
so intermittent self-catheterisation may be required. A study on female sexuality after cystectomy and
orthotopic ileal neobladder showed pain relief in all patients and improvement in sexual function items
in women who remained sexually active [499]. Pregnancies with subsequent lower-segment Caesarean
section have been reported after ileocystoplasty [500].

3. Subtrigonal or simple cystectomy refers to removal of the entire bladder at the level of the bladder neck.
This approach has the benefit of removing the trigone as a possible disease site, but at the cost of
requiring ureteric re-implantation. Trigonal disease is reported in 50% of patients and surgical failure has
been blamed on the trigone being left in place [501], especially in patients with non-lesion type disease
[502, 503]. However, in a previous study all patients were rendered symptom-free by supratrigonal
resection compared to 82% of those undergoing subtrigonal cystectomy. Voiding dysfunction is most likely
to occur following trigonal resection and patients considering augmentation and especially substitution
procedures must be capable of accepting, performing and tolerating self-catheterisation [504].

4. Cystectomy with formation of an ileal conduit is considered for patients with PBPS who develop recurrent
pain in the augmented bladder, continent pouch after enterocystoplasty or continent urinary diversion.
Re-tubularisation of a previously used bowel segment to form a urinary conduit has been recommended
[505].

Primary Prostate Pain Syndrome


There is no evidence for surgical management, including transurethral incision of the bladder neck, radical
transurethral resection of the prostate or, in particular, radical prostatectomy in the management of chronic
pain in patients with PPPS. Recently, a large Chinese randomised-controlled trial of circumcision combined
with a triple oral therapy (ciprofloxacin, ibuprofen, tamsulosin) vs. oral therapy alone has been published for
patients with PPPS (total n=774) [506]. It is hypothesised that there may be some immunological interaction
via pathogenic antigen presenting cells in the foreskin with CD4+ T cells causing auto-immunity to the

52 CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021


prostate gland. They reported an improvement in total NIH-CPSI score and subdomain scores at twelve
weeks. However, despite a large cohort, the study results are questionable because of the weak theoretical
background, and a potential large placebo effect lacking a sham control. In addition, no long-term effectiveness
has been reported. Before having an impact on recommendations, the results of this study have to be
independently confirmed and the treatment effect must persist.

Primary Testicular Pain Syndrome


Microsurgical denervation of the spermatic cord can be offered to patients with testicular pain. In a long-term
follow-up study, patients who had a positive result on blocking the spermatic cord were found to have a good
result following denervation [507, 508].

Chronic Primary Anal and Abdominal Pain Syndrome


Chronic primary anal pain syndrome after stapled procedures, such as hemorrhoidopexy (PPH) or stapled
transanal rectal resection (STARR) may respond to excision of the scarred staple line as shown in 21
consecutive patients with an overall improvement of pain in 85.7% of patients undergoing scar excision surgery
[509]. An early scar excision before three to six months after pain onset was associated with better pain relief.
Adhesiolysis is still in discussion in the pain management after laparotomy/laparoscopy for different surgical
indications in the pelvis and entire abdomen. A recent study has shown, that adhesiolysis is associated with
an increased risk of operative complications, and additional operations and increased health care costs as
compared to laparoscopy alone [510].

Primary Urethral Pain Syndrome


There is no specific treatment that can be advised. Management should be multi-disciplinary and multi-modal
[511]. Laser therapy of the trigonal region may be a specific treatment. One trial comparing two forms of
laser reported good results, but did not compare with sham treatment [512]. The majority of publications on
treatment of primary urethral pain syndrome have come from psychologists [168].

Presumed intra-abdominal adhesions


In gynaeocological patients with CPPPS and presumed adhesions, there is no consensus as to whether
adhesiolysis should be performed to improve pain [513].

Extensive surgery for endometriosis is challenging and is still considered to be controversial, as there is at least
one RCT showing no benefit in pain relief after the removal of early extensive endometriosis compared to sham
surgery [272, 514]. In patients with adenomyosis, the only curative surgery is hysterectomy but patients can
benefit from hormonal therapy and analgesics (see Section 5.2.2).

Pudendal neuralgia and surgery


Decompression of an entrapped or injured nerve is a routine approach and probably should apply to the
pudendal nerve as it applies to all other nerves. There are several approaches and the approach of choice
probably depends upon the nature of the pathology. The most traditional approach is transgluteal; however,
a transperineal approach may be an alternative, particularly if the nerve damage is thought to be related to
previous pelvic surgery [198, 260, 515-518]. Currently, there has been only one prospective randomised study
(transgluteal approach) [517]. This study suggests that, if the patient has had the pain for less than six years,
66% of patients will see some improvement with surgery (compared to 40% if the pain has been present for
more than six years). Surgery is not the answer for all patients. On talking to patients that have undergone
surgery, providing the diagnosis was clear-cut, most patients were grateful to have undergone surgery but
many still have symptoms that need management.

Chronic Pelvic Pain and Prolapse/Incontinence Mesh


Removing an existing mesh is a complex procedure [519]. Each patient is approached on an individual basis
depending on the type of mesh and extent of complications [520]. The complexity of surgery often involves
removal of dense scar tissue, reformation of inflamed vaginal skin and surgical reconstruction of the urethra
and bladder [521]. Such surgery requires specialist skills, best provided within a multi-disciplinary tertiary
setting. Possible complications as a result of this surgical removal include bleeding, infection, damage to
surrounding organs as well as lower urinary tract symptoms, persistent chronic pain and recurrent USI, which
occurs after mesh removal [522].

Removal of mesh, whilst complex, does have beneficial outcomes generally, which are also durable particularly
for chronic pain [523]. However, the long-term consequences after the mesh is removed still can include,
not only chronic persistent pain but also autoimmune responses and complex neuropathies affecting the

CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021 53


pelvis and the lower limbs [524, 525]. Some of these can be treated effectively using a multi-disciplinary pain
medicine approach [526]. In other cases, the residual symptoms may require the input of an immunologist,
rheumatologist or other symptom-defined specialist.

The alternative to continence and prolapse mesh surgery is dependent on the clinical findings at the
time. They include behavioural change, physiotherapy (for USI and Grade I-II uterovaginal prolapse) or
traditional surgical techniques. Studies have shown that over 70% who committed to physiotherapy for
stress urinary incontinence often did not need any further intervention [527]. Many clinicians are reverting to
conservative measures first, before re-considering surgery. Clinicians are also now retraining in traditional
continence surgical techniques, which existed in the pre-mesh era, such as the Burch colposuspension and
autologous fascial sling; as well as traditional utero-vaginal prolapse techniques such as vaginal hysterectomy,
sacrospinous fixation and fascial repair of vaginal wall prolapse.

5.4 Summary of evidence and recommendations: management

5.4.1 Management of primary prostate pain syndrome

Summary of evidence LE
Phenotypically directed treatment may improve treatment success. 3
α-blockers have moderate treatment effect regarding total pain, voiding, and QoL scores in PPPS. 1a
Antimicrobial therapy has a moderate effect on total pain, voiding, and QoL scores in PPPS. 1a
NSAIDs have moderate overall treatment effects on PPPS. 1a
Phytotherapy has some beneficial effect on pain and overall favourable treatment response in PPPS. 1a
Pentosane polysulphate improves global assessment and QoL score in PPPS. 1b
There are insufficient data on the effectiveness of muscle relaxants in PPPS. 2b
Pregabalin is not effective for the treatment of PPPS. 1b
BTX-A injection into the pelvic floor (or prostate) may have a modest effect in PPPS. 2b
Acupuncture is superior to sham acupuncture in improving symptoms and QoL. 1a
Posterior tibial nerve stimulation is probably effective for the treatment of PPPS. 1b
Extracorporeal shock wave therapy is probably effective over the short term. 1b
There are insufficient data supporting the use of other surgical treatments, such as transurethral 3
incision of the bladder neck, transurethral resection of the prostate, or radical prostatectomy in
patients with PPPS.
Cognitive behavioural therapy designed for PPPS may improve pain and QoL. 3

Recommendations Strength rating


Offer multimodal and phenotypically directed treatment options for Primary Prostate Pain Weak
Syndrome (PPPS).
Use antimicrobial therapy (quinolones or tetracyclines) over a minimum of six weeks in Strong
treatment-naïve patients with a duration of PPPS less than one year.
Use α-blockers for patients with a duration of PPPS less than one year. Strong
Offer high-dose oral pentosane polysulphate in PPPS. Weak
Offer acupuncture in PPPS. Strong
Offer non-steroidal anti-inflammatory drugs (NSAIDs) in PPPS, but long-term side-effects Weak
have to be considered.

5.4.2 Management of primary bladder pain syndrome

Summary of evidence LE
There is insufficient data for the long-term use of corticosteroids. 3
Limited data exist on effectiveness of cimetidine in PBPS. 2b
Amitriptyline is effective for pain and related symptoms of PBPS. 1b
Oral pentosane polysulphate is effective for pain and related symptoms of PBPS. 1a
Oral pentosane polysulphate plus subcutaneous heparin is effective for pain and related symptoms of 1b
PBPS, especially in initially low responders to pentosane polysulphate alone.
Intravesical lidocaine plus sodium bicarbonate is effective in the short term. 1b
Intravesical pentosane polysulphate is effective, based on limited data, and may enhance oral 1b
treatment.

54 CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021


There are limited data on the effectiveness of intravesical heparin. 3
Intravesical chondroitin sulphate may be effective. 2b
There is insufficient data for the use of bladder distension as a therapeutic intervention. 3
Hydrodistension plus BTX-A is superior to hydrodistension alone. 1b
Intravesical BCG is not effective in PBPS. 1b
Transurethral resection (coagulation and laser) may be effective in PBPS type 3 C. 3
Sacral neuromodulation may be effective in PBPS. 3
Pudendal nerve stimulation is superior to sacral neuromodulation for treatment of PBPS. 1b
Avoidance of certain foods and drink may reduce symptoms. 3
Outcome of cystectomy for PBPS is variable. 3

Recommendations Strength rating


Offer subtype and phenotype-oriented therapy for the treatment of Primary Bladder Pain Strong
Syndrome (PBPS).
Always consider offering multimodal behavioural, physical and psychological techniques Strong
alongside oral or invasive treatments of PBPS.
Offer dietary advice. Weak
Administer amitriptyline for treatment of PBPS. Strong
Offer oral pentosane polysulphate for the treatment of PBPS. Strong
Offer oral pentosane polysulphate plus subcutaneous heparin in low responders to Weak
pentosane polysulphate alone.
Do not recommend oral corticosteroids for long-term treatment. Strong
Offer intravesical hyaluronic acid or chondroitin sulphate before more invasive measures. Weak
Offer intravesical lidocaine plus sodium bicarbonate prior to more invasive methods. Weak
Offer intravesical heparin before more invasive measures alone or in combination treatment. Weak
Do not use bladder distension alone as a treatment of PBPS. Weak
Offer submucosal bladder wall and trigonal injection of botulinum toxin type A (BTX-A) plus Strong
hydrodistension if intravesical instillation therapies have failed.
Offer neuromodulation before more invasive interventions. Weak
Only undertake ablative and/or reconstructive surgery as the last resort and only by Strong
experienced and PBPS-knowledgeable surgeons, following a multi-disciplinary assessment
including pain management.
Offer transurethral resection (or coagulation or laser) of bladder lesions, but in PBPS type 3 C Strong
only.

5.4.3 Management of scrotal pain syndrome

Summary of evidence LE
Microsurgical denervation of the spermatic cord is an effective therapy for primary scrotal pain 2b
syndrome.
Vasovasostomy is effective in post-vasectomy pain. 2b

Recommendations Strength rating


Inform about the risk of post-vasectomy pain when counselling patients planned for Strong
vasectomy.
Do open instead of laparoscopic inguinal hernia repair, to reduce the risk of scrotal pain. Strong
In patients with testicular pain improving after spermatic block, offer microsurgical Weak
denervation of the spermatic cord.

5.4.4 Management of primary urethral pain syndrome

Summary of evidence LE
There is no specific treatment for primary urethral pain syndrome. 4

CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021 55


5.4.5 Management of gynaecological aspects of chronic pelvic pain

Summary of evidence LE
Therapeutic options, including pharmacotherapy and surgery, can treat endometriosis effectively. 1b
Psychological treatment (CBT or supportive psychotherapy) can improve pain and sexual and 1b
emotional function in vaginal and vulvar pain syndrome.
Most gynaecological pain conditions (including dysmenorrhea, post-mesh insertion and 3
gynaecological malignancy) can be treated effectively using pharmacotherapy.
All other gynaecological conditions (including obstetric injury, pelvic organ prolapse) can be treated 2
effectively using surgery.

Recommendations Strength rating


Involve a gynaecologist to provide therapeutic options such as hormonal therapy or surgery Strong
in well-defined disease states.
Provide a multi-disciplinary approach to pain management in persistent disease states. Strong
All patients who have developed complications after mesh insertion should be referred to a Strong
multi-disciplinary service (incorporating pain medicine and surgery).

5.4.6 Management of primary anorectal pain syndrome

Summary of evidence LE
Biofeedback is the preferred treatment for chronic primary anal pain syndrome. 1a
Electro-galvanic stimulation is less effective than biofeedback. 1b
Botulinum toxin type A is effective. 1b
Percutaneous tibial nerve stimulation is effective in anal pain. 3
Sacral neuromodulation is effective in anal pain. 3
Inhaled salbutamol is effective in intermittent chronic primary anal pain syndrome. 3

Recommendations Strength rating


Undertake biofeedback treatment in patients with chronic anal pain. Strong
Offer botulinum toxin type A in chronic primary anal pain syndrome. Weak
Offer percutaneous tibial nerve stimulation in chronic primary anal pain syndrome. Weak
Offer sacral neuromodulation in chronic primary anal pain syndrome. Weak
Offer inhaled salbutamol in intermittent chronic primary anal pain syndrome. Weak

5.4.7 Management of pudendal neuralgia

Summary of evidence LE
There are multiple treatment options with varying levels of evidence. 3

Recommendation Strength rating


Neuropathic pain guidelines are well-established. Use standard approaches to management Strong
of neuropathic pain.

5.4.8 Management of sexological aspects in chronic pelvic pain

Summary of evidence LE
Pelvic floor muscle physical therapy may offer relief of pain and reduction in sexual complaints. 2b

Recommendations Strength rating


Offer behavioural strategies to the patient and his/her partner to reduce sexual dysfunctions. Weak
Offer pelvic floor muscle therapy as part of the treatment plan to improve quality of life Weak
and sexual function.

56 CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021


5.4.9 Management of psychological aspects in chronic pelvic pain

Recommendation Strength rating


For chronic pelvic pain with significant psychological distress, refer patient for chronic pelvic Strong
pain-focused psychological treatment.

5.4.10 Management of pelvic floor dysfunction

Summary of evidence LE
Myofascial treatment is effective. 1b
Biofeedback improves the outcome of myofascial therapy. 1a

Recommendations Strength rating


Apply myofascial treatment as first-line treatment. Weak
Offer biofeedback as therapy adjuvant to muscle exercises, in patients with anal pain due to Strong
an overactive pelvic floor.

5.4.11 Management of chronic/non-acute urogenital pain by opioids

Recommendations Strength rating


Opioids and other drugs of addiction/dependency should only be prescribed following Strong
multi-disciplinary assessment and only after other reasonable treatments have been tried
and failed.
The decision to instigate long-term opioid therapy should be made by an appropriately- Strong
trained specialist in consultation with the patient and their family doctor.
Where there is a history or suspicion of drug abuse, involve a psychiatrist or psychologist Strong
with an interest in pain management and drug addiction.

6. EVALUATION OF TREATMENT RESULTS


6.1 Evaluation of treatment
For patients with chronic primary visceral pain, a visit to the clinician is important because they can ask
questions, talk about how the process is going and have some time with the caregiver who understands the
nature of their pain. First evaluation should take place after about six weeks to see if the treatment has been
successful or not. When necessary adaptations are made and a next evaluation is planned.

6.1.1 Treatment has not been effective


6.1.1.1 Alternative treatment
In cases where the treatment initiated did not have enough effect, an alternative approach is advised. The first
thing to do is a thorough evaluation of the patients’ or care providers’ adherence to the treatment that was
initiated. Ask the patient if they have taken the medication according to the prescription, if there were any side-
effects and if there were any changes in pain and function. Adjustment of medication or dose schemes might
help. Another important thing to do is to read the reports of other caregivers, for example, the physiotherapist
and the psychologist. Has the therapy been followed until the end, what was the opinion of the therapist about
the changes that were observed? In cases where the sessions had been terminated by the patient, ask the
patient why they made that decision. Check if the patient has understood the idea behind the therapy that had
been prematurely stopped.

6.1.1.2 Referral to next envelope of care


If patients and doctors conclude that none of the therapies given showed enough effect, then referral to a next
envelope of care is advised. Unfortunately, the terminology used to describe the nature and specialisation level
of centres providing specialised care for visceral pain patients is not standardised and is country-based. This
does not facilitate easy referral schemes. It is advised that patients are referred to a centre that is working with

CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021 57


a multi-disciplinary team and nationally recognised as specialised in pelvic pain. Such a centre will re-evaluate
what has been done and when available, provide specialised care.

6.1.1.3 Self-management and shared care


Patients who find themselves confronted with CPPPS, for which there is no specific treatment option available,
will have to live with their pain. They will need to manage their pain, meaning that they will have to find a way to
deal with the impact of their pain on daily activities in all domains of life. Self-help programmes may be advised
and can be of help. The patient may also benefit from shared care, which means that a caregiver is available for
supporting the self-management strategies. Together with this caregiver, the patient can optimise and use the
management strategies.

6.1.2 Treatment has been effective


In cases where treatment has been effective, the caregiver may pay attention to fall-back prevention. If the
patient feels the same pain again, it helps to start at an early stage with the self-management strategies that
he/she has learned during the former treatment. By doing so they will have the best chance of preventing the
re-development of pelvic pain syndromes.

7. REFERENCES
1. Fall, M., et al., EAU Guidelines on Chronic Pelvic Pain., in: EAU Guidelines on Chronic Pelvic Pain.
Presented at the 18th EAU Annual Congress Madrid 2003. 2003, European Association of Urology:
Arnhem.
https://uroweb.org/guideline/chronic-pelvic-pain/?type=archive
2. Fall, M., et al. EAU guidelines on chronic pelvic pain. Eur Urol, 2004. 46: 681.
https://pubmed.ncbi.nlm.nih.gov/15548433
3. Fall, M., et al., EAU Guidelines on Chronic Pelvic Pain., in EAU Guidelines on Chronic Pelvic Pain.
Presented at the 18th EAU Annual Congress Barcelona 2010. 2010, EAU: Arnhem.
https://uroweb.org/guideline/chronic-pelvic-pain/?type=archive
4. Fall, M., et al. EAU guidelines on chronic pelvic pain. Eur Urol, 2010. 57: 35.
https://pubmed.ncbi.nlm.nih.gov/19733958
5. Engeler, D.S., et al. The 2013 EAU guidelines on chronic pelvic pain: is management of chronic
pelvic pain a habit, a philosophy, or a science? 10 years of development. Eur Urol, 2013. 64: 431.
https://pubmed.ncbi.nlm.nih.gov/23684447
6. McMahon, S.B., et al. Visceral pain. Br J Anaesth, 1995. 75: 132.
https://pubmed.ncbi.nlm.nih.gov/7577247
7. Shoskes, D.A., et al. Clinical phenotyping of patients with chronic prostatitis/chronic pelvic pain
syndrome and correlation with symptom severity. Urology, 2009. 73: 538.
https://pubmed.ncbi.nlm.nih.gov/19118880
8. Magri, V., et al. Use of the UPOINT chronic prostatitis/chronic pelvic pain syndrome classification in
European patient cohorts: sexual function domain improves correlations. J Urol, 2010. 184: 2339.
https://pubmed.ncbi.nlm.nih.gov/20952019
9. World Health Organization. International Classification of Diseases (11th Revision). 2018.
https://icd.who.int/en
10. Aziz, Q., et al. The IASP classification of chronic pain for ICD-11: chronic secondary visceral pain.
Pain, 2019. 160: 69.
https://pubmed.ncbi.nlm.nih.gov/30586073
11. Häuser, W., et al. Taxonomies for chronic visceral pain. Pain, 2020. 161: 1129.
https://pubmed.ncbi.nlm.nih.gov/32032194
12. NICE, Chronic pain in over 16s: assessment and management guideline. 2020.
https://www.nice.org.uk/guidance/GID-NG10069/documents/draft-guideline
13. Merskey, H., et al., Classification of Chronic Pain. 1994, Seattle.
https://s3.amazonaws.com/rdcms-iasp/files/production/public/Content/ContentFolders/
Publications2/FreeBooks/Classification-of-Chronic-Pain.pdf
14. Krieger, J.N., et al. NIH consensus definition and classification of prostatitis. JAMA, 1999. 282: 236.
https://pubmed.ncbi.nlm.nih.gov/10422990

58 CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021


15. van de Merwe, J.P., et al. Diagnostic criteria, classification, and nomenclature for painful bladder
syndrome/interstitial cystitis: an ESSIC proposal. Eur Urol, 2008. 53: 60.
https://pubmed.ncbi.nlm.nih.gov/17900797
16. Longstreth, G.F., et al. Functional bowel disorders. Gastroenterology, 2006. 130: 1480.
https://pubmed.ncbi.nlm.nih.gov/16678561
17. Guyatt, G.H., et al. GRADE: an emerging consensus on rating quality of evidence and strength of
recommendations. BMJ, 2008. 336: 924.
https://pubmed.ncbi.nlm.nih.gov/18436948
18. Guyatt, G.H., et al. What is "quality of evidence" and why is it important to clinicians? BMJ, 2008.
336: 995.
https://pubmed.ncbi.nlm.nih.gov/18456631
19. Philips, B. et al. Modified from Oxford Centre for Evidence-based Medicine Levels of Evidence
(March 2009). Updated Jeremy Howick March 2009. Access date February 2014.
https://www.cebm.net/2009/06/oxford-centre-evidence-based-medicine-levels-evidence-
march-2009/
20. Global, regional, and national incidence, prevalence, and years lived with disability for 301 acute
and chronic diseases and injuries in 188 countries, 1990-2013: a systematic analysis for the Global
Burden of Disease Study 2013. Lancet, 2015. 386: 743.
https://pubmed.ncbi.nlm.nih.gov/26063472
21. Ayorinde, A.A., et al. Chronic pelvic pain in women of reproductive and post-reproductive age: a
population-based study. Eur J Pain, 2017. 21: 445.
https://pubmed.ncbi.nlm.nih.gov/ 27634190
22. Choung, R.S., et al. Irritable bowel syndrome and chronic pelvic pain: a population-based study.
J Clin Gastroenterol, 2010. 44: 696.
https://pubmed.ncbi.nlm.nih.gov/20375730
23. Fenton, B.W. Measuring quality of life in chronic pelvic pain syndrome. Exp Rev Obstet Gynecol,
2010. 5: 115.
https://pubmed.ncbi.nlm.nih.gov/336768
24. Baranowski, A.P. Chronic pelvic pain. Best Pract Res: Clin Gastroenterol, 2009. 23: 593.
https://pubmed.ncbi.nlm.nih.gov/19647692
25. Krieger, J., et al. Non-urological syndromes and severity of urological pain symptoms: Baseline
evaluation of the national institutes of health multidisciplinary approach to pelvic pain study. J Urol,
2013. 1): e181.
https://pubmed.ncbi.nlm.nih.gov/71031385
26. Chuang, Y.C., et al. Increased risks of healthcare-seeking behaviors of anxiety, depression and
insomnia among patients with bladder pain syndrome/interstitial cystitis: a nationwide population-
based study. Int Urol Nephrol, 2015. 47: 275.
https://pubmed.ncbi.nlm.nih.gov/25577231
27. Riedl, A., et al. Somatic comorbidities of irritable bowel syndrome: A systematic analysis.
J Psychosom Res, 2008. 64: 573.
https://pubmed.ncbi.nlm.nih.gov/18501257
28. Savidge, C.J., et al. Psychological aspects of chronic pelvic pain. J Psychosom Res, 1997. 42: 433.
https://pubmed.ncbi.nlm.nih.gov/9194016
29. Anda, R.F., et al. The enduring effects of abuse and related adverse experiences in childhood. A
convergence of evidence from neurobiology and epidemiology. Eur Arch Psychiatry Clin Neurosci,
2006. 256: 174.
https://pubmed.ncbi.nlm.nih.gov/16311898
30. Raphael, K.G., et al. Childhood victimization and pain in adulthood: a prospective investigation.
Pain, 2001. 92: 283.
https://pubmed.ncbi.nlm.nih.gov/11323150
31. Tunitsky, E., et al. Bladder pain syndrome/interstitial cystitis in twin sisters. J Urol, 2012. 187: 148.
https://pubmed.ncbi.nlm.nih.gov/22088343
32. Vehof, J., et al. Shared genetic factors underlie chronic pain syndromes. Pain, 2014. 155: 1562.
https://pubmed.ncbi.nlm.nih.gov/24879916
33. Dybowski, C., et al. Predictors of pain, urinary symptoms and quality of life in patients with chronic
pelvic pain syndrome (CPPS): A prospective 12-month follow-up study. J Psychosom Res, 2018.
112: 99.
https://pubmed.ncbi.nlm.nih.gov/30097143

CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021 59


34. Roth, R.S., et al. Patient beliefs about pain diagnosis in chronic pelvic pain: relation to pain
experience, mood and disability. J Reprod Med, 2011. 56: 123.
https://pubmed.ncbi.nlm.nih.gov/21542529
35. Berman, S.M., et al. Reduced brainstem inhibition during anticipated pelvic visceral pain correlates
with enhanced brain response to the visceral stimulus in women with irritable bowel syndrome.
J Neurosci, 2008. 28: 349.
https://pubmed.ncbi.nlm.nih.gov/18184777
36. Naliboff, B.D., et al. Clinical and Psychosocial Predictors of Urological Chronic Pelvic Pain Symptom
Change in 1 Year: A Prospective Study from the MAPP Research Network. J Urol, 2017. 198: 848.
https://pubmed.ncbi.nlm.nih.gov/28528930
37. Bajaj, P., et al. Endometriosis is associated with central sensitization: a psychophysical controlled
study. J Pain, 2003. 4: 372.
https://pubmed.ncbi.nlm.nih.gov/14622679
38. Vincent, K., et al. Dysmenorrhoea is associated with central changes in otherwise healthy women.
Pain, 2011. 152: 1966.
https://pubmed.ncbi.nlm.nih.gov/21524851
39. Savidge, C.J., et al. Women's Perspectives on their Experiences of Chronic Pelvic Pain and Medical
Care. J Health Psychol, 1998. 3: 103.
https://pubmed.ncbi.nlm.nih.gov/22021346
40. Zondervan, K.T., et al. The community prevalence of chronic pelvic pain in women and associated
illness behaviour. Br J Gen Pract, 2001. 51: 541.
https://pubmed.ncbi.nlm.nih.gov/11462313
41. Price, J., et al. Attitudes of women with chronic pelvic pain to the gynaecological consultation:
a qualitative study. BJOG, 2006. 113: 446.
https://pubmed.ncbi.nlm.nih.gov/16489938
42. Martin, C.E., et al. Catastrophizing: A predictor of persistent pain among women with endometriosis
at 1 year. Human Reprod, 2011. 26: 3078.
https://pubmed.ncbi.nlm.nih.gov/21900393
43. Riegel, B., et al. Assessing psychological factors, social aspects and psychiatric co-morbidity
associated with Chronic Prostatitis/Chronic Pelvic Pain Syndrome (CP/CPPS) in men - A systematic
review. J Psychosom Res, 2014. 77: 333.
https://pubmed.ncbi.nlm.nih.gov/25300538
44. Chung, S.D., et al. Association between chronic prostatitis/chronic pelvic pain syndrome and
anxiety disorder: a population-based study. PLoS ONE [Electronic Resource], 2013. 8.
https://pubmed.ncbi.nlm.nih.gov/23691256
45. Latthe, P., et al. Factors predisposing women to chronic pelvic pain: systematic review. BMJ, 2006.
332: 749.
https://pubmed.ncbi.nlm.nih.gov/16484239
46. Hilden, M., et al. A history of sexual abuse and health: a Nordic multicentre study. BJOG, 2004.
111: 1121.
https://pubmed.ncbi.nlm.nih.gov/15383115
47. McGowan, L., et al. Chronic pelvic pain: A meta-analytic review. Psychol Health, 1998. 13: 937.
https://www.tandfonline.com/doi/abs/10.1080/08870449808407441
48. Walker, E.A., et al. Psychiatric diagnoses and sexual victimization in women with chronic pelvic pain.
Psychosomatics, 1995. 36: 531.
https://pubmed.ncbi.nlm.nih.gov/7501783
49. Angst, J. Sexual problems in healthy and depressed persons. Int Clin Psychopharmacol, 1998.
13 Suppl 6: S1.
https://pubmed.ncbi.nlm.nih.gov/9728667
50. Nickel, J.C., et al. Childhood sexual trauma in women with interstitial cystitis/bladder pain
syndrome: a case control study. Can Urol Assoc J, 2011. 5: 410.
https://pubmed.ncbi.nlm.nih.gov/22154637
51. Schrepf, A., et al. Adverse Childhood Experiences and Symptoms of Urologic Chronic Pelvic Pain
Syndrome: A Multidisciplinary Approach to the Study of Chronic Pelvic Pain Research Network
Study. Ann Behav Med, 2018. 52: 865.
https://pubmed.ncbi.nlm.nih.gov/30212850
52. Paras, M.L., et al. Sexual abuse and lifetime diagnosis of somatic disorders: a systematic review
and meta-analysis. JAMA, 2009. 302: 550.
https://pubmed.ncbi.nlm.nih.gov/19654389

60 CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021


53. Campbell, R., et al. Gynecological health impact of sexual assault. Res Nurs Health, 2006. 29: 399.
https://pubmed.ncbi.nlm.nih.gov/16977640
54. Hu, J.C., et al. The association of abuse and symptoms suggestive of chronic prostatitis/chronic
pelvic pain syndrome: results from the Boston Area Community Health survey. J Gen Intern Med,
2007. 22: 1532.
https://pubmed.ncbi.nlm.nih.gov/17763912
55. Linley, J.E., et al. Understanding inflammatory pain: ion channels contributing to acute and chronic
nociception. Pflugers Arch, 2010. 459: 657.
https://pubmed.ncbi.nlm.nih.gov/20162302
56. Nickel, J.C., et al. Prevalence and impact of bacteriuria and/or urinary tract infection in interstitial
cystitis/painful bladder syndrome. Urology, 2010. 76: 799.
https://pubmed.ncbi.nlm.nih.gov/20573386
57. Tripp, D.A., et al. Sexual functioning, catastrophizing, depression, and pain, as predictors of quality
of life in women with interstitial cystitis/painful bladder syndrome. Urology, 2009. 73: 987.
https://pubmed.ncbi.nlm.nih.gov/19394494
58. Tripp, D.A., et al. Catastrophizing and pain-contingent rest predict patient adjustment in men with
chronic prostatitis/chronic pelvic pain syndrome. J Pain, 2006. 7: 697.
https://pubmed.ncbi.nlm.nih.gov/17018330
59. Whitaker, L.H., et al. An Exploratory Study into Objective and Reported Characteristics of
Neuropathic Pain in Women with Chronic Pelvic Pain. PLoS One, 2016. 11: e0151950.
https://pubmed.ncbi.nlm.nih.gov/27046128
60. Kutch, J.J., et al. Altered resting state neuromotor connectivity in men with chronic prostatitis/
chronic pelvic pain syndrome: A MAPP: Research Network Neuroimaging Study. Neuroimage Clin,
2015. 8: 493.
https://pubmed.ncbi.nlm.nih.gov/26106574
61. Abrams, P., et al. A new classification is needed for pelvic pain syndromes--are existing
terminologies of spurious diagnostic authority bad for patients? J Urol, 2006. 175: 1989.
https://pubmed.ncbi.nlm.nih.gov/16697782
62. Hanno, P., et al. Bladder Pain Syndrome Committee of the International Consultation on
Incontinence. Neurourol Urodyn, 2010. 29: 191.
https://pubmed.ncbi.nlm.nih.gov/20025029
63. Yoon, B.I., et al. Clinical courses following acute bacterial prostatitis. Prostate Int, 2013. 1: 89.
https://pubmed.ncbi.nlm.nih.gov/24223408
64. Giamberardino, M.A., et al. Viscero-visceral hyperalgesia: characterization in different clinical
models. Pain, 2010. 151: 307.
https://pubmed.ncbi.nlm.nih.gov/20638177
65. Pezet, S., et al. Neurotrophins: mediators and modulators of pain. Annu Rev Neurosci, 2006. 29: 507.
https://pubmed.ncbi.nlm.nih.gov/16776595
66. Cervero, F., et al. Understanding the signaling and transmission of visceral nociceptive events.
J Neurobiol, 2004. 61: 45.
https://pubmed.ncbi.nlm.nih.gov/15362152
67. Kobayashi, H., et al. Mechanism of pain generation for endometriosis-associated pelvic pain. Arch
Gynecol Obstet, 2014. 289: 13.
https://pubmed.ncbi.nlm.nih.gov/24121693
68. Melzack, R., et al. Central neuroplasticity and pathological pain. Ann N Y Acad Sci, 2001. 933: 157.
https://pubmed.ncbi.nlm.nih.gov/12000018
69. Fulbright, R.K., et al. Functional MR imaging of regional brain activation associated with the affective
experience of pain. AJR Am J Roentgenol, 2001. 177: 1205.
https://pubmed.ncbi.nlm.nih.gov/11641204
70. Rygh, L.J., et al. Cellular memory in spinal nociceptive circuitry. Scand J Psychol, 2002. 43: 153.
https://pubmed.ncbi.nlm.nih.gov/12004953
71. Malykhina, A.P. Neural mechanisms of pelvic organ cross-sensitization. Neuroscience, 2007. 149: 660.
https://pubmed.ncbi.nlm.nih.gov/17920206
72. Sanford, M.T., et al. The role of environmental stress on lower urinary tract symptoms. Curr Opin
Urol, 2017. 27: 268.
https://pubmed.ncbi.nlm.nih.gov/28376513
73. Binik, Y.M. The DSM diagnostic criteria for dyspareunia. Arch Sex Behav, 2010. 39: 292.
https://pubmed.ncbi.nlm.nih.gov/19830537

CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021 61


74. Bergeron, S., et al. Genital pain in women: Beyond interference with intercourse. Pain, 2011.
152: 1223.
https://pubmed.ncbi.nlm.nih.gov/21324589
75. Davis, S.N., et al. Sexual dysfunction and pelvic pain in men: a male sexual pain disorder? J Sex
Marital Ther, 2009. 35: 182.
https://pubmed.ncbi.nlm.nih.gov/19360518
76. Leserman, J., et al. Identification of diagnostic subtypes of chronic pelvic pain and how subtypes
differ in health status and trauma history. Am J Obstet Gynecol, 2006. 195: 554.
https://pubmed.ncbi.nlm.nih.gov/16769027
77. Meltzer-Brody, S., et al. Trauma and posttraumatic stress disorder in women with chronic pelvic
pain. Obstet Gynecol, 2007. 109: 902.
https://pubmed.ncbi.nlm.nih.gov/17400852
78. Iglesias-Rios, L., et al. Depression and Posttraumatic Stress Disorder Among Women with
Vulvodynia: Evidence from the Population-Based Woman to Woman Health Study. J Women's
Health, 2015. 24: 557.
https://pubmed.ncbi.nlm.nih.gov/25950702
79. Anderson, A.B., et al. Associations Between Penetration Cognitions, Genital Pain, and Sexual Well-
being in Women with Provoked Vestibulodynia. J Sex Med, 2016. 13: 444.
https://pubmed.ncbi.nlm.nih.gov/26853045
80. Roth, R.S., et al. Psychological factors and chronic pelvic pain in women: a comparative study with
women with chronic migraine headaches. Health Care Women Int, 2011. 32: 746.
https://pubmed.ncbi.nlm.nih.gov/21767098
81. Souza, P.P., et al. Qualitative research as the basis for a biopsychosocial approach to women with
chronic pelvic pain. J Psychosom Obstet Gynaecol, 2011. 32: 165.
https://pubmed.ncbi.nlm.nih.gov/21919820
82. Allaire, C., et al., History-taking, physical examination and psychological assessment. In: Jarrell JF,
Vilos GJ (editors) Consensus guidelines for the management of chronic pelvic pain., in J Obstet
Gynaecol Can. 2005. p. 869.
83. Toye, F., et al. A meta-ethnography of patients' experiences of chronic pelvic pain: struggling to
construct chronic pelvic pain as 'real'. J Adv Nurs, 2014. 70: 2713.
https://pubmed.ncbi.nlm.nih.gov/25081990
84. Dworkin, R.H., et al. Core outcome measures for chronic pain clinical trials: IMMPACT
recommendations. Pain, 2005. 113: 9.
https://pubmed.ncbi.nlm.nih.gov/15621359
85. Awad, S.A., et al. Long-term results and complications of augmentation ileocystoplasty for
idiopathic urge incontinence in women. Br J Urol, 1998. 81: 569.
https://pubmed.ncbi.nlm.nih.gov/9598629
86. Slocumb, J.C. Neurological factors in chronic pelvic pain: trigger points and the abdominal pelvic
pain syndrome. Am J Obstet Gynecol, 1984. 149: 536.
https://pubmed.ncbi.nlm.nih.gov/6234807
87. Barry, M.J., et al. Overlap of different urological symptom complexes in a racially and ethnically
diverse, community-based population of men and women. BJU Int, 2008. 101: 45.
https://pubmed.ncbi.nlm.nih.gov/17868419
88. Roberts, R.O., et al. Low agreement between previous physician diagnosed prostatitis and national
institutes of health chronic prostatitis symptom index pain measures. J Urol, 2004. 171: 279.
https://pubmed.ncbi.nlm.nih.gov/14665894
89. Krieger, J.N., et al. Epidemiology of prostatitis. Int J Antimicrob Agents, 2008. 31 Suppl 1: S85.
https://pubmed.ncbi.nlm.nih.gov/18164907
90. Mehik, A., et al. Epidemiology of prostatitis in Finnish men: a population-based cross-sectional
study. BJU Int, 2000. 86: 443.
https://pubmed.ncbi.nlm.nih.gov/10971269
91. Bade, J.J., et al. Interstitial cystitis in The Netherlands: prevalence, diagnostic criteria and
therapeutic preferences. J Urol, 1995. 154: 2035.
https://pubmed.ncbi.nlm.nih.gov/7500452
92. Burkman, R.T. Chronic pelvic pain of bladder origin: epidemiology, pathogenesis and quality of life.
J Reprod Med, 2004. 49: 225.
https://pubmed.ncbi.nlm.nih.gov/15088860
93. Curhan, G.C., et al. Epidemiology of interstitial cystitis: a population based study. J Urol, 1999.
161: 549.
https://pubmed.ncbi.nlm.nih.gov/9915446

62 CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021


94. Held, P., et al., Interstitial Cystitis. Epidemiology of interstitial cystitis. In: Hanno PM, Staskin DR,
Krane RJ, Wein AJ, eds. 1990, Springer Verlag: London.
95. Jones, C., et al. Prevalence of interstitial cystitisin the United States. Proc Am Urol Ass J Urol, 1994.
151 (Suppl). [No abstract available].
96. Leppilahti, M., et al. Prevalence of clinically confirmed interstitial cystitis in women: a population
based study in Finland. J Urol, 2005. 174: 581.
https://pubmed.ncbi.nlm.nih.gov/16006902
97. Oravisto, K.J. Epidemiology of interstitial cystitis. Ann Chir Gynaecol Fenn, 1975. 64: 75.
https://pubmed.ncbi.nlm.nih.gov/1137336
98. Parsons, C.L., et al. Prevalence of interstitial cystitis in young women. Urology, 2004. 64: 866.
https://pubmed.ncbi.nlm.nih.gov/15533465
99. Roberts, R.O., et al. Incidence of physician-diagnosed interstitial cystitis in Olmsted County: a
community-based study. BJU Int, 2003. 91: 181.
https://pubmed.ncbi.nlm.nih.gov/12581000
100. Temml, C., et al. Prevalence and correlates for interstitial cystitis symptoms in women participating
in a health screening project. Eur Urol, 2007. 51: 803.
https://pubmed.ncbi.nlm.nih.gov/16979286
101. Berry, S.H., et al. Prevalence of symptoms of bladder pain syndrome/interstitial cystitis among adult
females in the United States. J Urol, 2011. 186: 540.
https://pubmed.ncbi.nlm.nih.gov/21683389
102. Song, Y., et al. Prevalence and correlates of painful bladder syndrome symptoms in Fuzhou Chinese
women. Neurourol Urodyn, 2009. 28: 22.
https://pubmed.ncbi.nlm.nih.gov/18671294
103. Koziol, J.A., et al. Discrimination between the ulcerous and the nonulcerous forms of interstitial
cystitis by noninvasive findings. J Urol, 1996. 155: 87.
https://pubmed.ncbi.nlm.nih.gov/7490906
104. Messing, E.M., et al. Interstitial cystitis: early diagnosis, pathology, and treatment. Urology, 1978.
12: 381.
https://pubmed.ncbi.nlm.nih.gov/213864
105. Parsons, C. Interstitial cystitis: clinical manifestations and diagnostic criteria in over 200 cases.
Neurourol Urodyn, 1990. 9.
https://onlinelibrary.wiley.com/doi/abs/10.1002/nau.1930090302
106. Peeker, R., et al. Toward a precise definition of interstitial cystitis: further evidence of differences in
classic and nonulcer disease. J Urol, 2002. 167: 2470.
https://pubmed.ncbi.nlm.nih.gov/11992059
107. Mattox, T.F. Interstitial cystitis in adolescents and children: a review. J Pediatr Adolesc Gynecol,
2004. 17: 7.
https://pubmed.ncbi.nlm.nih.gov/15010032
108. Berghuis, J.P., et al. Psychological and physical factors involved in chronic idiopathic prostatitis.
J Psychosom Res, 1996. 41: 313.
https://pubmed.ncbi.nlm.nih.gov/8971661
109. Lee, S.W., et al. Adverse impact of sexual dysfunction in chronic prostatitis/chronic pelvic pain
syndrome. Urology, 2008. 71: 79.
https://pubmed.ncbi.nlm.nih.gov/18242370
110. Liang, C.Z., et al. Prevalence of sexual dysfunction in Chinese men with chronic prostatitis. BJU Int,
2004. 93: 568.
https://pubmed.ncbi.nlm.nih.gov/15008731
111. Bartoletti, R., et al. Prevalence, incidence estimation, risk factors and characterization of chronic
prostatitis/chronic pelvic pain syndrome in urological hospital outpatients in Italy: results of a
multicenter case-control observational study. J Urol, 2007. 178: 2411.
https://pubmed.ncbi.nlm.nih.gov/17937946
112. Gonen, M., et al. Prevalence of premature ejaculation in Turkish men with chronic pelvic pain
syndrome. J Androl, 2005. 26: 601.
https://pubmed.ncbi.nlm.nih.gov/16088036
113. Mehik, A., et al. Fears, sexual disturbances and personality features in men with prostatitis: a
population-based cross-sectional study in Finland. BJU Int, 2001. 88: 35.
https://pubmed.ncbi.nlm.nih.gov/11446842
114. Weidner, W., et al. Acute bacterial prostatitis and chronic prostatitis/chronic pelvic pain syndrome:
andrological implications. Andrologia, 2008. 40: 105.
https://pubmed.ncbi.nlm.nih.gov/18336460

CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021 63


115. van Ophoven, A., et al. Safety and efficacy of hyperbaric oxygen therapy for the treatment of
interstitial cystitis: a randomized, sham controlled, double-blind trial. J Urol, 2006. 176: 1442.
https://pubmed.ncbi.nlm.nih.gov/16952654
116. Rosen, R.C., et al. The international index of erectile function (IIEF): a multidimensional scale for
assessment of erectile dysfunction. Urology, 1997. 49: 822.
https://pubmed.ncbi.nlm.nih.gov/9187685
117. Anderson, R.U., et al. Sexual dysfunction in men with chronic prostatitis/chronic pelvic pain
syndrome: improvement after trigger point release and paradoxical relaxation training. J Urol, 2006.
176: 1534.
https://pubmed.ncbi.nlm.nih.gov/16952676
118. Trinchieri, A., et al. Prevalence of sexual dysfunction in men with chronic prostatitis/chronic pelvic
pain syndrome. Arch Ital Urol Androl, 2007. 79: 67.
https://pubmed.ncbi.nlm.nih.gov/17695411
119. Zondervan, K.T., et al. The prevalence of chronic pelvic pain in women in the United Kingdom:
a systematic review. Br J Obstet Gynaecol, 1998. 105: 93.
https://pubmed.ncbi.nlm.nih.gov/9442169
120. Grace, V., et al. Chronic pelvic pain in women in New Zealand: comparative well-being, comorbidity,
and impact on work and other activities. Health Care Women Int, 2006. 27: 585.
https://pubmed.ncbi.nlm.nih.gov/16844672
121. Pitts, M.K., et al. Prevalence and correlates of three types of pelvic pain in a nationally
representative sample of Australian women. Med J Aust, 2008. 189: 138.
https://pubmed.ncbi.nlm.nih.gov/18673099
122. Verit, F.F., et al. The prevalence of sexual dysfunction and associated risk factors in women with
chronic pelvic pain: a cross-sectional study. Arch Gynecol Obstet, 2006. 274: 297.
https://pubmed.ncbi.nlm.nih.gov/16705463
123. Florido, J., et al. Sexual behavior and findings on laparoscopy or laparotomy in women with severe
chronic pelvic pain. Eur J Obstet Gynecol Reprod Biol, 2008. 139: 233.
https://pubmed.ncbi.nlm.nih.gov/18403089
124. Ambler, N., et al. Sexual difficulties of chronic pain patients. Clin J Pain, 2001. 17: 138.
https://pubmed.ncbi.nlm.nih.gov/11444715
125. Loving, S., et al. Pelvic floor muscle dysfunctions are prevalent in female chronic pelvic pain: A
cross-sectional population-based study. Eur J Pain, 2014. 18: 1259.
https://pubmed.ncbi.nlm.nih.gov/24700500
126. Chiarioni, G., et al. Biofeedback is superior to electrogalvanic stimulation and massage for treatment
of levator ani syndrome. Gastroenterology, 2010. 138: 1321.
https://pubmed.ncbi.nlm.nih.gov/20044997
127. Rao, S.S., et al. ANMS-ESNM position paper and consensus guidelines on biofeedback therapy for
anorectal disorders. Neurogastroenterol Motil, 2015. 27: 594.
https://pubmed.ncbi.nlm.nih.gov/25828100
128. Zermann, D., et al. Chronic prostatitis: a myofascial pain syndrome? Infect Urol, 1999. 12: 84.
https://prostatitis.org/myofascial.html
129. Shoskes, D.A., et al. Muscle tenderness in men with chronic prostatitis/chronic pelvic pain
syndrome: the chronic prostatitis cohort study. J Urol, 2008. 179: 556.
https://pubmed.ncbi.nlm.nih.gov/18082223
130. Peters, K.M., et al. Prevalence of pelvic floor dysfunction in patients with interstitial cystitis. Urology,
2007. 70: 16.
https://pubmed.ncbi.nlm.nih.gov/17656199
131. Reissing, E.D., et al. Pelvic floor muscle functioning in women with vulvar vestibulitis syndrome.
J Psychosom Obstet Gynaecol, 2005. 26: 107.
https://pubmed.ncbi.nlm.nih.gov/16050536
132. Nickel, J.C., et al. Chronic Prostate Inflammation Predicts Symptom Progression in Patients with
Chronic Prostatitis/Chronic Pelvic Pain. J Urol, 2017. 198: 122.
https://pubmed.ncbi.nlm.nih.gov/28089730
133. Nickel , J., et al. Management of men diagnosed with chronic prostatitis/chronic pelvic pain
syndrome who have failed traditional management. Rev Urol, 2007. 9: 63.
https://pubmed.ncbi.nlm.nih.gov/17592539
134. Peters, K.M., et al. Childhood symptoms and events in women with interstitial cystitis/painful
bladder syndrome. Urology, 2009. 73: 258.
https://pubmed.ncbi.nlm.nih.gov/19036420

64 CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021


135. Rudick, C.N., et al. O-antigen modulates infection-induced pain states. PLoS One, 2012. 7: e41273.
https://pubmed.ncbi.nlm.nih.gov/22899994
136. Richter, B., et al. YKL-40 and mast cells are associated with detrusor fibrosis in patients diagnosed
with bladder pain syndrome/interstitial cystitis according to the 2008 criteria of the European
Society for the Study of Interstitial Cystitis. Histopathology, 2010. 57: 371.
https://pubmed.ncbi.nlm.nih.gov/20840668
137. Dundore, P.A., et al. Mast cell counts are not useful in the diagnosis of nonulcerative interstitial
cystitis. J Urol, 1996. 155: 885.
https://pubmed.ncbi.nlm.nih.gov/8583599
138. Peeker, R., et al. Recruitment, distribution and phenotypes of mast cells in interstitial cystitis. J Urol,
2000. 163: 1009.
https://pubmed.ncbi.nlm.nih.gov/10688040
139. Anderstrom, C.R., et al. Scanning electron microscopic findings in interstitial cystitis. Br J Urol,
1989. 63: 270.
https://pubmed.ncbi.nlm.nih.gov/2702424
140. Johansson, S.L., et al. Clinical features and spectrum of light microscopic changes in interstitial
cystitis. J Urol, 1990. 143: 1118.
https://pubmed.ncbi.nlm.nih.gov/2342171
141. Logadottir, Y.R., et al. Intravesical nitric oxide production discriminates between classic and
nonulcer interstitial cystitis. J Urol, 2004. 171: 1148.
https://pubmed.ncbi.nlm.nih.gov/14767289
142. Lokeshwar, V.B., et al. Urinary uronate and sulfated glycosaminoglycan levels: markers for interstitial
cystitis severity. J Urol, 2005. 174: 344.
https://pubmed.ncbi.nlm.nih.gov/15947687
143. Parsons, C.L., et al. Epithelial dysfunction in nonbacterial cystitis (interstitial cystitis). J Urol, 1991.
145: 732.
https://pubmed.ncbi.nlm.nih.gov/2005689
144. Parsons, C.L., et al. Successful therapy of interstitial cystitis with pentosanpolysulfate. J Urol, 1987.
138: 513.
https://pubmed.ncbi.nlm.nih.gov/2442417
145. Sanchez-Freire, V., et al. Acid-sensing channels in human bladder: expression, function and
alterations during bladder pain syndrome. J Urol, 2011. 186: 1509.
https://pubmed.ncbi.nlm.nih.gov/21855903
146. Hang, L., et al. Cytokine repertoire of epithelial cells lining the human urinary tract. J Urol, 1998.
159: 2185.
https://pubmed.ncbi.nlm.nih.gov/9598567
147. Parsons, C.L., et al. Cyto-injury factors in urine: a possible mechanism for the development of
interstitial cystitis. J Urol, 2000. 164: 1381.
https://pubmed.ncbi.nlm.nih.gov/10992419
148. Chelimsky, G., et al. Autonomic Testing in Women with Chronic Pelvic Pain. J Urol, 2016. 196: 429.
https://pubmed.ncbi.nlm.nih.gov/27026035
149. Charrua, A., et al. Can the adrenergic system be implicated in the pathophysiology of bladder pain
syndrome/interstitial cystitis? A clinical and experimental study. Neurourol Urodyn, 2015. 34: 489.
https://pubmed.ncbi.nlm.nih.gov/24375689
150. Alagiri, M., et al. Interstitial cystitis: unexplained associations with other chronic disease and pain
syndromes. Urology, 1997. 49: 52.
https://pubmed.ncbi.nlm.nih.gov/9146002
151. Buffington, C.A. Comorbidity of interstitial cystitis with other unexplained clinical conditions. J Urol,
2004. 172: 1242.
https://pubmed.ncbi.nlm.nih.gov/15371816
152. Erickson, D.R., et al. Nonbladder related symptoms in patients with interstitial cystitis. J Urol, 2001.
166: 557.
https://pubmed.ncbi.nlm.nih.gov/11458068
153. Warren, J.W., et al. Antecedent nonbladder syndromes in case-control study of interstitial cystitis/
painful bladder syndrome. Urology, 2009. 73: 52.
https://pubmed.ncbi.nlm.nih.gov/11378121
154. Weissman, M., et al. Interstitial Cystitis and Panic Disorder - A Potential Genetic Syndrome. Arch
Gen Psych, 2004. 61.
https://pubmed.ncbi.nlm.nih.gov/14993115

CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021 65


155. Warren, J.W., et al. Numbers and types of nonbladder syndromes as risk factors for interstitial
cystitis/painful bladder syndrome. Urology, 2011. 77: 313.
https://pubmed.ncbi.nlm.nih.gov/21295246
156. Peters, K.M., et al. Are ulcerative and nonulcerative interstitial cystitis/painful bladder syndrome 2
distinct diseases? A study of coexisting conditions. Urology, 2011. 78: 301.
https://pubmed.ncbi.nlm.nih.gov/21703668
157. Rab, M., et al. Anatomic variability of the ilioinguinal and genitofemoral nerve: implications for the
treatment of groin pain. Plast Reconstr Surg, 2001. 108: 1618.
https://pubmed.ncbi.nlm.nih.gov/11711938
158. Eklund, A., et al. Chronic pain 5 years after randomized comparison of laparoscopic and
Lichtenstein inguinal hernia repair. Br J Surg, 2010. 97: 600.
https://pubmed.ncbi.nlm.nih.gov/20186889
159. Nariculam, J., et al. A review of the efficacy of surgical treatment for and pathological changes in
patients with chronic scrotal pain. BJU Int, 2007. 99: 1091.
https://pubmed.ncbi.nlm.nih.gov/17244279
160. Manikandan, R., et al. Early and late morbidity after vasectomy: a comparison of chronic scrotal
pain at 1 and 10 years. BJU Int, 2004. 93: 571.
https://pubmed.ncbi.nlm.nih.gov/15008732
161. Leslie, T.A., et al. The incidence of chronic scrotal pain after vasectomy: a prospective audit. BJU
Int, 2007. 100: 1330.
https://pubmed.ncbi.nlm.nih.gov/17850378
162. Hallen, M., et al. Laparoscopic extraperitoneal inguinal hernia repair versus open mesh repair: long-
term follow-up of a randomized controlled trial. Surgery, 2008. 143: 313.
https://pubmed.ncbi.nlm.nih.gov/18291251
163. Grant, A.M., et al. Five-year follow-up of a randomized trial to assess pain and numbness after
laparoscopic or open repair of groin hernia. Br J Surg, 2004. 91: 1570.
https://pubmed.ncbi.nlm.nih.gov/15515112
164. Alfieri, S., et al. Influence of preservation versus division of ilioinguinal, iliohypogastric, and genital
nerves during open mesh herniorrhaphy: prospective multicentric study of chronic pain. Ann Surg,
2006. 243: 553.
https://pubmed.ncbi.nlm.nih.gov/16552209
165. Rönkä, K., et al. Role of orchiectomy in severe testicular pain after inguinal hernia surgery: audit of
the Finnish Patient Insurance Centre. Hernia, 2015. 19: 53.
https://pubmed.ncbi.nlm.nih.gov/23929499
166. Parsons, C.L. The role of a leaky epithelium and potassium in the generation of bladder symptoms
in interstitial cystitis/overactive bladder, urethral syndrome, prostatitis and gynaecological chronic
pelvic pain. BJU Int, 2011. 107: 370.
https://pubmed.ncbi.nlm.nih.gov/21176078
167. Parsons, C.L., et al. Intravesical potassium sensitivity in patients with interstitial cystitis and urethral
syndrome. Urology, 2001. 57: 428.
https://pubmed.ncbi.nlm.nih.gov/11248610
168. Kaur, H., et al. Urethral pain syndrome and its management. Obstet Gynecol Surv, 2007. 62: 348.
https://pubmed.ncbi.nlm.nih.gov/17425813
169. Gurel, H., et al. Urethral syndrome and associated risk factors related to obstetrics and gynecology.
Eur J Obstet Gynecol Reprod Biol, 1999. 83: 5.
https://pubmed.ncbi.nlm.nih.gov/10221602
170. Gornall, J. How mesh became a four letter word. BMJ, 2018. 363: k4137.
https://pubmed.ncbi.nlm.nih.gov/30305291
171. Heneghan, C., et al. Surgical mesh and patient safety. BMJ, 2018. 363: k4231.
https://pubmed.ncbi.nlm.nih.gov/30305286
172. Nilsson, C.G. Creating a gold standard surgical procedure: the development and implementation of
TVT. Int Urogynecol J, 2015. 26: 467.
https://pubmed.ncbi.nlm.nih.gov/25731721
173. Waltregny, D. TVT-O: a new gold standard surgical treatment of female stress urinary incontinence?
Eur Urol, 2013. 63: 879.
https://pubmed.ncbi.nlm.nih.gov/23352654
174. NICE. Urinary incontinence and pelvic organ prolapse in women: management. National Institute for
Health and Care Excellence Guideline, 2019. NG123.
https://www.nice.org.uk/guidance/ng123

66 CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021


175. Höfner, K., et al. [Use of synthetic slings and mesh implants in the treatment of female stress urinary
incontinence and prolapse : Statement of the Working Group on Urological Functional Diagnostics
and Female Urology of the Academy of the German Society of Urology]. Urologe A, 2020. 59: 65.
https://pubmed.ncbi.nlm.nih.gov/31741004
176. Keltie, K., et al. Complications following vaginal mesh procedures for stress urinary incontinence: an
8 year study of 92,246 women. Sci Rep, 2017. 7: 12015.
https://pubmed.ncbi.nlm.nih.gov/28931856
177. Wang, C., et al. Synthetic mid-urethral sling complications: Evolution of presenting symptoms over
time. Neurourol Urodyn, 2018.
https://pubmed.ncbi.nlm.nih.gov/29464783
178. Vancaillie, T., et al. Pain after vaginal prolapse repair surgery with mesh is a post-surgical neuropathy
which needs to be treated - and can possibly be prevented in some cases. Aust N Z J Obstet
Gynaecol, 2018.
https://pubmed.ncbi.nlm.nih.gov/29577243
179. Mellano, E.M., et al. The Role of Chronic Mesh Infection in Delayed-Onset Vaginal Mesh
Complications or Recurrent Urinary Tract Infections: Results From Explanted Mesh Cultures. Female
Pelvic Med Reconstr Surg, 2016. 22: 166.
https://pubmed.ncbi.nlm.nih.gov/26829350
180. Ubertazzi, E.P., et al. Long-term outcomes of transvaginal mesh (TVM) In patients with pelvic organ
prolapse: A 5-year follow-up. Eur J Obstet Gynecol Reprod Biol, 2018. 225: 90.
https://pubmed.ncbi.nlm.nih.gov/29680466
181. Mateu Arrom, L., et al. Pelvic Organ Prolapse Repair with Mesh: Mid-Term Efficacy and
Complications. Urol Int, 2018: 1.
https://pubmed.ncbi.nlm.nih.gov/29874667
182. Khatri, G., et al. Diagnostic Evaluation of Chronic Pelvic Pain. Phys Med Rehabil Clin N Am, 2017.
28: 477.
https://pubmed.ncbi.nlm.nih.gov/28676360
183. Bendavid, R., et al. A mechanism of mesh-related post-herniorrhaphy neuralgia. Hernia, 2016. 20: 357.
https://pubmed.ncbi.nlm.nih.gov/26597872
184. Treede, R.D., et al. A classification of chronic pain for ICD-11. Pain, 2015. 156: 1003.
https://pubmed.ncbi.nlm.nih.gov/25844555
185. Schug, S.A., et al. Risk stratification for the development of chronic postsurgical pain. Pain Rep,
2017. 2: e627.
https://pubmed.ncbi.nlm.nih.gov/29392241
186. Strik, C., et al. Risk of Pain and Gastrointestinal Complaints at 6Months After Elective Abdominal
Surgery. J Pain, 2019. 20: 38.
https://pubmed.ncbi.nlm.nih.gov/30107242
187. Bouman, E.A., et al. Reduced incidence of chronic postsurgical pain after epidural analgesia for
abdominal surgery. Pain Pract, 2014. 14: E76.
https://pubmed.ncbi.nlm.nih.gov/23758753
188. Mala, T., et al. Abdominal Pain After Roux-En-Y Gastric Bypass for Morbid Obesity. Scand J Surg,
2018. 107: 277.
https://pubmed.ncbi.nlm.nih.gov/29739280
189. Alsulaimy, M., et al. The Utility of Diagnostic Laparoscopy in Post-Bariatric Surgery Patients with
Chronic Abdominal Pain of Unknown Etiology. Obes Surg, 2017. 27: 1924.
https://pubmed.ncbi.nlm.nih.gov/28229315
190. Han, C., et al. Incidence and risk factors of chronic pain following hysterectomy among Southern
Jiangsu Chinese Women. BMC Anesthesiol, 2017. 17: 103.
https://pubmed.ncbi.nlm.nih.gov/28800726
191. Behera, M., et al. Laparoscopic findings, histopathologic evaluation, and clinical outcomes in
women with chronic pelvic pain after hysterectomy and bilateral salpingo-oophorectomy. J Minim
Invasive Gynecol, 2006. 13: 431.
https://pubmed.ncbi.nlm.nih.gov/16962527
192. Amid, P.K., et al. Surgical treatment of chronic groin and testicular pain after laparoscopic and open
preperitoneal inguinal hernia repair. J Am Coll Surg, 2011. 213: 531.
https://pubmed.ncbi.nlm.nih.gov/21784668
193. Hahn, L. Treatment of ilioinguinal nerve entrapment - a randomized controlled trial. Acta Obstet
Gynecol Scand, 2011. 90: 955.
https://pubmed.ncbi.nlm.nih.gov/21615360

CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021 67


194. Antolak, S.J., Jr., et al. Anatomical basis of chronic pelvic pain syndrome: the ischial spine and
pudendal nerve entrapment. Med Hypotheses, 2002. 59: 349.
https://pubmed.ncbi.nlm.nih.gov/12208168
195. Mahakkanukrauh, P., et al. Anatomical study of the pudendal nerve adjacent to the sacrospinous
ligament. Clin Anat, 2005. 18: 200.
https://pubmed.ncbi.nlm.nih.gov/15768420
196. Labat, J.J., et al. Diagnostic criteria for pudendal neuralgia by pudendal nerve entrapment (Nantes
criteria). Neurourol Urodyn, 2008. 27: 306.
https://pubmed.ncbi.nlm.nih.gov/17828787
197. Robert, R., et al. Anatomic basis of chronic perineal pain: role of the pudendal nerve. Surg Radiol
Anat, 1998. 20: 93.
https://pubmed.ncbi.nlm.nih.gov/9658526
198. Shafik, A. Pudendal canal syndrome as a cause of vulvodynia and its treatment by pudendal nerve
decompression. Eur J Obstet Gynecol Reprod Biol, 1998. 80: 215.
https://pubmed.ncbi.nlm.nih.gov/9846672
199. Amarenco, G., et al. Electrophysiological analysis of pudendal neuropathy following traction. Muscle
Nerve, 2001. 24: 116.
https://pubmed.ncbi.nlm.nih.gov/11150974
200. Goldet, R., et al. [Traction on the orthopedic table and pudendal nerve injury. Importance of
electrophysiologic examination]. Rev Chir Orthop Reparatrice Appar Mot, 1998. 84: 523.
https://pubmed.ncbi.nlm.nih.gov/9846326
201. Alevizon, S.J., et al. Sacrospinous colpopexy: management of postoperative pudendal nerve
entrapment. Obstet Gynecol, 1996. 88: 713.
https://pubmed.ncbi.nlm.nih.gov/8841264
202. Fisher, H.W., et al. Nerve injury locations during retropubic sling procedures. Int Urogynecol J, 2011.
22: 439.
https://pubmed.ncbi.nlm.nih.gov/21060989
203. Moszkowicz, D., et al. Where does pelvic nerve injury occur during rectal surgery for cancer?
Colorectal Dis, 2011. 13: 1326.
https://pubmed.ncbi.nlm.nih.gov/20718836
204. Ashton-Miller, J.A., et al. Functional anatomy of the female pelvic floor. Ann N Y Acad Sci, 2007.
1101: 266.
https://pubmed.ncbi.nlm.nih.gov/17416924
205. Amarenco, G., et al. [Perineal neuropathy due to stretching and urinary incontinence.
Physiopathology, diagnosis and therapeutic implications]. Ann Urol (Paris), 1990. 24: 463.
https://pubmed.ncbi.nlm.nih.gov/2176777
206. Fleming, M., et al. Sexuality and chronic pain. J Sex Educ Ther, 2001. 26: 204.
https://www.tandfonline.com/doi/abs/10.1080/01614576.2001.11074415
207. Chen, X., et al. The effect of chronic prostatitis/chronic pelvic pain syndrome (CP/CPPS) on erectile
function: A systematic review and meta-analysis. PLoS ONE, 2015. 10 (10) (no pagination).
https://pubmed.ncbi.nlm.nih.gov/26509575
208. Jacobsen, S.J., et al. Frequency of sexual activity and prostatic health: fact or fairy tale? Urology,
2003. 61: 348.
https://pubmed.ncbi.nlm.nih.gov/12597946
209. Tripp, D.A., et al. Prevalence, symptom impact and predictors of chronic prostatitis-like symptoms in
Canadian males aged 16-19 years. BJU Int, 2009. 103: 1080.
https://pubmed.ncbi.nlm.nih.gov/19007369
210. Pereira, R., et al. Sexual Functioning and Cognitions During Sexual Activity in Men With Genital
Pain: A Comparative Study. J Sex Marital Ther, 2016. 42: 602.
https://pubmed.ncbi.nlm.nih.gov/26548315
211. Muller, A., et al. Sexual dysfunction in the patient with prostatitis. Curr Opin Urol, 2005. 15: 404.
https://pubmed.ncbi.nlm.nih.gov/16205492
212. Smith, K.B., et al. Sexual and relationship functioning in men with chronic prostatitis/chronic pelvic
pain syndrome and their partners. Arch Sex Behav, 2007. 36: 301.
https://pubmed.ncbi.nlm.nih.gov/17186130
213. Gunter, J. Chronic pelvic pain: an integrated approach to diagnosis and treatment. Obstet Gynecol
Surv, 2003. 58: 615.
https://pubmed.ncbi.nlm.nih.gov/12972837

68 CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021


214. Latthe, P., et al. WHO systematic review of prevalence of chronic pelvic pain: a neglected
reproductive health morbidity. BMC Public Health, 2006. 6: 177.
https://pubmed.ncbi.nlm.nih.gov/16824213
215. Pearce, C., et al. A multidisciplinary approach to self care in chronic pelvic pain. Br J Nurs, 2007.
16: 82.
https://pubmed.ncbi.nlm.nih.gov/17353816
216. ter Kuile, M.M., et al. Sexual functioning in women with chronic pelvic pain: the role of anxiety and
depression. J Sex Med, 2010. 7: 1901.
https://pubmed.ncbi.nlm.nih.gov/19678881
217. Collett, B.J., et al. A comparative study of women with chronic pelvic pain, chronic nonpelvic pain
and those with no history of pain attending general practitioners. Br J Obstet Gynaecol, 1998.
105: 87.
https://pubmed.ncbi.nlm.nih.gov/9442168
218. McCabe, M.P., et al. Intercorrelations among general arousability, emerging and current sexual
desire, and severity of sexual dysfunction in women. Psychol Rep, 1989. 65: 147.
https://pubmed.ncbi.nlm.nih.gov/2780925
219. Flor, H., et al. The role of spouse reinforcement, perceived pain, and activity levels of chronic pain
patients. J Psychosom Res, 1987. 31: 251.
https://pubmed.ncbi.nlm.nih.gov/3585827
220. Paice, J. Sexuality and chronic pain. Am J Nurs, 2003. 103: 87.
https://pubmed.ncbi.nlm.nih.gov/12544064
221. Verit, F.F., et al. Validation of the female sexual function index in women with chronic pelvic pain.
J Sex Med, 2007. 4: 1635.
https://pubmed.ncbi.nlm.nih.gov/17888066
222. Maruta, T., et al. Chronic pain patients and spouses: marital and sexual adjustment. Mayo Clin Proc,
1981. 56: 307.
https://pubmed.ncbi.nlm.nih.gov/7230895
223. Hetrick, D.C., et al. Musculoskeletal dysfunction in men with chronic pelvic pain syndrome type III: a
case-control study. J Urol, 2003. 170: 828.
https://pubmed.ncbi.nlm.nih.gov/12913709
224. Clemens, J.Q., et al. Biofeedback, pelvic floor re-education, and bladder training for male chronic
pelvic pain syndrome. Urology, 2000. 56: 951.
https://pubmed.ncbi.nlm.nih.gov/11113739
225. Ishigooka, M., et al. Similarity of distributions of spinal c-Fos and plasma extravasation after acute
chemical irritation of the bladder and the prostate. J Urol, 2000. 164: 1751.
https://pubmed.ncbi.nlm.nih.gov/11025764
226. Liao, C.H., et al. Chronic Prostatitis/Chronic Pelvic Pain Syndrome is associated with Irritable Bowel
Syndrome: A Population-based Study. Sci Rep, 2016. 6: 26939.
https://pubmed.ncbi.nlm.nih.gov/27225866
227. Zondervan, K.T., et al. Prevalence and incidence of chronic pelvic pain in primary care: evidence
from a national general practice database. Br J Obstet Gynaecol, 1999. 106: 1149.
https://pubmed.ncbi.nlm.nih.gov/10549959
228. Drossman, D.A., et al. U.S. householder survey of functional gastrointestinal disorders. Prevalence,
sociodemography, and health impact. Dig Dis Sci, 1993. 38: 1569.
https://pubmed.ncbi.nlm.nih.gov/8359066
229. Prior, A., et al. Gynaecological consultation in patients with the irritable bowel syndrome. Gut, 1989.
30: 996.
https://pubmed.ncbi.nlm.nih.gov/2759494
230. Longstreth, G.F., et al. Irritable bowel syndrome in women having diagnostic laparoscopy or
hysterectomy. Relation to gynecologic features and outcome. Dig Dis Sci, 1990. 35: 1285.
https://pubmed.ncbi.nlm.nih.gov/2145139
231. Sperber, A.D., et al. Development of abdominal pain and IBS following gynecological surgery: a
prospective, controlled study. Gastroenterology, 2008. 134: 75.
https://pubmed.ncbi.nlm.nih.gov/18166349
232. Monnikes, H. Quality of life in patients with irritable bowel syndrome. J Clin Gastroenterol, 2011.
45 Suppl: S98.
https://pubmed.ncbi.nlm.nih.gov/21666428
233. Canavan, C., et al. Review article: the economic impact of the irritable bowel syndrome. Aliment
Pharmacol Ther, 2014. 40: 1023.
https://pubmed.ncbi.nlm.nih.gov/25199904

CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021 69


234. Morgan, C.J., et al. Ketamine use: a review. Addiction, 2012. 107: 27.
https://pubmed.ncbi.nlm.nih.gov/21777321
235. Lorencatto, C., et al. Depression in women with endometriosis with and without chronic pelvic pain.
Acta Obstet Gynecol Scand, 2006. 85: 88.
https://pubmed.ncbi.nlm.nih.gov/16521687
236. Howard, F.M. Chronic pelvic pain. Obstet Gynecol, 2003. 101: 594.
https://pubmed.ncbi.nlm.nih.gov/12636968
237. Fitzgerald, M.P., et al. Beyond the lower urinary tract: the association of urologic and sexual
symptoms with common illnesses. Eur Urol, 2007. 52: 407.
https://pubmed.ncbi.nlm.nih.gov/17382458
238. Davis, S.N., et al. Is a sexual dysfunction domain important for quality of life in men with urological
chronic pelvic pain syndrome? Signs "UPOINT" to yes. J Urol, 2013. 189: 146.
https://pubmed.ncbi.nlm.nih.gov/23164384
239. Cleeland, C.S. The Brief Pain Inventory User Guide. 2009.
https://www.mdanderson.org/documents/Departments-and-Divisions/Symptom-Research/BPI_
UserGuide.pdf
240. Turk, D.C., et al. Core outcome domains for chronic pain clinical trials: IMMPACT recommendations.
Pain, 2003. 106: 337.
https://pubmed.ncbi.nlm.nih.gov/14659516
241. Fall, M., et al. Chronic interstitial cystitis: a heterogeneous syndrome. J Urol, 1987. 137: 35.
https://pubmed.ncbi.nlm.nih.gov/3795363
242. Warren, J.W., et al. Evidence-based criteria for pain of interstitial cystitis/painful bladder syndrome in
women. Urology, 2008. 71: 444.
https://pubmed.ncbi.nlm.nih.gov/18342184
243. Rao, S.S., et al. Functional Anorectal Disorders. Gastroenterology, 2016.
https://pubmed.ncbi.nlm.nih.gov/27144630
244. Lacy, B.E., et al. Bowel Disorders. Gastroenterology, 2016. 150: 1393.
https://www.gastrojournal.org/article/S0016-5085(16)00222-5/abstract
245. McNaughton Collins, M., et al. Quality of life is impaired in men with chronic prostatitis: the Chronic
Prostatitis Collaborative Research Network. J Gen Intern Med, 2001. 16: 656.
https://pubmed.ncbi.nlm.nih.gov/11679032
246. Wenninger, K., et al. Sickness impact of chronic nonbacterial prostatitis and its correlates. J Urol,
1996. 155: 965.
https://pubmed.ncbi.nlm.nih.gov/8583619
247. Gerlinger, C., et al. Defining a minimal clinically important difference for endometriosis-associated
pelvic pain measured on a visual analog scale: analyses of two placebo-controlled, randomized
trials. Health Qual Life Outcomes, 2010. 8: 138.
https://pubmed.ncbi.nlm.nih.gov/21106059
248. Litwin, M.S., et al. The National Institutes of Health chronic prostatitis symptom index: development
and validation of a new outcome measure. Chronic Prostatitis Collaborative Research Network.
J Urol, 1999. 162: 369.
https://pubmed.ncbi.nlm.nih.gov/10411041
249. Mebust, W., et al., Symptom evaluation, quality of life and sexuality. In: 2ndConsultation on Benign
Prostatic Hyperplasia (BPH). 1993, Jersey, Channel Islands.
250. Lubeck, D.P., et al. Psychometric validation of the O'leary-Sant interstitial cystitis symptom index in
a clinical trial of pentosan polysulfate sodium. Urology, 2001. 57: 62.
https://pubmed.ncbi.nlm.nih.gov/11378052
251. Francis, C.Y., et al. The irritable bowel severity scoring system: a simple method of monitoring
irritable bowel syndrome and its progress. Aliment Pharmacol Ther, 1997. 11: 395.
https://pubmed.ncbi.nlm.nih.gov/9146781
252. Spiegel, B.M., et al. Characterizing abdominal pain in IBS: guidance for study inclusion criteria,
outcome measurement and clinical practice. Aliment Pharmacol Ther, 2010. 32: 1192.
https://pubmed.ncbi.nlm.nih.gov/20807217
253. Slieker-ten Hove, M.C., et al. Face validity and reliability of the first digital assessment scheme
of pelvic floor muscle function conform the new standardized terminology of the International
Continence Society. Neurourol Urodyn, 2009. 28: 295.
https://pubmed.ncbi.nlm.nih.gov/19090583
254. Wyndaele, J.J., et al. Reproducibility of digital testing of the pelvic floor muscles in men. Arch Phys
Med Rehabil, 1996. 77: 1179.
https://pubmed.ncbi.nlm.nih.gov/8931532

70 CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021


255. Davis, S.N., et al. Use of pelvic floor ultrasound to assess pelvic floor muscle function in urological
chronic pelvic pain syndrome in men. J Sex Med, 2011. 8: 3173.
https://pubmed.ncbi.nlm.nih.gov/21883952
256. Anderson, R.U., et al. Painful myofascial trigger points and pain sites in men with chronic prostatitis/
chronic pelvic pain syndrome. J Urol, 2009. 182: 2753.
https://pubmed.ncbi.nlm.nih.gov/19837420
257. Sanses, T.V., et al. The Pelvis and Beyond: Musculoskeletal Tender Points in Women With Chronic
Pelvic Pain. Clin J Pain, 2016. 32: 659.
https://pubmed.ncbi.nlm.nih.gov/26491938
258. Yang, C.C., et al. Physical Examination for Men and Women With Urologic Chronic Pelvic Pain
Syndrome: A MAPP (Multidisciplinary Approach to the Study of Chronic Pelvic Pain) Network Study.
Urology, 2018. 116: 23.
https://pubmed.ncbi.nlm.nih.gov/29604315
259. Antolak, S.J., Jr., et al. Therapeutic pudendal nerve blocks using corticosteroids cure pelvic pain
after failure of sacral neuromodulation. Pain Med, 2009. 10: 186.
https://pubmed.ncbi.nlm.nih.gov/19222779
260. Filler, A.G. Diagnosis and treatment of pudendal nerve entrapment syndrome subtypes: imaging,
injections, and minimal access surgery. Neurosurg Focus, 2009. 26: E9.
https://pubmed.ncbi.nlm.nih.gov/19323602
261. Labat, J.J., et al. [Electrophysiological studies of chronic pelvic and perineal pain]. Prog Urol, 2010.
20: 905.
https://pubmed.ncbi.nlm.nih.gov/21056364
262. Lee, J.C., et al. Neurophysiologic testing in chronic pelvic pain syndrome: a pilot study. Urology,
2001. 58: 246.
https://pubmed.ncbi.nlm.nih.gov/11489711
263. Lefaucheur, J.P., et al. What is the place of electroneuromyographic studies in the diagnosis and
management of pudendal neuralgia related to entrapment syndrome? Neurophysiol Clin, 2007.
37: 223.
https://pubmed.ncbi.nlm.nih.gov/17996810
264. Poldrack, R., et al. Scanning the Horizon: challenges and solutions for neuroimaging research.
bioRxiv, 2016.
https://pubmed.ncbi.nlm.nih.gov/28053326
265. Salomons, T.V., et al. THe “pain matrix” in pain-free individuals. JAMA Neurol, 2016. 73: 755.
https://pubmed.ncbi.nlm.nih.gov/27111250
266. Meares, E.M., et al. Bacteriologic localization patterns in bacterial prostatitis and urethritis. Invest
Urol, 1968. 5: 492.
https://pubmed.ncbi.nlm.nih.gov/4870505
267. Nickel, J.C. The Pre and Post Massage Test (PPMT): a simple screen for prostatitis. Tech Urol, 1997.
3: 38.
https://pubmed.ncbi.nlm.nih.gov/9170224
268. Nickel, J.C., et al. How does the pre-massage and post-massage 2-glass test compare to the
Meares-Stamey 4-glass test in men with chronic prostatitis/chronic pelvic pain syndrome? J Urol,
2006. 176: 119.
https://pubmed.ncbi.nlm.nih.gov/16753385
269. Nickel, J.C., et al. A randomized, placebo controlled, multicenter study to evaluate the safety and
efficacy of rofecoxib in the treatment of chronic nonbacterial prostatitis. J Urol, 2003. 169: 1401.
https://pubmed.ncbi.nlm.nih.gov/12629372
270. Manganaro, L., et al. Diffusion tensor imaging and tractography to evaluate sacral nerve root
abnormalities in endometriosis-related pain: a pilot study. Eur Radiol, 2014. 24: 95.
https://pubmed.ncbi.nlm.nih.gov/23982288
271. Howard, F.M. The role of laparoscopy as a diagnostic tool in chronic pelvic pain. Baillieres Best
Pract Res Clin Obstet Gynaecol, 2000. 14: 467.
https://pubmed.ncbi.nlm.nih.gov/10962637
272. Jacobson, T.Z., et al. Laparoscopic surgery for pelvic pain associated with endometriosis. Cochrane
Database Syst Rev, 2009: CD001300.
https://pubmed.ncbi.nlm.nih.gov/19821276
273. Porpora, M.G., et al. The role of laparoscopy in the management of pelvic pain in women of
reproductive age. Fertil Steril, 1997. 68: 765.
https://pubmed.ncbi.nlm.nih.gov/9389799

CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021 71


274. Seracchioli, R., et al. Cystoscopy-assisted laparoscopic resection of extramucosal bladder
endometriosis. J Endourol, 2002. 16: 663.
https://pubmed.ncbi.nlm.nih.gov/12490020
275. Wyndaele, J.J., et al. Cystoscopy and bladder biopsies in patients with bladder pain syndrome
carried out following ESSIC guidelines. Scand J Urol Nephrol, 2009. 43: 471.
https://pubmed.ncbi.nlm.nih.gov/19707951
276. Elcombe, S., et al. The psychological effects of laparoscopy on women with chronic pelvic pain.
Psychol Med, 1997. 27: 1041.
https://pubmed.ncbi.nlm.nih.gov/9300510
277. Onwude, J.L., et al. A randomised trial of photographic reinforcement during postoperative
counselling after diagnostic laparoscopy for pelvic pain. Eur J Obstet Gynecol Reprod Biol, 2004.
112: 89.
https://pubmed.ncbi.nlm.nih.gov/14687747
278. Peters, A.A., et al. A randomized clinical trial to compare two different approaches in women with
chronic pelvic pain. Obstet Gynecol, 1991. 77: 740.
https://pubmed.ncbi.nlm.nih.gov/1826544
279. Cole, E.E., et al. Are patient symptoms predictive of the diagnostic and/or therapeutic value of
hydrodistention? Neurourol Urodyn, 2005. 24: 638.
https://pubmed.ncbi.nlm.nih.gov/16208660
280. Lamale, L.M., et al. Symptoms and cystoscopic findings in patients with untreated interstitial
cystitis. Urology, 2006. 67: 242.
https://pubmed.ncbi.nlm.nih.gov/16442603
281. Ottem, D.P., et al. What is the value of cystoscopy with hydrodistension for interstitial cystitis?
Urology, 2005. 66: 494.
https://pubmed.ncbi.nlm.nih.gov/16140064
282. Shear, S., et al. Development of glomerulations in younger women with interstitial cystitis. Urology,
2006. 68: 253.
https://pubmed.ncbi.nlm.nih.gov/16904429
283. Tamaki, M., et al. Possible mechanisms inducing glomerulations in interstitial cystitis: relationship
between endoscopic findings and expression of angiogenic growth factors. J Urol, 2004. 172: 945.
https://pubmed.ncbi.nlm.nih.gov/15311005
284. Aihara, K., et al. Hydrodistension under local anesthesia for patients with suspected painful bladder
syndrome/interstitial cystitis: safety, diagnostic potential and therapeutic efficacy. Int J Urol, 2009.
16: 947.
https://pubmed.ncbi.nlm.nih.gov/19817916
285. Messing, E., et al. Associations among cystoscopic findings and symptoms and physical
examination findings in women enrolled in the Interstitial Cystitis Data Base (ICDB) Study. Urology,
1997. 49: 81.
https://pubmed.ncbi.nlm.nih.gov/9146006
286. Waxman, J.A., et al. Cystoscopic findings consistent with interstitial cystitis in normal women
undergoing tubal ligation. J Urol, 1998. 160: 1663.
https://pubmed.ncbi.nlm.nih.gov/9783927
287. Geurts, N., et al. Bladder pain syndrome: do the different morphological and cystoscopic features
correlate? Scand J Urol Nephrol, 2011. 45: 20.
https://pubmed.ncbi.nlm.nih.gov/20846081
288. Johansson, S.L., et al. Pathology of interstitial cystitis. Urol Clin North Am, 1994. 21: 55.
https://pubmed.ncbi.nlm.nih.gov/8284845
289. Ness, R.B., et al. Effectiveness of inpatient and outpatient treatment strategies for women with
pelvic inflammatory disease: results from the Pelvic Inflammatory Disease Evaluation and Clinical
Health (PEACH) Randomized Trial. Am J Obstet Gynecol, 2002. 186: 929.
https://pubmed.ncbi.nlm.nih.gov/12015517
290. Corey, L., et al. Genital herpes simplex virus infections: clinical manifestations, course, and
complications. Ann Intern Med, 1983. 98: 958.
https://pubmed.ncbi.nlm.nih.gov/6344712
291. Young, H., et al. Screening for treponemal infection by a new enzyme immunoassay. Genitourin
Med, 1989. 65: 72.
https://pubmed.ncbi.nlm.nih.gov/2666302
292. Culley, L., et al. The social and psychological impact of endometriosis on women's lives: A critical
narrative review. Human Reprod Update, 2013. 19: 625.
https://pubmed.ncbi.nlm.nih.gov/23884896

72 CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021


293. Souza, C.A., et al. Quality of life associated to chronic pelvic pain is independent of endometriosis
diagnosis--a cross-sectional survey. Health Qual Life Outcomes, 2011. 9.
https://pubmed.ncbi.nlm.nih.gov/21663624
294. Barri, P.N., et al. Endometriosis-associated infertility: surgery and IVF, a comprehensive therapeutic
approach. Reprod Biomed Online, 2010. 21: 179.
https://pubmed.ncbi.nlm.nih.gov/20541976
295. Fauconnier, A., et al. Relation between pain symptoms and the anatomic location of deep infiltrating
endometriosis. Fertil Steril, 2002. 78: 719.
https://pubmed.ncbi.nlm.nih.gov/12372446
296. Vercellini, P., et al. The effect of surgery for symptomatic endometriosis: the other side of the story.
Hum Reprod Update, 2009. 15: 177.
https://pubmed.ncbi.nlm.nih.gov/19136455
297. Vercellini, P., et al. Medical treatment for rectovaginal endometriosis: what is the evidence? Hum
Reprod, 2009. 24: 2504.
https://pubmed.ncbi.nlm.nih.gov/19574277
298. Walters, C., et al. Pelvic girdle pain in pregnancy. Aust J Gen Pract, 2018. 47: 439.
https://pubmed.ncbi.nlm.nih.gov/30114872
299. Khan, K.S., et al. MRI versus laparoscopy to diagnose the main causes of chronic pelvic pain in
women: a test-accuracy study and economic evaluation. Health Technol Assess, 2018. 22: 1.
https://pubmed.ncbi.nlm.nih.gov/30045805
300. Kaminski, P., et al. The usefulness of laparoscopy and hysteroscopy in the diagnostics and
treatment of infertility. Neuro Endocrinol Lett, 2006. 27: 813.
https://pubmed.ncbi.nlm.nih.gov/17187014
301. Hay-Smith, E.J. Therapeutic ultrasound for postpartum perineal pain and dyspareunia. Cochrane
Database Syst Rev, 2000: CD000495.
https://pubmed.ncbi.nlm.nih.gov/10796210
302. Cappell, J., et al. Clinical profile of persistent genito-pelvic postpartum pain. Midwifery, 2017. 50: 125.
https://pubmed.ncbi.nlm.nih.gov/28419979
303. Landau, R., et al. Chronic pain after childbirth. Int J Obstet Anesth, 2013. 22: 133.
https://pubmed.ncbi.nlm.nih.gov/23477888
304. Roovers, J.P., et al. A randomised controlled trial comparing abdominal and vaginal prolapse
surgery: effects on urogenital function. BJOG, 2004. 111: 50.
https://pubmed.ncbi.nlm.nih.gov/14687052
305. Niro, J., et al. [Postoperative pain after transvaginal repair of pelvic organ prolapse with or without
mesh]. Gynecol Obstet Fertil, 2010. 38: 648.
https://pubmed.ncbi.nlm.nih.gov/21030280
306. Vancaillie, T., et al. Sacral neuromodulation for pelvic pain and pelvic organ dysfunction: A case
series. Aust N Z J Obstet Gynaecol, 2018. 58: 102.
https://pubmed.ncbi.nlm.nih.gov/29218704
307. Eisenberg, V.H., et al. Ultrasound visualization of sacrocolpopexy polyvinylidene fluoride meshes
containing paramagnetic Fe particles compared with polypropylene mesh. Int Urogynecol J, 2018.
https://pubmed.ncbi.nlm.nih.gov/30083941
308. Kim, K.Y., et al. Translabial Ultrasound Evaluation of Pelvic Floor Structures and Mesh in the Urology
Office and Intraoperative Setting. Urology, 2018. 120: 267.
https://pubmed.ncbi.nlm.nih.gov/30031831
309. Sindhwani, N., et al. Short term post-operative morphing of sacrocolpopexy mesh measured by
magnetic resonance imaging. J Mech Behav Biomed Mater, 2018. 80: 269.
https://pubmed.ncbi.nlm.nih.gov/29455036
310. Zacharakis, D., et al. Pre- and postoperative magnetic resonance imaging (MRI) findings in patients
treated with laparoscopic sacrocolpopexy. Is it a safe procedure for all patients? Neurourol Urodyn,
2018. 37: 316.
https://pubmed.ncbi.nlm.nih.gov/28481045
311. Ford, A.C., et al. Irritable Bowel Syndrome. N Engl J Med, 2017. 376: 2566.
https://pubmed.ncbi.nlm.nih.gov/28657875
312. McGowan, L., et al. How do you explain a pain that can't be seen?: the narratives of women with
chronic pelvic pain and their disengagement with the diagnostic cycle. Br J Health Psychol, 2007.
12: 261.
https://pubmed.ncbi.nlm.nih.gov/17456285
313. European Association of Urology (EAU). EAU Survey: What do you tell your patients?
https://uroweb.org/news/?act=showfull&aid=246

CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021 73


314. Kanter, G., et al. Important role of physicians in addressing psychological aspects of interstitial
cystitis/bladder pain syndrome (IC/BPS): a qualitative analysis. Int Urogynecol J, 2017. 28: 249.
https://pubmed.ncbi.nlm.nih.gov/27581769
315. Loving, S., et al. Does evidence support physiotherapy management of adult female chronic pelvic
pain?. Scand J Pain, 2012. 3: 70.
https://pubmed.ncbi.nlm.nih.gov/29913781
316. Haugstad, G.K., et al. Mensendieck somatocognitive therapy as treatment approach to chronic
pelvic pain: results of a randomized controlled intervention study. Am J Obstet Gynecol, 2006.
194: 1303.
https://pubmed.ncbi.nlm.nih.gov/16647914
317. Fitzgerald, M.P., et al. Randomized multicenter feasibility trial of myofascial physical therapy for the
treatment of urological chronic pelvic pain syndromes. J Urol, 2013. 189: S75.
https://pubmed.ncbi.nlm.nih.gov/23234638
318. de las Penas, C., et al. Manual therapies in myofascial trigger point treatment: a systematic review.
J Bodyw Mov Ther, 2005. 9: 27.
https://www.somasimple.com/pdf_files/myofascial.pdf
319. Tough, E.A., et al. Acupuncture and dry needling in the management of myofascial trigger point
pain: a systematic review and meta-analysis of randomised controlled trials. Eur J Pain, 2009. 13: 3.
https://pubmed.ncbi.nlm.nih.gov/18395479
320. Oyama, I.A., et al. Modified Thiele massage as therapeutic intervention for female patients with
interstitial cystitis and high-tone pelvic floor dysfunction. Urology, 2004. 64: 862.
https://pubmed.ncbi.nlm.nih.gov/15533464
321. Langford, C.F., et al. Levator ani trigger point injections: An underutilized treatment for chronic pelvic
pain. Neurourol Urodyn, 2007. 26: 59.
https://pubmed.ncbi.nlm.nih.gov/17195176
322. FitzGerald, M.P., et al. Randomized multicenter clinical trial of myofascial physical therapy in
women with interstitial cystitis/painful bladder syndrome and pelvic floor tenderness. J Urol, 2012.
187: 2113.
https://pubmed.ncbi.nlm.nih.gov/22503015
323. Kellog-Spadt, S., et al., Role of the female urologist/urogynecologist. In: Women’s sexual function
and dysfunction: Study, diagnosis and treatment., 2006, Taylor and Francis: London.
324. Webster, D.C., et al. Use and effectiveness of physical self-care strategies for interstitial cystitis.
Nurse Pract, 1994. 19: 55.
https://pubmed.ncbi.nlm.nih.gov/7529390
325. Hayes, R.D., et al. What can prevalence studies tell us about female sexual difficulty and
dysfunction? J Sex Med, 2006. 3: 589.
https://pubmed.ncbi.nlm.nih.gov/16839314
326. Berghmans, B. Physiotherapy for pelvic pain and female sexual dysfunction: an untapped resource.
Int Urogynecol J, 2018. 29: 631.
https://pubmed.ncbi.nlm.nih.gov/29318334
327. Fuentes-Marquez, P., et al. Trigger Points, Pressure Pain Hyperalgesia, and Mechanosensitivity of
Neural Tissue in Women with Chronic Pelvic Pain. Pain Med, 2019. 20: 5.
https://pubmed.ncbi.nlm.nih.gov/29025041
328. Ghaderi, F., et al. Pelvic floor rehabilitation in the treatment of women with dyspareunia: a
randomized controlled clinical trial. Int Urogynecol J, 2019.
https://pubmed.ncbi.nlm.nih.gov/31286158
329. Rowe, E., et al. A prospective, randomized, placebo controlled, double-blind study of pelvic
electromagnetic therapy for the treatment of chronic pelvic pain syndrome with 1 year of followup.
J Urol, 2005. 173: 2044.
https://pubmed.ncbi.nlm.nih.gov/15879822
330. Kastner, C., et al. Cooled transurethral microwave thermotherapy for intractable chronic prostatitis--
results of a pilot study after 1 year. Urology, 2004. 64: 1149.
https://pubmed.ncbi.nlm.nih.gov/15596188
331. Montorsi, F., et al. Is there a role for transrectal microwave hyperthermia of the prostate in the
treatment of abacterial prostatitis and prostatodynia? Prostate, 1993. 22: 139.
https://pubmed.ncbi.nlm.nih.gov/8456052
332. Zimmermann, R., et al. Extracorporeal shock wave therapy for the treatment of chronic pelvic pain
syndrome in males: a randomised, double-blind, placebo-controlled study. Eur Urol, 2009. 56: 418.
https://pubmed.ncbi.nlm.nih.gov/19372000

74 CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021


333. Zeng, X.Y., et al. Extracorporeal shock wave treatment for non-inflammatory chronic pelvic pain
syndrome: A prospective, randomized and sham-controlled study. Chin Med J, 2012. 125: 114.
https://pubmed.ncbi.nlm.nih.gov/22340476
334. Vahdatpour B, et al. Efficacy of extracorporeal shock wave therapy for the treatment of chronic
pelvic pain syndrome: A randomized, controlled trial. ISRN Urology, 2013. 1.
https://pubmed.ncbi.nlm.nih.gov/24000311
335. Moayednia, A., et al. Long-term effect of extracorporeal shock wave therapy on the treatment of
chronic pelvic pain syndrome due to non bacterial prostatitis. J Res Med Sci, 2014. 19: 293.
https://pubmed.ncbi.nlm.nih.gov/25097599
336. Franco, J.V., et al. Non-pharmacological interventions for treating chronic prostatitis/chronic pelvic
pain syndrome. Cochrane Database Syst Rev, 2018. 1: Cd012551.
https://pubmed.ncbi.nlm.nih.gov/29372565
337. Lee, S.H., et al. Electroacupuncture relieves pain in men with chronic prostatitis/chronic pelvic pain
syndrome: three-arm randomized trial. Urology, 2009. 73: 1036.
https://pubmed.ncbi.nlm.nih.gov/19394499
338. Sahin, S., et al. Acupuncture relieves symptoms in chronic prostatitis/chronic pelvic pain syndrome:
A randomized, sham-controlled trial. Prostate Cancer Prostatic Dis, 2015. 18: 249.
https://pubmed.ncbi.nlm.nih.gov/25939517
339. Qin, Z., et al. Acupuncture for Chronic Prostatitis/Chronic Pelvic Pain Syndrome: A Randomized,
Sham Acupuncture Controlled Trial. J Urol, 2018. 200: 815.
https://pubmed.ncbi.nlm.nih.gov/29733836
340. Chang, S.C., et al. The efficacy of acupuncture in managing patients with chronic prostatitis/chronic
pelvic pain syndrome: A systemic review and meta-analysis. Neurourol Urodyn, 2016. 6: 6.
https://pubmed.ncbi.nlm.nih.gov/26741647
341. Qin, Z., et al. Systematic review of acupuncture for chronic prostatitis/chronic pelvic pain syndrome.
Medicine (United States), 2016. 95: e3095.
https://pubmed.ncbi.nlm.nih.gov/26986148
342. Nickel, J.C., et al. Sexual function is a determinant of poor quality of life for women with treatment
refractory interstitial cystitis. J Urol, 2007. 177: 1832.
https://pubmed.ncbi.nlm.nih.gov/17437831
343. Williams, A.C., et al. Psychological therapies for the management of chronic pain (excluding
headache) in adults. Cochrane Database Syst Rev, 2012. 11: CD007407.
https://pubmed.ncbi.nlm.nih.gov/23152245
344. Cheong, Y.C., et al. Non-surgical interventions for the management of chronic pelvic pain. Cochrane
Database Syst Rev, 2014. 3: CD008797.
https://pubmed.ncbi.nlm.nih.gov/24595586
345. Champaneria, R., et al. Psychological therapies for chronic pelvic pain: Systematic review of
randomized controlled trials. Acta Obstet Gynecol Scand, 2012. 91: 281.
https://pubmed.ncbi.nlm.nih.gov/22050516
346. Ariza-Mateos, M.J., et al. Effects of a Patient-Centered Graded Exposure Intervention Added to
Manual Therapy for Women With Chronic Pelvic Pain: A Randomized Controlled Trial. Arch Phys
Med Rehabil, 2019. 100: 9.
https://pubmed.ncbi.nlm.nih.gov/30312595
347. Brunahl, C.A., et al. Combined Cognitive-Behavioural and Physiotherapeutic Therapy for Patients
with Chronic Pelvic Pain Syndrome (COMBI-CPPS): study protocol for a controlled feasibility trial.
Trials, 2018. 19: 20.
https://pubmed.ncbi.nlm.nih.gov/29316946
348. Meissner, K., et al. Psychotherapy With Somatosensory Stimulation for Endometriosis-Associated
Pain: A Randomized Controlled Trial with 24-month follow-up. Obstet Gynecol Surv, 2017. 73: 163.
https://pubmed.ncbi.nlm.nih.gov/27741200
349. Mira, T.A.A., et al. Systematic review and meta-analysis of complementary treatments for women
with symptomatic endometriosis. Int J Gynaecol Obstet, 2018. 143: 2.
https://pubmed.ncbi.nlm.nih.gov/29944729
350. Farquhar, C.M., et al. A randomized controlled trial of medroxyprogesterone acetate and
psychotherapy for the treatment of pelvic congestion. Br J Obstet Gynaecol, 1989. 96: 1153.
https://pubmed.ncbi.nlm.nih.gov/2531611
351. Poleshuck, E.L., et al. Randomized controlled trial of interpersonal psychotherapy versus enhanced
treatment as usual for women with co-occurring depression and pelvic pain. J Psychosom Res,
2014. 77: 264.
https://pubmed.ncbi.nlm.nih.gov/25280823

CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021 75


352. Kanter, G., et al. Mindfulness-based stress reduction as a novel treatment for interstitial cystitis/
bladder pain syndrome: a randomized controlled trial. Int Urogynecol J, 2016. 26: 26.
https://pubmed.ncbi.nlm.nih.gov/27116196
353. Daniels, J.P., et al. Chronic pelvic pain in women. BMJ, 2010. 341: c4834.
https://pubmed.ncbi.nlm.nih.gov/20923840
354. Rosenbaum, T.Y. How well is the multidisciplinary model working? J Sex Med, 2011. 8: 2957.
https://pubmed.ncbi.nlm.nih.gov/22032406
355. Macea, D.D., et al. The efficacy of Web-based cognitive behavioral interventions for chronic pain: a
systematic review and meta-analysis. J Pain, 2010. 11: 917.
https://pubmed.ncbi.nlm.nih.gov/20650691
356. Shoskes, D.A., et al. Phenotypically directed multimodal therapy for chronic prostatitis/chronic
pelvic pain syndrome: a prospective study using UPOINT. Urology, 2010. 75: 1249.
https://pubmed.ncbi.nlm.nih.gov/20363491
357. Nickel, J.C., et al. Treatment of chronic prostatitis/chronic pelvic pain syndrome with tamsulosin: a
randomized double blind trial. J Urol, 2004. 171: 1594.
https://pubmed.ncbi.nlm.nih.gov/15017228
358. Zhao, W.P., et al. Celecoxib reduces symptoms in men with difficult chronic pelvic pain syndrome
(Category IIIA). Braz J Med Biol Res, 2009. 42: 963.
https://pubmed.ncbi.nlm.nih.gov/19787151
359. Bates, S.M., et al. A prospective, randomized, double-blind trial to evaluate the role of a short
reducing course of oral corticosteroid therapy in the treatment of chronic prostatitis/chronic pelvic
pain syndrome. BJU Int, 2007. 99: 355.
https://pubmed.ncbi.nlm.nih.gov/17313424
360. Cheah, P.Y., et al. Terazosin therapy for chronic prostatitis/chronic pelvic pain syndrome: a
randomized, placebo controlled trial. J Urol, 2003. 169: 592.
https://pubmed.ncbi.nlm.nih.gov/12544314
361. Gul, O., et al. Use of terazosine in patients with chronic pelvic pain syndrome and evaluation by
prostatitis symptom score index. Int Urol Nephrol, 2001. 32: 433.
https://pubmed.ncbi.nlm.nih.gov/11583367
362. Mehik, A., et al. Alfuzosin treatment for chronic prostatitis/chronic pelvic pain syndrome: a
prospective, randomized, double-blind, placebo-controlled, pilot study. Urology, 2003. 62: 425.
https://pubmed.ncbi.nlm.nih.gov/12946740
363. Evliyaoglu, Y., et al. Lower urinary tract symptoms, pain and quality of life assessment in chronic
non-bacterial prostatitis patients treated with alpha-blocking agent doxazosin; versus placebo. Int
Urol Nephrol, 2002. 34: 351.
https://pubmed.ncbi.nlm.nih.gov/12899226
364. Tugcu, V., et al. A placebo-controlled comparison of the efficiency of triple- and monotherapy in
category III B chronic pelvic pain syndrome (CPPS). Eur Urol, 2007. 51: 1113.
https://pubmed.ncbi.nlm.nih.gov/17084960
365. Chen, Y., et al. Effects of a 6-month course of tamsulosin for chronic prostatitis/chronic pelvic pain
syndrome: a multicenter, randomized trial. World J Urol, 2011. 29: 381.
https://pubmed.ncbi.nlm.nih.gov/20336302
366. Nickel, J.C., et al. A randomized placebo-controlled multicentre study to evaluate the safety and
efficacy of finasteride for male chronic pelvic pain syndrome (category IIIA chronic nonbacterial
prostatitis). BJU Int, 2004. 93: 991.
https://pubmed.ncbi.nlm.nih.gov/15142149
367. Anothaisintawee, T., et al. Management of chronic prostatitis/chronic pelvic pain syndrome: a
systematic review and network meta-analysis. JAMA, 2011. 305: 78.
https://pubmed.ncbi.nlm.nih.gov/21205969
368. Cohen, J.M., et al. Therapeutic intervention for chronic prostatitis/chronic pelvic pain syndrome (CP/
CPPS): A systematic review and meta-analysis. PLoS ONE, 2012. 7: e4194.
https://pubmed.ncbi.nlm.nih.gov/22870266
369. Nickel, J.C., et al. Alfuzosin and symptoms of chronic prostatitis-chronic pelvic pain syndrome. N
Engl J Med, 2008. 359: 2663.
https://pubmed.ncbi.nlm.nih.gov/19092152
370. Nickel, J.C., et al. Predictors of patient response to antibiotic therapy for the chronic prostatitis/
chronic pelvic pain syndrome: a prospective multicenter clinical trial. J Urol, 2001. 165: 1539.
https://pubmed.ncbi.nlm.nih.gov/11342913

76 CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021


371. Lee, J.C., et al. Prostate biopsy culture findings of men with chronic pelvic pain syndrome do not
differ from those of healthy controls. J Urol, 2003. 169: 584.
https://pubmed.ncbi.nlm.nih.gov/12544312
372. Alexander, R.B., et al. Ciprofloxacin or tamsulosin in men with chronic prostatitis/chronic pelvic pain
syndrome: a randomized, double-blind trial. Ann Intern Med, 2004. 141: 581.
https://pubmed.ncbi.nlm.nih.gov/15492337
373. Nickel, J.C., et al. Levofloxacin for chronic prostatitis/chronic pelvic pain syndrome in men: a
randomized placebo-controlled multicenter trial. Urology, 2003. 62: 614.
https://pubmed.ncbi.nlm.nih.gov/14550427
374. Zhou, Z., et al. Detection of nanobacteria infection in type III prostatitis. Urology, 2008. 71: 1091.
https://pubmed.ncbi.nlm.nih.gov/18538692
375. Thakkinstian, A., et al. alpha-blockers, antibiotics and anti-inflammatories have a role in the
management of chronic prostatitis/chronic pelvic pain syndrome. BJU Int, 2012. 110: 1014.
https://pubmed.ncbi.nlm.nih.gov/22471591
376. Leskinen, M., et al. Effects of finasteride in patients with inflammatory chronic pelvic pain syndrome:
a double-blind, placebo-controlled, pilot study. Urology, 1999. 53: 502.
https://pubmed.ncbi.nlm.nih.gov/10096374
377. Kaplan, S.A., et al. A prospective, 1-year trial using saw palmetto versus finasteride in the treatment
of category III prostatitis/chronic pelvic pain syndrome. J Urol, 2004. 171: 284.
https://pubmed.ncbi.nlm.nih.gov/14665895
378. Wagenlehner, F.M., et al. A pollen extract (Cernilton) in patients with inflammatory chronic prostatitis-
chronic pelvic pain syndrome: a multicentre, randomised, prospective, double-blind, placebo-
controlled phase 3 study. Eur Urol, 2009. 56: 544.
https://pubmed.ncbi.nlm.nih.gov/19524353
379. Cai, T., et al. Pollen extract in association with vitamins provides early pain relief in patients affected
by chronic prostatitis/chronic pelvic pain syndrome. Exp Ther Med, 2014. 8: 1032.
https://pubmed.ncbi.nlm.nih.gov/25187793
380. Cai, T., et al. The role of flower pollen extract in managing patients affected by chronic prostatitis/
chronic pelvic pain syndrome: a comprehensive analysis of all published clinical trials. BMC Urol,
2017. 17: 32.
https://pubmed.ncbi.nlm.nih.gov/28431537
381. Shoskes, D.A., et al. Quercetin in men with category III chronic prostatitis: a preliminary prospective,
double-blind, placebo-controlled trial. Urology, 1999. 54: 960.
https://pubmed.ncbi.nlm.nih.gov/10604689
382. Aboumarzouk, O.M., et al. Pregabalin for chronic prostatitis. Cochrane Database Syst Rev, 2012.
8: CD009063.
https://pubmed.ncbi.nlm.nih.gov/22895982
383. Pontari, M.A., et al. Pregabalin for the treatment of men with chronic prostatitis/chronic pelvic pain
syndrome: a randomized controlled trial. Arch Intern Med, 2010. 170: 1586.
https://pubmed.ncbi.nlm.nih.gov/20876412
384. Nickel, J.C., et al. Pentosan polysulfate sodium therapy for men with chronic pelvic pain syndrome:
a multicenter, randomized, placebo controlled study. J Urol, 2005. 173: 1252.
https://pubmed.ncbi.nlm.nih.gov/15758763
385. Gottsch, H.P., et al. A pilot study of botulinum toxin A for male chronic pelvic pain syndrome. Scand
J Urol Nephrol, 2011. 45: 72.
https://pubmed.ncbi.nlm.nih.gov/21062115
386. Falahatkar, S., et al. Transurethral intraprostatic injection of botulinum neurotoxin type A for the
treatment of chronic prostatitis/chronic pelvic pain syndrome: Results of a prospective pilot double-
blind and randomized placebo-controlled study. BJU Int, 2015. 116: 641.
https://pubmed.ncbi.nlm.nih.gov/25307409
387. Goldmeier, D., et al. Treatment of category III A prostatitis with zafirlukast: a randomized controlled
feasibility study. Int J STD AIDS, 2005. 16: 196.
https://pubmed.ncbi.nlm.nih.gov/15829018
388. Nickel, J.C., et al. Preliminary assessment of safety and efficacy in proof-of-concept, randomized
clinical trial of tanezumab for chronic prostatitis/chronic pelvic pain syndrome. Urology, 2012.
80: 1105.
https://pubmed.ncbi.nlm.nih.gov/23010344
389. McNaughton, C.O., et al. Allopurinol for chronic prostatitis. Cochrane Database Syst Rev, 2002:
CD001041.
https://pubmed.ncbi.nlm.nih.gov/12519549

CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021 77


390. Ziaee, A.M., et al. Effect of allopurinol in chronic nonbacterial prostatitis: a double blind randomized
clinical trial. Int Braz J Urol, 2006. 32: 181.
https://pubmed.ncbi.nlm.nih.gov/16650295
391. Theoharides, T.C. Hydroxyzine in the treatment of interstitial cystitis. Urol Clin North Am, 1994. 21: 113.
https://pubmed.ncbi.nlm.nih.gov/8284834
392. Seshadri, P., et al. Cimetidine in the treatment of interstitial cystitis. Urology, 1994. 44: 614.
https://pubmed.ncbi.nlm.nih.gov/7941209
393. Sant, G.R., et al. A pilot clinical trial of oral pentosan polysulfate and oral hydroxyzine in patients
with interstitial cystitis. J Urol, 2003. 170: 810.
https://pubmed.ncbi.nlm.nih.gov/12913705
394. Hanno, P.M., et al. Use of amitriptyline in the treatment of interstitial cystitis. J Urol, 1989. 141: 846.
https://pubmed.ncbi.nlm.nih.gov/2926877
395. Foster, H.E., Jr., et al. Effect of amitriptyline on symptoms in treatment naive patients with interstitial
cystitis/painful bladder syndrome. J Urol, 2010. 183: 1853.
https://pubmed.ncbi.nlm.nih.gov/20303115
396. Hwang, P., et al. Efficacy of pentosan polysulfate in the treatment of interstitial cystitis: a meta-
analysis. Urology, 1997. 50: 39.
https://pubmed.ncbi.nlm.nih.gov/9218016
397. Mulholland, S.G., et al. Pentosan polysulfate sodium for therapy of interstitial cystitis. A double-blind
placebo-controlled clinical study. Urology, 1990. 35: 552.
https://pubmed.ncbi.nlm.nih.gov/1693797
398. Fritjofsson, A., et al. Treatment of ulcer and nonulcer interstitial cystitis with sodium
pentosanpolysulfate: a multicenter trial. J Urol, 1987. 138: 508.
https://pubmed.ncbi.nlm.nih.gov/2442416
399. van Ophoven, A., et al. Safety and efficacy of concurrent application of oral pentosan polysulfate
and subcutaneous low-dose heparin for patients with interstitial cystitis. Urology, 2005. 66: 707.
https://pubmed.ncbi.nlm.nih.gov/16230121
400. Nickel, J.C., et al. Pentosan polysulfate sodium for treatment of interstitial cystitis/bladder pain
syndrome: insights from a randomized, double-blind, placebo controlled study. J Urol, 2015. 193: 857.
https://pubmed.ncbi.nlm.nih.gov/25245489
401. Oravisto, K.J., et al. Treatment of interstitial cystitis with immunosuppression and chloroquine
derivatives. Eur Urol, 1976. 2: 82.
https://pubmed.ncbi.nlm.nih.gov/971677
402. Forsell, T., et al. Cyclosporine in severe interstitial cystitis. J Urol, 1996. 155: 1591.
https://pubmed.ncbi.nlm.nih.gov/8627830
403. Moran, P.A., et al. Oral methotrexate in the management of refractory interstitial cystitis. Aust N Z J
Obstet Gynaecol, 1999. 39: 468.
https://pubmed.ncbi.nlm.nih.gov/10687766
404. Barua, J.M., et al. A systematic review and meta-analysis on the efficacy of intravesical therapy for
bladder pain syndrome/interstitial cystitis. Int Urogynecol J, 2016. 27: 1137.
https://pubmed.ncbi.nlm.nih.gov/26590137
405. Asklin, B., et al. Intravesical lidocaine in severe interstitial cystitis. Case report. Scand J Urol
Nephrol, 1989. 23: 311.
https://pubmed.ncbi.nlm.nih.gov/2595329
406. Giannakopoulos, X., et al. Chronic interstitial cystitis. Successful treatment with intravesical
idocaine. Arch Ital Urol Nefrol Androl, 1992. 64: 337.
https://pubmed.ncbi.nlm.nih.gov/1462157
407. Henry, R., et al. Absorption of alkalized intravesical lidocaine in normal and inflamed bladders: a
simple method for improving bladder anesthesia. J Urol, 2001. 165: 1900.
https://pubmed.ncbi.nlm.nih.gov/11371877
408. Parsons, C.L. Successful downregulation of bladder sensory nerves with combination of heparin
and alkalinized lidocaine in patients with interstitial cystitis. Urology, 2005. 65: 45.
https://pubmed.ncbi.nlm.nih.gov/15667861
409. Nickel, J.C., et al. Intravesical alkalinized lidocaine (PSD597) offers sustained relief from symptoms
of interstitial cystitis and painful bladder syndrome. BJU Int, 2009. 103: 910.
https://pubmed.ncbi.nlm.nih.gov/19021619
410. Cervigni, M., et al. A randomized, open-label, multicenter study of the efficacy and safety of
intravesical hyaluronic acid and chondroitin sulfate versus dimethyl sulfoxide in women with bladder
pain syndrome/interstitial cystitis. Neurourol Urodyn, 2017. 36: 1178.
https://pubmed.ncbi.nlm.nih.gov/27654012

78 CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021


411. Pyo, J.S., et al. Systematic Review and Meta-Analysis of Intravesical Hyaluronic Acid and
Hyaluronic Acid/Chondroitin Sulfate Instillation for Interstitial Cystitis/Painful Bladder Syndrome. Cell
Physiol Biochem, 2016. 39: 1618.
https://pubmed.ncbi.nlm.nih.gov/27627755
412. Parsons, C.L., et al. Treatment of interstitial cystitis with intravesical heparin. Br J Urol, 1994. 73: 504.
https://pubmed.ncbi.nlm.nih.gov/8012771
413. Kuo, H.C. Urodynamic results of intravesical heparin therapy for women with frequency urgency
syndrome and interstitial cystitis. J Formos Med Assoc, 2001. 100: 309.
https://pubmed.ncbi.nlm.nih.gov/11432309
414. Baykal, K., et al. Intravesical heparin and peripheral neuromodulation on interstitial cystitis. Urol Int,
2005. 74: 361.
https://pubmed.ncbi.nlm.nih.gov/15897705
415. Thilagarajah, R., et al. Oral cimetidine gives effective symptom relief in painful bladder disease: a
prospective, randomized, double-blind placebo-controlled trial. BJU Int, 2001. 87: 207.
https://pubmed.ncbi.nlm.nih.gov/11167643
416. Kelly, J.D., et al. Clinical response to an oral prostaglandin analogue in patients with interstitial
cystitis. Eur Urol, 1998. 34: 53.
https://pubmed.ncbi.nlm.nih.gov/9676414
417. Korting, G.E., et al. A randomized double-blind trial of oral L-arginine for treatment of interstitial
cystitis. J Urol, 1999. 161: 558.
https://pubmed.ncbi.nlm.nih.gov/9915448
418. Smith, S.D., et al. Improvement in interstitial cystitis symptom scores during treatment with oral
L-arginine. J Urol, 1997. 158: 703.
https://pubmed.ncbi.nlm.nih.gov/9258064
419. Wheeler, M.A., et al. Effect of long-term oral L-arginine on the nitric oxide synthase pathway in the
urine from patients with interstitial cystitis. J Urol, 1997. 158: 2045.
https://pubmed.ncbi.nlm.nih.gov/9366309
420. Lundberg, J.O., et al. Elevated nitric oxide in the urinary bladder in infectious and noninfectious
cystitis. Urology, 1996. 48: 700.
https://pubmed.ncbi.nlm.nih.gov/8911512
421. Cartledge, J.J., et al. A randomized double-blind placebo-controlled crossover trial of the efficacy of
L-arginine in the treatment of interstitial cystitis. BJU Int, 2000. 85: 421.
https://pubmed.ncbi.nlm.nih.gov/10691818
422. Ehren, I., et al. Effects of L-arginine treatment on symptoms and bladder nitric oxide levels in
patients with interstitial cystitis. Urology, 1998. 52: 1026.
https://pubmed.ncbi.nlm.nih.gov/9836549
423. Barbalias, G.A., et al. Interstitial cystitis: bladder training with intravesical oxybutynin. J Urol, 2000.
163: 1818.
https://pubmed.ncbi.nlm.nih.gov/10799190
424. van Ophoven, A., et al. The dual serotonin and noradrenaline reuptake inhibitor duloxetine for the
treatment of interstitial cystitis: results of an observational study. J Urol, 2007. 177: 552.
https://pubmed.ncbi.nlm.nih.gov/17222632
425. Shiraishi, K., et al. High Inguinal Microsurgical Denervation of the Spermatic Cord for Chronic
Scrotal Content Pain: A Novel Approach for Adult and Pediatric Patients. Urology, 2019. 131: 144.
https://pubmed.ncbi.nlm.nih.gov/31136771
426. Lee, J.Y., et al. Efficacy of vasectomy reversal according to patency for the surgical treatment of
postvasectomy pain syndrome. Int J Impot Res, 2012. 24: 202.
https://pubmed.ncbi.nlm.nih.gov/22622333
427. Hetta, D.F., et al. Pulsed Radiofrequency Treatment for Chronic Post-Surgical Orchialgia: A Double-
Blind, Sham-Controlled, Randomized Trial: Three-Month Results. Pain Physician, 2018. 21: 199.
https://pubmed.ncbi.nlm.nih.gov/29565950
428. West, A.F., et al. Epididymectomy is an effective treatment for scrotal pain after vasectomy. BJU Int,
2000. 85: 1097.
https://pubmed.ncbi.nlm.nih.gov/10848703
429. Sauvan, M., et al. [Medical treatment for the management of painful endometriosis without infertility:
CNGOF-HAS Endometriosis Guidelines]. Gynecol Obstet Fertil Senol, 2018. 46: 267.
https://pubmed.ncbi.nlm.nih.gov/29510966
430. Kamanli, A., et al. Comparison of lidocaine injection, botulinum toxin injection, and dry needling to
trigger points in myofascial pain syndrome. Rheumatol Int, 2005. 25: 604.
https://pubmed.ncbi.nlm.nih.gov/15372199

CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021 79


431. Ho, K.Y., et al. Botulinum toxin A for myofascial trigger point injection: a qualitative systematic
review. Eur J Pain, 2007. 11: 519.
https://pubmed.ncbi.nlm.nih.gov/17071119
432. Abbott, J.A., et al. Botulinum toxin type A for chronic pain and pelvic floor spasm in women: a
randomized controlled trial. Obstet Gynecol, 2006. 108: 915.
https://pubmed.ncbi.nlm.nih.gov/17012454
433. Zermann, D., et al. Perisphincteric injection of botulinum toxin type A. A treatment option for
patients with chronic prostatic pain? Eur Urol, 2000. 38: 393.
https://pubmed.ncbi.nlm.nih.gov/11025376
434. Jarvis, S.K., et al. Pilot study of botulinum toxin type A in the treatment of chronic pelvic pain
associated with spasm of the levator ani muscles. Aust N Z J Obstet Gynaecol, 2004. 44: 46.
https://pubmed.ncbi.nlm.nih.gov/15089868
435. Rao, S.S., et al. Clinical trial: effects of botulinum toxin on Levator ani syndrome--a double-blind,
placebo-controlled study. Aliment Pharmacol Ther, 2009. 29: 985.
https://pubmed.ncbi.nlm.nih.gov/19222415
436. Eckardt, V.F., et al. Treatment of proctalgia fugax with salbutamol inhalation. Am J Gastroenterol,
1996. 91: 686.
https://pubmed.ncbi.nlm.nih.gov/8677929
437. Atkin, G.K., et al. Patient characteristics and treatment outcome in functional anorectal pain. Dis
Colon Rectum, 2011. 54: 870.
https://pubmed.ncbi.nlm.nih.gov/21654255
438. Chey W.D., et al. Effects of 26 weeks of linaclotide treatment on adequate relief and IBS severity in
patients with irritable bowel syndrome with constipation. Gastroenterol, 2012. 142.
https://www.gastrojournal.org/article/S0016-5085(12)63175-8/abstract
439. de Vries, M., et al. Tetrahydrocannabinol Does Not Reduce Pain in Patients With Chronic Abdominal
Pain in a Phase 2 Placebo-controlled Study. Clin Gastroenterol Hepatol, 2017. 15: 1079.
https://pubmed.ncbi.nlm.nih.gov/27720917
440. Stones, R.W., et al. Interventions for treating chronic pelvic pain in women. Cochrane Database Syst
Rev, 2000: CD000387.
https://pubmed.ncbi.nlm.nih.gov/11034686
441. Remy, C., et al. State of the art of paracetamol in acute pain therapy. Curr Opin Anaesthesiol, 2006.
19: 562.
https://pubmed.ncbi.nlm.nih.gov/16960492
442. Moore, R.A., et al. Overview review: Comparative efficacy of oral ibuprofen and paracetamol
(acetaminophen) across acute and chronic pain conditions. Eur J Pain, 2015. 19: 1213.
https://pubmed.ncbi.nlm.nih.gov/25530283
443. Marjoribanks, J., et al. Nonsteroidal anti-inflammatory drugs for dysmenorrhoea. Cochrane
Database Syst Rev, 2010: CD001751.
https://pubmed.ncbi.nlm.nih.gov/20091521
444. Allen, C., et al. Nonsteroidal anti-inflammatory drugs for pain in women with endometriosis.
Cochrane Database Syst Rev, 2009: CD004753.
https://pubmed.ncbi.nlm.nih.gov/19370608
445. NICE, NCG 173. Neuropathic pain. The pharmacological management of neuropathic pain in adults
in non-specialist settings. 2013.
https://pubmed.ncbi.nlm.nih.gov/25577930
446. Baldessarini, R., Drugs and the treatment of psychiatric disorders. In: Goodman and Gilman’s the
pharmacological basis of therapeutics. Bunton L.L., Hilal-Dandan R., Knollmann B.C. eds. 1985,
New York.
447. Saarto, T., et al. Antidepressants for neuropathic pain. Cochrane Database Syst Rev, 2007:
CD005454.
https://pubmed.ncbi.nlm.nih.gov/17943857
448. Lunn, M.P., et al. Duloxetine for treating painful neuropathy or chronic pain. Cochrane Database Syst
Rev, 2009: CD007115.
https://pubmed.ncbi.nlm.nih.gov/19821395
449. Engel, C.C., Jr., et al. A randomized, double-blind crossover trial of sertraline in women with chronic
pelvic pain. J Psychosom Res, 1998. 44: 203.
https://pubmed.ncbi.nlm.nih.gov/9532549
450. England, P.H., Report of the review of the evidence for dependence on, and withdrawal from,
prescribed medicines. 2019.
https://www.gov.uk/government/publications/prescribed-medicines-review-report

80 CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021


451. Wiffen, P.J., et al. Carbamazepine for acute and chronic pain in adults. Cochrane Database Syst
Rev, 2011: CD005451.
https://pubmed.ncbi.nlm.nih.gov/21249671
452. Moore, R.A., et al. Gabapentin for chronic neuropathic pain and fibromyalgia in adults. Cochrane
Database Syst Rev, 2011: CD007938.
https://pubmed.ncbi.nlm.nih.gov/21412914
453. Sator-Katzenschlager, S.M., et al. Chronic pelvic pain treated with gabapentin and amitriptyline: a
randomized controlled pilot study. Wien Klin Wochenschr, 2005. 117: 761.
https://pubmed.ncbi.nlm.nih.gov/16416358
454. Lewis, S.C., et al. Gabapentin for the Management of Chronic Pelvic Pain in Women (GaPP1): A
Pilot Randomised Controlled Trial. PLoS ONE [Electronic Resource], 2016. 11.
https://pubmed.ncbi.nlm.nih.gov/27070434
455. Moore, R.A., et al. Pregabalin for acute and chronic pain in adults. Cochrane Database Syst Rev,
2009: CD007076.
https://pubmed.ncbi.nlm.nih.gov/19588419
456. IASP Statement on Opioids. 2019.
https://www.iasp-pain.org/Advocacy/OpioidPositionStatement?navItemNumber=7225
457. Noble, M., et al. Long-term opioid management for chronic noncancer pain. Cochrane Database
Syst Rev, 2010: CD006605.
https://pubmed.ncbi.nlm.nih.gov/20091598
458. Faculty of Pain Medicine, P., Opioids Aware: A resource for patients and healthcare professionals to
support prescribing of opioid 2015.
https://www.rcoa.ac.uk/faculty-of-pain-medicine/opioids-aware
459. Sandhu, H., et al. What interventions are effective to taper opioids in patients with chronic pain?
BMJ, 2018. 362: k2990.
https://pubmed.ncbi.nlm.nih.gov/30262590
460. Sign 136, Management of chronic pain A national clinical guideline 2019.
https://www.sign.ac.uk/our-guidelines/management-of-chronic-pain/
461. Mucke, M., et al. Cannabis-based medicines for chronic neuropathic pain in adults. Cochrane
Database Syst Rev, 2018. 3: CD012182.
https://pubmed.ncbi.nlm.nih.gov/29513392
462. Stockings, E., et al. Cannabis and cannabinoids for the treatment of people with chronic noncancer
pain conditions: a systematic review and meta-analysis of controlled and observational studies.
Pain, 2018. 159: 1932.
https://pubmed.ncbi.nlm.nih.gov/29847469
463. NICE, Cannabis-based medicinal products. NICE guideline [NG144] 2019.
https://www.nice.org.uk/guidance/ng144
464. Baranowski, A., et al., Urogenital Pain in Clinical Practice. 2008, New York.
465. Li, C.B., et al. The efficacy and safety of the ganglion impar block in chronic intractable pelvic and/
or perineal pain: A systematic review and meta-analysis. Int J Clin Exp Med, 2016. 9: 15746.
https://www.researchgate.net/publication/308138251
466. Eker, H.E., et al. Management of neuropathic pain with methylprednisolone at the site of nerve injury.
Pain Med, 2012. 13: 443.
https://pubmed.ncbi.nlm.nih.gov/22313580
467. Labat, J.J., et al. Adding corticosteroids to the pudendal nerve block for pudendal neuralgia: a
randomised, double-blind, controlled trial. Bjog, 2017. 124: 251.
https://pubmed.ncbi.nlm.nih.gov/27465823
468. Bolandard, F., et al. Nerve stimulator guided pudendal nerve blocks. Can J Anaesth, 2005. 52:
773; author reply 773.
https://pubmed.ncbi.nlm.nih.gov/16103396
469. Kim, S.H., et al. Nerve-stimulator-guided pudendal nerve block by pararectal approach. Colorectal
Dis, 2012. 14: 611.
https://pubmed.ncbi.nlm.nih.gov/21752174
470. Kovacs, P., et al. New, simple, ultrasound-guided infiltration of the pudendal nerve: ultrasonographic
technique. Dis Colon Rectum, 2001. 44: 1381.
https://pubmed.ncbi.nlm.nih.gov/11584221
471. Naja, M.Z., et al. Nerve-stimulator-guided repeated pudendal nerve block for treatment of pudendal
neuralgia. Eur J Anaesthesiol, 2006. 23: 442.
https://pubmed.ncbi.nlm.nih.gov/16573866

CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021 81


472. Peng, P.W., et al. Ultrasound-guided interventional procedures for patients with chronic pelvic pain -
a description of techniques and review of literature. Pain Phys, 2008. 11: 215.
https://pubmed.ncbi.nlm.nih.gov/18354713
473. Rigaud, J., et al. [Somatic nerve block in the management of chronic pelvic and perineal pain]. Prog
Urol, 2010. 20: 1072.
https://pubmed.ncbi.nlm.nih.gov/21056387
474. Romanzi, L. Techniques of pudendal nerve block. J Sex Med, 2010. 7: 1716.
https://pubmed.ncbi.nlm.nih.gov/20537059
475. Thoumas, D., et al. Pudendal neuralgia: CT-guided pudendal nerve block technique. Abdom
Imaging, 1999. 24: 309.
https://pubmed.ncbi.nlm.nih.gov/10227901
476. Rhame, E.E., et al. Successful treatment of refractory pudendal neuralgia with pulsed
radiofrequency. Pain Physician, 2009. 12: 633.
https://pubmed.ncbi.nlm.nih.gov/19461829
477. Fang, H., et al. Clinical effect and safety of pulsed radiofrequency treatment for pudendal neuralgia:
a prospective, randomized controlled clinical trial. J Pain Res, 2018. 11: 2367.
https://pubmed.ncbi.nlm.nih.gov/30410389
478. Fariello, J.Y., et al. Sacral neuromodulation stimulation for IC/PBS, chronic pelvic pain, and sexual
dysfunction. Int Urogynecol J, 2010. 21: 1553.
https://pubmed.ncbi.nlm.nih.gov/20972541
479. Cottrell, A.M., et al. Benefits and Harms of Electrical Neuromodulation for Chronic Pelvic Pain: A
Systematic Review. Eur Urol Focus, 2019.
https://pubmed.ncbi.nlm.nih.gov/31636030
480. Tutolo, M., et al. Efficacy and Safety of Sacral and Percutaneous Tibial Neuromodulation in Non-
neurogenic Lower Urinary Tract Dysfunction and Chronic Pelvic Pain: A Systematic Review of the
Literature. Eur Urol, 2018. 73: 406.
https://pubmed.ncbi.nlm.nih.gov/29336927
481. Smith, C.P., et al. Botulinum toxin a has antinociceptive effects in treating interstitial cystitis.
Urology, 2004. 64: 871.
https://pubmed.ncbi.nlm.nih.gov/15533466
482. Kuo, H.C., et al. Comparison of intravesical botulinum toxin type A injections plus hydrodistention
with hydrodistention alone for the treatment of refractory interstitial cystitis/painful bladder
syndrome. BJU Int, 2009. 104: 657.
https://pubmed.ncbi.nlm.nih.gov/19338543
483. Pinto, R., et al. Trigonal injection of botulinum toxin A in patients with refractory bladder pain
syndrome/interstitial cystitis. Eur Urol, 2010. 58: 360.
https://pubmed.ncbi.nlm.nih.gov/20227820
484. Kuo, Y.C., et al. Adverse Events of Intravesical Onabotulinum Toxin A Injection between Patients
with Overactive Bladder and Interstitial Cystitis-Different Mechanisms of Action of Botox on Bladder
Dysfunction? Toxins, 2016. 8.
https://pubmed.ncbi.nlm.nih.gov/26999201
485. Akiyama, Y., et al. Botulinum toxin type A injection for refractory interstitial cystitis: A randomized
comparative study and predictors of treatment response. Int J Urol, 2015. 22: 835.
https://pubmed.ncbi.nlm.nih.gov/26041274
486. Kuo, H.C., et al. Intravesical botulinum toxin-A injections reduce bladder pain of interstitial
cystitis/bladder pain syndrome refractory to conventional treatment - A prospective, multicenter,
randomized, double-blind, placebo-controlled clinical trial. Neurourol Urodyn, 2015. 24: 24.
https://pubmed.ncbi.nlm.nih.gov/25914337
487. Lee, C.L., et al. Long-term efficacy and safety of repeated intravescial onabotulinumtoxinA
injections plus hydrodistention in the treatment of interstitial cystitis/bladder pain syndrome. Toxins,
2015. 7: 4283.
https://pubmed.ncbi.nlm.nih.gov/26506388
488. Pinto, R., et al. Persistent therapeutic effect of repeated injections of onabotulinum toxin A in
refractory bladder pain syndrome/interstitial cystitis. J Urol, 2013. 189: 548.
https://pubmed.ncbi.nlm.nih.gov/23253961
489. Pinto, R.A., et al. Intratrigonal OnabotulinumtoxinA Improves Bladder Symptoms and Quality of Life
in Patients with Bladder Pain Syndrome/Interstitial Cystitis: A Pilot, Single Center, Randomized,
Double-Blind, Placebo Controlled Trial. J Urol, 2018. 199: 998.
https://pubmed.ncbi.nlm.nih.gov/29031769

82 CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021


490. Hanno, P.M., et al. Diagnosis and treatment of interstitial cystitis/bladder pain syndrome: AUA
guideline amendment. J Urol, 2015. 193: 1545.
https://pubmed.ncbi.nlm.nih.gov/25623737
491. Kerr, W.S., Jr. Interstitial cystitis: treatment by transurethral resection. J Urol, 1971. 105: 664.
https://pubmed.ncbi.nlm.nih.gov/4397018
492. Peeker, R., et al. Complete transurethral resection of ulcers in classic interstitial cystitis. Int
Urogynecol J Pelvic Floor Dysfunct, 2000. 11: 290.
https://pubmed.ncbi.nlm.nih.gov/11052564
493. Rofeim, O., et al. Use of the neodymium: YAG laser for interstitial cystitis: a prospective study.
J Urol, 2001. 166: 134.
https://pubmed.ncbi.nlm.nih.gov/11435840
494. Freiha, F.S., et al. The surgical treatment of intractable interstitial cystitis. J Urol, 1980. 123: 632.
https://pubmed.ncbi.nlm.nih.gov/7420547
495. Kim, H.J., et al. Efficacy and safety of augmentation ileocystoplasty combined with supratrigonal
cystectomy for the treatment of refractory bladder pain syndrome/interstitial cystitis with Hunner's
lesion. Int J Urol, 2014. 21 Suppl 1: 69.
https://pubmed.ncbi.nlm.nih.gov/24807503
496. Shirley, S.W., et al. Experiences with colocystoplasties, cecocystoplasties and ileocystoplasties in
urologic surgery: 40 patients. J Urol, 1978. 120: 165.
https://pubmed.ncbi.nlm.nih.gov/671623
497. von Garrelts, B. Interstitial cystitis: thirteen patients treated operatively with intestinal bladder
substitutes. Acta Chir Scand, 1966. 132: 436.
https://pubmed.ncbi.nlm.nih.gov/5972716
498. Webster, G.D., et al. The management of chronic interstitial cystitis by substitution cystoplasty.
J Urol, 1989. 141: 287.
https://pubmed.ncbi.nlm.nih.gov/2913346
499. Volkmer, B.G., et al. Cystectomy and orthotopic ileal neobladder: the impact on female sexuality.
J Urol, 2004. 172: 2353.
https://pubmed.ncbi.nlm.nih.gov/15538266
500. Shaikh, A., et al. Pregnancy after augmentation cystoplasty. J Pak Med Assoc, 2006. 56: 465.
https://pubmed.ncbi.nlm.nih.gov/17144396
501. Nurse, D.E., et al. The problems of substitution cystoplasty. Br J Urol, 1988. 61: 423.
https://pubmed.ncbi.nlm.nih.gov/3395801
502. Rössberger, J., et al. Long-term results of reconstructive surgery in patients with bladder pain
syndrome/interstitial cystitis: subtyping is imperative. Urology, 2007. 70: 638.
https://pubmed.ncbi.nlm.nih.gov/17991529
503. Peeker, R., et al. The treatment of interstitial cystitis with supratrigonal cystectomy and ileocystoplasty:
difference in outcome between classic and nonulcer disease. J Urol, 1998. 159: 1479.
https://pubmed.ncbi.nlm.nih.gov/9554337
504. Linn, J.F., et al. Treatment of interstitial cystitis: comparison of subtrigonal and supratrigonal
cystectomy combined with orthotopic bladder substitution. J Urol, 1998. 159: 774.
https://pubmed.ncbi.nlm.nih.gov/9474146
505. Elzawahri, A., et al. Urinary conduit formation using a retubularized bowel from continent urinary
diversion or intestinal augmentations: ii. Does it have a role in patients with interstitial cystitis?
J Urol, 2004. 171: 1559.
https://pubmed.ncbi.nlm.nih.gov/15017220
506. Zhao, Y., et al. Circumcision plus antibiotic, anti-inflammatory, and alpha-blocker therapy for
the treatment for chronic prostatitis/chronic pelvic pain syndrome: a prospective, randomized,
multicenter trial. World J Urol, 2015. 33: 617.
https://pubmed.ncbi.nlm.nih.gov/24980414
507. Chaudhari, R., et al. Microsurgical Denervation of Spermatic Cord for Chronic Idiopathic Orchialgia:
Long-Term Results from an Institutional Experience. World J Mens Health, 2019. 37: 78.
https://pubmed.ncbi.nlm.nih.gov/30209898
508. Oomen, R.J.A., et al. Prospective double-blind preoperative pain clinic screening before
microsurgical denervation of the spermatic cord in patients with testicular pain syndrome. Pain,
2014. 155: 1720.
https://pubmed.ncbi.nlm.nih.gov/24861586
509. Menconi, C., et al. Persistent anal and pelvic floor pain after PPH and STARR: surgical management
of the fixed scar staple line. Int J Colorectal Dis, 2016. 31: 41.
https://pubmed.ncbi.nlm.nih.gov/26248794

CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021 83


510. Molegraaf, M.J., et al. Twelve-year outcomes of laparoscopic adhesiolysis in patients with chronic
abdominal pain: A randomized clinical trial. Surgery, 2017. 161: 415.
https://pubmed.ncbi.nlm.nih.gov/27866713
511. Yoon, S.M., et al. Treatment of female urethral syndrome refractory to antibiotics. Yonsei Med J,
2002. 43: 644.
https://pubmed.ncbi.nlm.nih.gov/12402379
512. Costantini, E., et al. Treatment of urethral syndrome: a prospective randomized study with Nd:YAG
laser. Urol Int, 2006. 76: 134.
https://pubmed.ncbi.nlm.nih.gov/16493214
513. Ploteau, S., et al. [Minimal and mild endometriosis: Impact of the laparoscopic surgery on pelvic pain
and fertility. CNGOF-HAS Endometriosis Guidelines]. Gynecol Obstet Fertil Senol, 2018. 46: 273.
https://pubmed.ncbi.nlm.nih.gov/29510965
514. de Paula Andres, M., et al. The current management of deep endometriosis: a systematic review.
Minerva Ginecol, 2017. 69: 587.
https://pubmed.ncbi.nlm.nih.gov/28545293
515. Bautrant, E., et al. [Modern algorithm for treating pudendal neuralgia: 212 cases and 104
decompressions]. J Gynecol Obstet Biol Reprod (Paris), 2003. 32: 705.
https://pubmed.ncbi.nlm.nih.gov/15067894
516. Possover, M., et al. Laparoscopic neurolysis of the sacral plexus and the sciatic nerve for extensive
endometriosis of the pelvic wall. Minim Invasive Neurosurg, 2007. 50: 33.
https://pubmed.ncbi.nlm.nih.gov/17546541
517. Robert, R., et al. Decompression and transposition of the pudendal nerve in pudendal neuralgia: a
randomized controlled trial and long-term evaluation. Eur Urol, 2005. 47: 403.
https://pubmed.ncbi.nlm.nih.gov/15716208
518. Robert, R., et al. [Pudendal nerve surgery in the management of chronic pelvic and perineal pain].
Prog Urol, 2010. 20: 1084.
https://pubmed.ncbi.nlm.nih.gov/21056388
519. Duckett, J., et al. Mesh removal after vaginal surgery: what happens in the UK? Int Urogynecol J,
2017. 28: 989.
https://pubmed.ncbi.nlm.nih.gov/27924372
520. Lee, D., et al. Transvaginal mesh kits--how "serious" are the complications and are they reversible?
Urology, 2013. 81: 43.
https://pubmed.ncbi.nlm.nih.gov/23200966
521. Shah, K., et al. Surgical management of lower urinary mesh perforation after mid-urethral
polypropylene mesh sling: mesh excision, urinary tract reconstruction and concomitant pubovaginal
sling with autologous rectus fascia. Int Urogynecol J, 2013. 24: 2111.
https://pubmed.ncbi.nlm.nih.gov/23824269
522. Ramart, P., et al. The Risk of Recurrent Urinary Incontinence Requiring Surgery After Suburethral
Sling Removal for Mesh Complications. Urology, 2017. 106: 203.
https://pubmed.ncbi.nlm.nih.gov/28476681
523. Jong, K., et al. Is pain relief after vaginal mesh and/or sling removal durable long term? Int
Urogynecol J, 2018. 29: 859.
https://pubmed.ncbi.nlm.nih.gov/28695345
524. Hansen, B.L., et al. Long-term follow-up of treatment for synthetic mesh complications. Female
Pelvic Med Reconstr Surg, 2014. 20: 126.
https://pubmed.ncbi.nlm.nih.gov/24763152
525. Ridgeway, B., et al. Early experience with mesh excision for adverse outcomes after transvaginal
mesh placement using prolapse kits. Am J Obstet Gynecol, 2008. 199: 703 e1.
https://pubmed.ncbi.nlm.nih.gov/18845292
526. Gyang, A.N., et al. Managing chronic pelvic pain following reconstructive pelvic surgery with
transvaginal mesh. Int Urogynecol J, 2014. 25: 313.
https://pubmed.ncbi.nlm.nih.gov/24217793
527. Ferreira, M., et al. [Pelvic floor muscle training programmes: a systematic review]. Acta Med Port,
2011. 24: 309.
https://pubmed.ncbi.nlm.nih.gov/22011604

84 CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021


8. CONFLICT OF INTEREST
All members of the EAU Chronic Pelvic Pain Guidelines working panel have provided disclosure statements
on all relationships that they have that might be perceived to be a potential source of conflict of interest. This
information is publically accessible through the European Association of Urology website http://www.uroweb.
org/guidelines/. This document was developed with the financial support of the European Association of
Urology. No external sources of funding and support have been involved. The EAU is a non-profit organisation
and funding is limited to administrative assistance and travel and meeting expenses. No honoraria or other
reimbursements have been provided.

9. CITATION INFORMATION
The format in which to cite the EAU Guidelines will vary depending on the style guide of the journal in which the
citation appears. Accordingly, the number of authors or whether, for instance, to include the publisher, location,
or an ISBN number may vary.

The compilation of the complete Guidelines should be referenced as:


EAU Guidelines. Edn. presented at the EAU Annual Congress Milan 2021. ISBN 978-94-92671-13-4.

If a publisher and/or location is required, include:


EAU Guidelines Office, Arnhem, The Netherlands. http://uroweb.org/guidelines/compilations-of-all-guidelines/

References to individual guidelines should be structured in the following way:


Contributors’ names. Title of resource. Publication type. ISBN. Publisher and publisher location, year.

CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021 85


86 CHRONIC PELVIC PAIN - LIMITED UPDATE MARCH 2021
EAU Guidelines on
Neuro-Urology
B. Blok (Chair), D. Castro-Diaz,
G. Del Popolo, J. Groen, R. Hamid, G. Karsenty, T.M. Kessler,
J. Pannek (Vice-chair)
Guidelines Associates: H. Ecclestone, S. Musco,
B. Padilla-Fernández, A. Sartori, L.A. ‘t Hoen

© European Association of Urology 2020


TABLE OF CONTENTS PAGE
1. INTRODUCTION 4
1.1 Aim and objectives 4
1.2 Panel composition 4
1.3 Available publications 4
1.4 Publication history 4
1.5 Background 4

2. METHODS 4
2.1 Introduction 4
2.2 Review 5

3. THE GUIDELINE 5
3.1 Epidemiology, aetiology and pathophysiology 5
3.1.1 Introduction 5
3.2 Classification systems 7
3.2.1 Introduction 7
3.3 Diagnostic evaluation 7
3.3.1 Introduction 7
3.3.2 Classification systems 8
3.3.3 Timing of diagnosis and treatment 8
3.3.4 Patient history 8
3.3.4.1 Bladder diaries 9
3.3.5 Patient quality of life questionnaires 10
3.3.5.1 Available Questionnaires 10
3.3.6 Physical examination 11
3.3.6.1 Autonomic dysreflexia 11
3.3.6.2 Summary of evidence and recommendations for history taking and
physical examination 12
3.3.7 Urodynamics 13
3.3.7.1 Introduction 13
3.3.7.2 Urodynamic tests 13
3.3.7.3 Specialist uro-neurophysiological tests 14
3.3.7.4 Summary of evidence and recommendations for urodynamics
and uro-neurophysiology 14
3.3.8 Renal function 14
3.4 Disease management 15
3.4.1 Introduction 15
3.4.2 Non-invasive conservative treatment 15
3.4.2.1 Assisted bladder emptying - Credé manoeuvre, Valsalva manoeuvre,
triggered reflex voiding 15
3.4.2.2 Neuro-urological rehabilitation 15
3.4.2.2.1 Bladder rehabilitation including electrical stimulation 15
3.4.2.3 Drug treatment 16
3.4.2.3.1 Drugs for storage symptoms 16
3.4.2.3.2 Drugs for voiding symptoms 17
3.4.2.4 Summary of evidence and recommendations for drug treatments 17
3.4.2.5 Minimally invasive treatment 17
3.4.2.5.1 Catheterisation 17
3.4.2.5.2 Summary of evidence and recommendations for
catheterisation 18
3.4.2.5.3 Intravesical drug treatment 18
3.4.2.5.4 Summary of evidence and recommendations for
intravesical drug treatment 18
3.4.2.5.5 Botulinum toxin injections in the bladder 18
3.4.2.5.6 Bladder neck and urethral procedures 19
3.4.2.5.7 Summary of evidence and recommendations for botulinum
toxin A injections and bladder neck procedures 19
3.4.3 Surgical treatment 19

2 NEURO-UROLOGY - LIMITED UPDATE MARCH 2020


3.4.3.1 Bladder neck and urethral procedures 19
3.4.3.2 Denervation, deafferentation, sacral neuromodulation 21
3.4.3.3 Bladder covering by striated muscle 21
3.4.3.4 Bladder augmentation 21
3.4.3.5 Urinary diversion 21
3.4.3.6 Summary of evidence and recommendations for surgical treatment 22
3.5 Urinary tract infection in neuro-urological patients 22
3.5.1 Epidemiology, aetiology and pathophysiology 22
3.5.2 Diagnostic evaluation 23
3.5.3 Disease management 23
3.5.3.1 Recurrent UTI 23
3.5.3.2 Prevention 23
3.5.4 Summary of evidence and recommendations for the treatment of UTI 24
3.6 Sexual function and fertility 24
3.6.1 Erectile dysfunction 24
3.6.1.1 Phosphodiesterase type 5 inhibitors (PDE5Is) 24
3.6.1.2 Drug therapy other than PDE5Is 24
3.6.1.3 Mechanical devices 25
3.6.1.4 Intracavernous injections and intraurethral application 25
3.6.1.5 Sacral neuromodulation 25
3.6.1.6 Penile prostheses 25
3.6.1.7 Summary of evidence and recommendations for erectile dysfunction 25
3.6.2 Male fertility 25
3.6.2.1 Sperm quality and motility 26
3.6.2.2 Summary of evidence and recommendations for male fertility 26
3.6.3 Female sexuality 26
3.6.4 Female fertility 27
3.6.4.1 Summary of evidence and recommendation for female sexuality and
fertility 27
3.7 Follow-up 27
3.7.1 Introduction 27
3.7.2 Summary of evidence and recommendations for follow-up 28
3.8 Conclusions 28

4. REFERENCES 28

5. CONFLICT OF INTEREST 54

6. CITATION INFORMATION 54

NEURO-UROLOGY - LIMITED UPDATE MARCH 2020 3


1. INTRODUCTION
1.1 Aim and objectives
The European Association of Urology (EAU) Neuro-Urology Guidelines aim to provide information for clinical
practitioners on the incidence, definitions, diagnosis, therapy, and follow-up of neuro-urological disorders.
These Guidelines reflect the current opinion of experts in this specific pathology and represent a state-of-the
art reference for all clinicians, as of the publication date.
The terminology used and the diagnostic procedures advised throughout these Guidelines follow
the recommendations for investigations of the lower urinary tract (LUT) as published by the International
Continence Society (ICS) [1-3]. Readers are advised to consult other EAU Guidelines that may address different
aspects of the topics discussed in this document.
It must be emphasised that clinical guidelines present the best evidence available to the experts
but following guideline recommendations will not necessarily result in the best outcome. Guidelines can never
replace clinical expertise when making treatment decisions for individual patients, but rather help to focus
decisions - also taking personal values and preferences/individual circumstances of patients into account.
Guidelines are not mandates and do not purport to be a legal standard of care.

1.2 Panel composition


The EAU Neuro-Urology Guidelines Panel consists of an international multidisciplinary group of neuro-
urological experts. All experts involved in the production of this document have submitted potential conflict of
interest statements which can be viewed on the EAU website: http://www.uroweb.org/guideline/neuro-urology/.

1.3 Available publications


A quick reference document, the Pocket Guidelines, is available in print and as an app for iOS and Android
devices. These are abridged versions which may require consultation with the full text version. A guideline
summary has also been published in European Urology [4]. All are available through the EAU website:
­http://www.uroweb.org/guideline/neurourology/.

1.4 Publication history


The EAU published the first Neuro-Urology Guidelines in 2003 with updates in 2008, 2014, and 2017. This 2020
document represents a limited update of the 2019 publication. The literature was assessed for all chapters.

1.5 Background
The function of the LUT is mainly storage and voiding of urine, which is regulated by the nervous system that
coordinates the activity of the urinary bladder and bladder outlet. The part of the nervous system that regulates
LUT function is disseminated from the peripheral nerves in the pelvis to highly specialised cortical areas. Any
disturbance of the nervous system involved, can result in neuro-urological symptoms. The extent and location
of the disturbance will determine the type of LUT dysfunction, which can be symptomatic or asymptomatic.
Neuro-urological symptoms can cause a variety of long-term complications; the most significant being
deterioration of renal function. Since symptoms and long-term complications do not correlate [5], it is important
to identify patients with neuro-urological symptoms, and establish if they have a low or high risk of subsequent
complications. The risk of developing upper urinary tract (UUT) damage and renal failure is much lower in
patients with slowly progressive non-traumatic neurological disorders than in those with spinal cord injury or
spina bifida [6]. In summary, treatment and intensity of follow-up examinations are based on the type of neuro-
urological disorder and the underlying cause.

2. METHODS
2.1 Introduction
For the 2020 Neuro-Urology Guidelines, new and relevant evidence has been identified, collated and appraised
through a structured assessment of the literature. A broad and comprehensive literature search, covering all
sections of the Neuro-Urology Guidelines was performed. Databases searched included Medline, EMBASE,
and the Cochrane Libraries, covering a time frame between May 31st 2018 and 1st April 2019. A total of 754
unique records were identified, retrieved and screened for relevance. A detailed search strategy is available
online: http://uroweb.org/guideline/neuro-urology/?type=appendices-publications.

4 NEURO-UROLOGY - LIMITED UPDATE MARCH 2020


For each recommendation within the guidelines there is an accompanying online strength rating form, the
bases of which is a modified GRADE methodology [7, 8]. Each strength rating form addresses a number of key
elements namely:
1. the overall quality of the evidence which exists for the recommendation, references used in
this text are graded according to a classification system modified from the Oxford Centre for
Evidence-Based Medicine Levels of Evidence [9];
2. the magnitude of the effect (individual or combined effects);
3. the certainty of the results (precision, consistency, heterogeneity and other statistical or
study related factors);
4. the balance between desirable and undesirable outcomes;
5. the impact of patient values and preferences on the intervention;
6. the certainty of those patient values and preferences.

These key elements are the basis which panels use to define the strength rating of each recommendation. The
strength of each recommendation is represented by the words ‘strong’ or ‘weak’ [10]. The strength of each
recommendation is determined by the balance between desirable and undesirable consequences of alternative
management strategies, the quality of the evidence (including certainty of estimates), and nature and variability
of patient values and preferences.
Additional information can be found in the general Methodology section of this print, and online at
the EAU website; http://www.uroweb.org/guideline/. A list of associations endorsing the EAU Guidelines can
also be viewed online at the above address.

2.2 Review
Publications ensuing from panel-led systematic reviews have all been peer-reviewed. The 2015 Neuro-Urology
Guidelines were subject to peer review prior to publication.

3. THE GUIDELINE
3.1 Epidemiology, aetiology and pathophysiology
3.1.1 Introduction
Neuro-urological symptoms may be caused by a variety of diseases and events affecting the nervous system
controlling the LUT. The resulting neuro-urological symptoms depend predominantly on the location and the
extent of the neurological lesion. There are no exact figures on the overall prevalence of neuro-urological
disorders in the general population, but data are available on the prevalence of the underlying conditions and
the relative risk of these for the development of neuro-urological symptoms. It is important to note that the
majority of the data shows a very wide range of prevalence/incidence. This reflects the variability in the cohort
(e.g. early or late stage disease) and the frequently small sample sizes, resulting in a low level of evidence in
most published data (summarised in Table 1).

NEURO-UROLOGY - LIMITED UPDATE MARCH 2020 5


Table 1: Epidemiology of Neuro-Urological Disorders

Suprapontine and pontine lesions and diseases


Neurological Disease Frequency in General Population Type and Frequency of Neuro-
Urological Symptoms
Cerebrovascular accident 450 cases/100,000/yr (Europe) [11], Nocturia - overactive bladder (OAB)
(Strokes) 10% of cardiovascular mortality. - urgency urinary incontinence
(UUI) - detrusor overactivity (DO),
other patterns less frequent
[12]. 57-83% of neuro-urological
symptoms at 1 month post-stroke,
71-80% spontaneous recovery at 6
months [13]. Persistence of urinary
incontinence (UI) correlates with
poor prognosis [14].
Dementias: 6.4% of adults > 65 yrs [15]. OAB - UUI – DO 25% of
Alzheimer’s disease (80%), incontinence in Alzheimer’s
Vascular (10%), Other (10%). disease, > 25% in other dementias:
Lewy body, NPH, Binswanger,
Nasu-Hakola, Pick Disease [16].
Incontinence 3 times more frequent
in geriatric patients with dementia
than without [17].
Parkinsonian syndrome (PS) Second most prevalent Urinary symptoms affect 50%
Idiopathic Parkinson’s disease neurodegenerative disease after at onset, with urgency and
(IPD): 75-80% of PS. Alzheimer’s disease. nocturia being the most common.
Rising prevalence of IPD with age Patients with urinary symptoms at
[18]. presentation have worse disease
progression in Parkinson’s disease
[19].

Non-IPD: Parkinson’s-plus (18%): MSA is the most frequent non-IPD Infections account for a major cause
- Multiple system atrophy (MSA), PS. of mortality in MSA [20].
- Progressive supranuclear palsy,
- Corticobasal degeneration, Impaired detrusor contractility with
- Dementia with Lewy bodies. post-void residual (PVR) > 150 mL
Secondary Parkinson’s (2%) seems to be the urodynamic finding
distinguishing MSA from IPD [21-23].
Brain tumours 26.8/100,000/yr in adults (> 19 yrs), Incontinence occurs mainly in frontal
(17.9 benign, 8.9 malignant) [24]. location (part of frontal syndrome or
isolated in frontal location) [25].
Cerebral palsy Cerebral palsy: 3.1-3.6/1,000 in 46% of patients with cerebral palsy
children aged 8 yrs [26]. suffer from UI, with 85% of patients
having abnormal urodynamic studies
(NDO most common 59%). Upper
tract deterioration is rare (2.5%) [27].
Traumatic brain injury 235/100,000/yr [28]. 44% storage dysfunction,
38% voiding dysfunction,
60% urodynamic abnormalities [29].
Normal pressure hydrocephalus 0.5% of the population > 60, up to Classic triad of gait and cognitive
2.9% of those > 65 [30]. disturbance along with UI.
Incontinence affects 98-100% of
patients [30].

6 NEURO-UROLOGY - LIMITED UPDATE MARCH 2020


Lesions and diseases between caudal brainstem and sacral spinal cord
Spinal cord injury (SCI) Prevalence of traumatic SCI in Neurogenic detrusor overactivity
developed countries ranges from (NDO) and detrusor sphincter
280 to 906/million [31]. dyssynergia (DSD) (up to 95%) and
detrusor underactivity (up to 83%)
depending on the level of the lesion
[32].
Spina bifida (SB) Spina bifida 3-4/10,000 Bladder function is impaired in up
Lumbar and lumbosacral form are to 96% of SB patients [34]. Over
the most common (60%) [33]. 50% of patients are incontinent
[35]. Patients with open and closed
defects can have equally as severe
neurogenic lower urinary tract
dysfunction [36].
Lesions and diseases of the peripheral nervous system
Lumbar spine Male (5%) and female (3%) > 35 yr 26% difficulty to void and
Degenerative disease have had a lumbosciatic acontractile detrusor [37].
Disk prolapse episode related to disc prolapse. Detrusor underactivity (up to 83%)
Lumbar canal stenosis [32].

Incidence: approx. 5/100,000/yr Tarlov cysts: early sensation of


More common in females > 45 yr. filling (70%), NDO (33%), urethral
instability (33%) and stress urinary
incontinence (SUI) (33%) [38].
Iatrogenic pelvic nerve lesions Rectal cancer. After abdomino-perineal resection:
Cervical cancer (multimodal 50% urinary retention.
therapy, radiotherapy and surgery). After total mesorectal excision:
Endometriosis surgery. 10-30% voiding dysfunction [39].
Peripheral neuropathy Worldwide, prevalence of Urgency/frequency +/- incontinence
Diabetes pharmacologically treated diabetes [41].
8.3% [40]. Hyposensitive and detrusor
Other causes of peripheral underactivity at later phase [41].
neuropathy causing neuro-
urological symptoms: alcohol
abuse; lumbosacral zona and
genital herpes; Guillain Barré
syndrome.
Disseminated central diseases
Multiple sclerosis (MS) Prevalence: 83/100,000 in Europe 10% of MS patients present with
[42]. voiding dysfunction at disease onset,
75% of patients will develop it after
10 yrs of MS [43].
DO: 86% [43].
DSD: 35% [43].
Detrusor underactivity: 25% [43].

3.2 Classification systems


3.2.1 Introduction
Relevant definitions can be found in the general ICS standardisation reports [2, 3, 44]. Supplementary online
Tables S1 and S2 list the definitions from these references, partly adapted, and other definitions considered
useful for clinical practice: https://uroweb.org/guideline/neuro-urology/?type=appendices-publications.

3.3 Diagnostic evaluation


3.3.1 Introduction
The normal physiological function of the LUT depends on an intricate interplay between the sensory and motor
nervous systems. When diagnosing neuro-urological symptoms, the aim is to describe the type of dysfunction
involved. A thorough medical history, physical examination and bladder diary are mandatory before any
additional diagnostic investigations can be planned. Results of the initial evaluation are used to decide the
patient’s long-term treatment and follow-up.

NEURO-UROLOGY - LIMITED UPDATE MARCH 2020 7


3.3.2 Classification systems
The pattern of LUT dysfunction following neurological disease is determined by the site and nature of the
lesion. A very simple classification system for use in daily clinical practice to decide on the appropriate
therapeutic approach is provided in Figure 1 [6].

Figure 1: Patterns of lower urinary tract dysfunction following neurological disease

(A)

(B)

(C)

The pattern of LUT dysfunction following neurological disease is determined by the site and nature of the lesion.
Panel (A) denotes the region above the pons, panel (B) the region between the pons and the sacral cord and
panel (C) the sacral cord and infrasacral region. Figures on the right show the expected dysfunctional states of
the detrusor-sphincter system. Figure adapted from Panicker et al. [6] with permission from Elsevier.
PVR = post-void residual.

3.3.3 Timing of diagnosis and treatment


Early diagnosis and treatment are essential in both congenital and acquired neuro-urological disorders [45].
This helps to prevent irreversible changes within the LUT, even in the presence of normal reflexes [46, 47].
Furthermore, urological symptoms can be the presenting feature of neurological pathology [48, 49]. Early
intervention can prevent irreversible deterioration of the LUT and UUT [50]. Long term follow up (life-long) is
mandatory to assess risk of UUT damage, renal failure and bladder cancer [51-53].

3.3.4 Patient history


History taking should include past and present symptoms and disorders (Table 4). It is the cornerstone of
evaluation, as the answers will aid selection of diagnostic investigations and treatment options.

• In non-traumatic neuro-urological patients with an insidious onset, a detailed history may find that the
condition started in childhood or adolescence [54].
• Urinary history consists of symptoms associated with both urine storage and voiding.
• Bowel history is important because patients with neuro-urological symptoms may also have related
neurogenic bowel dysfunction [55].

8 NEURO-UROLOGY - LIMITED UPDATE MARCH 2020


• Sexual function may be impaired because of the neuro-urological condition [56].
• Special attention should be paid to possible warning signs and symptoms (e.g. pain, infection,
haematuria and fever) requiring further investigation.
• Patients with SCI usually find it difficult to report urinary tract infection (UTI)-related symptoms accurately
[57, 58].
• The presence of urinary, bowel and sexual symptoms without neurological symptoms could be
suggestive of an underlying neurological disease or condition.
• Ambulatory status after acute SCI does not predict presence or absence of unfavourable urodynamic
parameters [59].

Table 4: History taking in patients with suspected neuro-urological disorder

Past history
Childhood through to adolescence and into adulthood
Hereditary or familial risk factors
Specific female: menarche (age); this may suggest a metabolic disorder
Obstetric history
History of diabetes
Diseases, e.g. multiple sclerosis, parkinsonism, encephalitis, syphilis
Accidents and operations, especially those involving the spine and central nervous system
Present history
Present medication
Lifestyle (smoking, alcohol and drugs); may influence urinary, sexual and bowel function
Quality of life
Specific urinary history
Onset of urological history
Relief after voiding; to detect the extent of a neurological lesion in the absence of obstructive uropathy
Bladder sensation
Initiation of micturition (normal, precipitate, reflex, strain, Credé)
Interruption of micturition (normal, paradoxical, passive)
Enuresis
Mode and type of voiding (catheterisation)
Frequency, voided volume, incontinence, urgency episodes
Sexual history
Genital or sexual dysfunction symptoms
Sensation in genital area
Specific male: erection, (lack of) orgasm, ejaculation
Specific female: dyspareunia, (lack of) orgasm
Bowel history
Frequency and faecal incontinence
Desire to defecate
Defecation pattern
Rectal sensation
Initiation of defecation (digitation)
Neurological history
Acquired or congenital neurological condition
Mental status and comprehension
Neurological symptoms (somatic and sensory), with onset, evolution and any treatment
Spasticity or autonomic dysreflexia (especially in lesions at or above level Th 6)
Mobility and hand function

3.3.4.1 Bladder diaries


Bladder diaries provide data on the number of voids, voided volume, pad weight and incontinence and urgency
episodes [3, 60]. Although a 24-hour bladder diary (recording should be done for three consecutive days) is
reliable in women with UI [61, 62], no research has been done on bladder diaries in neuro-urological patients.
Nevertheless, bladder diaries are considered a valuable diagnostic tool.

NEURO-UROLOGY - LIMITED UPDATE MARCH 2020 9


3.3.5 Patient quality of life questionnaires
An assessment of the patient’s present and expected future quality of life (QoL) is important to evaluate the
effect of any therapy. Quality of life is an essential aspect of the overall management of neuro-urological
patients, for example when evaluating treatment related changes on a patient’s QoL [63]. The type of bladder
management has been shown to affect health-related QoL (HRQoL) in patients with SCI [64, 65] and MS [66],
as does the presence or absence of urinary and faecal incontinence [67]. Other research has also highlighted
the importance of urological treatment and its impact on the urodynamic functionality of the neuro-urological
patient in determining patient QoL [68].
In recent years a proliferation in the number of questionnaires to evaluate symptoms and QoL has
been seen. Condition-specific questionnaires can be used to assess symptom severity and the impact of
symptoms on QoL. A patient’s overall QoL can be assessed using generic questionnaires. It is important that
the questionnaire of choice has been validated in the neuro-urological population, and that it is available in the
language that it is to be used in.

3.3.5.1 Available Questionnaires


Three condition-specific questionnaires for urinary or bowel dysfunction and QoL have been developed
specifically for adult neuro-urological patients [69]. In MS and SCI patients the Qualiveen [70, 71] is validated
and can be used for urinary symptoms. A short form of the Qualiveen is available [70, 71] and it has been
translated into various languages [72-77]. Although several objective and subjective tools have been used to
assess the influence of neurogenic bladder on QoL in SCI, the Quality life index-SCI and Qualiveen are the only
validated condition-specific outcomes that have shown consistent sensitivity to neurogenic bladder [78]. The
Neurogenic Bladder Symptom Score (NBSS) has been validated in neurological patients to measure urinary
symptoms and their consequences [79, 80]. The QoL scoring tool related to Bowel Management (QoL-BM) [81]
can be used to assess bowel dysfunction in MS and SCI patients.
In addition, sixteen validated questionnaires that evaluate QoL and asses urinary symptoms as a
subscale or question in neuro-urological patients have been identified [82, 83] (Table 5). The condition-specific
Incontinence-Quality of Life (I-QoL) questionnaire which was initially developed for the non-neurological
population has now also been validated for neuro-urological patients [84].

A patient’s overall QoL can be assessed by generic HRQoL questionnaires, the most commonly used
being the I-QOL, King’s Health Questionnaire (KHQ), or the Short Form 36-item and 12-item Health Survey
Questionnaires (SF-36, SF-12) [69]. In addition, the quality-adjusted life year (QALY), quantifies outcomes, by
weighing years of life spent in a specified health state, adjusted by a factor representing the value placed by
society or patients on their specific health state [85].
No evidence was found for which validated questionnaires are the most appropriate for use, since
no quality criteria for validated questionnaires have been assessed [69].

Table 5: Patient questionnaires

Questionnaire Underlying neurological Bladder Bowel Sexual function


disorder
FAMS [86] MS X X
FILMS [87] MS X X
HAQUAMS [88] MS X X X
I-QOL [84] MS, SCI X X
MDS [89] MS X X
MSISQ-15 / MSISQ-19 [90, 91] MS X X X
MSQLI [92] MS X X X
MSQoL-54 [93] MS X X X
MSWDQ [94] MS X X
NBSS [95] MS, SCI, Congenital X
neurogenic bladder
QoL-BM [81] SCI X
Qualiveen/SF-Qualiveen [71, 96] MS, SCI X X
RAYS [97] MS X X
RHSCIR [98] SCI X X X
Fransceschini [97] SCI X X X

10 NEURO-UROLOGY - LIMITED UPDATE MARCH 2020


3.3.6 Physical examination
In addition to a detailed patient history, attention should be paid to possible physical and intellectual disabilities
with respect to the planned investigations [99, 100]. Neuro-urological status should be described as completely
as possible (Figure 2) [6]. Patients with a high spinal cord lesion or supraspinal neurological lesions may suffer
from a significant drop in blood pressure when moved into a sitting or standing position. All sensations and
reflexes in the urogenital area must be tested [6]. Furthermore, detailed testing of the anal sphincter and pelvic
floor functions must be performed (Figure 2) [6, 101]. It is essential to have this clinical information to reliably
interpret later diagnostic investigations.
Additionally, urinalysis, blood chemistry, ultrasonography, residual and free flowmetry and
incontinence quantification should be performed as part of the routine assessment of neuro-urological patients
[6, 102].

3.3.6.1 Autonomic dysreflexia


Autonomic dysreflexia (AD) is a sudden and exaggerated autonomic response to various stimuli in patients with
SCI or spinal dysfunction. It generally manifests at or above level Th 6. The stimulus can be distended bladder
or bowel. For example, iatrogenic stimuli during cystoscopy or urodynamics can trigger AD [103]. It can also
be secondary to sexual stimulation or a noxious stimulus, e.g. infected toe nail or pressure sore. Autonomic
dysreflexia is defined by an increase in systolic blood pressure > 20 mmHg from baseline [104] and can have
life-threatening consequences if not properly managed.

Figure 2: Lumbosacral dermatomes, cutaneous nerves, and reflexes

The physical examination includes testing sensations and reflexes mediated through the lower spinal cord.
Abnormal findings would suggest a lesion affecting the lumbosacral segments; mapping out distinct areas of
sensory impairment helps to further localise the site of the lesion. Distribution of dermatomes (areas of skin
mainly supplied by a single spinal nerve) and cutaneous nerves over the perianal region and back of the upper
thigh (A), the perineum [105] (B), male external genitalia [106] (C) and root values of lower spinal cord reflexes
(D). Figure adapted from Panicker et al. [6] with parts A-C adapted from Standring [107], both with permission
from Elsevier.

NEURO-UROLOGY - LIMITED UPDATE MARCH 2020 11


Table 6: Neurological items to be specified

Sensation S2-S5 (both sides)


Presence (increased/normal/reduced/absent)
Type (light touch/pin prick)
Affected dermatomes
Reflexes (increased/normal/reduced/absent)
Bulbocavernous reflex
Perianal/anal reflex
Knee and ankle reflexes
Plantar responses (Babinski)
Anal sphincter tone
Presence (increased/normal/reduced/absent)
Voluntary contractions of anal sphincter and pelvic muscles (increased/normal/reduced/absent)
Prostate palpation
Descensus (prolapse) of pelvic organs

3.3.6.2 Summary of evidence and recommendations for history taking and physical examination

Summary of evidence LE
Early diagnosis and treatment are essential in both congenital and acquired neuro-urological disorders 4
to prevent irreversible changes within the LUT.
An extensive general history is the basis of evaluation focusing on past and present symptoms 4
including urinary, sexual, bowel and neurological functions.
Assessment of present and expected future QoL is an essential aspect of the overall management of 2a
neuro-urological patients and is important to evaluate the effect of any therapy.
Quality of life assessment should be completed with validated QoL questionnaires for neuro-urological 1a
patients.
Bladder diaries provide data on the number of voids, voided volume, pad weight and incontinence and 3
urgency episodes.

Recommendations Strength rating


History taking
Take an extensive general history, concentrating on past and present symptoms. Strong
Take a specific history for each of the four mentioned functions - urinary, bowel, sexual and Strong
neurological.
Pay special attention to the possible existence of alarm signs (e.g. pain, infection, Strong
haematuria, fever) that warrant further specific diagnosis.
Assess quality of life when evaluating and treating the neuro-urological patient. Strong
Use available validated tools including the Qualiveen and I-QoL for urinary symptoms and Strong
the QoL-BM for bowel dysfunction in multiple sclerosis and spinal cord injury patients. In
addition, generic (SF-36 or KHQ) questionnaires can be used.
Use MSISQ-15 and MSISQ-19 to evaluate sexual function in multiple sclerosis patients. Strong
Physical examination
Acknowledge individual patient disabilities when planning further investigations. Strong
Describe the neurological status as completely as possible, sensations and reflexes in the Strong
urogenital area must all be tested.
Test the anal sphincter and pelvic floor functions. Strong
Perform urinalysis, blood chemistry, bladder diary, residual and free flowmetry, incontinence Strong
quantification and urinary tract imaging.
I-QoL = Incontinence Quality of Life Instrument; QoL-BM = Quality of Life Bowel Management scoring tool;
KHQ = King’s Health Questionnaire; SF-36 = Short Form 36-item Health Survey Questionnaires; MSISQ 15/19 =
Multiple Sclerosis Intimacy and Sexuality Questionnaire 15/19 question version.

12 NEURO-UROLOGY - LIMITED UPDATE MARCH 2020


3.3.7 Urodynamics
3.3.7.1 Introduction
Urodynamic investigation is the only method that can objectively assess the function and dysfunction of the
LUT. In neuro-urological patients, invasive urodynamic investigation is even more challenging than in general
patients. Any technical source of artefacts must be critically considered. It is essential to maintain the quality of
the urodynamic recording and its interpretation [1]. Same session repeat urodynamic investigations are crucial
in clinical decision making, since repeat measurements may yield completely different results [108].
In patients at risk of AD, it is advisable to measure blood pressure during the urodynamic study
[109, 110]. The rectal ampulla should be empty of stool before the start of the investigation. All urodynamic
findings must be reported in detail and performed, according to the ICS technical recommendations and
standards [1, 111].

3.3.7.2 Urodynamic tests


Free uroflowmetry and assessment of residual urine: Provides a first impression of the voiding function and
is compulsory prior to planning any invasive urodynamics in patients able to void. For reliable information, it
should be repeated at least two to three times [1]. Possible pathological findings include a low flow rate, low
voided volume, intermittent flow, hesitancy and residual urine. Care must be taken when assessing the results
in patients unable to void in a normal position, as both flow pattern and rate may be modified by inappropriate
positions.

Filling cystometry: This test is the only method for quantifying the patient’s filling function. The status of LUT
(C)
function must be documented during the filling phase. However, this technique has limited use as a solitary
procedure. It is much more effective combined with bladder pressure measurement during micturition and is
even more effective in video-urodynamics.
The bladder should be empty at the start of filling. A physiological filling rate should be used with
body-warm saline. Possible pathological findings include DO, low bladder compliance, abnormal bladder
sensations, incontinence, and an incompetent or relaxing urethra. There is some evidence that a bladder
capacity < 200 mL and detrusor pressures over 75 cm H2O are independent risk factors for UUT damage in
patients with SCI [51].

Detrusor leak point pressure [112]: Appears to have no use as a diagnostic tool. Some positive findings have
been reported [113-115], but sensitivity is too low to estimate the risk to the UUT or for secondary bladder
damage [116, 117].

Pressure flow study: Reflects the coordination between detrusor and urethra or pelvic floor during the voiding
phase. It is even more effective if combined with filling cystometry and video-urodynamics. Lower urinary
tract function must be recorded during the voiding phase. Possible pathological findings include detrusor
underactivity, bladder outlet obstruction (BOO), DSD, a high urethral resistance, and residual urine.
Most types of obstruction caused by neuro-urological disorders are due to DSD [118, 119], non-
relaxing urethra, or non-relaxing bladder neck [120, 121]. Pressure-flow analysis mainly assesses the amount of
mechanical obstruction caused by the urethra’s inherent mechanical and anatomical properties and has limited
value in patients with neuro-urological disorders.

Electromyography (EMG): Reflects the activity of the external urethral sphincter, the peri-urethral striated
musculature, the anal sphincter and the striated pelvic floor muscles. Correct interpretation may be difficult due
to artefacts introduced by other equipment. In the urodynamic setting, an EMG is useful as a gross indication
of the patient’s ability to control the pelvic floor. Possible pathological findings include inadequate recruitment
upon specific stimuli (e.g. bladder filling, involuntary detrusor contractions, onset of voiding, coughing, Valsalva
manoeuvre) suggesting a diagnosis of DSD [122].

Urethral pressure measurement: Has a very limited role in neuro-urological disorders. There is no consensus on
parameters indicating pathological findings [123].

Video-urodynamics: Is the combination of filling cystometry and pressure flow studies with imaging. It is
the optimum procedure for urodynamic investigation in neuro-urological disorders [5]. Possible pathological
findings include all those described in the cystometry and the pressure flow study sections, and any
morphological pathology of the LUT and reflux to the UUT [124].

Ambulatory urodynamics: This is the functional investigation of the urinary tract, which predominantly uses the
natural filling of the urinary tract to reproduce the patient’s normal activity. Although this type of study might be

NEURO-UROLOGY - LIMITED UPDATE MARCH 2020 13


considered when conventional urodynamics does not reproduce the patient’s symptoms, its role in the neuro-
urological patient still needs to be determined [125, 126].

Triggered tests during urodynamics: Lower urinary tract function can be provoked by coughing, triggered
voiding, or anal stretch. Fast-filling cystometry with cooled saline (the ‘ice water test’) will discriminate between
upper and lower motor neuron lesions [127, 128]. Patients with upper motor neuron lesions develop a detrusor
contraction if the detrusor is intact, while patients with lower motor neuron lesions do not. However, the test
does not seem to be fully discriminative in other types of patients [129].

Previously, a positive bethanechol test [130] (detrusor contraction > 25 cm H2O) was thought to indicate
detrusor denervation hypersensitivity and the muscular integrity of an acontractile detrusor. However, in
practice, the test has given equivocal results. A variation of this method was reported using intravesical
electromotive administration of the bethanechol [131], but there was no published follow-up. Currently, there is
no indication for this test.

3.3.7.3 Specialist uro-neurophysiological tests


The following tests are advised as part of the neurological work-up [132]:
• electromyography (in a neurophysiological setting) of pelvic floor muscles, urethral sphincter and/or anal
sphincter;
• nerve conduction studies of pudendal nerve;
• reflex latency measurements of bulbocavernosus and anal reflex arcs;
• evoked responses from clitoris or glans penis;
• sensory testing on bladder and urethra.

Other elective tests, for specific conditions, may become obvious during the work-up and urodynamic
investigations.

3.3.7.4 Summary of evidence and recommendations for urodynamics and uro-neurophysiology

Summary of evidence LE
Urodynamic investigation is the only method that can objectively assess the (dys-)function of the LUT. 2a
Video-urodynamics is the optimum procedure for urodynamic investigation in neuro-urological 4
disorders.
Specific uro-neurophysiological tests are elective procedures and should only be carried out in 4
specialised settings.

Recommendations Strength rating


Perform a urodynamic investigation to detect and specify lower urinary tract (dys-)function, Strong
use same session repeat measurement as it is crucial in clinical decision making.
Non-invasive testing is mandatory before invasive urodynamics is planned. Strong
Use video-urodynamics for invasive urodynamics in neuro-urological patients. If this is not Strong
available, then perform a filling cystometry continuing into a pressure flow study.
Use a physiological filling rate and body-warm saline. Strong

3.3.8 Renal function


In many patients with neuro-urological disorders, the UUT is at risk, particularly in patients who develop high
detrusor pressure during the filling phase. Although effective treatment can reduce this risk, there is still a
relatively high incidence of renal morbidity [133, 134]. Patients with SCI or SB have a higher risk of developing
renal failure compared with patients with slowly progressive non-traumatic neurological disorders, such as MS
and Parkinson’s disease (PD) [135].
Caregivers must be informed of this risk and instructed to watch carefully for any signs or
symptoms of a possible deterioration in the patient’s renal function. In patients with poor muscle mass cystatin
C based glomerular filtration rate (GFR) is more accurate in detecting chronic kidney disease than serum
creatinine estimated GFR [136, 137]. There are no high level evidence publications available which show the
optimal management to preserve renal function in these patients [138].

14 NEURO-UROLOGY - LIMITED UPDATE MARCH 2020


3.4 Disease management
3.4.1 Introduction
The primary aims for treatment of neuro-urological symptoms, and their priorities, are [139, 140]:
• protection of the UUT;
• achievement (or maintenance) of urinary continence;
• restoration of LUT function;
• improvement of the patient’s QoL.

Further considerations are the patient’s disability, cost-effectiveness, technical complexity and possible
complications [140].

Renal failure is the main mortality factor in SCI patients who survive the trauma [141, 142]. Keeping the
detrusor pressure during both the filling and voiding phases within safe limits significantly reduces the mortality
from urological causes in these patients [143-145] and has consequently become the top priority in the
treatment of patients with neuro-urological symptoms [139, 140].
In patients with high detrusor pressure during the filling phase (DO, low bladder compliance),
treatment is aimed primarily at conversion of an overactive, high-pressure bladder into a low-pressure reservoir
despite the resulting residual urine [139]. Reduction of the detrusor pressure contributes to urinary continence,
and consequently to social rehabilitation and QoL. It is also pivotal in preventing UTIs [146, 147]. However,
complete continence cannot always be obtained.

3.4.2 Non-invasive conservative treatment


3.4.2.1 Assisted bladder emptying - Credé manoeuvre, Valsalva manoeuvre, triggered reflex voiding
Incomplete bladder emptying is a serious risk factor for UTI, high intravesical pressure and incontinence.
Methods to improve the voiding process should therefore be practiced.

Bladder expression: The downwards movement of the lower abdomen by suprapubic compression (Credé)
or by abdominal straining (Valsalva) leads to an increase in intravesical pressure, and generally also causes a
reflex sphincter contraction [148, 149]. The latter may increase bladder outlet resistance and lead to inefficient
emptying. The high pressures created during these procedures are hazardous for the urinary tract [150, 151].
Therefore, their use should be discouraged unless urodynamics show that the intravesical pressure remains
within safe limits [140].
Long-term complications are unavoidable for both methods of bladder emptying [149]. The already
weak pelvic floor function may be further impaired, thus introducing or exacerbating already existing stress
urinary incontinence [151].

Triggered reflex voiding: Stimulation of the sacral or lumbar dermatomes in patients with a upper motor
neuron lesion can elicit a reflex detrusor contraction [151]. The risk of high pressure voiding is present and
interventions to decrease outlet resistance may be necessary [152]. Triggering can induce AD, especially in
patients with high level SCI (at or above Th 6) [153]. All assisted bladder emptying techniques require low outlet
resistance. Even then, high detrusor pressures may still be present. Hence, patients need dedicated education
and close urodynamic and urological surveillance [151, 154, 155].

Note: In the literature, including some of the references cited here, the concept “reflex voiding” is sometimes
used to cover all three assisted voiding techniques described in this section.

External appliances: Social continence may be achieved by collecting urine during incontinence, for instance
using pads. Condom catheters with urine collection devices are a practical method for men [140]. The
penile clamp is absolutely contraindicated in case of NDO or low bladder compliance because of the risk of
developing high intravesical pressure and pressure sores/necrosis in cases of altered/absent sensations.

3.4.2.2 Neuro-urological rehabilitation


3.4.2.2.1 Bladder rehabilitation including electrical stimulation
The term bladder rehabilitation summarises treatment options that aim to re-establish bladder function in
patients with neuro-urological symptoms. Strong contraction of the urethral sphincter and/or pelvic floor, as
well as anal dilatation, manipulation of the genital region, and physical activity inhibit micturition in a reflex
manner [140, 156]. The first mechanism is affected by activation of efferent nerve fibres, and the latter ones are
produced by activation of afferent fibres [116]. Electrical stimulation of the pudendal nerve afferents strongly
inhibits the micturition reflex and detrusor contraction [157]. This stimulation might then support the restoration
of the balance between excitatory and inhibitory inputs at the spinal or supraspinal level [140, 158]. Evidence

NEURO-UROLOGY - LIMITED UPDATE MARCH 2020 15


for bladder rehabilitation using electrical stimulation in neurological patients is mainly based on small non-
comparative studies with high risk of bias.

Peripheral temporary electrostimulation: Tibial nerve stimulation and transcutaneous electrical nerve stimulation
(TENS) might be effective and safe for treating neurogenic LUT dysfunction, but more reliable evidence from
well-designed randomised controlled trials (RCTs) is required to reach definitive conclusions [158-160]. In post-
stroke patients TENS has been shown to effectively improve urodynamic findings and QoL [161-163].

Peripheral temporary electrostimulation combined with pelvic floor muscle training and biofeedback: In MS
patients, combining active neuromuscular electrical stimulation with pelvic floor muscle training and EMG
biofeedback can achieve a substantial reduction of neuro-urological symptoms [164, 165]. This treatment
combination seems to be more effective than either therapy alone [166, 167]. However, the combination of
intravaginal electrostimulation and pelvic floor muscle training was not superior to pelvic floor muscle training
alone in reducing urinary incontinence in women with incomplete spinal cord injury [168].

Intravesical electrostimulation: Intravesical electrostimulation can increase bladder capacity and improve
bladder filling sensation in patients with incomplete SCI or myelomeningocele (MMC) [169]. In patients with
neurogenic detrusor underactivity, intravesical electrostimulation may also improve voiding and reduce residual
volume [170, 171].

Repetitive transcranial magnetic stimulation: Although improvement of neuro-urological symptoms has been
described in PD and MS patients, this technique is still under investigation [172, 173].

Summary: To date, bladder rehabilitation techniques are mainly based on electrical or magnetic stimulation;
however, there is a lack of well-designed studies.

3.4.2.3 Drug treatment


A single, optimal, medical therapy for neuro-urological symptoms is not always available. Commonly, a
combination of different therapies (e.g. intermittent catheterisation and antimuscarinic drugs) is advised to
prevent urinary tract damage and improve long-term outcomes, particularly in patients with a suprasacral SCI
or MS [151, 174-176].

3.4.2.3.1 Drugs for storage symptoms


Antimuscarinic drugs: Are the first-line choice for treating NDO, increasing bladder capacity and reducing
episodes of UI secondary to NDO by the inhibition of parasympathetic pathways [140, 177-183].
Antimuscarinic drugs have been used for many years to treat patients with NDO [181, 182, 184], and the
responses of individual patients to antimuscarinic treatment are variable. Despite a meta-analysis confirming
the clinical and urodynamic efficacy of antimuscarinic therapy compared to placebo in adult NDO, a more
recent integrative review has indicated that the information provided is still too limited for clinicians to be able
to match trial data to the needs of individual patients with SCI, mainly due to the lack of use of standardised
clinical evaluation tools such as the American Spinal Injury Association bladder diary and validated symptoms
score [182, 185].
Higher doses or a combination of antimuscarinic agents may be an option to maximise outcomes
in neurological patients [178, 179, 186-189]. However, these drugs have a high incidence of adverse events,
which may lead to early discontinuation of therapy. Despite this, NDO patients have generally shown better
treatment adherence compared to idiopathic DO patients [190].

Choice of antimuscarinic agent: Oxybutynin [140, 178, 179, 181, 182, 191], trospium [182, 188, 192], tolterodine
[193] and propiverine [182, 194] are established, effective and well tolerated treatments even in long-term use
[181, 182, 195, 196]. Darifenacin [197, 198] and solifenacin [199] have been evaluated in NDO secondary to
SCI and MS [182, 197-199] with results similar to other antimuscarinic drugs. A pilot study using solifenacin in
NDO due to PD showed an improvement in UI [200]. Fesoterodine, an active metabolite of tolterodine, has also
been introduced; to date there has been no published clinical evidence for its use in the treatment of neuro-
urological disorders. Favourable results with the new drug imidafenacin have been reported in suprapontine as
well as SCI patients [201, 202].

Side effects: Controlled-release antimuscarinics have some minor side effects, e.g. dry mouth [203]. It has been
suggested that different ways of administration may help to reduce side effects [204]. Imidafenacine has been
safely used in neurological patients with no worsening of cognitive function [201].

16 NEURO-UROLOGY - LIMITED UPDATE MARCH 2020


Beta-3-adrenergic receptor agonists
The role of mirabegron in neuro-urological patients is still unclear. In MS and SCI patients, with very short
follow up, mirabegron has not demonstrated any significant effect on detrusor pressure or cystometric
capacity despite the reported improvement in LUTS [205, 206]. A significant subjective improvement in OAB
symptoms has also been reported using lower dosages of mirabegron in patients affected by CNS lesions
without any negative effects on voiding function [207]. A standard dosage of 50 mg has been found effective
with no worsening of cognitive function in patients with PD [208]. Combination therapy with mirabegron and
desmopressin in MS patients has shown promising results; however, clinical experience is still very limited in
neuro-urological populations [209].

Other drugs
In preliminary studies, improvements in daily incontinence rates, nocturia, daytime and 24 hour voids, as well
as the low risk of adverse events, suggest that cannabinoids may be effective and safe in MS patients [210].
A concomitant improvement in OAB symptoms has been reported in male MS patients using daily tadalafil to
treat neurogenic erectile dysfunction (ED) [211]. A systematic review found that desmopressin may be effective
for treating nocturia in MS patients; however, adverse events were common, with the included studies being
heterogeneous and of low quality [212].

3.4.2.3.2 Drugs for voiding symptoms


Detrusor underactivity: Cholinergic drugs, such as bethanechol and distigmine, have been considered to
enhance detrusor contractility and promote bladder emptying, but are not frequently used in clinical practice
[213]. Only preclinical studies have documented the potential benefits of cannabinoid agonists for improving
detrusor contractility when administered intravesically [214, 215].

Decreasing bladder outlet resistance: α-blockers (e.g. tamsulosin, naftopidil and silodosin) seem to be effective
for decreasing bladder outlet resistance, PVR and AD [216-218].

Increasing bladder outlet resistance: Several drugs have shown efficacy in selected cases of mild SUI, but there
are no high-level evidence studies in neurological patients [140].

3.4.2.4 Summary of evidence and recommendations for drug treatments

Summary of evidence LE
Long-term efficacy and safety of antimuscarinic therapy for NDO is well documented. 1a
Mirabegron does not improve urodynamic outcomes in NDO patients. 1b
Maximise outcomes for neurogenic detrusor overactivity by considering a combination therapy. 3

Recommendations Strength rating


Use antimuscarinic therapy as the first-line medical treatment for neurogenic detrusor Strong
overactivity.
Prescribe α-blockers to decrease bladder outlet resistance. Strong
Do not prescribe parasympathomimetics for underactive detrusor. Strong

3.4.2.5 Minimally invasive treatment


3.4.2.5.1 Catheterisation
Intermittent self- or third-party catheterisation [219, 220] is the preferred management for neuro-urological
patients who cannot effectively empty their bladders [140]. Sterile IC, as originally proposed by Guttmann
and Frankel [219], significantly reduces the risk of UTI and bacteriuria [140, 221, 222], compared with clean
IC introduced by Lapides et al. [220]. However, it has not yet been established whether or not the incidence
of UTI, other complications and user satisfaction are affected by either sterile or clean IC, coated or uncoated
catheters or by any other strategy [223].

Sterile IC cannot be considered a routine procedure [140, 222]. Aseptic IC is an alternative to sterile IC [224].
The use of hydrophilic catheters was associated with a lower rate of UTI [225].

Contributing factors to contamination are insufficient patient education and the inherently greater risk of UTI in
neuro-urological patients [140, 226-230]. The average frequency of catheterisations per day is four to six times
[231] and the catheter size most often used is between 12-16 Fr. In aseptic IC, an optimum frequency of five

NEURO-UROLOGY - LIMITED UPDATE MARCH 2020 17


times showed a reduction of UTI [231]. Ideally, bladder volume at catheterisation should, as a rule, not exceed
400-500 mL.
Indwelling transurethral catheterisation and, to a lesser extent, suprapubic cystostomy are
associated with a range of complications as well as an enhanced risk for UTI [140, 232-239]; therefore, both
procedures should be avoided, when possible. Silicone catheters are preferred as they are less susceptible to
encrustation and because of the high incidence of latex allergy in the neuro-urological patient population [240].

3.4.2.5.2 Summary of evidence and recommendations for catheterisation

Summary of evidence LE
Intermittent catheterisation is the standard treatment for patients who are unable to empty their 3
bladder.
Indwelling transurethral catheterisation and suprapubic cystostomy are associated with a range of 3
complications as well as an enhanced risk for UTI.

Recommendations Strength rating


Use intermittent catheterisation, whenever possible aseptic technique, as a standard Strong
treatment for patients who are unable to empty their bladder.
Thoroughly instruct patients in the technique and risks of intermittent catheterisation. Strong
Avoid indwelling transurethral and suprapubic catheterisation whenever possible. Strong

3.4.2.5.3 Intravesical drug treatment


To reduce DO, antimuscarinics can also be administered intravesically [204, 241-244]. The efficacy, safety and
tolerability of intravesical administration of 0.1% oxybutynin hydrochloride compared to its oral administration
for treatment of NDO has been demonstrated in a recent randomised controlled study [204]. This approach
may reduce adverse effects due to the fact that the antimuscarinic drug is metabolised differently [241] and a
greater amount is sequestered in the bladder, even more than with electromotive administration [242].

The vanilloids, capsaicin and resiniferatoxin, desensitise the C-fibres for a period of a few months [245, 246].
Clinical studies have shown that resiniferatoxin has limited clinical efficacy compared to botulinum toxin A
injections in the detrusor [245].

Although preliminary data suggest that intravesical vanilloids might be effective for treating neurogenic LUT
dysfunction, their safety profile appears to be unfavourable [247]. Currently, there is no indication for the use of
these substances, which are not licensed for intravesical treatment.

3.4.2.5.4 Summary of evidence and recommendations for intravesical drug treatment

Summary of evidence LE
A significant reduction in adverse events was observed for intravesical administration of oxybutynine 1a
compared to oral administration.

Recommendation Strength rating


Offer intravesical oxybutynin to neurogenic detrusor overactivity patients with poor Strong
tolerance to the oral route.

3.4.2.5.5 Botulinum toxin injections in the bladder


Botulinum toxin A causes a long-lasting but reversible chemical denervation that lasts for about nine months
[248, 249]. The toxin injections are mapped over the detrusor in a dosage that depends on the preparation
used. Botulinum toxin A has been proven effective in patients with neuro-urological disorders due to MS, SCI
and PD in multiple RCTs and meta-analyses [250-252]. Urodynamic studies might be necessary after treatment
in patients with maximal filling pressure of > 40 cm H2O cm in order to monitor the effect of the injections
on bladder pressure [253]. Repeated injections seem to be possible without loss of efficacy, even after initial
low response rates, based on years of follow up [248, 254-257]. The clinical efficacy of botulinum toxin A
injection in patients with low morbidity after failure of augmentation enterocystoplasty has been demonstrated
[258]. A switch between different toxin variations may improve responsiveness [259]. The most frequent side

18 NEURO-UROLOGY - LIMITED UPDATE MARCH 2020


effects are UTIs, urinary retention and haematuria [260]. Intermittent catheterisation may become necessary,
this is especially relevant in MS patients as they do not often perform IC prior to intravesical botulinum toxin
injections. However, a lower dose of botulinum toxin A (100 U) may reduce the rate of clean IC in MS patients
[261]. Rare complications include generalised muscle weakness and AD [260]. Current research focuses on
different delivery approaches to injection such as liposome encapsulated botulinum toxin to decrease side
effects [262]. Neuro-urological patients with an indwelling catheter and concomitant bladder pain and/or
catheter bypass leakage could benefit from intravesical botulinum injections [263].

3.4.2.5.6 Bladder neck and urethral procedures


Reduction of the bladder outlet resistance may be necessary to protect the UUT. This can be achieved by
chemical denervation of the sphincter or by surgical interventions (bladder neck or sphincter incision or urethral
stent – Section 3.4.3.1). Incontinence may result and can be managed by external devices (Section 3.4.2.1).

Botulinum toxin A: This can be used to treat DSD effectively by injecting the sphincter at a dose that depends
on the preparation used. The dyssynergia is abolished only for a few months, necessitating repeat injections.
The efficacy of this treatment has been reported to be high with few adverse effects [264-266]. However, a
recent Cochrane report concluded that, because of limited evidence, future RCTs assessing the effectiveness
of botulinum toxin A injections also need to address the uncertainty about the optimal dose and mode of
injection [267]. In addition, this therapy is not licensed.

Balloon dilatation: Favourable immediate results were reported [268], but there have been no further reports
since 1994; therefore, this method is no longer recommended.

Increasing bladder outlet resistance: This can improve the continence condition. Despite early positive results
with urethral bulking agents, a relative early loss of continence is reported in patients with neuro-urological
disorders [140, 269, 270].

Urethral inserts: Urethral plugs or valves for the management of (female) stress incontinence have not been
applied in neuro-urological patients. The experience with active pumping urethral prosthesis for treatment of
the underactive or acontractile detrusor were disappointing [271].

3.4.2.5.7 S
 ummary of evidence and recommendations for botulinum toxin A injections and bladder neck
procedures

Summary of evidence LE
Botulinum toxin A has been proven effective in patients with neuro-urological disorders due to MS or 1a
SCI in multiple RCTs and meta-analyses.
Bladder neck incision is indicated only for secondary changes (fibrosis) at the bladder neck. 4

Recommendations Strength rating


Use botulinum toxin injection in the detrusor to reduce neurogenic detrusor overactivity in Strong
multiple sclerosis or spinal cord injury patients if antimuscarinic therapy is ineffective.
Bladder neck incision is effective in a fibrotic bladder neck. Strong

3.4.3 Surgical treatment


There is considerable heterogeneity in outcome parameters and definitions of cure used to report on outcome
of surgical interventions for SUI in neuro-urological patients [272]. The heterogeneity of outcome reporting
makes it difficult to interpret and compare different studies and therapies. A consistent comparison of the
outcomes of therapy can only be made after standardisation of outcome parameters and definitions of cure
or success; therefore, it would seem prudent to develop a core outcome set (COS) for use in UI research in
neuro-urological patients [272]. Until such a COS is developed it would seem feasible to use both a subjective
and objective outcome parameter and the combination of both to define cure [272]. Due to the importance of
QoL for neuro-urological patients a disease-specific QoL questionnaire or a bother questionnaire validated for
neuro-urological patients should be used as the subjective outcome parameter [272].

3.4.3.1 Bladder neck and urethral procedures


Increasing the bladder outlet resistance has the inherent risk of causing high intravesical pressure. Procedures
to treat sphincteric incontinence are therefore suitable only when the detrusor activity can be controlled and
when no significant reflux is present. A simultaneous bladder augmentation and IC may be necessary [140].

NEURO-UROLOGY - LIMITED UPDATE MARCH 2020 19


Urethral sling: Various materials have been used for this procedure with enduring positive results. The
procedure is established in women with the ability to self-catheterise [140, 273-276]. There is growing evidence
that synthetic slings can be used effectively with acceptable medium to long-term results and minimal
morbidity in neurological patients [277, 278]. Besides the pubovaginal sling, which has been considered the
procedure of choice in this subgroup of patients, recent reports suggest that both the transobturator and the
retropubic approaches may also be considered, with similar failure rates and a reduction in the need for IC.
However, for both approaches a higher incidence of de novo urgency was reported [278, 279]. In men, both
autologous and synthetic slings may also be an alternative [280-284].

Artificial urinary sphincter (AUS): This device was introduced by Light and Scott for patients with neuro-
urological disorders [285]. It has stood the test of time and acceptable long-term outcomes can be obtained
[286]. However, the complication rates and re-operation rates are higher than in non-neurogenic patient groups
(up to 60%), so it is advisable that patients are conscientiously informed about the success rates as well as
the complications that may occur after the procedure [287, 288]. In a case series with 25 years follow up only
7.1% of patients were revision free at 20 years [289]. Re-interventions are commonly due to infection, urethral
atrophy or erosion and mechanical failure.

There is growing interest in the use of this device in women with development of laparoscopic and robot-
assisted approaches which appear to reduce infection and erosion rates [290-293]. Long-term surgical
and patient-reported outcomes are needed to determine the role of AUS placement in female patients with
neurogenic SUI [294].

Adjustable continence device - ProACT/ACT®: The efficacy of this device has been reported mainly in post-
prostatectomy incontinence. A marginally lower cure rate has been reported in neurological patients when
compared to non-neurological patients [295]. A retrospective study in neuro-urological patients reported a low
rate of efficacy and high complication rate for this device [296].

Functional sphincter augmentation: Transposing the gracilis muscle to the bladder neck [297] or proximal
urethra [298], can enable the possible creation of a functional autologous sphincter by electrical stimulation
[297-299]; therefore, raising the prospect of restoring control over the urethral closure.

Bladder neck and urethra reconstruction: The classical Young-Dees-Leadbetter procedure [300] for bladder
neck reconstruction in children with bladder exstrophy, and Kropp urethra lengthening [301] improved by Salle
[302], are established methods to restore continence provided that IC is practiced and/or bladder augmentation
is performed [140, 303].

Endoscopic techniques for treating anatomic bladder outlet obstruction [304]:


• Transurethral resection of the prostate is indicated in male patients with refractory LUT symptoms due to
benign prostatic obstruction. Special consideration should be given to pre-operative abnormal sphincter
function which can lead to de novo or persistent UI [305, 306].
• Bladder neck resection is indicated in patients with high PVR and when a prominent obstruction of
the sclerotic ring in the bladder neck is identified during cystoscopy. The resection can be performed
between the three or nine o’clock position or full circle [307].
• Urethrotomy is indicated in patients with urethral strictures. Cold knife or neodymium:YAG contact laser
urethrotomy at the twelve o’clock position can be performed [308, 309]. In recurrent strictures, open
surgery should be considered.
• Sphincterotomy has been shown to be an efficient technique for the resolution of AD, hydronephrosis
and recurrent UTI, and for decreasing detrusor pressures, PVR and vesicoureteral reflux. It is irreversible
and should be limited to men who are able to wear a condom catheter. By staged incision, bladder
outlet resistance can be reduced without completely losing the closure function of the urethra [139, 140,
310]. The incision with less complications is the twelve o’clock sphincterotomy with cold knife [311] or
neodymium:YAG laser [312]. Sphincterotomy needs to be repeated at regular intervals in many patients
[313], but it is efficient and does not cause severe adverse effects [139, 268]. Secondary narrowing of the
bladder neck may occur, for which combined bladder neck incision might be considered [314].

Bladder neck incision: This is indicated only for secondary changes at the bladder neck (fibrosis) [139, 315].
This procedure is not recommended in patients with detrusor hypertrophy, which causes thickening of the
bladder neck [139].

20 NEURO-UROLOGY - LIMITED UPDATE MARCH 2020


Stents: Implantation of urethral stents results in continence being dependent on adequate closure of the
bladder neck [140]. The results are comparable with sphincterotomy and the stenting procedure has a
shorter duration of surgery and hospital stay [316, 317]. However, the costs [139], possible complications and
re-interventions [318, 319] are limiting factors in their use [320-323].

3.4.3.2 Denervation, deafferentation, sacral neuromodulation


Sacral anterior root stimulation (SARS) is aimed at producing detrusor contraction. The technique was
developed by Brindley [324] and is only applicable to complete lesions above the implant location, as its
stimulation amplitude is over the pain threshold. The urethral sphincter efferents are also stimulated, but
because the striated muscle relaxes faster than the smooth muscle of the detrusor, so-called “post-stimulus
voiding” occurs. This approach has been successful in highly selected patients [325-327]. Although it has
been shown that detrusor pressure during SARS decreases over time, the changes do not seem to be
clinically relevant during the first decade after surgery [328]. By changing the stimulation parameters, this
method can also induce defecation or erection. A recent study reported that Charcot spinal arthropathy
should be considered as a potential long-term complication of SARS, leading to spinal instability and to SARS
dysfunction [329].
Sacral rhizotomy, also known as sacral deafferentation, has achieved some success in reducing DO
[330-332], but nowadays, it is used mostly as an adjuvant to SARS [325, 333-336]. Alternatives to rhizotomy
are sought in this treatment combination [337-339].
Sacral neuromodulation [340] might be effective and safe for treating neuro-urological symptoms,
but there is a lack of RCTs and it is unclear which neurological patients are most suitable [341-344].

3.4.3.3 Bladder covering by striated muscle


When the bladder is covered by striated muscle, that can be stimulated electrically, or ideally that can be
contracted voluntarily, voiding function can be restored to an acontractile bladder. The rectus abdominis [345]
and latissimus dorsi [346] have been used successfully in patients with neuro-urological symptoms [347, 348].

3.4.3.4 Bladder augmentation


The aim of auto-augmentation (detrusor myectomy) is to reduce DO or improve low bladder compliance. The
advantages are: low surgical burden, low rate of long-term adverse effects, positive effect on patient QoL, and
it does not preclude further interventions [139, 140, 349-352].
Replacing or expanding the bladder by intestine or other passive expandable coverage will improve
bladder compliance and at least reduce the pressure effect of DO [353, 354]. Improved QoL and stable renal
function has been reported during long-term follow-up [355]. Patients performing IC with augmentation
cystoplasty had better urinary function and satisfaction with their urinary symptoms compared to patients
performing IC with or without botulinum toxin treatment [356]. Long-term complications included bladder
perforation (1.9%), mucus production (12.5%), metabolic abnormalities (3.35%), bowel dysfunction (15%), and
stone formation (10%) [355].
The procedure should be used with caution in patients with neuro-urological symptoms, but
may become necessary if all less-invasive treatment methods have failed. Special attention should be paid
to patients with pre-operative renal scars since metabolic acidosis can develop [357]. Bladder substitution
with bowel after performing a supratrigonal cystectomy [354], to create a low-pressure reservoir, is indicated
in patients with a severely thick and fibrotic bladder wall [140]. Intermittent catheterisation may become
necessary after this procedure. The long-term scientific evidence shows that bladder augmentation is a highly
successful procedure that stabilises renal function and prevents anatomical deterioration; however, lifelong
follow-up is essential in this patient group given the significant morbidity associated with this procedure [355,
358].

3.4.3.5 Urinary diversion


When no other therapy is successful, urinary diversion must be considered for the protection of the UUT and
for the patient’s QoL [140].

Continent diversion: This should be the first choice for urinary diversion. Patients with limited dexterity may
prefer a stoma instead of using the urethra for catheterisation. For cosmetic reasons, the umbilicus is often
used for the stoma site [359-364]. A systematic review of the literature concluded that continent catheterisable
tubes/stomas are an effective treatment option in neuro-urological patients unable to perform intermittent self-
catheterisation through the urethra [365]. However, the complication rates were significant with 85/213 post-
operative events requiring re-operation [365]. Tube stenosis occurred in 4-32% of the cases. Complications
related to concomitant procedures (augmentation cystoplasty, pouch) included neovesicocutaneous fistulae
(3.4%), bladder stones (20-25%), and bladder perforations (40%). In addition, data comparing HRQoL before
and after surgery were not reported [365].

NEURO-UROLOGY - LIMITED UPDATE MARCH 2020 21


Incontinent diversion: If catheterisation is impossible, incontinent diversion with a urine-collecting device
is indicated. Ultimately, it could be considered in patients who are wheelchair bound or bed-ridden with
intractable and untreatable incontinence, in patients with LUT destruction, when the UUT is severely
compromised, and in patients who refuse other therapy [140]. An ileal segment is used for the deviation in most
cases [140, 366-369]. Patients gain better functional status and QoL after surgery [370].

Undiversion: Long-standing diversions may be successfully undiverted or an incontinent diversion changed


to a continent one with the emergence of new and better techniques for control of detrusor pressure
and incontinence [140]. The patient must be carefully counselled and must comply meticulously with the
instructions [140]. Successful undiversion can then be performed [371].

3.4.3.6 Summary of evidence and recommendations for surgical treatment

Summary of evidence LE
Bladder augmentation is an effective option to decrease detrusor pressure and increase bladder 3
capacity, when all less-invasive treatment methods have failed.
Urethral sling placement is an established procedure, with acceptable medium- to long-term results, in 3
women with the ability to self-catheterise.
Artificial urinary sphincter insertion is a viable option, with acceptable long-term outcomes, in males. 3
The complication and re-operation rates are higher in neuro-urological patients; therefore, patients
must be adequately informed regarding the success rates as well as the complications that may occur
following the procedure.

Recommendations Strength rating


Perform bladder augmentation in order to treat refractory neurogenic detrusor overactivity. Strong
Place an autologous urethral sling in female patients with neurogenic stress urinary Strong
incontinence who are able to self-catheterise.
Insert an artificial urinary sphincter in male patients with neurogenic stress urinary Strong
incontinence.

3.5 Urinary tract infection in neuro-urological patients


3.5.1 Epidemiology, aetiology and pathophysiology
Urinary tract infection is the onset of signs and/or symptoms accompanied by laboratory findings of a UTI
(bacteriuria, leukocyturia and positive urine culture) [360]. There are no evidence-based cut-off values for the
quantification of these findings. The published consensus is that a significant bacteriuria in persons performing
IC is present with > 102 cfu/mL, > 104 cfu/mL in clean-void specimens and any detectable concentration
in suprapubic aspirates. Regarding leukocyturia, ten or more leukocytes in centrifuged urine samples per
microscopic field (400x) are regarded as significant [360].

The pathogenesis of UTI in neuro-urological patients is multifactorial. Male gender seems to be a risk factor
for febrile UTI [372]. Several etiological factors have been described: altered intrinsic defence mechanisms,
impaired washout and catheterisation [373]. Poor glycemic control has been established as a risk factor for UTI
in women with type 1 diabetes [374]. However, the exact working mechanisms remain unknown. The presence
of asymptomatic bacteriuria in SCI patients is higher than in the general population, and varies depending
on bladder management. Prevalence of bacteriuria in those performing clean IC varies from 23-89% [375].
Sphincterotomy and condom catheter drainage has a 57% prevalence [376]. Asymptomatic bacteria should not
be routinely screened for in this population [377].

Individuals with neuro-urological symptoms, especially those with SCI, may have other signs and symptoms in
addition to or instead of traditional signs and symptoms of a UTI in able-bodied individuals. Other problems,
such as AD, may develop or worsen due to a UTI [225]. The most common signs and symptoms suspicious
of a UTI in those with neuro-urological disorders are fever, new onset or increase in incontinence, including
leaking around an indwelling catheter, increased spasticity, malaise, lethargy or sense of unease, cloudy urine
with increased urine odour, discomfort or pain over the kidney or bladder, dysuria, or AD [225, 378]. New
incontinence is the most specific symptom, whereas cloudy and foul smelling urine has the highest positive
predictive value for UTI diagnosis [379].

22 NEURO-UROLOGY - LIMITED UPDATE MARCH 2020


3.5.2 Diagnostic evaluation
Urine culture and urinalysis are the optimum tests for the diagnosis of UTI in neuro-urological patients.
A dipstick test may be more useful to exclude than to prove UTI [380, 381]. As bacterial strains and
resistance patterns in persons with neuro-urological disorders may differ from those of able-bodied patients,
microbiologic testing is mandatory [382].

3.5.3 Disease management


Bacteriuria in patients with neuro-urological disorders should not be treated. Treatment of asymptomatic
bacteriuria results in significantly more resistant bacterial strains without improving the outcome [383]. Urinary
tract infections in persons with neuro-urological disorders are by definition a complicated UTI; therefore,
single-dose treatment is not advised. There is no consensus in the literature about the duration of treatment
as it depends on the severity of the UTI and the involvement of kidneys and the prostate. Generally, a five to
seven day course of antibiotic treatment is advised, which can be extended up to fourteen days according
to the extent of the infection [383]. The choice of antibiotic therapy should be based on the results of
the microbiologic testing. If immediate treatment is mandatory (e.g. fever, septicaemia, intolerable clinical
symptoms, extensive AD), the choice of treatment should be based on local and individual resistance profiles
[384]. In patients with afebrile UTI, an initial non-antibiotic treatment may be justified [385, 386].

3.5.3.1 Recurrent UTI


Recurrent UTI in patients with neuro-urological disorders may indicate suboptimal management of the
underlying functional problem, e.g. high bladder pressure during storage and voiding, incomplete voiding or
bladder stones. The improvement of bladder function, by treating DO by botulinum toxin A injection in the
detrusor [387], and the removal of bladder stones or other direct supporting factors, especially indwelling
catheters, as early as possible, are mandatory [382].

3.5.3.2 Prevention
If the improvement of bladder function and removal of foreign bodies/stones is not successful, additional UTI
prevention strategies should be utilised. In a meta-analysis the use of hydrophilic catheters was associated
with a lower rate of UTI [225]. Bladder irrigation has not been proven effective [388].

Various medical approaches have been tested for UTI prophylaxis in patients with neuro-urological disorders.
The benefit of cranberry juice or probiotics for the prevention of UTI could not be demonstrated in RCTs [389,
390]. Methenamine hippurate is not effective in individuals with neuro-urological symptoms [391]. There is no
sufficient evidence to support the use of L-methionine for urine acidification to prevent recurrent UTI [392].
There is only weak evidence that oral immunotherapy reduces bacteriuria in patients with SCI [393] and that
recurrent UTIs are reduced [394]. Low-dose, long-term, antibiotic prophylaxis can reduce UTI frequency, but
increases bacterial resistance and is therefore not recommended [383].
Weekly cycling of antibiotic prophylaxis provided long-term positive results, but the results of
this trial need to be confirmed in further studies [395]. Another possible future option, the inoculation of
apathogenic Escherichia coli strains into the bladder, has provided positive results in initial studies, but
because of the paucity of data [396], cannot be recommended as a treatment option. There is initial evidence
that homeopathic treatment can decrease UTI frequency [397]. In addition, intravesical gentamycin instillations
can reduce UTI frequency without increasing the number of multi-resistant bacteria [398].

In summary, based on the criteria of evidence-based medicine, there is currently no preventive measure
for recurrent UTI in patients with neuro-urological disorders that can be recommended without limitations.
Therefore, individualised concepts should be taken into consideration, including immunostimulation,
phytotherapy and complementary medicine [399]. Prophylaxis in patients with neuro-urological disorders is
important to pursue, but since there are no data favouring one approach over another, prophylaxis is essentially
a trial and error approach.

NEURO-UROLOGY - LIMITED UPDATE MARCH 2020 23


3.5.4 Summary of evidence and recommendations for the treatment of UTI

Summary of evidence LE
Treatment of asymptomatic bacteriuria results in significantly more resistant bacterial strains without 1a
improving patient outcome.
Low-dose, long-term, antibiotic prophylaxis does not reduce UTI frequency, but increases bacterial 2a
resistance.
Recurrent UTI in patients with neuro-urological disorders may indicate suboptimal management of the 3
underlying functional problem. Improvement of bladder function as early as possible is mandatory.
There is currently no preventive measure for recurrent UTI in patients with neuro-urological disorders 3
that can be recommended without limitations.

Recommendations Strength rating


Do not screen for or treat asymptomatic bacteriuria in patients with neuro-urological Strong
disorders.
Avoid the use of long-term antibiotics for recurrent urinary tract infections (UTIs). Strong
In patients with recurrent UTI, optimise treatment of neuro-urological symptoms and remove Strong
foreign bodies (e.g. stones, indwelling catheters) from the urinary tract.
Individualise UTI prophylaxis in patients with neuro-urological disorders as there is no Strong
optimal prophylactic measure available.

3.6 Sexual function and fertility


These Guidelines specifically focus on sexual dysfunction and infertility in patients with a neurological disease
[400, 401]. Non-neurogenic, male sexual dysfunction and infertility are covered in separate EAU Guidelines
[402, 403]. In neuro-urological patients sexual problems can be identified at three levels: primary (direct
neurological damage), secondary (general physical disabilities) and tertiary (psychosocial and emotional
issues) sexual dysfunction [404]. Adopting a systematic approach, such as the PLISSIT model (Permission,
Limited Information, Specific Suggestions and Intensive Therapy) [405], provides a framework for counselling
and treatment involving a stepwise approach to the management of neurogenic sexual dysfunction. Sexual
dysfunction is associated with neurogenic LUT dysfunction in patients with MS [406] and SB [407]. Although
various patient reported outcome measures (PROMs) are available to evaluate sexual function, the evidence for
good PROMs is limited and studies with high methodological quality are needed [408].

3.6.1 Erectile dysfunction


3.6.1.1 Phosphodiesterase type 5 inhibitors (PDE5Is)
Phosphodiesterase type 5 inhibitors (PDE5Is) are recommended as first-line treatment in neurogenic ED [400,
401]. In SCI patients, tadalafil, vardenafil and sildenafil have all improved retrograde ejaculation and improved
erectile function and satisfaction on IIEF-15. Tadalafil 10 mg was shown to be more effective than sildenafil
50 mg. All currently available PDE5Is appear to be effective and safe, although there are no high level evidence
studies in neuro-urological patients investigating the efficacy and side effects across different PDE5Is, dosages
and formulations [409].

For MS patients two studies reported significant improvement in ED when using sildenafil and tadalafil. One
study; however, showed no improvement in ED with sildenafil.
In PD normal erectile function was described in over half of the patients using sildenafil 100 mg
and a significant improvement in IIEF-15 score was found compared to placebo. While most neuro-urological
patients require long-term therapy for ED some have a low compliance rate or stop therapy because of side
effects [410, 411], most commonly headache and flushing [401]. In addition, PDE5Is may induce relevant
hypotension in patients with tetraplegia/high-level paraplegia and multiple system atrophy [410, 411]. As a
prerequisite for successful PDE5I-therapy, some residual nerve function is required to induce erection. Since
many patients with SCI use on-demand nitrates for the treatment of AD, they must be counselled that PDE5Is
are contraindicated when using nitrate medication.

3.6.1.2 Drug therapy other than PDE5Is


Fampridine to treat neurogenic spasticity has been shown to be beneficial in improving ED in two domains
of the IIEF-15 in SCI and MS patients, however, with a significant discontinuation rate due to severe adverse

24 NEURO-UROLOGY - LIMITED UPDATE MARCH 2020


events [412]. Sublingual apomorfine was shown to have poor results on ED in SCI patients and side-effects in
half of the patients [413]. In PD pergolide mesylate showed a significant improvement in IIEF-15 scores up to
twelve months follow-up [414].

3.6.1.3 Mechanical devices


Mechanical devices (vacuum tumescence devices and penile rings) may be effective but are less popular [415-419].

3.6.1.4 Intracavernous injections and intraurethral application


Patients not responding to oral drugs may be offered intracavernous injections (alprostadil, papaverine and
phentolamine) that have been shown to be effective in a number of neurological conditions, including SCI,
MS, and diabetes mellitus [420-426], but their use requires careful dose titration and some precautions.
Complications of intracavernous drugs include pain, priapism and corpora cavernosa fibrosis.
Intracavernous vasoactive drug injection is the first-line therapeutic option in patients taking
nitrate medications, as well as those with concerns about drug interactions with PDE5Is, or in whom PDE5Is
are ineffective. The impact of intracavernous injections on ejaculation and orgasmic function, their early
use for increasing the recovery rate of a spontaneous erection, and their effectiveness and tolerability in the
long-term are unclear [410]. Intra-urethral alprostadil application is an alternative, but a less effective, route of
administration [422, 427].

3.6.1.5 Sacral neuromodulation


Sacral neuromodulation for LUT dysfunction may improve sexual function; however, high level evidence studies
are lacking.

3.6.1.6 Penile prostheses


Penile prostheses may be considered for treatment of neurogenic ED when all conservative treatments have
failed. At a mean follow-up of seven years 83.7% of patients with SCI were able to have sexual intercourse
[401]. Serious complications, including infection and prosthesis perforation, may occur in about 10% of
patients, depending on implant type [428-430].

3.6.1.7 Summary of evidence and recommendations for erectile dysfunction

Summary of evidence LE
The long-term efficacy and safety of oral PDE5Is for the treatment of ED is well documented. 1b
Intracavernous vasoactive drug injections have been shown to be effective in a number of neurological 3
conditions, including SCI and MS; however, their use requires careful dose titration and precautions.
Mechanical devices (vacuum tumescence devices and penile rings) may be effective but are less 3
popular.
Reserve penile prostheses for selected patients, those in which all conservative treatments have 4
failed, with neurogenic ED.

Recommendations Strength rating


Prescribe oral phosphodiesterase type 5 inhibitors as first-line medical treatment in Strong
neurogenic erectile dysfunction (ED).
Give intracavernous injections of vasoactive drugs (alone or in combination) as second-line Strong
medical treatment in neurogenic ED.
Offer mechanical devices such as vacuum devices and rings to patients with neurogenic ED. Strong

3.6.2 Male fertility


Male fertility can be compromised in the neurological patient by ED, ejaculation disorder, impaired sperm
quality or various combinations of these three disorders. Among the major conditions contributing to
neurogenic infertility are pelvic and retroperitoneal surgery, diabetes mellitus, SB, MS and SCI [431].
Erectile dysfunction is managed as described previously. Retrograde ejaculation may be reversed by
sympathomimetic agents contracting the bladder neck, including imipramine, ephedrine, pseudoephedrine,
and phenylpropanolamine [431]. The use of a balloon catheter to obstruct the bladder neck may be effective in
obtaining antegrade ejaculation [432]. If antegrade ejaculation is not achieved, the harvest of semen from the
urine may be considered [433].

NEURO-UROLOGY - LIMITED UPDATE MARCH 2020 25


Prostatic massage is safe and easy to use for obtaining semen in men with lesions above Th 10 [434]. In
several patients, vibrostimulation or transrectal electroejaculation are needed for sperm retrieval [431, 435,
436]. Semen retrieval is more likely with vibrostimulation in men with lesions above Th 10 [437-439]. In men
with SCI, especially at or above Th 6, AD might occur during sexual activity and ejaculation [440, 441]; patients
at risk and fertility clinics must be informed and aware of this potentially life-threatening condition. In SCI
patients the use of oral midodrine can improve sperm retrieval at vibrostimulation [442].

In men with MS, use of disease modifying drugs during the conception phase, has not been associated with
altered pregnancy outcomes [443]. Surgical procedures, such as, microsurgical epididymal sperm aspiration
(MESA) or testicular sperm extraction (TESE), may be used if vibrostimulation and electroejaculation are not
successful [444, 445]. Pregnancy rates in patients with SCI are lower than in the general population, but since
the introduction of intracytoplasmic sperm injection (ICSI), men with SCI now have a good chance of becoming
biological fathers [446-448].

3.6.2.1 Sperm quality and motility


The following has been reported on sperm quality and motility:
• bladder management with clean IC may improve semen quality compared to indwelling catheterisation,
reflex voiding or bladder expression [449];
• in SCI patients sperm quality decreases at the early post traumatic phase demonstrating lower
spermatozoid vitality (necrospermia), reduced motility (asthenospermia) and leucospermia [444];
• long-term valproate treatment for epilepsy negatively influences sperm count and motility [450];
• vibrostimulation produces samples with better sperm motility than electrostimulation [451, 452];
• electroejaculation with interrupted current produces better sperm motility than continuous current [453];
• freezing of sperm is unlikely to improve fertility rates in men with SCI [454].

3.6.2.2 Summary of evidence and recommendations for male fertility

Summary of evidence LE
Vibrostimulation and transrectal electroejaculation have been shown to be effective for sperm retrieval 1b
in neuro-urological patients.
Surgical procedures, such as, microsurgical epididymal sperm aspiration or testicular sperm 3
extraction, may be used if vibrostimulation and electroejaculation are not successful.
In men with SCI at or above Th 6, AD might occur during sexual activity and ejaculation. 3

Recommendations Strength rating


Perform vibrostimulation and transrectal electroejaculation for sperm retrieval in men with Strong
spinal cord injury.
Perform microsurgical epididymal sperm aspiration, testicular sperm extraction Strong
and intracytoplasmic sperm injection after failed vibrostimulation and/or transrectal
electroejaculation in men with spinal cord injury.
Counsel men with spinal cord injury at or above Th 6 and fertility clinics about the potentially Strong
life-threatening condition of autonomic dysreflexia.

3.6.3 Female sexuality


The most relevant publications on neurogenic female sexual dysfunction are in women with SCI and MS. After
SCI, about 65-80% of women continue to be sexually active, but to a much lesser extent than before the injury,
and about 25% report a decreased satisfaction with their sexual life [455-458]. Although sexual dysfunction is
very common in women with MS, it is still often overlooked by medical professionals [459, 460].
The greatest physical barrier to sexual activity is UI. A correlation has been found between the
urodynamic outcomes of low bladder capacity, compliance and high maximum detrusor pressure and sexual
dysfunction in MS patients. Problems with positioning and spasticity affect mainly tetraplegic patients. Peer
support may help to optimise the sexual adjustment of women with SCI in achieving a more positive self-
image, self-esteem and feelings of being attractive to themselves and others [455, 461-463].

The use of specific drugs for sexual dysfunction is indicated to treat inadequate lubrication. Data on sildenafil
for treating female sexual dysfunction are poor and controversial [401]. Although good evidence exists that
psychological interventions are effective in the treatment of female hypoactive sexual desire disorder and
female orgasmic disorder [464], there is a lack of high-level evidence studies in the neurological population.

26 NEURO-UROLOGY - LIMITED UPDATE MARCH 2020


Neurophysiological studies have shown that women with the ability to perceive Th 11-L2 pin-prick sensations
may have psychogenic genital vasocongestion. Reflex lubrication and orgasm is more prevalent in women with
SCI who have preserved the sacral reflex arc (S2-S5), even when it has not been shown in an individual woman
that a specific level and degree of lesion is the cause of a particular sexual dysfunction. In SCI women with a
complete lesion of the sacral reflex, arousal and orgasm may be evoked through stimulation of other erogenous
zones above the level of lesions [465-467].
Sacral neuromodulation for LUT dysfunction may improve sexual function but high-evidence studies
are lacking [401].
Women with SCI reported dissatisfaction with the quality and quantity of sexuality-related
rehabilitation services and were less likely to receive sexual information than men [465, 468, 469].

3.6.4 Female fertility


There are few studies on female fertility in neurological patients. More than a third (38%) of women with
epilepsy had infertility and the relevant predictors were exposure to multiple (three or more) antiepileptic drugs,
older age and lower education [470].

Although it seems that the reproductive capacity of women with SCI is only temporarily affected by SCI with
cessation of menstruation for approximately six months after SCI [471], there are no high-level evidence
studies. About 70% of sexually active women use some form of contraception after injury, but fewer women
use the birth control pill compared to before their injury [472].
Women with SCI are more likely to suffer complications during pregnancy, labour and delivery
compared to able-bodied women. Complications of labour and delivery include bladder problems, spasticity,
pressure sores, anaemia, and AD [473-477]. Obstetric outcomes include higher rates of Caesarean sections
and an increased incidence of low birth-weight babies [472, 475-477].

Epidural anaesthesia is chosen and effective for most patients with AD during labour and delivery [478, 479].

There is very little published data on women’s experience of the menopause following SCI [480]. Women with
MS who plan a pregnancy should evaluate their current drug treatment with their treating physician [481-483].
Clinical management should be individualised to optimise both the mother’s reproductive outcomes and MS
course [482-484].

3.6.4.1 Summary of evidence and recommendation for female sexuality and fertility

Summary of evidence LE
Data on specific drugs for treating female sexual dysfunction are poor and controversial. 4
There are limited numbers of studies on female fertility in neurological patients, clinical management 4
should be individualised to optimise both the mother’s reproductive outcomes and medical condition.

Recommendations Strength rating


Do not offer medical therapy for the treatment of neurogenic sexual dysfunction in women. Strong
Take a multidisciplinary approach, tailored to individual patient’s needs and preferences, in Strong
the management of fertility, pregnancy and delivery in women with neurological diseases.

3.7 Follow-up
3.7.1 Introduction
Neuro-urological disorders are often unstable and the symptoms may vary considerably, even within a relatively
short period. Regular follow-up is therefore necessary to assess the UUT [138].

Depending on the type of the underlying neurological pathology and the current stability of the neuro-urological
symptoms, the interval between initial investigations and control diagnostics may vary and in many cases
should not exceed one to two years. In high-risk neuro-urological patients this interval should be much shorter.
Urinalysis should be performed only when patients present with symptoms [485]. The UUT should be checked
by ultrasonography at regular intervals in high-risk patients; about once every six months [6, 486]. In these
patients, physical examination and urine laboratory should take place every year [6, 486]. In MS patients higher
scores on the Expanded Disability Status Scale (EDSS) are associated with risk factors for UUT deterioration
[53]. A urodynamic investigation should be performed as a diagnostic baseline, and repeated during follow-
up, more frequently in high-risk patients [6, 486]. In addition, bladder wall thickness can be measured on

NEURO-UROLOGY - LIMITED UPDATE MARCH 2020 27


ultrasonography as an additional risk assessment for upper tract damage [487], although a ‘safe’ cut-off
threshold for this has not been agreed [488]. The utility of DMSA for follow-up of neuro-urological patients has
not been fully evaluated [489]. Any significant clinical change warrants further, specialised, investigation [6,
486]. However, there is a lack of high level evidence studies on this topic and every recommendation must be
viewed critically in each individual neuro-urological patient [138].

The increased prevalence of muscle invasive bladder cancer in neuro-urological patients also warrants long-
term follow-up [490]. The exact frequency of cystoscopy with or without cytology remains unknown, but
presence of risk factors similar to the general population should trigger further investigation [485].

Adolescent patients with neurological pathology are at risk of being lost to follow-up during the transition to
adulthood. It is important that a standardised approach during this transition is adopted to improve follow-up
and specific treatment during adult life [491].

3.7.2 Summary of evidence and recommendations for follow-up

Summary of evidence LE
Neuro-urological disorders are often unstable and the symptoms may vary considerably, therefore, 4
regular follow-up is necessary.

Recommendations Strength rating


Assess the upper urinary tract at regular intervals in high-risk patients. Strong
Perform a physical examination and urine laboratory every year in high-risk patients. Strong
Any significant clinical changes should instigate further, specialised, investigation. Strong
Perform urodynamic investigation as a mandatory baseline diagnostic intervention in high- Strong
risk patients at regular intervals.

3.8 Conclusions
Neuro-urological disorders have a multi-faceted pathology. They require an extensive and specific diagnosis
before one can embark on an individualised therapy, which takes into account the medical and physical
condition of the patient and the patient’s expectations about their future. The urologist can select from a
wealth of therapeutic options, each with its own pros and cons. Notwithstanding the success of any therapy
embarked upon, close surveillance is necessary for the patient’s entire life.
These Guidelines offer you expert advice on how to define the patient’s neuro-urological symptoms
as precisely as possible and how to select, together with the patient, the appropriate therapy. This last choice,
as always, is governed by the golden rule: as effective as needed, as non-invasive as possible.

4. REFERENCES
1. Schafer, W., et al. Good urodynamic practices: uroflowmetry, filling cystometry, and pressure-flow
studies. Neurourol Urodyn, 2002. 21: 261.
https://www.ncbi.nlm.nih.gov/pubmed/11948720
2. Abrams, P., et al. Reviewing the ICS 2002 terminology report: the ongoing debate. Neurourol
Urodyn, 2009. 28: 287.
https://www.ncbi.nlm.nih.gov/pubmed/19350662
3. Abrams, P., et al. The standardisation of terminology of lower urinary tract function: report from the
Standardisation Sub-committee of the International Continence Society. Neurourol Urodyn, 2002.
21: 167.
https://www.ncbi.nlm.nih.gov/pubmed/11857671
4. Groen, J., et al. Summary of European Association of Urology (EAU) Guidelines on Neuro-Urology.
Eur Urol, 2015.
https://www.ncbi.nlm.nih.gov/pubmed/26304502
5. Nosseir, M., et al. Clinical usefulness of urodynamic assessment for maintenance of bladder
function in patients with spinal cord injury. Neurourol Urodyn, 2007. 26: 228.

28 NEURO-UROLOGY - LIMITED UPDATE MARCH 2020


https://www.ncbi.nlm.nih.gov/pubmed/16998859
6. Panicker, J.N., et al. Lower urinary tract dysfunction in the neurological patient: clinical assessment
and management. Lancet Neurol, 2015. 14: 720.
https://www.ncbi.nlm.nih.gov/pubmed/26067125
7. Guyatt, G.H., et al. GRADE: an emerging consensus on rating quality of evidence and strength of
recommendations. BMJ, 2008. 336: 924.
https://www.ncbi.nlm.nih.gov/pubmed/18436948
8. Guyatt, G.H., et al. What is “quality of evidence” and why is it important to clinicians? BMJ, 2008.
336: 995.
https://www.ncbi.nlm.nih.gov/pubmed/18456631
9. Phillips B, et al. Oxford Centre for Evidence-based Medicine Levels of Evidence. Updated by
Jeremy Howick March 2009. 1998.
https://www.cebm.net/2009/06/oxford-centre-evidence-based-medicine-levels-evidence-
march-2009/
10. Guyatt, G.H., et al. Going from evidence to recommendations. BMJ, 2008. 336: 1049.
https://www.ncbi.nlm.nih.gov/pubmed/18467413
11. Townsend, N., et al. Cardiovascular disease in Europe - epidemiological update 2015. Eur Heart J,
2015.
https://www.ncbi.nlm.nih.gov/pubmed/26306399
12. Tibaek, S., et al. Prevalence of lower urinary tract symptoms (LUTS) in stroke patients: a cross-
sectional, clinical survey. Neurourol Urodyn, 2008. 27: 763.
https://www.ncbi.nlm.nih.gov/pubmed/18551565
13. Marinkovic, S.P., et al. Voiding and sexual dysfunction after cerebrovascular accidents. J Urol, 2001.
165: 359.
https://www.ncbi.nlm.nih.gov/pubmed/11176374
14. Rotar, M., et al. Stroke patients who regain urinary continence in the first week after acute first-ever
stroke have better prognosis than patients with persistent lower urinary tract dysfunction. Neurourol
Urodyn, 2011. 30: 1315.
https://www.ncbi.nlm.nih.gov/pubmed/21488096
15. Lobo, A., et al. Prevalence of dementia and major subtypes in Europe: A collaborative study of
population-based cohorts. Neurologic Diseases in the Elderly Research Group. Neurology, 2000. 54:
S4.
https://www.ncbi.nlm.nih.gov/pubmed/10854354
16. Na, H.R., et al. Urinary incontinence in Alzheimer’s disease is associated with Clinical Dementia
Rating-Sum of Boxes and Barthel Activities of Daily Living. Asia Pac Psychiatry, 2015. 7: 113.
https://www.ncbi.nlm.nih.gov/pubmed/23857871
17. Grant, R.L., et al. First diagnosis and management of incontinence in older people with and without
dementia in primary care: a cohort study using The Health Improvement Network primary care
database. PLoS Med, 2013. 10: e1001505.
https://www.ncbi.nlm.nih.gov/pubmed/24015113
18. Pringsheim, T., et al. The prevalence of Parkinson’s disease: a systematic review and meta-analysis.
Mov Disord, 2014. 29: 1583.
https://www.ncbi.nlm.nih.gov/pubmed/24976103
19. Picillo, M., et al. The PRIAMO study: urinary dysfunction as a marker of disease progression in early
Parkinson’s disease. Eur J Neurol, 2017. 24: 788.
https://www.ncbi.nlm.nih.gov/pubmed/28425642
20. Papatsoris, A.G., et al. Urinary and erectile dysfunction in multiple system atrophy (MSA). Neurourol
Urodyn, 2008. 27: 22.
https://www.ncbi.nlm.nih.gov/pubmed/17563111
21. Kim, M., et al. Impaired detrusor contractility is the pathognomonic urodynamic finding of multiple
system atrophy compared to idiopathic Parkinson’s disease. Parkinsonism Relat Disord, 2015. 21:
205.
https://www.ncbi.nlm.nih.gov/pubmed/25534084
22. Sakakibara, R., et al. A guideline for the management of bladder dysfunction in Parkinson’s disease
and other gait disorders. Neurourol Urodyn, 2015.
https://www.ncbi.nlm.nih.gov/pubmed/25810035
23. Yamamoto, T., et al. Postvoid residual predicts the diagnosis of multiple system atrophy in
Parkinsonian syndrome. J Neurol Sci, 2017. 381: 230.
https://www.ncbi.nlm.nih.gov/pubmed/28991688
24. Dolecek, T.A., et al. CBTRUS statistical report: primary brain and central nervous system tumors

NEURO-UROLOGY - LIMITED UPDATE MARCH 2020 29


diagnosed in the United States in 2005-2009. Neuro Oncol, 2012. 14 Suppl 5: v1.
https://www.ncbi.nlm.nih.gov/pubmed/23095881
25. Maurice-Williams, R.S. Micturition symptoms in frontal tumours. J Neurol Neurosurg Psychiatry,
1974. 37: 431.
https://www.ncbi.nlm.nih.gov/pubmed/4365244
26. Christensen, D., et al. Prevalence of cerebral palsy, co-occurring autism spectrum disorders, and
motor functioning - Autism and Developmental Disabilities Monitoring Network, USA, 2008. Dev
Med Child Neurol, 2014. 56: 59.
https://www.ncbi.nlm.nih.gov/pubmed/24117446
27. Samijn, B., et al. Lower urinary tract symptoms and urodynamic findings in children and adults with
cerebral palsy: A systematic review. Neurourol Urodyn, 2017. 36: 541.
https://www.ncbi.nlm.nih.gov/pubmed/26894322
28. Tagliaferri, F., et al. A systematic review of brain injury epidemiology in Europe. Acta Neurochir
(Wien), 2006. 148: 255.
https://www.ncbi.nlm.nih.gov/pubmed/16311842
29. Kulakli, F., et al. Relationship between urinary dysfunction and clinical factors in patients with
traumatic brain injury. Brain Inj, 2014. 28: 323.
https://www.ncbi.nlm.nih.gov/pubmed/24377376
30. Aruga, S., et al. Effect of cerebrospinal fluid shunt surgery on lower urinary tract dysfunction in
idiopathic normal pressure hydrocephalus. Neurourol Urodyn, 2018. 37: 1053.
https://www.ncbi.nlm.nih.gov/pubmed/28892272
31. Singh, A., et al. Global prevalence and incidence of traumatic spinal cord injury. Clin Epidemiol,
2014. 6: 309.
https://www.ncbi.nlm.nih.gov/pubmed/25278785
32. Weld, K.J., et al. Association of level of injury and bladder behavior in patients with post-traumatic
spinal cord injury. Urology, 2000. 55: 490.
https://www.ncbi.nlm.nih.gov/pubmed/10736489
33. Kondo, A., et al. Neural tube defects: prevalence, etiology and prevention. Int J Urol, 2009. 16: 49.
https://www.ncbi.nlm.nih.gov/pubmed/19120526
34. Sawin, K.J., et al. The National Spina Bifida Patient Registry: profile of a large cohort of participants
from the first 10 clinics. J Pediatr, 2015. 166: 444.
https://www.ncbi.nlm.nih.gov/pubmed/25444012
35. Wiener, J.S., et al. Bladder Management and Continence Outcomes in Adults with Spina Bifida:
Results from the National Spina Bifida Patient Registry, 2009 to 2015. J Urol, 2018. 200: 187.
https://www.ncbi.nlm.nih.gov/pubmed/29588216
36. Peyronnet, B., et al. Comparison of neurogenic lower urinary tract dysfunctions in open versus
closed spinal dysraphism: A prospective cross-sectional study of 318 patients. Neurourol Urodyn,
2018. 37: 2818.
https://www.ncbi.nlm.nih.gov/pubmed/30070396
37. Bartolin, Z., et al. Relationship between clinical data and urodynamic findings in patients with
lumbar intervertebral disk protrusion. Urol Res, 2002. 30: 219.
https://www.ncbi.nlm.nih.gov/pubmed/12202938
38. Baker, M., et al. Urogenital symptoms in women with Tarlov cysts. J Obstet Gynaecol Res, 2018. 44:
1817.
https://www.ncbi.nlm.nih.gov/pubmed/29974579
39. Lange, M.M., et al. Urinary and sexual dysfunction after rectal cancer treatment. Nat Rev Urol, 2011.
8: 51.
https://www.ncbi.nlm.nih.gov/pubmed/21135876
40. Federation, I.D., IDF Diabetes Atlas, 6th edn. 2013, International Diabetes Federation: Brussels,
Belgium.
https://www.idf.org/e-library/epidemiology-research/diabetes-atlas/19-atlas-6th-edition.html
41. Yuan, Z., et al. Diabetic cystopathy: A review. J Diabetes, 2015. 7: 442.
https://www.ncbi.nlm.nih.gov/pubmed/25619174
42. Pugliatti, M., et al. The epidemiology of multiple sclerosis in Europe. Eur J Neurol, 2006. 13: 700.
https://www.ncbi.nlm.nih.gov/pubmed/16834700
43. de Seze, M., et al. The neurogenic bladder in multiple sclerosis: review of the literature and proposal
of management guidelines. Mult Scler, 2007. 13: 915.
https://www.ncbi.nlm.nih.gov/pubmed/17881401
44. Gajewski, J.B., et al. An International Continence Society (ICS) report on the terminology for adult

30 NEURO-UROLOGY - LIMITED UPDATE MARCH 2020


neurogenic lower urinary tract dysfunction (ANLUTD). Neurourol Urodyn, 2018. 37: 1152.
https://www.ncbi.nlm.nih.gov/pubmed/29149505
45. Del Popolo, G., et al. Diagnosis and therapy for neurogenic bladder dysfunctions in multiple
sclerosis patients. Neurol Sci, 2008. 29 Suppl 4: S352.
https://www.ncbi.nlm.nih.gov/pubmed/19089675
46. Satar, N., et al. The effects of delayed diagnosis and treatment in patients with an occult spinal
dysraphism. J Urol, 1995. 154: 754.
https://www.ncbi.nlm.nih.gov/pubmed/7609171
47. Watanabe, T., et al. High incidence of occult neurogenic bladder dysfunction in neurologically intact
patients with thoracolumbar spinal injuries. J Urol, 1998. 159: 965.
https://www.ncbi.nlm.nih.gov/pubmed/9474194
48. Ahlberg, J., et al. Neurological signs are common in patients with urodynamically verified
“idiopathic” bladder overactivity. Neurourol Urodyn, 2002. 21: 65.
https://www.ncbi.nlm.nih.gov/pubmed/11835426
49. Bemelmans, B.L., et al. Evidence for early lower urinary tract dysfunction in clinically silent multiple
sclerosis. J Urol, 1991. 145: 1219.
https://www.ncbi.nlm.nih.gov/pubmed/2033697
50. Klausner, A.P., et al. The neurogenic bladder: an update with management strategies for primary
care physicians. Med Clin North Am, 2011. 95: 111.
https://www.ncbi.nlm.nih.gov/pubmed/21095415
51. Cetinel, B., et al. Risk factors predicting upper urinary tract deterioration in patients with spinal cord
injury: A retrospective study. Neurourol Urodyn, 2017. 36: 653.
https://www.ncbi.nlm.nih.gov/pubmed/26934371
52. Elmelund, M., et al. Renal deterioration after spinal cord injury is associated with length of detrusor
contractions during cystometry-A study with a median of 41 years follow-up. Neurourol Urodyn,
2016.
https://www.ncbi.nlm.nih.gov/pubmed/27813141
53. Ineichen, B.V., et al. High EDSS can predict risk for upper urinary tract damage in patients with
multiple sclerosis. Multiple Sclerosis, 2017. 1352458517703801: 01.
https://www.ncbi.nlm.nih.gov/pubmed/28367674
54. Bors, E., et al. History and physical examination in neurological urology. J Urol, 1960. 83: 759.
https://www.ncbi.nlm.nih.gov/pubmed/13802958
55. Cameron, A.P., et al. The Severity of Bowel Dysfunction in Patients with Neurogenic Bladder. J Urol,
2015.
https://www.ncbi.nlm.nih.gov/pubmed/25956470
56. Vodusek, D.B. Lower urinary tract and sexual dysfunction in neurological patients. Eur Neurol, 2014.
72: 109.
https://www.ncbi.nlm.nih.gov/pubmed/24993182
57. Linsenmeyer, T.A., et al. Accuracy of individuals with spinal cord injury at predicting urinary tract
infections based on their symptoms. J Spinal Cord Med, 2003. 26: 352.
https://www.ncbi.nlm.nih.gov/pubmed/14992336
58. Massa, L.M., et al. Validity, accuracy, and predictive value of urinary tract infection signs and
symptoms in individuals with spinal cord injury on intermittent catheterization. J Spinal Cord Med,
2009. 32: 568.
https://www.ncbi.nlm.nih.gov/pubmed/20025153
59. Bellucci, C.H.S., et al. Acute spinal cord injury - Do ambulatory patients need urodynamic
investigations? J Urol, 2013. 189: 1369.
https://www.ncbi.nlm.nih.gov/pubmed/23069382
60. Kessler, T.M. Diagnosis of urinary incontinence. JAMA, 2008. 300: 283; author reply 283.
https://www.ncbi.nlm.nih.gov/pubmed/18632541
61. Honjo, H., et al. Impact of convenience void in a bladder diary with urinary perception grade to
assess overactive bladder symptoms: a community-based study. Neurourol Urodyn, 2010. 29: 1286.
https://www.ncbi.nlm.nih.gov/pubmed/20878998
62. Naoemova, I., et al. Reliability of the 24-h sensation-related bladder diary in women with urinary
incontinence. Int Urogynecol J Pelvic Floor Dysfunct, 2008. 19: 955.
https://www.ncbi.nlm.nih.gov/pubmed/18235981
63. Henze, T. Managing specific symptoms in people with multiple sclerosis. Int MS J, 2005. 12: 60.
https://www.ncbi.nlm.nih.gov/pubmed/16417816
64. Liu, C.W., et al. The relationship between bladder management and health-related quality of life in
patients with spinal cord injury in the UK. Spinal Cord, 2010. 48: 319.

NEURO-UROLOGY - LIMITED UPDATE MARCH 2020 31


https://www.ncbi.nlm.nih.gov/pubmed/19841636
65. Myers, J.B., et al. Patient reported bladder-related symptoms and quality of life after spinal cord
injury with different bladder management strategies. J Urol, 2019: 202: 574.
https://www.ncbi.nlm.nih.gov/pubmed/30958741
66. Khalaf, K.M., et al. The impact of lower urinary tract symptoms on health-related quality of life
among patients with multiple sclerosis. Neurourol Urodyn, 2016. 35: 48.
https://www.ncbi.nlm.nih.gov/pubmed/25327401
67. Szymanski, K.M., et al. All Incontinence is Not Created Equal: Impact of Urinary and Fecal
Incontinence on Quality of Life in Adults with Spina Bifida. J Urol, 2017. Part 2. 197: 885.
https://www.ncbi.nlm.nih.gov/pubmed/28131501
68. Pannek, J., et al. Does optimizing bladder management equal optimizing quality of life? Correlation
between health-related quality of life and urodynamic parameters in patients with spinal cord
lesions. Urology, 2009. 74: 263.
https://www.ncbi.nlm.nih.gov/pubmed/19428089
69. Patel, D.P., et al. Patient reported outcomes measures in neurogenic bladder and bowel: A
systematic review of the current literature. Neurourol Urodyn, 2016. 35: 8.
https://www.ncbi.nlm.nih.gov/pubmed/25327455
70. Bonniaud, V., et al. Qualiveen, a urinary-disorder specific instrument: 0.5 corresponds to the minimal
important difference. J Clin Epidemiol, 2008. 61: 505.
https://www.ncbi.nlm.nih.gov/pubmed/18394545
71. Bonniaud, V., et al. Development and validation of the short form of a urinary quality of life
questionnaire: SF-Qualiveen. J Urol, 2008. 180: 2592.
https://www.ncbi.nlm.nih.gov/pubmed/18950816
72. Bonniaud, V., et al. Italian version of Qualiveen-30: cultural adaptation of a neurogenic urinary
disorder-specific instrument. Neurourol Urodyn, 2011. 30: 354.
https://www.ncbi.nlm.nih.gov/pubmed/21305589
73. Ciudin, A., et al. Quality of life of multiple sclerosis patients: translation and validation of the Spanish
version of Qualiveen. Neurourol Urodyn, 2012. 31: 517.
https://www.ncbi.nlm.nih.gov/pubmed/22396437
74. D’Ancona, C.A., et al. Quality of life of neurogenic patients: translation and validation of the
Portuguese version of Qualiveen. Int Urol Nephrol, 2009. 41: 29.
https://www.ncbi.nlm.nih.gov/pubmed/18528780
75. Pannek, J., et al. [Quality of life in German-speaking patients with spinal cord injuries and bladder
dysfunctions. Validation of the German version of the Qualiveen questionnaire]. Urologe A, 2007. 46:
1416.
https://www.ncbi.nlm.nih.gov/pubmed/17605119
76. Reuvers, S.H.M., et al. The urinary-specific quality of life of multiple sclerosis patients: Dutch
translation and validation of the SF-Qualiveen. Neurourol Urodyn, 2017. 36: 1629.
https://www.ncbi.nlm.nih.gov/pubmed/27794179
77. Reuvers, S.H.M., et al. The validation of the Dutch SF-Qualiveen, a questionnaire on urinary-specific
quality of life, in spinal cord injury patients. BMC Urology, 2017. 17: 88.
https://www.ncbi.nlm.nih.gov/pubmed/28927392
78. Best, K.L., et al. Identifying and classifying quality of life tools for neurogenic bladder function after
spinal cord injury: A systematic review. J Spinal Cord Med, 2017. 40: 505.
https://www.ncbi.nlm.nih.gov/pubmed/27734771
79. Welk, B., et al. The conceptualization and development of a patient-reported neurogenic bladder
symptom score. Res Rep Urol, 2013. 5: 129.
https://www.ncbi.nlm.nih.gov/pubmed/24400244
80. Welk, B., et al. The Neurogenic Bladder Symptom Score (NBSS): A secondary assessment of its
validity, reliability among people with a spinal cord injury. Spinal Cord, 2018. 56: 259.
https://www.ncbi.nlm.nih.gov/pubmed/29184133
81. Gulick, E.E. Bowel management related quality of life in people with multiple sclerosis: psychometric
evaluation of the QoL-BM measure. Int J Nurs Stud, 2011. 48: 1066.
https://www.ncbi.nlm.nih.gov/pubmed/21377677
82. Tsang, B., et al. A systematic review and comparison of questionnaires in the management of spinal
cord injury, multiple sclerosis and the neurogenic bladder. Neurourol Urodyn, 2015.
https://www.ncbi.nlm.nih.gov/pubmed/25620137
83. Akkoc, Y., et al. Assessment of voiding dysfunction in Parkinson’s disease: Reliability and validity of
the Turkish version of the Danish Prostate Symptom Score. Neurourol Urodyn, 2017. 36: 1903.

32 NEURO-UROLOGY - LIMITED UPDATE MARCH 2020


https://www.ncbi.nlm.nih.gov/pubmed/28139847
84. Schurch, B., et al. Reliability and validity of the Incontinence Quality of Life questionnaire in patients
with neurogenic urinary incontinence. Arch Phys Med Rehabil, 2007. 88: 646.
https://www.ncbi.nlm.nih.gov/pubmed/17466735
85. Hollingworth, W., et al. Exploring the impact of changes in neurogenic urinary incontinence
frequency and condition-specific quality of life on preference-based outcomes. Qual Life Res, 2010.
19: 323.
https://www.ncbi.nlm.nih.gov/pubmed/20094804
86. Cella, D.F., et al. Validation of the functional assessment of multiple sclerosis quality of life
instrument. Neurology, 1996. 47: 129.
https://www.ncbi.nlm.nih.gov/pubmed/8710066
87. Wesson, J.M., et al. The functional index for living with multiple sclerosis: development and
validation of a new quality of life questionnaire. Mult Scler, 2009. 15: 1239.
https://www.ncbi.nlm.nih.gov/pubmed/19737850
88. Gold, S.M., et al. Disease specific quality of life instruments in multiple sclerosis: validation of the
Hamburg Quality of Life Questionnaire in Multiple Sclerosis (HAQUAMS). Mult Scler, 2001. 7: 119.
https://www.ncbi.nlm.nih.gov/pubmed/11424632
89. Goodin, D.S. A questionnaire to assess neurological impairment in multiple sclerosis. Mult Scler,
1998. 4: 444.
https://www.ncbi.nlm.nih.gov/pubmed/9839306
90. Foley, F.W., et al. The Multiple Sclerosis Intimacy and Sexuality Questionnaire -- re-validation and
development of a 15-item version with a large US sample. Mult Scler, 2013. 19: 1197.
https://www.ncbi.nlm.nih.gov/pubmed/23369892
91. Sanders, A.S., et al. The Multiple Sclerosis Intimacy and Sexuality Questionnaire-19 (MSISQ-19).
Sex Disabil, 2000. 18: 3.
https://link.springer.com/article/10.1023/A:1005421627154
92. Marrie, R.A., et al. Validity and reliability of the MSQLI in cognitively impaired patients with multiple
sclerosis. Mult Scler, 2003. 9: 621.
https://www.ncbi.nlm.nih.gov/pubmed/14664477
93. Vickrey, B.G., et al. A health-related quality of life measure for multiple sclerosis. Qual Life Res,
1995. 4: 187.
https://www.ncbi.nlm.nih.gov/pubmed/7613530
94. Honan, C.A., et al. The multiple sclerosis work difficulties questionnaire (MSWDQ): development of a
shortened scale. Disabil Rehabil, 2014. 36: 635.
https://www.ncbi.nlm.nih.gov/pubmed/23786346
95. Welk, B., et al. The validity and reliability of the neurogenic bladder symptom score. J Urol, 2014.
192: 452.
https://www.ncbi.nlm.nih.gov/pubmed/24518764
96. Bonniaud, V., et al. Measuring quality of life in multiple sclerosis patients with urinary disorders using
the Qualiveen questionnaire. Arch Phys Med Rehabil, 2004. 85: 1317.
https://www.ncbi.nlm.nih.gov/pubmed/15295759
97. Franceschini, M., et al. Follow-up in persons with traumatic spinal cord injury: questionnaire
reliability. Eura Medicophys, 2006. 42: 211.
https://www.ncbi.nlm.nih.gov/pubmed/17039217
98. Noreau, L., et al. Development and assessment of a community follow-up questionnaire for the Rick
Hansen spinal cord injury registry. Arch Phys Med Rehabil, 2013. 94: 1753.
https://www.ncbi.nlm.nih.gov/pubmed/23529142
99. Husmann, D.A. Mortality following augmentation cystoplasty: A transitional urologist’s viewpoint. J
Pediatr Urol, 2017.
https://www.ncbi.nlm.nih.gov/pubmed/28645552
100. Yang, C.C., et al. Bladder management in women with neurologic disabilities. Phys Med Rehabil Clin
N Am, 2001. 12: 91.
https://www.ncbi.nlm.nih.gov/pubmed/11853041
101. Podnar, S., et al. Protocol for clinical neurophysiologic examination of the pelvic floor. Neurourol
Urodyn, 2001. 20: 669.
https://www.ncbi.nlm.nih.gov/pubmed/11746548
102. Harrison, S., et al. Urinary incontinence in neurological disease: assessment and management. NICE
Clinical Guideline 2012. [CG148].
https://www.nice.org.uk/guidance/cg148

NEURO-UROLOGY - LIMITED UPDATE MARCH 2020 33


103. Liu, N., et al. Autonomic dysreflexia severity during urodynamics and cystoscopy in individuals with
spinal cord injury. Spinal Cord, 2013. 51: 863.
https://www.ncbi.nlm.nih.gov/pubmed/24060768
104. Krassioukov, A., et al. International standards to document remaining autonomic function after
spinal cord injury. J Spinal Cord Med, 2012. 35: 201.
https://www.ncbi.nlm.nih.gov/pubmed/22925746
105. Labat, J.J., et al. Diagnostic criteria for pudendal neuralgia by pudendal nerve entrapment (Nantes
criteria). Neurourol Urodyn, 2008. 27: 306.
https://www.ncbi.nlm.nih.gov/pubmed/17828787
106. Brown, D., Atlas of regional anesthesia. 3rd. ed. 2006, Philadelphia
107. Standring, S., Gray’s anatomy, . 40th ed. 2008.
108. Bellucci, C.H., et al. Neurogenic lower urinary tract dysfunction--do we need same session repeat
urodynamic investigations? J Urol, 2012. 187: 1318.
https://www.ncbi.nlm.nih.gov/pubmed/22341264
109. Walter, M., et al. Autonomic dysreflexia and repeatability of cardiovascular changes during same
session repeat urodynamic investigation in women with spinal cord injury. World J Urol, 2015.
https://www.ncbi.nlm.nih.gov/pubmed/26055644
110. Walter, M., et al. Prediction of autonomic dysreflexia during urodynamics: A prospective cohort
study. BMC Med, 2018. 16: 53.
https://www.ncbi.nlm.nih.gov/pubmed/29650001
111. Gammie, A., et al. International Continence Society guidelines on urodynamic equipment
performance. Neurourol Urodyn, 2014. 33: 370.
https://www.ncbi.nlm.nih.gov/pubmed/24390971
112. McGuire, E.J., et al. Leak-point pressures. Urol Clin North Am, 1996. 23: 253.
https://www.ncbi.nlm.nih.gov/pubmed/8659025
113. Ozkan, B., et al. Which factors predict upper urinary tract deterioration in overactive neurogenic
bladder dysfunction? Urology, 2005. 66: 99.
https://www.ncbi.nlm.nih.gov/pubmed/15992868
114. Wang, Q.W., et al. Is it possible to use urodynamic variables to predict upper urinary tract dilatation
in children with neurogenic bladder-sphincter dysfunction? BJU Int, 2006. 98: 1295.
https://www.ncbi.nlm.nih.gov/pubmed/17034510
115. Musco, S., et al. Value of urodynamic findings in predicting upper urinary tract damage in neuro-
urological patients: A systematic review. Neurourol Urodyn, 2018.
https://www.ncbi.nlm.nih.gov/pubmed/29392753
116. Linsenmeyer, T.A., et al. The impact of urodynamic parameters on the upper tracts of spinal cord
injured men who void reflexly. J Spinal Cord Med, 1998. 21: 15.
https://www.ncbi.nlm.nih.gov/pubmed/9541882
117. McGuire, E.J., et al. Prognostic value of urodynamic testing in myelodysplastic patients. J Urol,
1981. 126: 205.
https://www.ncbi.nlm.nih.gov/pubmed/7196460
118. Krongrad, A., et al. Bladder neck dysynergia in spinal cord injury. Am J Phys Med Rehabil, 1996. 75:
204.
https://www.ncbi.nlm.nih.gov/pubmed/8663928
119. Weld, K.J., et al. Clinical significance of detrusor sphincter dyssynergia type in patients with post-
traumatic spinal cord injury. Urology, 2000. 56: 565.
https://www.ncbi.nlm.nih.gov/pubmed/11018603
120. Rossier, A.B., et al. 5-microtransducer catheter in evaluation of neurogenic bladder function.
Urology, 1986. 27: 371.
https://www.ncbi.nlm.nih.gov/pubmed/3962062
121. Al-Ali, M., et al. A 10 year review of the endoscopic treatment of 125 spinal cord injured patients
with vesical outlet obstruction: does bladder neck dyssynergia exist? Paraplegia, 1996. 34: 34.
https://www.ncbi.nlm.nih.gov/pubmed/8848321
122. Bacsu, C.D., et al. Diagnosing detrusor sphincter dyssynergia in the neurological patient. BJU Int,
2012. 109 Suppl 3: 31.
https://www.ncbi.nlm.nih.gov/pubmed/22458490
123. Lose, G., et al. Standardisation of urethral pressure measurement: report from the Standardisation
Sub-Committee of the International Continence Society. Neurourol Urodyn, 2002. 21: 258.
https://www.ncbi.nlm.nih.gov/pubmed/11948719
124. Marks, B.K., et al. Videourodynamics: indications and technique. Urol Clin North Am, 2014. 41: 383.
https://www.ncbi.nlm.nih.gov/pubmed/25063594

34 NEURO-UROLOGY - LIMITED UPDATE MARCH 2020


125. Virseda, M., et al. Reliability of ambulatory urodynamics in patients with spinal cord injuries.
Neurourol Urodyn, 2013. 32: 387.
https://www.ncbi.nlm.nih.gov/pubmed/23002043
126. Virseda-Chamorro, M., et al. Comparison of ambulatory versus video urodynamics in patients with
spinal cord injury. Spinal Cord, 2014. 52: 551.
https://www.ncbi.nlm.nih.gov/pubmed/24663000
127. Geirsson, G., et al. The ice-water test--a simple and valuable supplement to routine cystometry. Br J
Urol, 1993. 71: 681.
https://www.ncbi.nlm.nih.gov/pubmed/8343894
128. Geirsson, G., et al. Pressure, volume and infusion speed criteria for the ice-water test. Br J Urol,
1994. 73: 498.
https://www.ncbi.nlm.nih.gov/pubmed/8012770
129. Al-Hayek, S., et al. The 50-year history of the ice water test in urology. J Urol, 2010. 183: 1686.
https://www.ncbi.nlm.nih.gov/pubmed/20299050
130. Lapides, J. Neurogenic bladder. Principles of treatment. Urol Clin North Am, 1974. 1: 81.
https://www.ncbi.nlm.nih.gov/pubmed/4428540
131. Riedl, C.R., et al. Electromotive administration of intravesical bethanechol and the clinical impact on
acontractile detrusor management: introduction of a new test. J Urol, 2000. 164: 2108.
https://www.ncbi.nlm.nih.gov/pubmed/11061937
132. Podnar, S., et al. Lower urinary tract dysfunction in patients with peripheral nervous system lesions.
Handb Clin Neurol, 2015. 130: 203.
https://www.ncbi.nlm.nih.gov/pubmed/26003246
133. Ouyang, L., et al. Characteristics and survival of patients with end stage renal disease and spina
bifida in the United States renal data system. J Urol, 2015. 193: 558.
https://www.ncbi.nlm.nih.gov/pubmed/25167993
134. Lane, G.I., et al. Clinical outcomes of non-surgical management of detrusor leak point pressures
above 40 cm water in adults with congenital neurogenic bladder. Neurourol Urodyn, 2018. 37: 1943.
https://www.ncbi.nlm.nih.gov/pubmed/29488655
135. Lawrenson, R., et al. Renal failure in patients with neurogenic lower urinary tract dysfunction.
Neuroepidemiology, 2001. 20: 138.
https://www.ncbi.nlm.nih.gov/pubmed/11359083
136. Dangle, P.P., et al. Cystatin C-calculated Glomerular Filtration Rate-A Marker of Early Renal
Dysfunction in Patients With Neuropathic Bladder. Urology, 2017. 100: 213.
https://www.ncbi.nlm.nih.gov/pubmed/27542858
137. Mingat, N., et al. Prospective study of methods of renal function evaluation in patients with
neurogenic bladder dysfunction. Urology, 2013. 82: 1032.
https://www.ncbi.nlm.nih.gov/pubmed/24001705
138. Averbeck, M.A., et al. Follow-up of the neuro-urological patient: a systematic review. BJU Int, 2015.
115 Suppl 6: 39.
https://www.ncbi.nlm.nih.gov/pubmed/25891319
139. Stöhrer, M., et al. Diagnosis and treatment of bladder dysfunction in spinal cord injury patients. Eur
Urol Update Series 1994. 3: 170. [No abstract available].
140. Apostolidis, A., et al., Neurologic Urinary and Faecal Incontinence, In: Incontinence 6th Edition, P.
Abrams, L. Cardozo, S. Khoury & A. Wein, Editors. 2017.
141. Chamberlain, J.D., et al. Mortality and longevity after a spinal cord injury: systematic review and
meta-analysis. Neuroepidemiology, 2015. 44: 182.
https://www.ncbi.nlm.nih.gov/pubmed/25997873
142. Game, X., et al. Botulinum toxin A detrusor injections in patients with neurogenic detrusor
overactivity significantly decrease the incidence of symptomatic urinary tract infections. Eur Urol,
2008. 53: 613.
https://www.ncbi.nlm.nih.gov/pubmed/17804150
143. Frankel, H.L., et al. Long-term survival in spinal cord injury: a fifty year investigation. Spinal Cord,
1998. 36: 266.
https://www.ncbi.nlm.nih.gov/pubmed/9589527
144. Jamil, F. Towards a catheter free status in neurogenic bladder dysfunction: a review of bladder
management options in spinal cord injury (SCI). Spinal Cord, 2001. 39: 355.
https://www.ncbi.nlm.nih.gov/pubmed/11464308
145. Thietje, R., et al. Mortality in patients with traumatic spinal cord injury: descriptive analysis of 62
deceased subjects. J Spinal Cord Med, 2011. 34: 482.

NEURO-UROLOGY - LIMITED UPDATE MARCH 2020 35


https://www.ncbi.nlm.nih.gov/pubmed/22118255
146. Hackler, R.H. A 25-year prospective mortality study in the spinal cord injured patient: comparison
with the long-term living paraplegic. J Urol, 1977. 117: 486.
https://www.ncbi.nlm.nih.gov/pubmed/850323
147. Rodrigues, P., et al. Involuntary detrusor contraction is a frequent finding in patients with recurrent
urinary tract infections. Urol Int, 2014. 93: 67.
https://www.ncbi.nlm.nih.gov/pubmed/25011551
148. Bauer, S.B. Neurogenic bladder: etiology and assessment. Pediatr Nephrol, 2008. 23: 541.
https://www.ncbi.nlm.nih.gov/pubmed/18270749
149. Barbalias, G.A., et al. Critical evaluation of the Crede maneuver: a urodynamic study of 207 patients.
J Urol, 1983. 130: 720.
https://www.ncbi.nlm.nih.gov/pubmed/6887405
150. Reinberg, Y., et al. Renal rupture after the Crede maneuver. J Pediatr, 1994. 124: 279.
https://www.ncbi.nlm.nih.gov/pubmed/8301439
151. Wyndaele, J.J., et al. Neurologic urinary incontinence. Neurourol Urodyn, 2010. 29: 159.
https://www.ncbi.nlm.nih.gov/pubmed/20025021
152. Menon, E.B., et al. Bladder training in patients with spinal cord injury. Urology, 1992. 40: 425.
https://www.ncbi.nlm.nih.gov/pubmed/1441039
153. Furusawa, K., et al. Incidence of symptomatic autonomic dysreflexia varies according to the bowel
and bladder management techniques in patients with spinal cord injury. Spinal Cord, 2011. 49: 49.
https://www.ncbi.nlm.nih.gov/pubmed/20697419
154. Outcomes following traumatic spinal cord injury: clinical practice guidelines for health-care
professionals. J Spinal Cord Med, 2000. 23: 289.
https://www.ncbi.nlm.nih.gov/pubmed/17536300
155. El-Masri, W.S., et al. Long-term follow-up study of outcomes of bladder management in spinal cord
injury patients under the care of the Midlands Centre for Spinal Injuries in Oswestry. Spinal Cord,
2012. 50: 14.
https://www.ncbi.nlm.nih.gov/pubmed/21808256
156. Fall, M., et al. Electrical stimulation. A physiologic approach to the treatment of urinary incontinence.
Urol Clin North Am, 1991. 18: 393.
https://www.ncbi.nlm.nih.gov/pubmed/2017820
157. Vodusek, D.B., et al. Detrusor inhibition induced by stimulation of pudendal nerve afferents.
Neurourol Urodyn, 1986. 5: 381.
https://onlinelibrary.wiley.com/doi/abs/10.1002/nau.1930050404
158. Gross, T., et al. Transcutaneous Electrical Nerve Stimulation for Treating Neurogenic Lower Urinary
Tract Dysfunction: A Systematic Review. Eur Urol, 2016. 69: 1102.
https://www.ncbi.nlm.nih.gov/pubmed/26831506
159. Schneider, M.P., et al. Tibial Nerve Stimulation for Treating Neurogenic Lower Urinary Tract
Dysfunction: A Systematic Review. Eur Urol, 2015.
https://www.ncbi.nlm.nih.gov/pubmed/26194043
160. Booth, J., et al. The effectiveness of transcutaneous tibial nerve stimulation (TTNS) for adults with
overactive bladder syndrome: A systematic review. Neurourol Urodyn, 2018. 37: 528.
https://www.ncbi.nlm.nih.gov/pubmed/28731583
161. Liu, Y., et al. Effects of Transcutaneous Electrical Nerve Stimulation at Two Frequencies on Urinary
Incontinence in Poststroke Patients: A Randomized Controlled Trial. Am J Phys Med Rehabil, 2016.
95: 183.
https://www.ncbi.nlm.nih.gov/pubmed/26259053
162. Guo, G.Y., et al. Effectiveness of neuromuscular electrical stimulation therapy in patients with urinary
incontinence after stroke: A randomized sham controlled trial. Medicine, 2018. 97: e13702.
https://www.ncbi.nlm.nih.gov/pubmed/30593142
163. Shen, S.X., et al. A retrospective study of neuromuscular electrical stimulation for treating women
with post-stroke incontinence. Medicine (United States), 2018. 97: e11264.
https://www.ncbi.nlm.nih.gov/pubmed/29952999
164. McClurg, D., et al. Neuromuscular electrical stimulation and the treatment of lower urinary tract
dysfunction in multiple sclerosis--a double blind, placebo controlled, randomised clinical trial.
Neurourol Urodyn, 2008. 27: 231.
https://www.ncbi.nlm.nih.gov/pubmed/17705160
165. Ferreira, A.P.S., et al. A Controlled Clinical Trial On The Effects Of Exercise On Lower Urinary Tract
Symptoms In Women With Multiple Sclerosis. Am J Phys Med Rehabil, 2019.
https://www.ncbi.nlm.nih.gov/pubmed/30932917

36 NEURO-UROLOGY - LIMITED UPDATE MARCH 2020


166. McClurg, D., et al. Comparison of pelvic floor muscle training, electromyography biofeedback, and
neuromuscular electrical stimulation for bladder dysfunction in people with multiple sclerosis: a
randomized pilot study. Neurourol Urodyn, 2006. 25: 337.
https://www.ncbi.nlm.nih.gov/pubmed/16637070
167. Ferreira, A.P., et al. Impact of a Pelvic Floor Training Program Among Women with Multiple Sclerosis:
A Controlled Clinical Trial. Am J Phys Med Rehabil, 2016. 95: 1.
https://www.ncbi.nlm.nih.gov/pubmed/25888662
168. Elmelund, M., et al. The effect of pelvic floor muscle training and intravaginal electrical stimulation
on urinary incontinence in women with incomplete spinal cord injury: an investigator-blinded parallel
randomized clinical trial. Int Urogynecol J, 2018. 29: 1597.
https://www.ncbi.nlm.nih.gov/pubmed/29574482
169. Hagerty, J.A., et al. Intravesical electrotherapy for neurogenic bladder dysfunction: a 22-year
experience. J Urol, 2007. 178: 1680.
https://www.ncbi.nlm.nih.gov/pubmed/17707024
170. Primus, G., et al. Restoration of micturition in patients with acontractile and hypocontractile detrusor
by transurethral electrical bladder stimulation. Neurourol Urodyn, 1996. 15: 489.
https://www.ncbi.nlm.nih.gov/pubmed/8857617
171. Lombardi, G., et al. Clinical efficacy of intravesical electrostimulation on incomplete spinal cord
patients suffering from chronic neurogenic non-obstructive retention: a 15-year single centre
retrospective study. Spinal Cord, 2013. 51: 232.
https://www.ncbi.nlm.nih.gov/pubmed/23147136
172. Brusa, L., et al. Effects of inhibitory rTMS on bladder function in Parkinson’s disease patients. Mov
Disord, 2009. 24: 445.
https://www.ncbi.nlm.nih.gov/pubmed/19133657
173. Centonze, D., et al. Effects of motor cortex rTMS on lower urinary tract dysfunction in multiple
sclerosis. Mult Scler, 2007. 13: 269.
https://www.ncbi.nlm.nih.gov/pubmed/17439897
174. Thomas, L.H., et al. Treatment of urinary incontinence after stroke in adults. Cochrane Database
Syst Rev, 2008: CD004462.
https://www.ncbi.nlm.nih.gov/pubmed/18254050
175. Yeo, L., et al. Urinary tract dysfunction in Parkinson’s disease: a review. Int Urol Nephrol, 2012. 44:
415.
https://www.ncbi.nlm.nih.gov/pubmed/21553114
176. Phe, V., et al. Management of neurogenic bladder in patients with multiple sclerosis. Nature Reviews
Urology, 2016. 13: 275.
https://www.ncbi.nlm.nih.gov/pubmed/27030526
177. Andersson, K.E. Antimuscarinic mechanisms and the overactive detrusor: an update. Eur Urol,
2011. 59: 377.
https://www.ncbi.nlm.nih.gov/pubmed/21168951
178. Bennett, N., et al. Can higher doses of oxybutynin improve efficacy in neurogenic bladder? J Urol,
2004. 171: 749.
https://www.ncbi.nlm.nih.gov/pubmed/14713802
179. Horstmann, M., et al. Neurogenic bladder treatment by doubling the recommended antimuscarinic
dosage. Neurourol Urodyn, 2006. 25: 441.
https://www.ncbi.nlm.nih.gov/pubmed/16847942
180. Kennelly, M.J., et al. Overactive bladder: pharmacologic treatments in the neurogenic population.
Rev Urol, 2008. 10: 182.
https://www.ncbi.nlm.nih.gov/pubmed/18836537
181. Madersbacher, H., et al. Neurogenic detrusor overactivity in adults: a review on efficacy, tolerability
and safety of oral antimuscarinics. Spinal Cord, 2013. 51: 432.
https://www.ncbi.nlm.nih.gov/pubmed/23743498
182. Madhuvrata, P., et al. Anticholinergic drugs for adult neurogenic detrusor overactivity: a systematic
review and meta-analysis. Eur Urol, 2012. 62: 816.
https://www.ncbi.nlm.nih.gov/pubmed/22397851
183. Stohrer, M., et al. EAU guidelines on neurogenic lower urinary tract dysfunction. Eur Urol, 2009. 56:
81.
https://www.ncbi.nlm.nih.gov/pubmed/19403235
184. Mehnert, U., et al. The management of urinary incontinence in the male neurological patient. Curr
Opin Urol, 2014. 24: 586.
https://www.ncbi.nlm.nih.gov/pubmed/25389549
185. Stothers, L., et al. An integrative review of standardized clinical evaluation tool utilization in

NEURO-UROLOGY - LIMITED UPDATE MARCH 2020 37


anticholinergic drug trials for neurogenic lower urinary tract dysfunction. Spinal Cord, 2016. 31: 31.
https://www.ncbi.nlm.nih.gov/pubmed/27241452
186. Amend, B., et al. Effective treatment of neurogenic detrusor dysfunction by combined high-dosed
antimuscarinics without increased side-effects. Eur Urol, 2008. 53: 1021.
https://www.ncbi.nlm.nih.gov/pubmed/18243516
187. Cameron, A.P. Pharmacologic therapy for the neurogenic bladder. Urol Clin North Am, 2010. 37:
495.
https://www.ncbi.nlm.nih.gov/pubmed/20955901
188. Menarini, M., et al. Trospium chloride in patients with neurogenic detrusor overactivity: is dose
titration of benefit to the patients? Int J Clin Pharmacol Ther, 2006. 44: 623.
https://www.ncbi.nlm.nih.gov/pubmed/17190372
189. Nardulli, R., et al. Combined antimuscarinics for treatment of neurogenic overactive bladder. Int J
Immunopathol Pharmacol, 2012. 25: 35s.
https://www.ncbi.nlm.nih.gov/pubmed/22652160
190. Tijnagel, M.J., et al. Real life persistence rate with antimuscarinic treatment in patients with
idiopathic or neurogenic overactive bladder: a prospective cohort study with solifenacin. BMC
Urology, 2017. 17: 13.
https://www.ncbi.nlm.nih.gov/pubmed/28403849
191. Cameron, A.P., et al. Combination drug therapy improves compliance of the neurogenic bladder. J
Urol, 2009. 182: 1062.
https://www.ncbi.nlm.nih.gov/pubmed/19616807
192. Isik, A.T., et al. Trospium and cognition in patients with late onset Alzheimer disease. J Nutr Health
Aging, 2009. 13: 672.
https://www.ncbi.nlm.nih.gov/pubmed/19657549
193. Ethans, K.D., et al. Efficacy and safety of tolterodine in people with neurogenic detrusor overactivity.
J Spinal Cord Med, 2004. 27: 214.
https://www.ncbi.nlm.nih.gov/pubmed/15478523
194. McKeage, K. Propiverine: A review of its use in the treatment of adults and children with overactive
bladder associated with idiopathic or neurogenic detrusor overactivity, and in men with lower urinary
tract symptoms. Clin Drug Invest, 2013. 33: 71.
https://www.ncbi.nlm.nih.gov/pubmed/23288694
195. Nicholas, R.S., et al. Anticholinergics for urinary symptoms in multiple sclerosis. Cochrane Database
Syst Rev, 2009: CD004193.
https://www.ncbi.nlm.nih.gov/pubmed/19160231
196. van Rey, F., et al. Solifenacin in multiple sclerosis patients with overactive bladder: a prospective
study. Adv Urol, 2011. 2011: 834753.
https://www.ncbi.nlm.nih.gov/pubmed/21687581
197. Bycroft, J., et al. The effect of darifenacin on neurogenic detrusor overactivity in patients with spinal
cord injury. Neurourol Urodyn 2003. 22: A190.
https://pdfs.semanticscholar.org/ba7c/06ce8114149cfabffebf3f8a090ec06d5432.pdf
198. Carl, S., et al. Darifenacin is also effective in neurogenic bladder dysfunction (multiple sclerosis).
Urology, 2006. 68 250. [No abstract available].
199. Amarenco, G., et al. Solifenacin is effective and well tolerated in patients with neurogenic detrusor
overactivity: Results from the double-blind, randomized, active- and placebo-controlled SONIC
urodynamic study. Neurourol Urodyn, 2015. 29: 29.
https://www.ncbi.nlm.nih.gov/pubmed/26714009
200. Zesiewicz, T.A., et al. Randomized, controlled pilot trial of solifenacin succinate for overactive
bladder in Parkinson’s disease. Parkinsonism Rel Disord, 2015. 21: 514.
https://www.ncbi.nlm.nih.gov/pubmed/25814050
201. Sakakibara, R., et al. Imidafenacin on bladder and cognitive function in neurologic OAB patients.
Clin Auton Res, 2013. 23: 189.
https://www.ncbi.nlm.nih.gov/pubmed/23820664
202. Sugiyama, H., et al. Effect of imidafenacin on the urodynamic parameters of patients with indwelling
bladder catheters due to spinal cord injury. Spinal Cord, 2017. 55: 187.
https://www.ncbi.nlm.nih.gov/pubmed/27897185
203. Stohrer, M., et al. Efficacy and tolerability of propiverine hydrochloride extended-release compared
with immediate-release in patients with neurogenic detrusor overactivity. Spinal Cord, 2013. 51: 419.
https://www.ncbi.nlm.nih.gov/pubmed/23338657
204. Schroder, A., et al. Efficacy, safety, and tolerability of intravesically administered 0.1% oxybutynin

38 NEURO-UROLOGY - LIMITED UPDATE MARCH 2020


hydrochloride solution in adult patients with neurogenic bladder: A randomized, prospective,
controlled multi-center trial. Neurourol Urodyn, 2016. 35: 582.
https://www.ncbi.nlm.nih.gov/pubmed/25754454
205. Krhut, J., et al. Efficacy and safety of mirabegron for the treatment of neurogenic detrusor
overactivity-Prospective, randomized, double-blind, placebo-controlled study. Neurourol Urodyn,
2018. 37: 2226.
https://www.ncbi.nlm.nih.gov/pubmed/29603781
206. Welk, B., et al. A pilot randomized-controlled trial of the urodynamic efficacy of mirabegron for
patients with neurogenic lower urinary tract dysfunction. Neurourol Urodyn, 2018. 37: 2810.
https://www.ncbi.nlm.nih.gov/pubmed/30168626
207. Chen, S.F., et al. Therapeutic efficacy of low-dose (25mg) mirabegron therapy for patients with mild
to moderate overactive bladder symptoms due to central nervous system diseases. LUTS: Lower
Urinary Tract Symptoms, 2018.
https://www.ncbi.nlm.nih.gov/pubmed/29380517
208. Peyronnet, B., et al. Mirabegron in patients with Parkinson disease and overactive bladder
symptoms: A retrospective cohort. Parkinsonism Rel Disord, 2018. 57: 22.
https://www.ncbi.nlm.nih.gov/pubmed/30037689
209. Zachariou, A., et al. Effective treatment of neurogenic detrusor overactivity in multiple sclerosis
patients using desmopressin and mirabegron. Can J Urol, 2017. 24: 9107.
https://www.ncbi.nlm.nih.gov/pubmed/29260636
210. Abo Youssef, N., et al. Cannabinoids for treating neurogenic lower urinary tract dysfunction in
patients with multiple sclerosis: a systematic review and meta-analysis. BJU Int, 2017. 119: 515.
https://www.ncbi.nlm.nih.gov/pubmed/28058780
211. Francomano, D., et al. Effects of daily tadalafil on lower urinary tract symptoms in young men with
multiple sclerosis and erectile dysfunction: a pilot study. J Endocrinol Invest, 2017. 40: 275.
https://www.ncbi.nlm.nih.gov/pubmed/27752863
212. Phe, V., et al. Desmopressin for treating nocturia in patients with multiple sclerosis: A systematic
review: A report from the Neuro-Urology Promotion Committee of the International Continence
Society (ICS). Neurourol Urodyn, 2019. 38: 563.
https://www.ncbi.nlm.nih.gov/pubmed/30653737
213. Barendrecht, M.M., et al. Is the use of parasympathomimetics for treating an underactive urinary
bladder evidence-based? BJU Int, 2007. 99: 749.
https://www.ncbi.nlm.nih.gov/pubmed/17233798
214. Apostolidis, A. Taming the cannabinoids: new potential in the pharmacologic control of lower urinary
tract dysfunction. Eur Urol, 2012. 61: 107.
https://www.ncbi.nlm.nih.gov/pubmed/21996529
215. Gratzke, C., et al. Effects of cannabinor, a novel selective cannabinoid 2 receptor agonist, on
bladder function in normal rats. Eur Urol, 2010. 57: 1093.
https://www.ncbi.nlm.nih.gov/pubmed/20207474
216. Abrams, P., et al. Tamsulosin: efficacy and safety in patients with neurogenic lower urinary tract
dysfunction due to suprasacral spinal cord injury. J Urol, 2003. 170: 1242.
https://www.ncbi.nlm.nih.gov/pubmed/14501734
217. Gomes, C.M., et al. Neurological status predicts response to alpha-blockers in men with voiding
dysfunction and Parkinson’s disease. Clinics, 2014. 69: 817.
https://www.ncbi.nlm.nih.gov/pubmed/25627993
218. Moon, K.H., et al. A 12-week, open label, multi-center study to evaluate the clinical efficacy and
safety of silodosin on voiding dysfunction in patients with neurogenic bladder. LUTS: Lower Urinary
Tract Symptoms, 2015. 7: 27.
https://www.ncbi.nlm.nih.gov/pubmed/26663648
219. Guttmann, L., et al. The value of intermittent catheterisation in the early management of traumatic
paraplegia and tetraplegia. Paraplegia, 1966. 4: 63.
https://www.ncbi.nlm.nih.gov/pubmed/5969402
220. Lapides, J., et al. Clean, intermittent self-catheterization in the treatment of urinary tract disease. J
Urol, 1972. 107: 458.
https://www.ncbi.nlm.nih.gov/pubmed/5010715
221. Wyndaele, J.J. Intermittent catheterization: which is the optimal technique? Spinal Cord, 2002. 40:
432.
https://www.ncbi.nlm.nih.gov/pubmed/12185603
222. Prieto-Fingerhut, T., et al. A study comparing sterile and nonsterile urethral catheterization in
patients with spinal cord injury. Rehabil Nurs, 1997. 22: 299.

NEURO-UROLOGY - LIMITED UPDATE MARCH 2020 39


https://www.ncbi.nlm.nih.gov/pubmed/9416190
223. Prieto, J., et al. Intermittent catheterisation for long-term bladder management. Cochrane Database
Syst Rev, 2014: CD006008.
https://www.ncbi.nlm.nih.gov/pubmed/25208303
224. Kiddoo, D., et al. Randomized Crossover Trial of Single Use Hydrophilic Coated vs Multiple Use
Polyvinylchloride Catheters for Intermittent Catheterization to Determine Incidence of Urinary
Infection. J Urol, 2015. 194: 174.
https://www.ncbi.nlm.nih.gov/pubmed/25584995
225. Goetz, L.L., et al. International Spinal Cord Injury Urinary Tract Infection Basic Data Set. Spinal Cord,
2013. 51: 700.
https://www.ncbi.nlm.nih.gov/pubmed/23896666
226. Bakke, A., et al. Physical predictors of infection in patients treated with clean intermittent
catheterization: a prospective 7-year study. Br J Urol, 1997. 79: 85.
https://www.ncbi.nlm.nih.gov/pubmed/9043503
227. Günther, M., et al. Auswirkungen des aseptischen intermittierenden Katheterismus auf die männliche
Harnröhre. Der Urologe B, 2001. 41: 359.
https://link.springer.com/article/10.1007%2Fs001310170044
228. Kurze, I., et al. Intermittent Catheterisation and Prevention of Urinary Tract Infections in Patients
with Neurogenic Lower Urinary Tract Dysfunction - Best PracticeAn Overview. [German]. Aktuelle
Neurologie, 2015. 42: 515. [No abstract available].
229. Waller, L., et al. Clean intermittent catheterization in spinal cord injury patients: long-term followup of
a hydrophilic low friction technique. J Urol, 1995. 153: 345.
https://www.ncbi.nlm.nih.gov/pubmed/7815579
230. Wyndaele, J.J. Complications of intermittent catheterization: their prevention and treatment. Spinal
Cord, 2002. 40: 536.
https://www.ncbi.nlm.nih.gov/pubmed/12235537
231. Woodbury, M.G., et al. Intermittent catheterization practices following spinal cord injury: a national
survey. Can J Urol, 2008. 15: 4065.
https://www.ncbi.nlm.nih.gov/pubmed/18570710
232. Bennett, C.J., et al. Comparison of bladder management complication outcomes in female spinal
cord injury patients. J Urol, 1995. 153: 1458.
https://www.ncbi.nlm.nih.gov/pubmed/7714965
233. Chao, R., et al. Fate of upper urinary tracts in patients with indwelling catheters after spinal cord
injury. Urology, 1993. 42: 259.
https://www.ncbi.nlm.nih.gov/pubmed/8379025
234. Larsen, L.D., et al. Retrospective analysis of urologic complications in male patients with spinal cord
injury managed with and without indwelling urinary catheters. Urology, 1997. 50: 418.
https://www.ncbi.nlm.nih.gov/pubmed/9301708
235. Mitsui, T., et al. Is suprapubic cystostomy an optimal urinary management in high quadriplegics?. A
comparative study of suprapubic cystostomy and clean intermittent catheterization. Eur Urol, 2000.
38: 434.
https://www.ncbi.nlm.nih.gov/pubmed/11025382
236. Weld, K.J., et al. Effect of bladder management on urological complications in spinal cord injured
patients. J Urol, 2000. 163: 768.
https://www.ncbi.nlm.nih.gov/pubmed/10687973
237. Weld, K.J., et al. Influences on renal function in chronic spinal cord injured patients. J Urol, 2000.
164: 1490.
https://www.ncbi.nlm.nih.gov/pubmed/11025689
238. West, D.A., et al. Role of chronic catheterization in the development of bladder cancer in patients
with spinal cord injury. Urology, 1999. 53: 292.
https://www.ncbi.nlm.nih.gov/pubmed/9933042
239. Lavelle, R.S., et al. Quality of life after suprapubic catheter placement in patients with neurogenic
bladder conditions. Neurourol Urodyn, 2016. 35: 831.
https://www.ncbi.nlm.nih.gov/pubmed/26197729
240. Hollingsworth, J.M., et al. Determining the noninfectious complications of indwelling urethral
catheters: a systematic review and meta-analysis. Ann Intern Med, 2013. 159: 401.
https://www.ncbi.nlm.nih.gov/pubmed/24042368
241. Buyse, G., et al. Intravesical oxybutynin for neurogenic bladder dysfunction: less systemic side
effects due to reduced first pass metabolism. J Urol, 1998. 160: 892.
https://www.ncbi.nlm.nih.gov/pubmed/9720583
242. Di Stasi, S.M., et al. Intravesical oxybutynin: mode of action assessed by passive diffusion and

40 NEURO-UROLOGY - LIMITED UPDATE MARCH 2020


electromotive administration with pharmacokinetics of oxybutynin and N-desethyl oxybutynin. J
Urol, 2001. 166: 2232.
https://www.ncbi.nlm.nih.gov/pubmed/11696741
243. Haferkamp, A., et al. Dosage escalation of intravesical oxybutynin in the treatment of neurogenic
bladder patients. Spinal Cord, 2000. 38: 250.
https://www.ncbi.nlm.nih.gov/pubmed/10822396
244. Pannek, J., et al. Combined intravesical and oral oxybutynin chloride in adult patients with spinal
cord injury. Urology, 2000. 55: 358.
https://www.ncbi.nlm.nih.gov/pubmed/10699610
245. Giannantoni, A., et al. Intravesical resiniferatoxin versus botulinum-A toxin injections for neurogenic
detrusor overactivity: a prospective randomized study. J Urol, 2004. 172: 240.
https://www.ncbi.nlm.nih.gov/pubmed/15201783
246. Kim, J.H., et al. Intravesical resiniferatoxin for refractory detrusor hyperreflexia: a multicenter,
blinded, randomized, placebo-controlled trial. J Spinal Cord Med, 2003. 26: 358.
https://www.ncbi.nlm.nih.gov/pubmed/14992337
247. Phe, V., et al. Intravesical vanilloids for treating neurogenic lower urinary tract dysfunction in patients
with multiple sclerosis: A systematic review and meta-analysis. A report from the Neuro-Urology
Promotion Committee of the International Continence Society (ICS). Neurourol Urodyn, 2018. 37: 67.
https://www.ncbi.nlm.nih.gov/pubmed/28618110
248. Del Popolo, G., et al. Neurogenic detrusor overactivity treated with english botulinum toxin a: 8-year
experience of one single centre. Eur Urol, 2008. 53: 1013.
https://www.ncbi.nlm.nih.gov/pubmed/17950989
249. Reitz, A., et al. European experience of 200 cases treated with botulinum-A toxin injections into the
detrusor muscle for urinary incontinence due to neurogenic detrusor overactivity. Eur Urol, 2004. 45:
510.
https://www.ncbi.nlm.nih.gov/pubmed/15041117
250. Yuan, H., et al. Efficacy and Adverse Events Associated With Use of OnabotulinumtoxinA for
Treatment of Neurogenic Detrusor Overactivity: A Meta-Analysis. Int Neurourol J, 2017. 21: 53.
https://www.ncbi.nlm.nih.gov/pubmed/28361515
251. Cheng, T., et al. Efficacy and Safety of OnabotulinumtoxinA in Patients with Neurogenic Detrusor
Overactivity: A Systematic Review and Meta-Analysis of Randomized Controlled Trials. PLoS One,
2016. 11: e0159307.
https://www.ncbi.nlm.nih.gov/pubmed/27463810
252. Wagle Shukla, A., et al. Botulinum Toxin Therapy for Parkinson’s Disease. Seminars in Neurology,
2017. 37: 193.
https://www.ncbi.nlm.nih.gov/pubmed/28511260
253. Koschorke, M., et al. Intradetrusor onabotulinumtoxinA injections for refractory neurogenic detrusor
overactivity incontinence: do we need urodynamic investigation for outcome assessment? BJU
International, 2017. 120: 848.
https://www.ncbi.nlm.nih.gov/pubmed/28771936
254. Ginsberg, D., et al. Phase 3 efficacy and tolerability study of onabotulinumtoxinA for urinary
incontinence from neurogenic detrusor overactivity. J Urol, 2012. 187: 2131.
https://www.ncbi.nlm.nih.gov/pubmed/22503020
255. Grosse, J., et al. Success of repeat detrusor injections of botulinum a toxin in patients with severe
neurogenic detrusor overactivity and incontinence. Eur Urol, 2005. 47: 653.
https://www.ncbi.nlm.nih.gov/pubmed/15826758
256. Rovner, E., et al. Long-Term Efficacy and Safety of OnabotulinumtoxinA in Patients with Neurogenic
Detrusor Overactivity Who Completed 4 Years of Treatment. J Urol, 2016.
https://www.ncbi.nlm.nih.gov/pubmed/27091236
257. Ni, J., et al. Is repeat Botulinum Toxin A injection valuable for neurogenic detrusor overactivity-A
systematic review and meta-analysis. Neurourol Urodyn, 2018. 37: 542.
https://www.ncbi.nlm.nih.gov/pubmed/28745818
258. Michel, F., et al. Botulinum toxin type A injection after failure of augmentation enterocystoplasty
performed for neurogenic detrusor overactivity: preliminary results of a salvage strategy. The
ENTEROTOX study. Urology, 2019.
https://www.ncbi.nlm.nih.gov/pubmed/30926380
259. Bottet, F., et al. Switch to Abobotulinum toxin A may be useful in the treatment of neurogenic
detrusor overactivity when intradetrusor injections of Onabotulinum toxin A failed. Neurourol
Urodyn, 2017. 21: 21.
https://www.ncbi.nlm.nih.gov/pubmed/28431196

NEURO-UROLOGY - LIMITED UPDATE MARCH 2020 41


260. Leu, R., et al. Complications of Botox and their Management. Curr Urol Rep, 2018. 19: 90.
https://www.ncbi.nlm.nih.gov/pubmed/30194497
261. Tullman, M., et al. Low-dose onabotulinumtoxinA improves urinary symptoms in noncatheterizing
patients with MS. Neurology, 2018. 91: e657.
https://www.ncbi.nlm.nih.gov/pubmed/30030330
262. Tyagi, P., et al. Past, Present and Future of Chemodenervation with Botulinum Toxin in the Treatment
of Overactive Bladder. J Urol, 2017. 197: 982.
https://www.ncbi.nlm.nih.gov/pubmed/27871929
263. Young, M.J., et al. Another Therapeutic Role for Intravesical Botulinum Toxin: Patients with Long-
stay Catheters and Refractory Bladder Pain and Catheter Bypass Leakage. Eur Urol Focus, 2018.
https://www.ncbi.nlm.nih.gov/pubmed/30392867
264. Dykstra, D.D., et al. Treatment of detrusor-sphincter dyssynergia with botulinum A toxin: a double-
blind study. Arch Phys Med Rehabil, 1990. 71: 24.
https://www.ncbi.nlm.nih.gov/pubmed/2297305
265. Schurch, B., et al. Botulinum-A toxin as a treatment of detrusor-sphincter dyssynergia: a
prospective study in 24 spinal cord injury patients. J Urol, 1996. 155: 1023.
https://www.ncbi.nlm.nih.gov/pubmed/8583552
266. Huang, M., et al. Effects of botulinum toxin A injections in spinal cord injury patients with detrusor
overactivity and detrusor sphincter dyssynergia. J Rehabil Med, 2016. 48: 683.
https://www.ncbi.nlm.nih.gov/pubmed/27563834
267. Utomo, E., et al. Surgical management of functional bladder outlet obstruction in adults with
neurogenic bladder dysfunction. Cochrane Database Syst Rev, 2014. 5: CD004927.
https://www.ncbi.nlm.nih.gov/pubmed/24859260
268. Chancellor, M.B., et al. Prospective comparison of external sphincter balloon dilatation and
prosthesis placement with external sphincterotomy in spinal cord injured men. Arch Phys Med
Rehabil, 1994. 75: 297.
https://www.ncbi.nlm.nih.gov/pubmed/8129583
269. Bennett, J.K., et al. Collagen injections for intrinsic sphincter deficiency in the neuropathic urethra.
Paraplegia, 1995. 33: 697.
https://www.ncbi.nlm.nih.gov/pubmed/
270. Block, C.A., et al. Long-term efficacy of periurethral collagen injection for the treatment of urinary
incontinence secondary to myelomeningocele. J Urol, 2003. 169: 327.
https://www.ncbi.nlm.nih.gov/pubmed/12478183
271. Schurch, B., et al. Intraurethral sphincter prosthesis to treat hyporeflexic bladders in women: does it
work? BJU Int, 1999. 84: 789.
https://www.ncbi.nlm.nih.gov/pubmed/10532973
272. Reuvers, S.H.M., et al. Heterogeneity in reporting on urinary outcome and cure after surgical
interventions for stress urinary incontinence in adult neuro-urological patients: A systematic review.
Neurourol Urodyn, 2018. 37: 554.
https://www.ncbi.nlm.nih.gov/pubmed/28792081
273. Barthold, J.S., et al. Results of the rectus fascial sling and wrap procedures for the treatment of
neurogenic sphincteric incontinence. J Urol, 1999. 161: 272.
https://www.ncbi.nlm.nih.gov/pubmed/10037423
274. Gormley, E.A., et al. Pubovaginal slings for the management of urinary incontinence in female
adolescents. J Urol, 1994. 152: 822.
https://www.ncbi.nlm.nih.gov/pubmed/8022024
275. Kakizaki, H., et al. Fascial sling for the management of urinary incontinence due to sphincter
incompetence. J Urol, 1995. 153: 644.
https://www.ncbi.nlm.nih.gov/pubmed/7861504
276. Mingin, G.C., et al. The rectus myofascial wrap in the management of urethral sphincter
incompetence. BJU Int, 2002. 90: 550.
https://www.ncbi.nlm.nih.gov/pubmed/12230615
277. Abdul-Rahman, A., et al. Long-term outcome of tension-free vaginal tape for treating stress
incontinence in women with neuropathic bladders. BJU Int, 2010. 106: 827.
https://www.ncbi.nlm.nih.gov/pubmed/20132201
278. Losco, G.S., et al. Long-term outcome of transobturator tape (TOT) for treatment of stress urinary
incontinence in females with neuropathic bladders. Spinal Cord, 2015. 53: 544.
https://www.ncbi.nlm.nih.gov/pubmed/25917951
279. El-Azab, A.S., et al. Midurethral slings versus the standard pubovaginal slings for women with
neurogenic stress urinary incontinence. Int Urogynecol J, 2015. 26: 427.

42 NEURO-UROLOGY - LIMITED UPDATE MARCH 2020


https://www.ncbi.nlm.nih.gov/pubmed/25315169
280. Athanasopoulos, A., et al. Treating stress urinary incontinence in female patients with neuropathic
bladder: the value of the autologous fascia rectus sling. Int Urol Nephrol, 2012. 44: 1363.
https://www.ncbi.nlm.nih.gov/pubmed/22821050
281. Groen, L.A., et al. The AdVance male sling as a minimally invasive treatment for intrinsic sphincter
deficiency in patients with neurogenic bladder sphincter dysfunction: a pilot study. Neurourol
Urodyn, 2012. 31: 1284.
https://www.ncbi.nlm.nih.gov/pubmed/22847896
282. Mehnert, U., et al. Treatment of neurogenic stress urinary incontinence using an adjustable
continence device: 4-year followup. J Urol, 2012. 188: 2274.
https://www.ncbi.nlm.nih.gov/pubmed/23083648
283. Daneshmand, S., et al. Puboprostatic sling repair for treatment of urethral incompetence in adult
neurogenic incontinence. J Urol, 2003. 169: 199.
https://www.ncbi.nlm.nih.gov/pubmed/12478135
284. Herschorn, S., et al. Fascial slings and bladder neck tapering in the treatment of male neurogenic
incontinence. J Urol, 1992. 147: 1073.
https://www.ncbi.nlm.nih.gov/pubmed/1552586
285. Light, J.K., et al. Use of the artificial urinary sphincter in spinal cord injury patients. J Urol, 1983.
130: 1127.
https://www.ncbi.nlm.nih.gov/pubmed/6644893
286. Farag, F., et al. Surgical treatment of neurogenic stress urinary incontinence: A systematic review of
quality assessment and surgical outcomes. Neurourol Urodyn, 2016. 35: 21.
https://www.ncbi.nlm.nih.gov/pubmed/25327633
287. Kim, S.P., et al. Long-term durability and functional outcomes among patients with artificial urinary
sphincters: a 10-year retrospective review from the University of Michigan. J Urol, 2008. 179: 1912.
https://www.ncbi.nlm.nih.gov/pubmed/18353376
288. Wang, R., et al. Long-term outcomes after primary failures of artificial urinary sphincter implantation.
Urology, 2012. 79: 922.
https://www.ncbi.nlm.nih.gov/pubmed/22305763
289. Guillot-Tantay, C., et al. [Male neurogenic stress urinary incontinence treated by artificial urinary
sphincter AMS 800TM (Boston Scientific, Boston, USA): Very long-term results (>25 years)].
Traitement de l’incontinence urinaire masculine neurologique par le sphincter urinaire artificiel AMS
800TM (Boston Scientific, Boston, Etats-Unis) : resultats a tres long terme (>25 ans). 2018. 28: 39.
https://www.ncbi.nlm.nih.gov/pubmed/29102375
290. Fournier, G., et al. Robotic-assisted laparoscopic implantation of artificial urinary sphincter in
women with intrinsic sphincter deficiency incontinence: initial results. Urology, 2014. 84: 1094.
https://www.ncbi.nlm.nih.gov/pubmed/25443911
291. Biardeau, X., et al. Robot-assisted laparoscopic approach for artificial urinary sphincter implantation
in 11 women with urinary stress incontinence: surgical technique and initial experience. Eur Urol,
2015. 67: 937.
https://www.ncbi.nlm.nih.gov/pubmed/25582931
292. Peyronnet, B., et al. Artificial urinary sphincter implantation in women with stress urinary
incontinence: preliminary comparison of robot-assisted and open approaches. Int Urogynecol J,
2016. 27: 475.
https://www.ncbi.nlm.nih.gov/pubmed/26431841
293. Phe, V., et al. Stress urinary incontinence in female neurological patients: long-term functional
outcomes after artificial urinary sphincter (AMS 800(TM) ) implantation. Neurourol Urodyn, 2017. 36:
764.
https://www.ncbi.nlm.nih.gov/pubmed/27080729
294. Scott, K.A., et al. Use of Artificial Urinary Sphincter and Slings to Manage Neurogenic Bladder
Following Spinal Cord Injury-Is It Safe? Curr Bladder Dysf Rep, 2017. 12: 311.
https://link.springer.com/article/10.1007/s11884-017-0449-9
295. Ammirati, E., et al. Management of male and female neurogenic stress urinary incontinence in spinal
cord injured (SCI) patients using adjustable continence therapy. Urologia, 2017. 0: 16.
https://www.ncbi.nlm.nih.gov/pubmed/28525663
296. Ronzi, Y., et al. Neurogenic stress urinary incontinence: is there a place for Adjustable Continence
Therapy (ACTTM and ProACTTM, Uromedica, Plymouth, MN, USA)? A retrospective multicenter
study. Spinal Cord, 2019.
https://www.nature.com/articles/s41393-018-0219-3
297. Janknegt, R.A., et al. Electrically stimulated gracilis sphincter for treatment of bladder sphincter
incontinence. Lancet, 1992. 340: 1129.

NEURO-UROLOGY - LIMITED UPDATE MARCH 2020 43


https://www.ncbi.nlm.nih.gov/pubmed/1359213
298. Chancellor, M.B., et al. Gracilis muscle transposition with electrical stimulation for sphincteric
incontinence: a new approach. World J Urol, 1997. 15: 320.
https://www.ncbi.nlm.nih.gov/pubmed/9372585
299. Chancellor, M.B., et al. Gracilis urethromyoplasty--an autologous urinary sphincter for neurologically
impaired patients with stress incontinence. Spinal Cord, 1997. 35: 546.
https://www.ncbi.nlm.nih.gov/pubmed/9267922
300. Donnahoo, K.K., et al. The Young-Dees-Leadbetter bladder neck repair for neurogenic incontinence.
J Urol, 1999. 161: 1946.
https://www.ncbi.nlm.nih.gov/pubmed/10332478
301. Kropp, K.A., et al. Urethral lengthening and reimplantation for neurogenic incontinence in children. J
Urol, 1986. 135: 533.
https://www.ncbi.nlm.nih.gov/pubmed/3944902
302. Salle, J.L., et al. Urethral lengthening with anterior bladder wall flap (Pippi Salle procedure):
modifications and extended indications of the technique. J Urol, 1997. 158: 585.
https://www.ncbi.nlm.nih.gov/pubmed/9224369
303. Rawashdeh, Y.F., et al. International Children’s Continence Society’s recommendations for
therapeutic intervention in congenital neuropathic bladder and bowel dysfunction in children.
Neurourol Urodyn, 2012. 31: 615.
https://www.ncbi.nlm.nih.gov/pubmed/22532368
304. Wyndaele, J.-J., et al. Surgical management of the neurogenic bladder after spinal cord injury. World
J Urol, 2018.
https://www.ncbi.nlm.nih.gov/pubmed/29680953
305. Moisey, C.U., et al. Results of transurethral resection of prostate in patients with cerebrovascular
disease. Br J Urol, 1978. 50: 539.
https://www.ncbi.nlm.nih.gov/pubmed/88982
306. Roth, B., et al. Benign prostatic obstruction and parkinson’s disease--should transurethral resection
of the prostate be avoided? J Urol, 2009. 181: 2209.
https://www.ncbi.nlm.nih.gov/pubmed/19296974
307. Elsaesser, E., et al. Urological operations for improvement of bladder voiding in paraplegic patients.
Paraplegia, 1972. 10: 68.
https://www.ncbi.nlm.nih.gov/pubmed/5039331
308. Cornejo-Davila, V., et al. Incidence of Urethral Stricture in Patients With Spinal Cord Injury Treated
With Clean Intermittent Self-Catheterization. Urology, 2017. 99: 260.
https://www.ncbi.nlm.nih.gov/pubmed/27566143
309. Perkash, I. Ablation of urethral strictures using contact chisel crystal firing neodymium:YAG laser. J
Urol, 1997. 157: 809.
https://www.ncbi.nlm.nih.gov/pubmed/9072572
310. Schurch, B., et al. Botulinum toxin type a is a safe and effective treatment for neurogenic urinary
incontinence: results of a single treatment, randomized, placebo controlled 6-month study. J Urol,
2005. 174: 196.
https://www.ncbi.nlm.nih.gov/pubmed/15947626
311. Madersbacher, H., et al. Twelve o’clock sphincterotomy: technique, indications, results. (Abbreviated
report). Urol Int, 1975. 30: 75.
https://www.ncbi.nlm.nih.gov/pubmed/1118951
312. Perkash, I. Laser sphincterotomy and ablation of the prostate using a sapphire chisel contact tip
firing neodymium:YAG laser. J Urol, 1994. 152: 2020.
https://www.ncbi.nlm.nih.gov/pubmed/7966667
313. Noll, F., et al. Transurethral sphincterotomy in quadriplegic patients: long-term-follow-up. Neurourol
Urodyn, 1995. 14: 351.
https://www.ncbi.nlm.nih.gov/pubmed/7581471
314. Derry, F., et al. Audit of bladder neck resection in spinal cord injured patients. Spinal Cord, 1998. 36:
345.
https://www.ncbi.nlm.nih.gov/pubmed/9601115
315. Perkash, I. Use of contact laser crystal tip firing Nd:YAG to relieve urinary outflow obstruction in
male neurogenic bladder patients. J Clin Laser Med Surg, 1998. 16: 33.
https://www.ncbi.nlm.nih.gov/pubmed/9728128
316. Chancellor, M.B., et al. Long-term followup of the North American multicenter UroLume trial for the
treatment of external detrusor-sphincter dyssynergia. J Urol, 1999. 161: 1545.
https://www.ncbi.nlm.nih.gov/pubmed/10210393

44 NEURO-UROLOGY - LIMITED UPDATE MARCH 2020


317. Seoane-Rodriguez, S., et al. Long-term follow-up study of intraurethral stents in spinal cord injured
patients with detrusor-sphincter dyssynergia. Spinal Cord, 2007. 45: 621.
https://www.ncbi.nlm.nih.gov/pubmed/17211463
318. Gajewski, J.B., et al. Removal of UroLume endoprosthesis: experience of the North American Study
Group for detrusor-sphincter dyssynergia application. J Urol, 2000. 163: 773.
https://www.ncbi.nlm.nih.gov/pubmed/10687974
319. Wilson, T.S., et al. UroLume stents: lessons learned. J Urol, 2002. 167: 2477.
https://www.ncbi.nlm.nih.gov/pubmed/11992061
320. Abdul-Rahman, A., et al. A 20-year follow-up of the mesh wallstent in the treatment of detrusor
external sphincter dyssynergia in patients with spinal cord injury. BJU Int, 2010. 106: 1510.
https://www.ncbi.nlm.nih.gov/pubmed/20500511
321. Pannek, J., et al. Clinical usefulness of the memokath stent as a second-line procedure after
sphincterotomy failure. J Endourol, 2011. 25: 335.
https://www.ncbi.nlm.nih.gov/pubmed/20977372
322. Polguer, T., et al. [Treatment of detrusor-striated sphincter dyssynergia with permanent nitinol
urethral stent: results after a minimum follow-up of 2 years]. Prog Urol, 2012. 22: 1058.
https://www.ncbi.nlm.nih.gov/pubmed/23182120
323. van der Merwe, A., et al. Outcome of dual flange metallic urethral stents in the treatment of
neuropathic bladder dysfunction after spinal cord injury. J Endourol, 2012. 26: 1210.
https://www.ncbi.nlm.nih.gov/pubmed/22519741
324. Brindley, G.S. An implant to empty the bladder or close the urethra. J Neurol Neurosurg Psychiatry,
1977. 40: 358.
https://www.ncbi.nlm.nih.gov/pubmed/406364
325. Krasmik, D., et al. Urodynamic results, clinical efficacy, and complication rates of sacral intradural
deafferentation and sacral anterior root stimulation in patients with neurogenic lower urinary tract
dysfunction resulting from complete spinal cord injury. Neurourol Urodyn, 2014. 33: 1202.
https://www.ncbi.nlm.nih.gov/pubmed/24038405
326. Benard, A., et al. Comparative cost-effectiveness analysis of sacral anterior root stimulation for
rehabilitation of bladder dysfunction in spinal cord injured patients. Neurosurgery, 2013. 73: 600.
https://www.ncbi.nlm.nih.gov/pubmed/23787880
327. Martens, F.M., et al. Quality of life in complete spinal cord injury patients with a Brindley bladder
stimulator compared to a matched control group. Neurourol Urodyn, 2011. 30: 551.
https://www.ncbi.nlm.nih.gov/pubmed/21328472
328. Krebs, J., et al. Long-term course of sacral anterior root stimulation in spinal cord injured individuals:
The fate of the detrusor. Neurourol Urodyn, 2017. 36: 1596.
https://www.ncbi.nlm.nih.gov/pubmed/27778371
329. Krebs, J., et al. Charcot arthropathy of the spine in spinal cord injured individuals with sacral
deafferentation and anterior root stimulator implantation. Neurourol Urodyn, 2016. 35: 241.
https://www.ncbi.nlm.nih.gov/pubmed/25524388
330. Nagib, A., et al. Successful control of selective anterior sacral rhizotomy for treatment of spastic
bladder and ureteric reflux in paraplegics. Med Serv J Can, 1966. 22: 576.
https://www.ncbi.nlm.nih.gov/pubmed/5966992
331. Schneidau, T., et al. Selective sacral rhizotomy for the management of neurogenic bladders in spina
bifida patients: long-term followup. J Urol, 1995. 154: 766.
https://www.ncbi.nlm.nih.gov/pubmed/7609174
332. Young, B., et al. Percutaneous sacral rhizotomy for neurogenic detrusor hyperreflexia. J Neurosurg,
1980. 53: 85.
https://www.ncbi.nlm.nih.gov/pubmed/7411212
333. Koldewijn, E.L., et al. Bladder compliance after posterior sacral root rhizotomies and anterior sacral
root stimulation. J Urol, 1994. 151: 955.
https://www.ncbi.nlm.nih.gov/pubmed/8126835
334. Singh, G., et al. Intravesical oxybutynin in patients with posterior rhizotomies and sacral anterior root
stimulators. Neurourol Urodyn, 1995. 14: 65.
https://www.ncbi.nlm.nih.gov/pubmed/7742851
335. Van Kerrebroeck, P.E., et al. Results of the treatment of neurogenic bladder dysfunction in spinal
cord injury by sacral posterior root rhizotomy and anterior sacral root stimulation. J Urol, 1996. 155:
1378.
https://www.ncbi.nlm.nih.gov/pubmed/8632580
336. Kutzenberger, J.S. Surgical therapy of neurogenic detrusor overactivity (hyperreflexia) in paraplegic
patients by sacral deafferentation and implant driven micturition by sacral anterior root stimulation:

NEURO-UROLOGY - LIMITED UPDATE MARCH 2020 45


methods, indications, results, complications, and future prospects. Acta Neurochir, 2007. 97: 333.
https://www.ncbi.nlm.nih.gov/pubmed/17691394
337. Bhadra, N., et al. Selective suppression of sphincter activation during sacral anterior nerve root
stimulation. Neurourol Urodyn, 2002. 21: 55.
https://www.ncbi.nlm.nih.gov/pubmed/11835425
338. Kirkham, A.P., et al. Neuromodulation through sacral nerve roots 2 to 4 with a Finetech-Brindley
sacral posterior and anterior root stimulator. Spinal Cord, 2002. 40: 272.
https://www.ncbi.nlm.nih.gov/pubmed/12037708
339. Schumacher, S., et al. Extradural cold block for selective neurostimulation of the bladder:
development of a new technique. J Urol, 1999. 161: 950.
https://www.ncbi.nlm.nih.gov/pubmed/10022732
340. Wollner, J., et al. Surgery Illustrated - surgical atlas sacral neuromodulation. BJU Int, 2012. 110: 146.
https://www.ncbi.nlm.nih.gov/pubmed/22691023
341. Kessler, T.M., et al. Sacral neuromodulation for neurogenic lower urinary tract dysfunction:
systematic review and meta-analysis. Eur Urol, 2010. 58: 865.
https://www.ncbi.nlm.nih.gov/pubmed/20934242
342. Lombardi, G., et al. Sacral neuromodulation for neurogenic non-obstructive urinary retention in
incomplete spinal cord patients: a ten-year follow-up single-centre experience. Spinal Cord, 2014.
52: 241.
https://www.ncbi.nlm.nih.gov/pubmed/24394604
343. Lay, A.H., et al. The role of neuromodulation in patients with neurogenic overactive bladder. Curr
Urol Rep, 2012. 13: 343.
https://www.ncbi.nlm.nih.gov/pubmed/22865208
344. Puccini, F., et al. Sacral neuromodulation: an effective treatment for lower urinary tract symptoms in
multiple sclerosis. Int Urogynecol J Pelvic Floor Dysf, 2016. 27: 347.
https://www.ncbi.nlm.nih.gov/pubmed/26156206
345. Zhang, Y.H., et al. Enveloping the bladder with displacement of flap of the rectus abdominis muscle
for the treatment of neurogenic bladder. J Urol, 1990. 144: 1194.
https://www.ncbi.nlm.nih.gov/pubmed/2146404
346. Stenzl, A., et al. Restoration of voluntary emptying of the bladder by transplantation of innervated
free skeletal muscle. Lancet, 1998. 351: 1483.
https://www.ncbi.nlm.nih.gov/pubmed/9605805
347. Gakis, G., et al. Functional detrusor myoplasty for bladder acontractility: long-term results. J Urol,
2011. 185: 593.
https://www.ncbi.nlm.nih.gov/pubmed/21168866
348. Ninkovic, M., et al. The latissimus dorsi detrusor myoplasty for functional treatment of bladder
acontractility. Clin Plast Surg, 2012. 39: 507.
https://www.ncbi.nlm.nih.gov/pubmed/23036300
349. Duel, B.P., et al. Alternative techniques for augmentation cystoplasty. J Urol, 1998. 159: 998.
https://www.ncbi.nlm.nih.gov/pubmed/9474216
350. Snow, B.W., et al. Bladder autoaugmentation. Urol Clin North Am, 1996. 23: 323.
https://www.ncbi.nlm.nih.gov/pubmed/8659030
351. Stohrer, M., et al. Bladder auto-augmentation--an alternative for enterocystoplasty: preliminary
results. Neurourol Urodyn, 1995. 14: 11.
https://www.ncbi.nlm.nih.gov/pubmed/7742844
352. Stohrer, M., et al. Bladder autoaugmentation in adult patients with neurogenic voiding dysfunction.
Spinal Cord, 1997. 35: 456.
https://www.ncbi.nlm.nih.gov/pubmed/9232751
353. Vainrib, M., et al. Differences in urodynamic study variables in adult patients with neurogenic
bladder and myelomeningocele before and after augmentation enterocystoplasty. Neurourol Urodyn,
2013. 32: 250.
https://www.ncbi.nlm.nih.gov/pubmed/22965686
354. Krebs, J., et al. Functional outcome of supratrigonal cystectomy and augmentation ileocystoplasty
in adult patients with refractory neurogenic lower urinary tract dysfunction. Neurourol Urodyn, 2016.
35.
https://www.ncbi.nlm.nih.gov/pubmed/25524480
355. Hoen, L., et al. Long-term effectiveness and complication rates of bladder augmentation in patients
with neurogenic bladder dysfunction: A systematic review. Neurourol Urodyn, 2017. 07: 07.
https://www.ncbi.nlm.nih.gov/pubmed/28169459
356. Myers, J.B., et al. The effects of augmentation cystoplasty and botulinum toxin injection on patient-

46 NEURO-UROLOGY - LIMITED UPDATE MARCH 2020


reported bladder function and quality of life among individuals with spinal cord injury performing
clean intermittent catheterization. Neurourol Urodyn, 2019. 38: 285.
https://www.ncbi.nlm.nih.gov/pubmed/30375055
357. Mitsui, T., et al. Preoperative renal scar as a risk factor of postoperative metabolic acidosis following
ileocystoplasty in patients with neurogenic bladder. Spinal Cord, 2014. 52: 292.
https://www.ncbi.nlm.nih.gov/pubmed/24469144
358. Perrouin-Verbe, M.A., et al. Long-term functional outcomes of augmentation cystoplasty in adult
spina bifida patients: A single-center experience in a multidisciplinary team. Neurourol Urodyn,
2019. 38: 330.
https://www.ncbi.nlm.nih.gov/pubmed/30350892
359. Moreno, J.G., et al. Improved quality of life and sexuality with continent urinary diversion in
quadriplegic women with umbilical stoma. Arch Phys Med Rehabil, 1995. 76: 758.
https://www.ncbi.nlm.nih.gov/pubmed/7632132
360. Peterson, A.C., et al. Urinary diversion in patients with spinal cord injury in the United States.
Urology, 2012. 80: 1247.
https://www.ncbi.nlm.nih.gov/pubmed/23206770
361. Sylora, J.A., et al. Intermittent self-catheterization by quadriplegic patients via a catheterizable
Mitrofanoff channel. J Urol, 1997. 157: 48.
https://www.ncbi.nlm.nih.gov/pubmed/8976213
362. Van Savage, J.G., et al. Transverse retubularized sigmoidovesicostomy continent urinary diversion
to the umbilicus. J Urol, 2001. 166: 644.
https://www.ncbi.nlm.nih.gov/pubmed/11458110
363. Vanni, A.J., et al. Ileovesicostomy for the neurogenic bladder patient: outcome and cost comparison
of open and robotic assisted techniques. Urology, 2011. 77: 1375.
https://www.ncbi.nlm.nih.gov/pubmed/21146864
364. Wiener, J.S., et al. Bladder augmentation versus urinary diversion in patients with spina bifida in the
United States. J Urol, 2011. 186: 161.
https://www.ncbi.nlm.nih.gov/pubmed/21575969
365. Phe, V., et al. Continent catheterizable tubes/stomas in adult neuro-urological patients: A systematic
review. Neurourol Urodyn, 2017.
https://www.ncbi.nlm.nih.gov/pubmed/28139848
366. Atan, A., et al. Advantages and risks of ileovesicostomy for the management of neuropathic bladder.
Urology, 1999. 54: 636.
https://www.ncbi.nlm.nih.gov/pubmed/10510920
367. Cass, A.S., et al. A 22-year followup of ileal conduits in children with a neurogenic bladder. J Urol,
1984. 132: 529.
https://www.ncbi.nlm.nih.gov/pubmed/6471190
368. Hald, T., et al. Vesicostomy--an alternative urine diversion operation. Long term results. Scand J
Urol Nephrol, 1978. 12: 227.
https://www.ncbi.nlm.nih.gov/pubmed/725543
369. Schwartz, S.L., et al. Incontinent ileo-vesicostomy urinary diversion in the treatment of lower urinary
tract dysfunction. J Urol, 1994. 152: 99.
https://www.ncbi.nlm.nih.gov/pubmed/8201699
370. Sakhri, R., et al. [Laparoscopic cystectomy and ileal conduit urinary diversion for neurogenic
bladders and related conditions. Morbidity and better quality of life]. Prog Urol, 2015. 25: 342.
https://www.ncbi.nlm.nih.gov/pubmed/25726693
371. Herschorn, S., et al. Urinary undiversion in adults with myelodysplasia: long-term followup. J Urol,
1994. 152: 329.
https://www.ncbi.nlm.nih.gov/pubmed/8015064
372. Mukai, S., et al. Retrospective study for risk factors for febrile UTI in spinal cord injury patients with
routine concomitant intermittent catheterization in outpatient settings. Spinal Cord, 2016. 54: 69.
https://www.ncbi.nlm.nih.gov/pubmed/26458969
373. Vasudeva, P., et al. Factors implicated in pathogenesis of urinary tract infections in neurogenic
bladders: some revered, few forgotten, others ignored. Neurourol Urodyn, 2014. 33: 95.
https://www.ncbi.nlm.nih.gov/pubmed/23460489
374. Lenherr, S.M., et al. Glycemic Control and Urinary Tract Infections in Women with Type 1 Diabetes:
Results from the DCCT/EDIC. J Urol, 2016. 196: 1129.
https://www.ncbi.nlm.nih.gov/pubmed/27131462
375. Bakke, A., et al. Bacteriuria in patients treated with clean intermittent catheterization. Scand J Infect
Dis, 1991. 23: 577.

NEURO-UROLOGY - LIMITED UPDATE MARCH 2020 47


https://www.ncbi.nlm.nih.gov/pubmed/1767253
376. Waites, K.B., et al. Epidemiology and risk factors for urinary tract infection following spinal cord
injury. Arch Phys Med Rehabil, 1993. 74: 691.
https://www.ncbi.nlm.nih.gov/pubmed/8328888
377. Nicolle, L.E., et al. Infectious Diseases Society of America guidelines for the diagnosis and treatment
of asymptomatic bacteriuria in adults. Clin Infect Dis, 2005. 40: 643.
https://www.ncbi.nlm.nih.gov/pubmed/15714408
378. Pannek, J. Treatment of urinary tract infection in persons with spinal cord injury: guidelines,
evidence, and clinical practice. A questionnaire-based survey and review of the literature. J Spinal
Cord Med, 2011. 34: 11.
https://www.ncbi.nlm.nih.gov/pubmed/21528621
379. Alavinia, S.M., et al. Enhancing quality practice for prevention and diagnosis of urinary tract infection
during inpatient spinal cord rehabilitation. J Spinal Cord Med, 2017. 40: 803.
https://www.ncbi.nlm.nih.gov/pubmed/28872426
380. Deville, W.L., et al. The urine dipstick test useful to rule out infections. A meta-analysis of the
accuracy. BMC Urol, 2004. 4: 4.
https://www.ncbi.nlm.nih.gov/pubmed/15175113
381. Hoffman, J.M., et al. Nitrite and leukocyte dipstick testing for urinary tract infection in individuals
with spinal cord injury. J Spinal Cord Med, 2004. 27: 128.
https://www.ncbi.nlm.nih.gov/pubmed/15162883
382. Biering-Sorensen, F., et al. Urinary tract infections in patients with spinal cord lesions: treatment and
prevention. Drugs, 2001. 61: 1275.
https://www.ncbi.nlm.nih.gov/pubmed/11511022
383. Everaert, K., et al. Urinary tract infections in spinal cord injury: prevention and treatment guidelines.
Acta Clin Belg, 2009. 64: 335.
https://www.ncbi.nlm.nih.gov/pubmed/19810421
384. Clark, R., et al. The ability of prior urinary cultures results to predict future culture results in
neurogenic bladder patients. Neurourol Urodyn, 2018. 37: 2645.
https://www.ncbi.nlm.nih.gov/pubmed/29799144
385. Pannek, J., et al. Treatment of Complicated Urinary Tract Infections in Individuals with Chronic
Neurogenic Lower Urinary Tract Dysfunction: Are Antibiotics Mandatory? Urologia Int, 2018.
https://www.ncbi.nlm.nih.gov/pubmed/29649808
386. Del Popolo, G., et al. Recurrent bacterial symptomatic cystitis: A pilot study on a new natural option
for treatment. Arch Ital Urol Androl, 2018. 9: 101.
https://www.ncbi.nlm.nih.gov/pubmed/29974728
387. Jia, C., et al. Detrusor botulinum toxin A injection significantly decreased urinary tract infection in
patients with traumatic spinal cord injury. Spinal Cord, 2013. 51: 487.
https://www.ncbi.nlm.nih.gov/pubmed/23357928
388. Waites, K.B., et al. Evaluation of 3 methods of bladder irrigation to treat bacteriuria in persons with
neurogenic bladder. J Spinal Cord Med, 2006. 29: 217.
https://www.ncbi.nlm.nih.gov/pubmed/16859225
389. Gallien, P., et al. Cranberry versus placebo in the prevention of urinary infections in multiple
sclerosis: a multicenter, randomized, placebo-controlled, double-blind trial. Mult Scler, 2014. 20:
1252.
https://www.ncbi.nlm.nih.gov/pubmed/24402038
390. Toh, S.L., et al. Probiotics [LGG-BB12 or RC14-GR1] versus placebo as prophylaxis for urinary tract
infection in persons with spinal cord injury [ProSCIUTTU]: a randomised controlled trial. Spinal Cord,
2019. 57: 550.
https://www.ncbi.nlm.nih.gov/pubmed/30814670
391. Lee, B.S., et al. Methenamine hippurate for preventing urinary tract infections. Cochrane Database
Syst Rev, 2012. 10: CD003265.
https://www.ncbi.nlm.nih.gov/pubmed/23076896
392. Günther, M., et al. Harnwegsinfektprophylaxe. Urinansäuerung mittels L-Methionin bei neurogener
Blasenfunktionsstörung. Urologe B, 2002. 42: 218.
https://link.springer.com/article/10.1007/s00131-002-0207-x
393. Hachen, H.J. Oral immunotherapy in paraplegic patients with chronic urinary tract infections: a
double-blind, placebo-controlled trial. J Urol, 1990. 143: 759.
https://www.ncbi.nlm.nih.gov/pubmed/2179584
394. Krebs, J., et al. Effects of oral immunomodulation therapy on urinary tract infections in individuals
with chronic spinal cord injury-A retrospective cohort study. Neurourol Urodyn, 2018.

48 NEURO-UROLOGY - LIMITED UPDATE MARCH 2020


https://www.ncbi.nlm.nih.gov/pubmed/30350886
395. Poirier, C., et al. Prevention of urinary tract infections by antibiotic cycling in spinal cord injury
patients and low emergence of multidrug resistant bacteria. Medecine et Maladies Infectieuses,
2016. 16: 16.
https://www.ncbi.nlm.nih.gov/pubmed/27321478
396. Darouiche, R.O., et al. Multicenter randomized controlled trial of bacterial interference for prevention
of urinary tract infection in patients with neurogenic bladder. Urology, 2011. 78: 341.
https://www.ncbi.nlm.nih.gov/pubmed/21683991
397. Pannek, J., et al. Usefulness of classical homeopathy for the prophylaxis of recurrent urinary tract
infections in individuals with chronic neurogenic lower urinary tract dysfunction. J Spinal Cord Med,
2018: 1.
https://www.ncbi.nlm.nih.gov/pubmed/29485355
398. Cox, L., et al. Gentamicin bladder instillations decrease symptomatic urinary tract infections in
neurogenic bladder patients on intermittent catheterization. Can Urol Assoc J, 2017. 11: E350.
https://www.ncbi.nlm.nih.gov/pubmed/29382457
399. Pannek, J., et al. Usefulness of classical homoeopathy for the prevention of urinary tract infections
in patients with neurogenic bladder dysfunction: A case series. Indian J Res Homoeopathy, 2014. 8:
31.
http://www.ijrh.org/article.asp?issn=0974-7168;year=2014;volume=8;issue=1;spage=31;epage=36;
aulast=Pannek
400. Rees, P.M., et al. Sexual function in men and women with neurological disorders. Lancet, 2007. 369:
512.
https://www.ncbi.nlm.nih.gov/pubmed/17292771
401. Lombardi, G., et al. Management of sexual dysfunction due to central nervous system disorders: a
systematic review. BJU Int, 2015. 115 Suppl 6: 47.
https://www.ncbi.nlm.nih.gov/pubmed/26193811
402. Jungwirth, A., et al., EAU Guidelines on Male Infertility, in Presented at the 30th Annual Congress in
Madrid. 2015.
https://uroweb.org/guideline/male-infertility/?type=archive
403. Hatzimouratidis, K., et al., EAU guidelines on Male Sexual Dysfunction and Premature Ejaculation.,
in Presented at the 30th Annual Congress in Madrid. 2014.
https://uroweb.org/guideline/male-sexual-dysfunction/?type=archive
404. Foley, F.W., Sexuality, In: Multiple Sclerosis: A Guide for Families. Kalb. RC., Editor. 2006, Demos
Medical Publishing: New York, USA.
405. Annon, J.S., PLISSIT Therapy, In: Handbook of Innovative Psychotherapies. R. Corsini, Editor. 1981,
Wiley & Sons: New York.
406. Fragala, E., et al. Relationship between urodynamic findings and sexual function in multiple sclerosis
patients with lower urinary tract dysfunction. Eur J Neurol, 2015. 22: 485.
https://www.ncbi.nlm.nih.gov/pubmed/25410608
407. Game, X., et al. Sexual function of young women with myelomeningocele. J Pediatr Urol, 2014. 10:
418.
https://www.ncbi.nlm.nih.gov/pubmed/23992838
408. ‘t Hoen, A., et al. A Quality Assessment of Patient-Reported Outcome Measures for Sexual
Function in Neurologic Patients Using the Consensus-based Standards for the Selection of Health
Measurement Instruments Checklist: A Systematic Review. Eur Urol Focus, 2016.
https://www.ncbi.nlm.nih.gov/pubmed/28753768
409. Chen, L., et al. Phosphodiesterase 5 Inhibitors for the Treatment of Erectile Dysfunction: A Trade-off
Network Meta-analysis. Eur Urol, 2015. 68: 674.
https://www.ncbi.nlm.nih.gov/pubmed/25817916
410. Lombardi, G., et al. Ten years of phosphodiesterase type 5 inhibitors in spinal cord injured patients.
J Sex Med, 2009. 6: 1248.
https://www.ncbi.nlm.nih.gov/pubmed/19210710
411. Lombardi, G., et al. Treating erectile dysfunction and central neurological diseases with oral
phosphodiesterase type 5 inhibitors. Review of the literature. J Sex Med, 2012. 9: 970.
https://www.ncbi.nlm.nih.gov/pubmed/22304626
412. Cardenas, D.D., et al. Two phase 3, multicenter, randomized, placebo-controlled clinical trials of
fampridine-SR for treatment of spasticity in chronic spinal cord injury. Spinal Cord, 2014. 52: 70.
https://www.ncbi.nlm.nih.gov/pubmed/24216616
413. Strebel, R.T., et al. Apomorphine sublingual as primary or secondary treatment for erectile
dysfunction in patients with spinal cord injury. BJU Int, 2004. 93: 100.

NEURO-UROLOGY - LIMITED UPDATE MARCH 2020 49


https://www.ncbi.nlm.nih.gov/pubmed/14678378
414. Pohanka, M., et al. The long-lasting improvement of sexual dysfunction in patients with advanced,
fluctuating Parkinson’s disease induced by pergolide: evidence from the results of an open,
prospective, one-year trial. Parkinsonism Relat Disord, 2005. 11: 509.
https://www.ncbi.nlm.nih.gov/pubmed/15994112
415. Chancellor, M.B., et al. Prospective comparison of topical minoxidil to vacuum constriction device
and intracorporeal papaverine injection in treatment of erectile dysfunction due to spinal cord injury.
Urology, 1994. 43: 365.
https://www.ncbi.nlm.nih.gov/pubmed/8134992
416. Cookson, M.S., et al. Long-term results with vacuum constriction device. J Urol, 1993. 149: 290.
https://www.ncbi.nlm.nih.gov/pubmed/8426404
417. Denil, J., et al. Vacuum erection device in spinal cord injured men: patient and partner satisfaction.
Arch Phys Med Rehabil, 1996. 77: 750.
https://www.ncbi.nlm.nih.gov/pubmed/8702367
418. Levine, L.A. External devices for treatment of erectile dysfunction. Endocrine, 2004. 23: 157.
https://www.ncbi.nlm.nih.gov/pubmed/15146095
419. Levine, L.A., et al. Vacuum constriction and external erection devices in erectile dysfunction. Urol
Clin North Am, 2001. 28: 335.
https://www.ncbi.nlm.nih.gov/pubmed/11402585
420. Bella, A.J., et al. Intracavernous pharmacotherapy for erectile dysfunction. Endocrine, 2004. 23: 149.
https://www.ncbi.nlm.nih.gov/pubmed/15146094
421. Bodner, D.R., et al. The application of intracavernous injection of vasoactive medications for
erection in men with spinal cord injury. J Urol, 1987. 138: 310.
https://www.ncbi.nlm.nih.gov/pubmed/3599245
422. Deforge, D., et al. Male erectile dysfunction following spinal cord injury: a systematic review. Spinal
Cord, 2006. 44: 465.
https://www.ncbi.nlm.nih.gov/pubmed/16317419
423. Dinsmore, W.W., et al. Treating men with predominantly nonpsychogenic erectile dysfunction with
intracavernosal vasoactive intestinal polypeptide and phentolamine mesylate in a novel auto-injector
system: a multicentre double-blind placebo-controlled study. BJU Int, 1999. 83: 274.
https://www.ncbi.nlm.nih.gov/pubmed/10233493
424. Hirsch, I.H., et al. Use of intracavernous injection of prostaglandin E1 for neuropathic erectile
dysfunction. Paraplegia, 1994. 32: 661.
https://www.ncbi.nlm.nih.gov/pubmed/7831071
425. Kapoor, V.K., et al. Intracavernous papaverine for impotence in spinal cord injured patients.
Paraplegia, 1993. 31: 675.
https://www.ncbi.nlm.nih.gov/pubmed/8259331
426. Vidal, J., et al. Intracavernous pharmacotherapy for management of erectile dysfunction in multiple
sclerosis patients. Rev Neurol, 1995. 23: 269.
https://www.ncbi.nlm.nih.gov/pubmed/7497173
427. Bodner, D.R., et al. Intraurethral alprostadil for treatment of erectile dysfunction in patients with
spinal cord injury. Urology, 1999. 53: 199.
https://www.ncbi.nlm.nih.gov/pubmed/9886612
428. Gross, A.J., et al. Penile prostheses in paraplegic men. Br J Urol, 1996. 78: 262.
https://www.ncbi.nlm.nih.gov/pubmed/8813925
429. Kimoto, Y., et al. Penile prostheses for the management of the neuropathic bladder and sexual
dysfunction in spinal cord injury patients: long term follow up. Paraplegia, 1994. 32: 336.
https://www.ncbi.nlm.nih.gov/pubmed/8058351
430. Zermann, D.H., et al. Penile prosthetic surgery in neurologically impaired patients: long-term
followup. J Urol, 2006. 175: 1041.
https://www.ncbi.nlm.nih.gov/pubmed/16469612
431. Fode, M., et al. Male sexual dysfunction and infertility associated with neurological disorders. Asian
J Androl, 2012. 14: 61.
https://www.ncbi.nlm.nih.gov/pubmed/22138899
432. Lim, T.C., et al. A simple technique to prevent retrograde ejaculation during assisted ejaculation.
Paraplegia, 1994. 32: 142.
https://www.ncbi.nlm.nih.gov/pubmed/8008416
433. Philippon, M., et al. Successful pregnancies and healthy live births using frozen-thawed sperm
retrieved by a new modified Hotchkiss procedure in males with retrograde ejaculation: first case
series. Basic Clin Androl, 2015. 25: 5.

50 NEURO-UROLOGY - LIMITED UPDATE MARCH 2020


https://www.ncbi.nlm.nih.gov/pubmed/26034605
434. Arafa, M.M., et al. Prostatic massage: a simple method of semen retrieval in men with spinal cord
injury. Int J Androl, 2007. 30: 170.
https://www.ncbi.nlm.nih.gov/pubmed/17298549
435. Kolettis, P.N., et al. Fertility outcomes after electroejaculation in men with spinal cord injury. Fertil
Steril, 2002. 78: 429.
https://www.ncbi.nlm.nih.gov/pubmed/12137889
436. Chehensse, C., et al. The spinal control of ejaculation revisited: a systematic review and meta-
analysis of anejaculation in spinal cord injured patients. Hum Reprod Update, 2013. 19: 507.
https://www.ncbi.nlm.nih.gov/pubmed/23820516
437. Beretta, G., et al. Reproductive aspects in spinal cord injured males. Paraplegia, 1989. 27: 113.
https://www.ncbi.nlm.nih.gov/pubmed/2717193
438. Brackett, N.L., et al. Application of 2 vibrators salvages ejaculatory failures to 1 vibrator during
penile vibratory stimulation in men with spinal cord injuries. J Urol, 2007. 177: 660.
https://www.ncbi.nlm.nih.gov/pubmed/17222653
439. Sonksen, J., et al. Ejaculation induced by penile vibratory stimulation in men with spinal cord
injuries. The importance of the vibratory amplitude. Paraplegia, 1994. 32: 651.
https://www.ncbi.nlm.nih.gov/pubmed/7831070
440. Claydon, V.E., et al. Cardiovascular responses to vibrostimulation for sperm retrieval in men with
spinal cord injury. J Spinal Cord Med, 2006. 29: 207.
https://www.ncbi.nlm.nih.gov/pubmed/16859224
441. Ekland, M.B., et al. Incidence of autonomic dysreflexia and silent autonomic dysreflexia in men with
spinal cord injury undergoing sperm retrieval: implications for clinical practice. J Spinal Cord Med,
2008. 31: 33.
https://www.ncbi.nlm.nih.gov/pubmed/18533409
442. Soler, J.M., et al. Midodrine improves ejaculation in spinal cord injured men. J Urol, 2007. 178: 2082.
https://www.ncbi.nlm.nih.gov/pubmed/17869290
443. Pecori, C., et al. Paternal therapy with disease modifying drugs in multiple sclerosis and pregnancy
outcomes: a prospective observational multicentric study. BMC Neurol, 2014. 14: 114.
https://www.ncbi.nlm.nih.gov/pubmed/24884599
444. Brackett, N.L., et al. Treatment of infertility in men with spinal cord injury. Nat Rev Urol, 2010. 7: 162.
https://www.ncbi.nlm.nih.gov/pubmed/20157304
445. Raviv, G., et al. Testicular sperm retrieval and intra cytoplasmic sperm injection provide favorable
outcome in spinal cord injury patients, failing conservative reproductive treatment. Spinal Cord,
2013. 51: 642.
https://www.ncbi.nlm.nih.gov/pubmed/23689394
446. Schatte, E.C., et al. Treatment of infertility due to anejaculation in the male with electroejaculation
and intracytoplasmic sperm injection. J Urol, 2000. 163: 1717.
https://www.ncbi.nlm.nih.gov/pubmed/10799167
447. Shieh, J.Y., et al. A protocol of electroejaculation and systematic assisted reproductive technology
achieved high efficiency and efficacy for pregnancy for anejaculatory men with spinal cord injury.
Arch Phys Med Rehabil, 2003. 84: 535.
https://www.ncbi.nlm.nih.gov/pubmed/12690592
448. Taylor, Z., et al. Contribution of the assisted reproductive technologies to fertility in males suffering
spinal cord injury. Aust N Z J Obstet Gynaecol, 1999. 39: 84.
https://www.ncbi.nlm.nih.gov/pubmed/10099757
449. Rutkowski, S.B., et al. The influence of bladder management on fertility in spinal cord injured males.
Paraplegia, 1995. 33: 263.
https://www.ncbi.nlm.nih.gov/pubmed/7630651
450. Hamed, S.A., et al. Seminal fluid analysis and testicular volume in adults with epilepsy receiving
valproate. J Clin Neurosci, 2015. 22: 508.
https://www.ncbi.nlm.nih.gov/pubmed/25636832
451. Ohl, D.A., et al. Electroejaculation versus vibratory stimulation in spinal cord injured men: sperm
quality and patient preference. J Urol, 1997. 157: 2147.
https://www.ncbi.nlm.nih.gov/pubmed/9146603
452. Brackett, N.L., et al. Semen quality of spinal cord injured men is better when obtained by vibratory
stimulation versus electroejaculation. J Urol, 1997. 157: 151.
https://www.ncbi.nlm.nih.gov/pubmed/8976239
453. Brackett, N.L., et al. Semen retrieval in men with spinal cord injury is improved by interrupting
current delivery during electroejaculation. J Urol, 2002. 167: 201.

NEURO-UROLOGY - LIMITED UPDATE MARCH 2020 51


https://www.ncbi.nlm.nih.gov/pubmed/11743305
454. DeForge, D., et al. Fertility following spinal cord injury: a systematic review. Spinal Cord, 2005. 43:
693.
https://www.ncbi.nlm.nih.gov/pubmed/15951744
455. Ferreiro-Velasco, M.E., et al. Sexual issues in a sample of women with spinal cord injury. Spinal
Cord, 2005. 43: 51.
https://www.ncbi.nlm.nih.gov/pubmed/15303115
456. Kreuter, M., et al. Sexuality and sexual life in women with spinal cord injury: a controlled study. J
Rehabil Med, 2008. 40: 61.
https://www.ncbi.nlm.nih.gov/pubmed/18176739
457. Kreuter, M., et al. Sexual adjustment and quality of relationship in spinal paraplegia: a controlled
study. Arch Phys Med Rehabil, 1996. 77: 541.
https://www.ncbi.nlm.nih.gov/pubmed/8831469
458. Szymanski, K.M., et al. Sexual identity and orientation in adult men and women with spina bifida. J
Pediatr Rehabil Med, 2017. 10: 313.
https://www.ncbi.nlm.nih.gov/pubmed/29125522
459. Kessler, T.M., et al. Sexual dysfunction in multiple sclerosis. Expert Rev Neurother, 2009. 9: 341.
https://www.ncbi.nlm.nih.gov/pubmed/19271943
460. Lew-Starowicz, M., et al. Prevalence of Sexual Dysfunctions Among Women with Multiple Sclerosis.
Sex Disabil, 2013. 31: 141.
https://www.ncbi.nlm.nih.gov/pubmed/23704801
461. Reitz, A., et al. Impact of spinal cord injury on sexual health and quality of life. Int J Impot Res, 2004.
16: 167.
https://www.ncbi.nlm.nih.gov/pubmed/14973522
462. Harrison, J., et al. Factors associated with sexual functioning in women following spinal cord injury.
Paraplegia, 1995. 33: 687.
https://www.ncbi.nlm.nih.gov/pubmed/8927405
463. Westgren, N., et al. Sexuality in women with traumatic spinal cord injury. Acta Obstet Gynecol
Scand, 1997. 76: 977.
https://www.ncbi.nlm.nih.gov/pubmed/9435740
464. Fruhauf, S., et al. Efficacy of psychological interventions for sexual dysfunction: a systematic review
and meta-analysis. Arch Sex Behav, 2013. 42: 915.
https://www.ncbi.nlm.nih.gov/pubmed/23559141
465. Alexander, M., et al. Spinal cord injuries and orgasm: a review. J Sex Marital Ther, 2008. 34: 308.
https://www.ncbi.nlm.nih.gov/pubmed/18576233
466. Sipski, M.L., et al. Physiologic parameters associated with sexual arousal in women with incomplete
spinal cord injuries. Arch Phys Med Rehabil, 1997. 78: 305.
https://www.ncbi.nlm.nih.gov/pubmed/9084355
467. Sipski, M.L., et al. Sexual arousal and orgasm in women: effects of spinal cord injury. Ann Neurol,
2001. 49: 35.
https://www.ncbi.nlm.nih.gov/pubmed/11198294
468. McAlonan, S. Improving sexual rehabilitation services: the patient’s perspective. Am J Occup Ther,
1996. 50: 826.
https://www.ncbi.nlm.nih.gov/pubmed/8947375
469. Schopp, L.H., et al. Impact of comprehensive gynecologic services on health maintenance
behaviours among women with spinal cord injury. Disabil Rehabil, 2002. 24: 899.
https://www.ncbi.nlm.nih.gov/pubmed/12519485
470. Sukumaran, S.C., et al. Polytherapy increases the risk of infertility in women with epilepsy.
Neurology, 2010. 75: 1351.
https://www.ncbi.nlm.nih.gov/pubmed/20938026
471. Axel, S.J. Spinal cord injured women’s concerns: menstruation and pregnancy. Rehabil Nurs, 1982.
7: 10.
https://www.ncbi.nlm.nih.gov/pubmed/6921826
472. Jackson, A.B., et al. A multicenter study of women’s self-reported reproductive health after spinal
cord injury. Arch Phys Med Rehabil, 1999. 80: 1420.
https://www.ncbi.nlm.nih.gov/pubmed/10569436
473. Baker, E.R., et al. Pregnancy in spinal cord injured women. Arch Phys Med Rehabil, 1996. 77: 501.
https://www.ncbi.nlm.nih.gov/pubmed/8629929
474. Baker, E.R., et al. Risks associated with pregnancy in spinal cord-injured women. Obstet Gynecol,
1992. 80: 425.

52 NEURO-UROLOGY - LIMITED UPDATE MARCH 2020


https://www.ncbi.nlm.nih.gov/pubmed/1495699
475. Bertschy, S., et al. Delivering care under uncertainty: Swiss providers’ experiences in caring for
women with spinal cord injury during pregnancy and childbirth - an expert interview study. BMC
Pregnancy Childbirth, 2016. 16: 181.
https://www.ncbi.nlm.nih.gov/pubmed/27443838
476. Le Liepvre, H., et al. Pregnancy in spinal cord-injured women, a cohort study of 37 pregnancies in
25 women. Spinal Cord, 2017. 55: 167.
https://www.ncbi.nlm.nih.gov/pubmed/27670808
477. Skowronski, E., et al. Obstetric management following traumatic tetraplegia: case series and
literature review. Aust N Z J Obstet Gynaecol, 2008. 48: 485.
https://www.ncbi.nlm.nih.gov/pubmed/19032665
478. Cross, L.L., et al. Pregnancy, labor and delivery post spinal cord injury. Paraplegia, 1992. 30: 890.
https://www.ncbi.nlm.nih.gov/pubmed/1287543
479. Hughes, S.J., et al. Management of the pregnant woman with spinal cord injuries. Br J Obstet
Gynaecol, 1991. 98: 513.
https://www.ncbi.nlm.nih.gov/pubmed/1873238
480. Dannels, A., et al. The perimenopause experience for women with spinal cord injuries. SCI Nurs,
2004. 21: 9.
https://www.ncbi.nlm.nih.gov/pubmed/15176344
481. Vukusic, S., et al. Multiple sclerosis and pregnancy in the ‘treatment era’. Nat Rev Neurol, 2015. 11:
280.
https://www.ncbi.nlm.nih.gov/pubmed/25896084
482. Amato, M.P., et al. Management of pregnancy-related issues in multiple sclerosis patients: the need
for an interdisciplinary approach. Neurol Sci, 2017. 38: 1849.
https://www.ncbi.nlm.nih.gov/pubmed/28770366
483. Delaney, K.E., et al. Multiple sclerosis and sexual dysfunction: A need for further education and
interdisciplinary care. NeuroRehabilitation, 2017. 41: 317.
https://www.ncbi.nlm.nih.gov/pubmed/29036844
484. Bove, R., et al. Management of multiple sclerosis during pregnancy and the reproductive years: a
systematic review. Obstet Gynecol, 2014. 124: 1157.
https://www.ncbi.nlm.nih.gov/pubmed/25415167
485. Przydacz, M., et al. Recommendations for urological follow-up of patients with neurogenic bladder
secondary to spinal cord injury. Int Urol Nephrol, 2018. 50: 1005.
https://www.ncbi.nlm.nih.gov/pubmed/29569211
486. Abrams, P., et al. A proposed guideline for the urological management of patients with spinal cord
injury. BJU Int, 2008. 101: 989.
https://www.ncbi.nlm.nih.gov/pubmed/18279449
487. Pannek, J., et al. Clinical usefulness of ultrasound assessment of detrusor wall thickness in patients
with neurogenic lower urinary tract dysfunction due to spinal cord injury: Urodynamics made easy?
World J Urol, 2013. 31: 659.
https://www.ncbi.nlm.nih.gov/pubmed/23073657
488. Silva, J.A., et al. Association between the bladder wall thickness and urodynamic findings in
patients with spinal cord injury. World J Urol, 2015. 33: 131.
https://www.ncbi.nlm.nih.gov/pubmed/24573904
489. Veenboer, P.W., et al. Diagnostic accuracy of Tc-99m DMSA scintigraphy and renal ultrasonography
for detecting renal scarring and relative function in patients with spinal dysraphism. Neurourol
Urodyn, 2015. 34: 513.
https://www.ncbi.nlm.nih.gov/pubmed/24706504
490. Ismail, S., et al. Prevalence, management, and prognosis of bladder cancer in patients with
neurogenic bladder: A systematic review. Neurourol Urodyn, 2018. 37: 1386.
https://www.ncbi.nlm.nih.gov/pubmed/29168217
491. Lewis, J., et al. A framework for transitioning patients from pediatric to adult health settings for
patients with neurogenic bladder. Neurourol Urodyn, 2017. 36: 973.
https://www.ncbi.nlm.nih.gov/pubmed/27276694

NEURO-UROLOGY - LIMITED UPDATE MARCH 2020 53


5. CONFLICT OF INTEREST
All members of the Neuro-urology working group have provided disclosure statements of all relationships that
they have that might be perceived as a potential source of a conflict of interest. This information is publically
accessible through the European Association of Urology website: http://uroweb.org/guideline. This guidelines
document was developed with the financial support of the European Association of Urology. No external
sources of funding and support have been involved. The EAU is a non-profit organisation and funding is limited
to administrative and travel and meeting expenses. No honoraria or other reimbursements have been provided.

6. CITATION INFORMATION
The format in which to cite the EAU Guidelines will vary depending on the style guide of the journal in which the
citation appears. Accordingly, the number of authors or whether, for instance, to include the publisher, location,
or an ISBN number may vary.

The compilation of the complete Guidelines should be referenced as:


EAU Guidelines. Edn. presented at the EAU Annual Congress Amsterdam 2020. ISBN 978-94-92671-07-3.

If a publisher and/or location is required, include:


EAU Guidelines Office, Arnhem, The Netherlands. http://uroweb.org/guidelines/compilations-of-all-guidelines/

References to individual guidelines should be structured in the following way:


Contributors’ names. Title of resource. Publication type. ISBN. Publisher and publisher location, year.

54 NEURO-UROLOGY - LIMITED UPDATE MARCH 2020


EAU Guidelines on
Management of
Non-Neurogenic
Female Lower
Urinary Tract
Symptoms (LUTS)
C.K. Harding (Chair), M.C. Lapitan (Vice-chair), S. Arlandis,
K. Bø, E. Costantini, J. Groen, A.K. Nambiar, M.I. Omar,
V. Phé, C.H. van der Vaart
Guidelines Associates: F. Farag, M. Karavitakis, M. Manso,
S. Monagas, A. Nic an Riogh, E. O’Connor, B. Peyronnet,
V. Sakalis, N. Sihra, L. Tzelves

© European Association of Urology 2021


TABLE OF CONTENTS PAGE
1. INTRODUCTION 10
1.1 Aim and objectives 10
1.2 Panel composition 10
1.3 Available publications 10
1.4 Publication history 11

2. METHODS 11
2.1 Introduction 11
2.2 Review 11
2.3 Future goals 11

3. DIAGNOSTIS 12
3.1 History and physical examination 12
3.1.1 Summary of evidence and recommendation for history taking and physical
examination 12
3.2 Patient questionnaires 12
3.2.1 Summary of evidence and recommendations for patient questionnaires 13
3.3 Bladder diaries 13
3.3.1 Summary of evidence and recommendations for bladder diaries 13
3.4 Urinalysis and urinary tract infection 14
3.4.1 Summary of evidence and recommendations for urinalysis 14
3.5 Post-void residual volume 14
3.5.1 Summary of evidence and recommendations for post-void residual 14
3.6 Urodynamics 15
3.6.1 Variability 15
3.6.2 Diagnostic accuracy 15
3.6.3 Predictive value 16
3.6.4 Summary of evidence and recommendations for urodynamics 17
3.7 Pad testing 17
3.7.1 Summary of evidence and recommendations for pad testing 17
3.8 Imaging 18
3.8.1 Ultrasound 18
3.8.2 Detrusor wall thickness 18
3.8.3 MRI 18
3.8.4 Summary of evidence and recommendations for imaging 19

4. DISEASE MANAGEMENT 19
4.1 Overactive bladder 19
4.1.1 Epidemiology, aetiology, pathophysiology 19
4.1.2 Classification 19
4.1.3 Diagnostic evaluation 19
4.1.3.1 Bladder diaries 19
4.1.3.2 Urodynamics 19
4.1.3.3 Summary of evidence and recommendations regarding associated
conditions 20
4.1.4 Disease management 20
4.1.4.1 Conservative management 20
4.1.4.1.1 Addressing underlying disease/cognitive impairment 20
4.1.4.1.1.1 Summary of evidence and recommendation
regarding associated conditions 20
4.1.4.1.2 Adjustment of other medication 20
4.1.4.1.2.1 Summary of evidence and recommendations
for adjustment of non-LUTS medication 21
4.1.4.1.3 Urinary containment 21
4.1.4.1.3.1 Summary of evidence and recommendations
for urinary containment 22
4.1.4.1.4 Lifestyle interventions 22

2 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021


4.1.4.1.4.1 Caffeine intake 22
4.1.4.1.4.2 Fluid intake 22
4.1.4.1.4.3 Obesity and weight loss 22
4.1.4.1.4.4 Smoking 22
4.1.4.1.4.5 Summary of evidence and recommendations
for lifestyle interventions 23
4.1.4.1.5 Behavioural and physical therapies 23
4.1.4.1.5.1 Prompted voiding and timed voiding 23
4.1.4.1.5.2 Bladder Training 23
4.1.4.1.5.3 Pelvic floor muscle training 23
4.1.4.1.5.4 Electrical stimulation 24
4.1.4.1.5.5 Acupuncture 24
4.1.4.1.5.6 Posterior tibial nerve stimulation 24
4.1.4.1.5.6.1 Percutaneous posterior tibial nerve
stimulation 24
4.1.4.1.5.6.2 Transcutaneous posterior tibial
nerve stimulation 24
4.1.4.1.5.7 Summary of evidence and recommendations
for behavioural and physical therapies 24
4.1.4.2 Pharmacological management 25
4.1.4.2.1 Anticholinergic drugs 25
4.1.4.2.1.1 Comparison of different anticholinergic agents 26
4.1.4.2.1.2 Anticholinergic drugs vs. conservative treatment 27
4.1.4.2.1.3 Anticholinergic drugs: adherence and persistence 27
4.1.4.2.1.4 Summary of evidence and recommendations
for anticholinergic drugs 28
4.1.4.2.2 Beta-3 agonists 28
4.1.4.2.2.1 Summary of evidence and recommendation
for mirabegron 29
4.1.4.2.3 Anticholinergics and beta-3 agonists: the elderly and cognition 29
4.1.4.2.3.1 Applicability of evidence to the general elderly
population 30
4.1.4.2.3.2 Anticholinergic burden 30
4.1.4.2.3.3 Summary of evidence and additional
recommendations for use of anticholinergic
drugs in the elderly 30
4.1.4.2.4 Oestrogens 31
4.1.4.2.4.1 Summary of evidence and recommendation
for oestrogen therapy 31
4.1.4.3 Surgical management 31
4.1.4.3.1 Bladder wall injection of botulinum toxin A 31
4.1.4.3.1.1 Summary of evidence and recommendations
for bladder wall injection of botulinum toxin A 32
4.1.4.3.2 Sacral nerve stimulation 32
4.1.4.3.2.1 Summary of evidence and recommendation
for sacral nerve stimulation 33
4.1.4.3.3 Cystoplasty/urinary diversion 33
4.1.4.3.3.1 Augmentation cystoplasty 33
4.1.4.3.3.2 Detrusor myectomy (bladder auto-augmentation) 34
4.1.4.3.3.3 Urinary diversion 34
4.1.4.3.3.4 Summary of evidence and recommendations
for cystoplasty/urinary diversion 35
4.1.5 Follow-up 35
4.1.5.1 Recommendations for follow-up of patients with overactive bladder 35
4.2 Stress Urinary Incontinence 36
4.2.1 Epidemiology, aetiology, pathophysiology 36
4.2.2 Classification 36
4.2.3 Diagnostic evaluation 36
4.2.3.1 History and physical examination 36

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 3


4.2.3.1.1 Summary of evidence and recommendation for history and
physical examination for SUI 37
4.2.3.2 Patient questionnaires 37
4.2.3.2.1 Summary of evidence and recommendation for patient
questionnaires 37
4.2.3.3 Post-void residual volume 37
4.2.3.3.1 Summary of evidence and recommendations for post-void
residual volume 37
4.2.3.4 Urodynamics 38
4.2.3.4.1 Summary of evidence and recommendations for
urodynamics 38
4.2.3.5 Pad testing 39
4.2.3.5.1 Summary of evidence and recommendations for pad testing 39
4.2.3.6 Imaging 39
4.2.3.6.1 Summary of evidence and recommendation for imaging 40
4.2.4
Disease management 40
4.2.4.1 Conservative management 40
4.2.4.1.1 Obesity and weight loss 40
4.2.4.1.1.1 Summary of evidence and recommendation for
obesity and weight loss 40
4.2.4.1.2 Urinary containment 40
4.2.4.1.2.1 Summary of evidence and recommendations
for urinary containment 40
4.2.4.1.3 Pelvic floor muscle training 41
4.2.4.1.3.1 Efficacy of pelvic floor muscle training in
stress urinary incontinence 41
4.2.4.1.3.2 Efficacy of electrical stimulation 41
4.2.4.1.3.3 Long-term efficacy of pelvic floor muscle training 42
4.2.4.1.3.4 Efficacy of pelvic floor muscle training in
childbearing women 42
4.2.4.1.3.5 Pelvic floor muscle training in the elderly 42
4.2.4.1.3.6 Summary of evidence and recommendations
for pelvic floor muscle training 43
4.2.4.1.4 Electromagnetic stimulation 43
4.2.4.2 Pharmacological management 43
4.2.4.2.1 Oestrogen 43
4.2.4.2.1.1 Summary of evidence and recommendations for
oestrogens 44
4.2.4.2.2 Duloxetine 44
4.2.4.2.2.1 Summary of evidence and recommendations for
duloxetine 44
4.2.4.3 Surgical management 45
4.2.4.3.1 General considerations 45
4.2.4.3.1.1 Recommendations for surgical treatment of SUI 46
4.2.4.3.2 Surgery for women with uncomplicated stress urinary
incontinence 46
4.2.4.3.2.1 Open- and laparoscopic colposuspension surgery 46
4.2.4.3.2.1.1 Summary of evidence and
recommendation for open- and
laparoscopic colposuspension surgery
for stress urinary incontinence 47
4.2.4.3.2.2 Autologous sling 47
4.2.4.3.2.2.1 Summary of evidence and
recommendation for autologous
sling 47
4.2.4.3.2.3 Urethral bulking agents 48
4.2.4.3.2.3.1 Summary of evidence and
recommendations for urethral
bulking agents 48
4.2.4.3.2.4 Mid-urethral slings 49

4 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021


4.2.4.3.2.4.1 Summary of evidence and
recommendations for mid-urethral
slings 52
4.2.4.3.2.5 Other treatments for uncomplicated SUI 53
4.2.4.3.2.5.1 Summary of evidence and
recommendations for other
treatments for uncomplicated SUI 53
4.2.4.3.3 Surgery for women with complicated stress urinary
incontinence 53
4.2.4.3.3.1 Colposuspension or sling (synthetic or
autologous) following failed primary SUI surgery 54
4.2.4.3.3.1.1 Summary of evidence for surgery
in those with recurrent SUI following
failed primary SUI surgery 54
4.2.4.3.3.2 Adjustable slings 55
4.2.4.3.3.2.1 Summary of evidence for
adjustable slings 55
4.2.4.3.3.3 External compression devices 55
4.2.4.3.3.3.1 Summary of evidence for external
compression devices 56
4.2.4.3.3.4 Recommendations for complicated stress
urinary incontinence 56
4.2.4.3.4 Surgery for stress urinary incontinence in special patient
groups 56
4.2.4.3.4.1 Stress urinary incontinence surgery in obese
women 56
4.2.4.3.4.2 Stress urinary incontinence surgery in elderly
women 56
4.2.4.3.4.3 Summary of evidence and recommendations
for SUI surgery in special patient groups 57
4.2.5 Follow-up 57
4.3 Mixed urinary incontinence 57
4.3.1 Epidemiology, aetiology and pathophysiology 57
4.3.2 Diagnostic evaluation 58
4.3.2.1 Summary of evidence and recommendations for the diagnosis of mixed
urinary incontinence 58
4.3.3 Disease Management 58
4.3.3.1 Conservative management 58
4.3.3.1.1 Pelvic floor muscle training in mixed urinary incontinence 58
4.3.3.1.2 Bladder training 59
4.3.3.1.3 Electrical stimulation 59
4.3.3.2 Summary of evidence and recommendations for conservative
management in MUI 59
4.3.3.3 Pharmacological management 59
4.3.3.3.1 Tolterodine 59
4.3.3.3.2 Duloxetine 60
4.3.3.3.3 Summary of evidence and recommendations for
pharmacological management of MUI 60
4.3.3.4 Surgical management 60
4.3.3.4.1 Summary of evidence and recommendations for surgery in
patients with MUI 61
4.4 Underactive bladder 61
4.4.1 Epidemiology, aetiology, pathophysiology 61
4.4.1.1 Epidemiology 61
4.4.1.2 Aetiology 62
4.4.1.3 Pathophysiology 62
4.4.2 Classification 63
4.4.3 Diagnostic evaluation 63
4.4.3.1 Symptoms associated with detrusor underactivity 63
4.4.3.2 Urodynamic studies 63

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 5


4.4.4 Disease management 65
4.4.4.1 Conservative management 65
4.4.4.1.1 Behavioural interventions 65
4.4.4.1.2 Pelvic floor muscle relaxation training with biofeedback 65
4.4.4.1.3 Clean intermittent self-catheterisation 65
4.4.4.1.4 Indwelling catheter 65
4.4.4.1.5 Intravesical electrical stimulation 65
4.4.4.1.6 Intraurethral insert 65
4.4.4.2 Pharmacology management 65
4.4.4.2.1 Parasympathomimetics 65
4.4.4.2.2 Alpha-blockers 66
4.4.4.2.3 Prostaglandins 66
4.4.4.3 Surgical management 66
4.4.4.3.1 Sacral nerve stimulation 66
4.4.4.3.2 OnabotulinumtoxinA 66
4.4.4.3.3 Transurethral incision of the bladder neck 66
4.4.4.3.4 Reduction cystoplasty 67
4.4.4.3.5 Myoplasty 67
4.4.4.4 Summary of therapeutic evidence on detrusor underactivity 67
4.4.4.4.1 Summary of evidence and recommendations for underactive
bladder 67
4.4.5 Follow-up 67
4.5 Bladder outlet obstruction 68
4.5.1 Introduction 68
4.5.2 Epidemiology, aetiology, pathophysiology 68
4.5.2.1 Epidemiology 68
4.5.2.2 Pathophysiology 68
4.5.2.3 Aetiology 68
4.5.3 Classification 69
4.5.3.1 Anatomic bladder outlet obstruction 69
4.5.3.2 Functional bladder outlet obstruction 69
4.5.3.3 Recommendation for the classification of bladder outlet obstruction 69
4.5.4 Diagnostic evaluation 69
4.5.4.1 Clinical history 69
4.5.4.2 Clinical examination 70
4.5.4.3 Uroflowmetry and post-void residual volume 70
4.5.4.4 Ultrasound 70
4.5.4.5 Magnetic resonance imaging 70
4.5.4.6 Electromyography 70
4.5.4.7 Cystourethroscopy 71
4.5.4.8 Urodynamics and video-urodynamics 71
4.5.4.9 Summary of evidence and recommendations for the diagnosis of bladder
outlet obstruction 72
4.5.5 Disease management 72
4.5.5.1 Conservative management 72
4.5.5.1.1 Behavioural modification 72
4.5.5.1.2 Pelvic floor muscle training +/- biofeedback 72
4.5.5.1.3 Electrical stimulation 73
4.5.5.1.4 Use of vaginal pessary 73
4.5.5.1.5 Urinary containment devices 73
4.5.5.1.6 Urinary catheterisation 73
4.5.5.1.7 Intra-urethral inserts 73
4.5.5.1.8 Extracorporeal magnetic stimulation 74
4.5.5.1.9 Summary of evidence and recommendations for conservative
treatment of bladder outlet obstruction 74
4.5.6 Pharmacologic management 74
4.5.6.1 Alpha-adrenergic blockers 74
4.5.6.2 Striated muscle relaxants 75
4.5.6.3 Oestrogens 75
4.5.6.4 Sildenafil 75

6 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021


4.5.6.5 Thyrotropin-releasing hormone 76
4.5.6.6 Summary of evidence and recommendations for pharmacologic treatment 76
4.5.7 Surgical treatment 76
4.5.7.1 Intra-sphincteric botulinum toxin injection 76
4.5.7.2 Sacral nerve stimulation 77
4.5.7.3 Pelvic organ prolapse surgery 77
4.5.7.4 Urethral dilatation 77
4.5.7.5 Urethrotomy 78
4.5.7.6 Bladder neck incision/resection 78
4.5.7.7 Urethroplasty/urethral reconstruction 79
4.5.7.8 Urethrolysis 79
4.5.7.9 Removal/excision/section/loosening of mid-urethral sling 80
4.5.7.9.1 Timing of sling revision 80
4.5.7.10 Summary of evidence and recommendations for surgical management
of BOO 80
4.5.8 Follow up 81
4.6 Nocturia 81
4.6.1 Epidemiology, aetiology, pathophysiology 81
4.6.2 Classification 82
4.6.3 Diagnostic evaluation 82
4.6.3.1 Summary of evidence and recommendations for the diagnosis of nocturia 82
4.6.4 Disease management 83
4.6.4.1 Conservative management 83
4.6.4.1.1 Summary of evidence and recommendations for the
conservative management of nocturia 84
4.6.4.2 Pharmacology management 84
4.6.4.2.1 Desmopressin 84
4.6.4.2.2 Anticholinergics 84
4.6.4.2.3 Oestrogens 85
4.6.4.2.4 Diuretic treatment 85
4.6.4.3 Surgical management 85
4.6.4.4 Summary of evidence and recommendations for the pharmacological
management of nocturia 85
4.6.5 Follow-up 86
4.7 Pelvic organ prolapse and LUTS 86
4.7.1 Epidemiology, aetiology, pathophysiology 86
4.7.2 Classification 86
4.7.3 Diagnostic evaluation 87
4.7.3.1 Summary of evidence and recommendation for the detection of SUI in
women with POP 88
4.7.3.2 Urodynamics in women with POP and LUTS (without stress urinary
incontinence). 89
4.7.4 Disease management 89
4.7.4.1 Conservative treatment of pelvic organ prolapse 89
4.7.4.1.1 Pelvic floor muscle training versus lifestyle advice 89
4.7.4.1.2 Pelvic floor muscle training versus pelvic floor muscle
training with pessary 90
4.7.4.1.3 Pelvic floor muscle training versus pessary only 90
4.7.4.1.4 Surgery versus surgery with pelvic floor muscle training 90
4.7.4.1.5 Summary of evidence and guidelines for the conservative
treatment of pelvic organ prolapse and lower urinary tract
symptoms 90
4.7.4.2 Pelvic organ prolapse surgery and overactive bladder 91
4.7.4.3 Pelvic organ prolapse surgery and bladder outlet obstruction 91
4.7.4.4 Pelvic organ prolapse surgery and stress urinary incontinence 91
4.7.4.4.1 Vaginal pelvic organ prolapse surgery in women with stress
urinary incontinence 91
4.7.4.4.2 Abdominal pelvic organ prolapse surgery in women with
stress urinary incontinence 91

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 7


4.7.4.4.3 Vaginal POP surgery in women with prolapse and occult stress
urinary incontinence 91
4.7.4.5 Vaginal pelvic organ prolapse surgery in continent women 92
4.7.4.5.1 Abdominal pelvic organ prolapse surgery in continent women 92
4.7.4.6 Adverse events associated with combined pelvic organ prolapse and
stress urinary incontinence surgery 92
4.7.5 Summary of evidence and recommendations for surgery in women with both pelvic
organ prolapse and stress urinary incontinence 92
4.8 Urinary fistula 92
4.8.1 Epidemiology, aetiology and pathophysiology 93
4.8.1.1 Obstetric fistula 93
4.8.1.2 Iatrogenic fistula 93
4.8.1.2.1 Post-gynaecological surgery 93
4.8.1.2.2 Radiation fistula 93
4.8.1.2.3 Rare causes of vesico-vaginal fistula 93
4.8.1.3 Summary of evidence for epidemiology, aetiology and pathophysiology of
urinary fistula 93
4.8.2 Classification 93
4.8.2.1 Recommendation for the classification of urinary fistula 94
4.8.3 Diagnostic evaluation 94
4.8.4 Management of fistula 94
4.8.4.1 Management of vesico-vaginal fistula 94
4.8.4.1.1 Conservative management 94
4.8.4.1.1.1 Spontaneous closure 94
4.8.4.1.1.2 Pharmacotherapies 94
4.8.4.1.1.3 Palliation and skin care 94
4.8.4.1.1.4 Nutrition 95
4.8.4.1.1.5 Physiotherapy 95
4.8.4.1.1.6 Antimicrobial therapy 95
4.8.4.1.1.7 Counselling 95
4.8.4.1.2 Surgical management 95
4.8.4.1.2.1 Timing of surgery 95
4.8.4.1.2.2 Surgical approaches 95
4.8.4.1.3 Management of complications of vesico-vaginal fistulae 95
4.8.4.2 Management of radiation fistulae 96
4.8.4.3 Management of ureteric fistulae 96
4.8.4.3.1 General principles 96
4.8.4.3.2 Uretero-vaginal fistulae 96
4.8.4.3.3 Management of urethrovaginal fistulae 96
4.8.4.3.3.1 Aetiology 96
4.8.4.3.3.2 Diagnosis 97
4.8.4.3.3.3 Surgical management 97
4.8.4.3.3.4 Flaps and neo-urethra 97
4.8.4.3.3.5 Martius flap 97
4.8.4.3.3.6 Rectus muscle flap 97
4.8.4.3.3.7 Alternative approaches 97
4.8.4.4 Summary of evidence and recommendations for the management of
urinary fistula 97
4.9 Urethral diverticulum 98
4.9.1 Epidemiology, aetiology, pathophysiology 98
4.9.2 Classification 99
4.9.3 Diagnosis 99
4.9.3.1 Associated voiding dysfunction 100
4.9.4 Disease management 100
4.9.4.1 Surgical treatment 100
4.9.4.2 Management of concomitant stress urinary incontinence 101
4.9.4.3 Pathological findings 102
4.9.5 Summary of evidence and recommendations for urethral diverticulum 102

8 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021


5. REFERENCES 102

6. CONFLICT OF INTEREST 144

7. CITATION INFORMATION 144

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 9


1. INTRODUCTION
Lower urinary tract symptoms (LUTS) is the overarching term encompassing storage, voiding and post-
micturition symptoms [1]. Storage symptoms commonly described include frequency, urgency, nocturia and
urinary incontinence (UI) (stress UI [SUI], urgency UI [UUI] and mixed UI [MUI]). Voiding symptoms include
hesitancy, intermittency, slow stream, straining, splitting or spraying of the urinary stream and terminal dribble.
Post-micturition symptoms include post-void dribbling and feeling of incomplete bladder emptying. Lower
urinary tract symptoms are often broadly classified into clinical syndromes/entities such as overactive bladder
(OAB), underactive bladder (UAB), UI, nocturia, dysfunctional voiding, or genito-urinary fistulae.

Lower urinary tract symptoms are an extremely common complaint in the female population in every part of
the world [2-5]. It causes a great deal of distress and embarrassment [6], as well as significant costs to both
individuals and societies. Estimates of prevalence vary according to the definition and the population studied.
However, there is universal agreement about the importance of the problem in terms of human suffering and
economic cost [7].

1.1 Aim and objectives


These Guidelines from the European Association of Urology (EAU) Working Panel on Non-neurogenic Female
LUTS are written by a multidisciplinary group, primarily for urologists, but are likely to be referred to by other
professional groups. They aim to provide sensible and practical evidence-based guidance on the clinical
problems associated with female LUTS rather than an exhaustive narrative review. Such reviews for UI and
other LUT syndromes are already available from the International Consultation on Incontinence (ICI) [8] and
other sources, so these EAU Guidelines do not describe the causation, basic science, epidemiology and
psychology of LUTS/UI in detail. The focus of these Guidelines is entirely on assessment and treatment,
reflecting clinical practice. These guidelines also do not consider women with LUTS caused by neurological
disease, or LUTS occurring in children, as this is covered by complementary EAU Guidelines [9, 10].

The current Guidelines provide:


• A clear description of the assessment and treatment of common clinical problems. This can provide the
basis for thinking through a patient’s management and for planning and designing clinical services;
• A brief but authoritative summary of the current state of evidence on clinical topics, complete with
references to the original sources;
• Clear guidance on what to do or not to do, in most clinical circumstances. This should be particularly
helpful in those areas of practice for which there is little, or no, high-quality evidence.

The latest edition of the guidelines has seen a significant expansion of scope from ‘urinary incontinence’ to ‘non-
neurogenic female LUTS’. The primary consideration here was to include the significant population of women with
functional urological conditions not necessarily associated with UI that were hitherto not accounted for in previous
guidelines. Secondary considerations were to align more cohesively with the existing Non-neurogenic Male LUTS
Guideline. As a consequence of the anatomical and physiological differences between the male and female LUT,
the prevalence, pathophysiology, diagnostic approach and management of male and female LUTS differ widely.
For that reason, the EAU Guidelines Office decided to provide gender-specific guidelines on LUTS and UI going
forward. As a result, the section on post-prostatectomy UI has been moved to the Non-neurogenic Male LUTS
Guideline. This reconfiguration has also seen some additional sections added to this Guideline (including non-
obstetric fistulae, female bladder outlet obstruction [BOO], UAB and nocturia) and over the course of the next two
or three iterations the scope is likely to widen further.

1.2 Panel composition


The EAU Non-neurogenic Female LUTS Panel consists of a multidisciplinary group of experts, including
urologists, a uro-gynaecologist, a urodynamic scientist and a physiotherapist. All experts involved in the
production of this document have submitted potential conflict of interest statements which can be viewed on
the EAU website: https://uroweb.org/guideline/non-neurogenic-female-luts/.

The Panel acknowledge the support of Mrs. M. de Heide (Bekkenbodem4all), Mrs. T. van den Bos
(Bekkenbodem4All), Mrs. M.L van Poelgeest-Pomfret (World Federation for Incontinence and Pelvic Problems
[WFIP]) and Dr. H. Cobussen-Boekhorst (nurse practitioner) in the development of these guidelines.

1.3 Available publications


A quick reference document (Pocket Guidelines) is available, both in print and as an app for iOS and Android
devices. These are abridged versions which require consideration together with the full text versions. All

10 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021


documents are accessible through the EAU website: https://uroweb.org/guideline/non-neurogenic-female-luts/.

1.4 Publication history


The first EAU Urinary Incontinence Guidelines were published in 2001. The guideline has been modified since
to broaden its scope specifically to include other female LUTS as of 2021.

2. METHODS
2.1 Introduction
For the 2021 Non-neurogenic Female LUTS Guideline, the existing text of the 2018 Urinary Incontinence
Guidelines was re-structured and significantly expanded. The PICO question-based format of the text was
modified to improve readability, although the underlying PICO structure still informs search strategies.

Databases searched for the 2021 update included Medline, EMBASE, and the Cochrane Libraries, covering
a time frame between Jan 1st, 2000 and June 16th, 2020 with a focus on high level evidence only (systematic
reviews and meta-analyses). Detailed search strategies are available online:
https://uroweb.org/guideline/non-neurogenic-female-luts/?type=appendices-publications.

For the 2021 edition of the Guidelines a number of de novo systematic reviews have been undertaken by the
Panel on the subjects of OAB syndrome and the diagnosis and treatment of female BOO [11, 12]. Full publication
of the systematic review results are pending; however, the preliminary results have informed the corresponding
sections of this Guidelines update.

For each recommendation within the guidelines there is an accompanying online strength rating form, the basis
of which is a modified GRADE methodology [13, 14]. Each strength rating form addresses a number of key
elements, namely:
1. the overall quality of the evidence which exists for the recommendation; references used in
this text are graded according to a classification system modified from the Oxford Centre for
Evidence-Based Medicine Levels of Evidence [15];
2. the magnitude of the effect (individual or combined effects);
3. the certainty of the results (precision, consistency, heterogeneity and other statistical or
study related factors);
4. the balance between desirable and undesirable outcomes;
5. the impact of patient values and preferences on the intervention;
6. the certainty of those patient values and preferences.

These key elements are the basis which panels use to define the strength rating of each recommendation. The
strength of each recommendation is represented by the words ‘strong’ or ‘weak’ [16]. The strength rating forms
will be available online.
Additional information can be found in the general Methodology section of this document, and
online at the EAU website: https://uroweb.org/guidelines/policies-and-methodological-documents/.

A list of Associations endorsing the EAU Guidelines can also be viewed online at the above address.

2.2 Review
The guideline has been peer reviewed prior to publication in 2021.

2.3 Future goals


• A systematic review on the topic of Pelvic Organ Prolapse (POP);
• A systematic review on synthetic/mesh-related mid-urethral sling (MUS) complications;
• A systematic review on the diagnosis and treatment of underactive bladder (UAB) in women.

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 11


3. DIAGNOSTIS
3.1 History and physical examination
Taking a thorough clinical history is fundamental to the process of clinical evaluation. Despite the lack of high-
level evidence to support it, there is universal agreement that taking a history should be the first step in the
assessment of anyone with LUTS. The history should include a full evaluation of LUT symptoms (storage,
voiding and post-micturition symptoms), sexual, gastrointestinal and neurological symptoms. Details of
urgency episodes, the type, timing and severity of UI, and some attempt to quantify symptoms should also be
made. The history should help to categorise LUTS as storage, voiding and post-void symptoms, and classify UI
as SUI, UUI, MUI or overflow incontinence, the latter being defined as ‘the complaint of UI in the symptomatic
presence of an excessively (over-) full bladder (no cause identified)’ [17]. It should also identify patients who
need referral to an appropriate clinic/specialist. These may include patients with associated pain, haematuria,
a history of recurrent urinary tract infection (UTI), pelvic surgery or radiotherapy, constant leakage suggesting
a fistula (see Section 4.8), new-onset enuresis or suspected neurological disease. A neurological, obstetric and
gynaecological history may help to understand the underlying cause and identify factors that may impact on
treatment decisions. Guidance on history taking and diagnosis in relation to UTIs, neuro-urological conditions
and chronic pelvic pain can be found in the relevant EAU Guidelines [9, 18, 19]. The patient should also be
asked about other co-morbidities as well as smoking status, previous surgical procedures and for the details of
current medications, as these may impact on symptoms of LUTS.

Similarly, there is little evidence from clinical trials that carrying out a clinical examination improves outcomes,
but wide consensus suggests that it remains an essential part of assessment of patients with LUTS. It should
include abdominal examination, to detect an enlarged urinary bladder or other abdominal mass, and digital
examination of the vagina and/or rectum. Examination of the perineum in women includes an assessment of
oestrogen status, pelvic floor muscle (PFM) function and a careful assessment of any associated POP. A cough
stress test is necessary to look for SUI. The urethral mobility can be assessed visually and with an Ulmsten/
Pinch or Marshall/Bonney test. Pelvic floor contraction strength can also be assessed digitally. A focused
neuro-urological examination should also be routinely undertaken.

3.1.1 Summary of evidence and recommendation for history taking and physical examination

Summary of evidence LE
History taking including symptoms and comorbidities and a focussed physical examination is an 4
essential part of the evaluation of a woman with LUTS.

Recommendation Strength rating


Take a complete medical history including symptoms and comorbidities and a focused Strong
physical examination in the evaluation of women with lower urinary tract symptoms.

3.2 Patient questionnaires


This section includes symptom scores, symptom questionnaires/scales/indices, patient-reported outcome
measures (PROMs) and health-related quality of life (HRQoL) measures. The latter include generic or condition-
specific measures. Questionnaires should have been validated for the language in which they are being
used, and, if used for outcome evaluation, should have been shown to be sensitive to change. The US Food
and Drug Administration (FDA) published guidance for industry on patient-reported outcome instruments
(questionnaires) in 2009 [20].

Although many studies have investigated the validity and reliability of urinary symptom questionnaires and
PROMs most of these studies include mixed populations (men and women). This limits the extent to which
results and conclusions from these studies can be applied to particular LUT syndromes in women. Some
questionnaires (ICIQ-FLUTS, QUID, 3IQ, ICIQ-SF) have potential to discriminate UI types in women [21-23].
Others have been developed to measure symptoms and bother in OAB (OABQ-SF, B-SAQ) and other specific
conditions. Some questionnaires are responsive to change and may be used to measure outcomes, though
evidence on their sensitivity is inconsistent [24, 25]. No evidence was found to indicate whether use of QoL or
condition-specific questionnaires has an impact on outcome of treatment.
Detailed description of the different urinary symptoms questionnaires and PROMs is beyond the
scope of this guideline. For more information we recommend the 6th ICI review on patient reported outcomes
assessment [26]. To date, there is no one questionnaire that fulfils all requirements for assessment of women

12 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021


with LUTS. Clinicians must evaluate the tools that exist, for use alone or in combination, for assessment
and monitoring of treatment outcome [27]. The questionnaires can be found on the following websites:
www.iciq.net, https://eprovide.mapi-trust.org, www.pfizerpcoa.com, www.ncbi.nlm.nih.gov.

3.2.1 Summary of evidence and recommendation for patient questionnaires

Summary of evidence LE
Validated condition-specific symptom scores assist in the screening for, and categorisation of LUTS. 3
Validated symptom scores measure the severity of UI and LUTS. 3
Both condition-specific and general health status questionnaires measure current health status and 3
change following treatment.
Patient questionnaires cannot replace a detailed patient consultation and should only be used as part 4
of a complete medical history.

Recommendation Strength rating


Use a validated and appropriate questionnaire as part of the standardised assessment of Strong
female lower urinary tract symptoms.

3.3 Bladder diaries


Measurement of the frequency and severity of LUTS is an important step in the evaluation and management
of lower urinary tract (LUT) dysfunction. Bladder diaries are a semi-objective method of quantifying symptoms,
such as frequency of UI events, number of nocturia episodes etc. They also quantify urodynamic variables,
such as voided volume, 24-hour urine volume or nocturnal total urine volume. Voiding diaries are also known as
micturition time charts, frequency/volume charts and bladder diaries.
Discrepancy between diary recordings and the patient rating of symptoms, e.g. frequency of UI,
can be useful for patient counselling. In addition, fluid intake and voided volume measurement can be used to
support diagnoses and management planning, for example in OAB, and for identifying polyuria, either 24-hour
or nocturnal polyuria. Diaries can also be used to monitor treatment response and are widely used in clinical
trials. In patients with severe UI, a bladder diary is unlikely to accurately report 24-hour urine output.

Consensus terminology is now well-defined and widely accepted [1, 28]. However, the terms micturition diary,
frequency volume chart, bladder diary and voiding diary, have been used interchangeably for many years, but
only bladder diaries include information on fluid intake, times of voiding, voided volumes, UI episodes, pad
usage, degree of urgency and degree of UI recorded for at least 24 hours. When reviewing the evidence all
synonymous search terms have been included.
Two studies have demonstrated the reproducibility of diaries in both men and women [29, 30].
Another two studies have shown the feasibility, reliability and validity of the bladder diary [31, 32]. Further
studies have demonstrated variability of diary data within a 24-hour period and compared voided volumes
recorded in diaries with those recorded by uroflowmetry [33, 34]. Another study found that keeping a bladder
diary had a therapeutic benefit [35].
A number of observational studies have demonstrated a close correlation between data obtained
from bladder diaries and standard symptom evaluation [36-39]. The optimum number of days required for
bladder diaries appears to be based on a balance between accuracy and compliance [40, 41]. Diary durations
between 3 and 7 days are routinely used in the literature.

3.3.1 Summary of evidence and recommendations for bladder diaries

Summary of evidence LE
Bladder diaries of three to seven days duration are a reliable tool for the objective measurement of 2b
mean voided volume, day- and night-time frequency, and UI episode frequency.
Bladder diaries are sensitive to change and are a reliable outcome measure. 2b

Recommendations Strength rating


Ask patients with lower urinary tract symptoms to complete a bladder diary as part of the Strong
standardised assessment of female LUTS.
Use a bladder diary with a duration of at least three days. Strong

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 13


3.4 Urinalysis and urinary tract infection
Reagent strip (‘dipstick’) urinalysis may indicate proteinuria, haematuria or glycosuria or suggest UTI requiring
further assessment. Refer to the Urological Infections Guidelines for diagnosis and treatment of UTI [18].

Urine dipstick testing is a useful adjunct to clinical evaluation in patients in whom urinary symptoms are
suspected to be due to UTI. Urinalysis negative for nitrite and leucocyte esterase may exclude bacteriuria in
women with LUTS [42], and should be included, with urine culture when necessary, in the evaluation of all
patients with LUTS. Urinary incontinence or worsening of LUTS may occur during a UTI [43] and existing UI
may worsen during UTI [44]. The rate and severity of UI was unchanged after eradication of asymptomatic
bacteriuria in nursing home residents [45].

3.4.1 Summary of evidence and recommendations for urinalysis

Summary of evidence LE
Urinalysis negative for nitrite and leucocyte esterase may exclude bacteriuria in women with LUTS. 3
Urinary incontinence may be a symptom during a UTI, and LUTS may deteriorate during a UTI. 3
The presence of a UTI worsens existing symptoms of UI. 3
Elderly nursing home patients with UI do not benefit from treatment of asymptomatic bacteriuria. 2

Recommendations Strength rating


Perform urinalysis as a part of the initial assessment of a patient LUTS. Strong
If a urinary tract infection is present with LUTS, reassess the patient after treatment. Strong
Do not routinely treat asymptomatic bacteriuria in elderly patients to improve urinary Strong
incontinence.

3.5 Post-void residual volume


Post-void residual (PVR) volume is the amount of urine that remains in the bladder after voiding. It indicates
poor voiding efficiency, which may result from a number of contributing factors. Assessment of PVR is
important because it may worsen symptoms and, more rarely, may be associated with UTI, upper urinary
tract (UUT) dilatation and renal insufficiency. Both BOO and detrusor underactivity (DU) potentially contribute
to the development of PVR. Post-void residual can be measured by catheterisation or ultrasound (US). The
prevalence of PVR in patients with LUTS is uncertain, partly because of the lack of a standard definition of an
abnormal PVR volume. Bladder voiding efficiency (BVE) is the proportion of the total bladder volume that is
voided by the patient. The BVE can be defined as a percentage: BVE = (voided volume/[VV]+PVR) × 100. This
may be a more reliable parameter to evaluate poor voiding [46].

Most studies investigating PVR assessed mixed populations. Although some studies have included women
with UI and men and women with LUTS, they have also included children and adults with neurogenic UI.
In general, the data on PVR can be applied with caution to women with non-neurogenic LUTS. The results
of studies investigating the best method of measuring PVR [45, 47-51] have led to the consensus that US
measurement of PVR is preferable to catheterisation due to its favourable risk-benefit profile.

In peri- and post-menopausal women without significant LUTS or pelvic organ symptoms, 95% of women had
a PVR < 100 mL [52]. In women with UUI, a PVR > 100 mL was found in only 10% of cases [53]. Other research
has found that a high PVR is associated with POP, voiding symptoms and an absence of SUI [52, 54-56]. In
women with SUI, the mean PVR was 39 mL measured by catheterisation and 63 mL measured by US, with
16% of women having a PVR > 100 mL [57]. Some authors have suggested that it is reasonable to consider a
PVR of > 100 mL to be significant, although many women may remain asymptomatic and hence it is imperative
to consider the clinical context [53]. There is no consensus on what constitutes a significant PVR in women;
therefore, the Panel suggests the additional use of BVE.

3.5.1 Summary of evidence and recommendations for post-void residual

Summary of evidence LE
Lower urinary tract symptoms are associated with a higher PVR compared to asymptomatic 2
population groups.

14 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021


Recommendations Strength rating
Measure post-void residual volume (PVR) in patients with LUTS during initial assessment. Strong
Use ultrasound to measure PVR. Strong
Monitor PVR in patients receiving treatments that may cause or worsen voiding dysfunction. Strong
Provide Bladder Voiding Efficiency as an additional parameter when measuring PVR. Weak

3.6 Urodynamics
Urodynamic testing is widely used as an adjunct to clinical diagnosis, in the belief that it may help to provide
or confirm diagnosis, predict treatment outcome or facilitate discussion during counselling. The simplest
form of urodynamic evaluation is uroflowmetry. The maximum flow rate (Qmax), the volume voided and the
shape of the curve in addition to the PVR volume (see above) are the most important aspects to be assessed
[26]. The bladder should be sufficiently full because of the volume dependency of Qmax [58, 59]. A minimum
voided volume of 150 mL is advised in males, but there is very little evidence to suggest a volume threshold in
females. It is of relevance to ask the patient whether or not the voiding was representative.

Invasive urodynamic tests are often performed prior to invasive treatment of LUTS. These tests include
multichannel cystometry and pressure-flow studies, ambulatory monitoring and video-urodynamics, and
different tests of urethral function, such as urethral pressure profilometry and Valsalva leak point pressure
(VLPP). The ICS and the United Kingdom Continence Society provide standards to optimise urodynamic
test performance and reporting [60, 61]. A characteristic of a good urodynamic study is that the patient’s
symptoms are replicated, recordings are checked for quality control and results interpreted in the context of the
clinical problem, remembering that there may be physiological variability within the same individual [60]. Non-
invasive alternatives for measurement of detrusor pressure and BOO include transabdominal wall near-infrared
spectroscopy and US detrusor wall thickness analysis, but as yet these techniques have not been adopted into
routine clinical practice [26].

In common with most physiological tests there is variability in urodynamic results. This has consequences for
the reproducibility, diagnostic accuracy and predictive value of urodynamic testing (UDS). It has been stated
that, at least in the case of cystometry and pressure-flow studies, one set of measurements suffice, but only if
the patient’s symptoms have been replicated [60]. It can nevertheless be sometimes necessary or advisable to
repeat the urodynamic measurements. The risk-benefit profile of repeating urodynamic testing should always
be considered.

Further condition-specific information regarding the role of urodynamic testing in OAB, SUI, BOO and UAB can
be found in respective sections of this Guideline.

3.6.1 Variability
Contradictory findings were reported in studies assessing same-session repeatability of cystometric and
pressure-flow studies [62, 63]. There is also conflicting evidence about the reproducibility of maximum urethral
closure pressure (MUCP) measurement [64, 65]. One method of recording MUCP cannot be compared
meaningfully to another [65, 66]. Valsalva leak point pressure measurement is not standardised and there is
minimal evidence about its reproducibility. No studies on the reproducibility of ambulatory monitoring in non-
neurological patients have been published [26].

3.6.2 Diagnostic accuracy


Clinical diagnosis and cystometric findings often do not correlate [67, 68] and asymptomatic women may have
abnormalities on urodynamic testing. As the urodynamic diagnosis is often taken as the benchmark in the
assessment of LUT function, this implies that the evaluation of other tests of LUT function may be biased as
a result. The diagnostic accuracy of urethral pressure profilometry [69] and urethral retro-resistance pressure
measurement in SUI is generally poor [26]. Valsalva leak point pressure did not reliably assess UI severity in a
cohort of women selected for surgical treatment of SUI [70]. Urethral pressure reflectometry may have greater
diagnostic accuracy but its clinical role remains unclear [71]. Ambulatory urodynamics may detect unexpected
physiological variance from normal more often than conventional cystometry, but the clinical relevance of this is
also uncertain [72, 73].

A pressure-flow study, that is, the simultaneous measurement of flow rate and detrusor pressure during
voiding, can reveal whether a poor flow rate and PVR’s are due to BOO, poor bladder contraction strength
(DU) or a combination of both. Also, it may provide information on the degree of pelvic floor relaxation and
thus diagnose dysfunctional voiding. As with uroflowmetry, its representativeness must be verified. Several
proposals to define BOO in women have been made. These definitions are based on detrusor pressure values,

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 15


either the value at Qmax or the maximum value, and the Qmax value, either during the pressure-flow study or
during uroflowmetry; sometimes combined with the findings during fluoroscopic imaging (see Section 4.5.4.8
for more detailed information) [74, 75]. Unlike the situation in men, there is no generally accepted definition
or nomogram in women. Bladder contraction strength parameters are derived from the detrusor pressure and
flow rate values during a pressure flow study or from stop tests [75], but again validation is poor. In addition,
while these parameters estimate the strength of the contraction, they ignore its speed and persistence (see
Section 4.4.3.2 for more detailed information) [76]. A video-urodynamic study can be useful to detect the site of
obstructed voiding, which may be anatomical or functional [77]. Also, video-urodynamics may detect a bladder
diverticulum or gross reflux as a pressure absorbing reservoir.

3.6.3 Predictive value


Performing urodynamic evaluation is only useful if it leads to more effective clinical care and better outcomes.
A Cochrane review of eight randomised controlled trials (RCTs) showed that use of urodynamic tests in
women with UI increased the likelihood of prescribing drugs and did not increase the likelihood of undergoing
surgery. However, there was no evidence that this influence on decision-making altered the clinical outcome
of treatment [78]. Most RCTs addressed the utility of urodynamic tests on SUI only, include women with
uncomplicated SUI. A meta-analysis including four RCTs comparing surgical outcomes in women with self-
reported SUI (or stress-predominant MUI) who were investigated via urodynamics with women who had office
evaluation only found that there was no difference in cure and complication rate [79]. On the other hand, in a
large retrospective multicentre study it was found that only 36% of patients were defined as ‘‘uncomplicated’’
according to the definitions used in large RCTs [80]. The urodynamic observations were not consistent with the
pre-urodynamic diagnosis in 1,276 out of 2,053 patients (62.2%). Voiding dysfunctions were urodynamically
diagnosed in 394 patients (19.2%) and planned surgery was cancelled or modified in 304 patients (19.2%), due
to the urodynamic findings [81].

The predictive value of urethral function tests remains unclear. In observational studies, there was no consistent
correlation between the results of these tests and subsequent success or failure of SUI surgery [37-39, 82]. The
same was true in a secondary analysis of an RCT [83].

The presence of pre-operative DO in women with stress-predominant MUI has been associated with post-
operative UUI but did not predict overall treatment failure following MUS surgery or colposuspension [83]. The
urodynamic diagnosis of detrusor overactivity (DO) had no predictive value for treatment response in studies on
fesoterodine, onabotulinumtoxinA and sacral nerve stimulation (SNS) in patients with OAB symptoms [84-87].
Augmentation cystoplasty aims to abolish DO, improve bladder compliance and increase functional bladder
capacity but there is no evidence to guide whether or not pre-operative urodynamics is predictive of outcome.
Most clinicians would however consider pre-operative urodynamics as essential prior to contemplating
augmentation cystoplasty.

A pressure-flow study is capable of discriminating BOO from DU as a cause of voiding dysfunction. The
predictive value of parameters derived from such a study for voiding dysfunction after a surgical procedure
for SUI is however low. A low pre-operative flow rate and a low detrusor voiding pressure have been shown
to correlate with voiding dysfunction after a tension-free vaginal tape (TVT) and an autologous fascial sling
procedure, respectively [88-90]. Bladder contraction strength parameters combining flow rate and detrusor
pressure values only poorly predicted voiding dysfunction after autologous fascial sling [91]. Post-hoc analysis
of two high-quality surgical trials on TVT, Burch colposuspension and autologous fascial sling showed that no
pre-operative urodynamic parameter could predict post-operative voiding dysfunction in a selected population
of women with low pre-operative PVR [92, 93].

The Panel recognises that it may be valuable to use urodynamic test results to select the optimum
management strategy; however, there is inconsistent evidence regarding the predictive value of such tests.
When urodynamics and clinical assessment (i.e. by history and examination) are in disagreement, there needs
to be a careful re-evaluation of the clinical symptoms of the patient and the investigation results to ensure that
the diagnosis is correct before invasive treatments are contemplated.

16 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021


3.6.4 Summary of evidence and recommendations for urodynamics

Summary of evidence LE
Most urodynamic parameters show variability within the same session and over time, and this may 3
limit their clinical interpretation.
Different techniques of measuring urethral function may have good test-retest reliability, but do not 3
consistently correlate to other urodynamic tests or to the severity of UI.
There may be inconsistency between history and urodynamic results. 3
Urodynamic diagnosis of DO does not influence treatment outcomes in patients with OAB. 1a
Pre-operative urodynamics in women with uncomplicated, clinically demonstrable, SUI does not 1b
improve the outcome of surgery for SUI.
There is no consistent correlation between the result of urethral function tests and subsequent 3
success or failure of SUI surgery.
There is no consistent evidence that pre-operative DO is associated with surgical failure of MUS in 3
women.
The presence of pre-operative DO may be associated with persistence of urgency post-operatively. 3

Recommendations Strength rating


Adhere to ‘Good Urodynamic Practice’ standards as described by the International Strong
Continence Society when performing urodynamics in patients with LUTS.
Do not routinely carry out urodynamics when offering treatment for uncomplicated stress Strong
urinary incontinence.
Do not routinely carry out urodynamics when offering first-line treatment to patients with Strong
uncomplicated overactive bladder symptoms.
Perform urodynamics if the findings may change the choice of invasive treatment. Weak
Do not use urethral pressure profilometry or leak point pressure to grade severity of urinary Strong
incontinence as they are primarily tests of urethral function.

3.7 Pad testing


Measurement of urine loss using an absorbent pad worn over a set period of time or during a protocol of
physical exercise can be used to quantify the presence and severity of UI, as well as a patient’s response to
treatment.

The clinical utility of pad tests for people with UI has been assessed in two systematic reviews [94, 95]. A
one-hour pad test using a standardised exercise protocol and a diagnostic threshold of 1.4 g shows good
specificity but lower sensitivity for symptoms of SUI and MUI. A 24-hour pad test using a threshold of 4.4 g is
more reproducible but is difficult to standardise with variation according to activity level [96]. A pad test with a
specific short graded exercise protocol also has diagnostic value but a negative test should be repeated with
the degree of provocation increased [97]. The usefulness of pad tests in quantifying severity and predicting
outcome of treatment is uncertain [94, 98, 99]. Pad tests are responsive to change following successful
treatment [100]. Pad testing using a standardised bladder volume (50% of cystometric capacity) has been
suggested to allow for a more reliable assessment of UI in a small study including 25 women [101]. There is no
evidence that one type of pad test is superior to another.

3.7.1 Summary of evidence and recommendations for pad testing

Summary of evidence LE
A pad test can diagnose UI accurately. 2
Standardisation of bladder volume and degree of provocation improves reproducibility. 2
Twenty-four hours is sufficient duration for home-based testing balancing diagnostic accuracy and 2
adherence.
Change in leaked urine volume on pad tests can be used to measure treatment outcome. 2
Pad tests can be a useful tool in the research setting and are an optional investigation in clinical 4
practice.

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 17


Recommendations Strength rating
Use a pad test of standardised duration and activity protocol. Strong
Use a pad test when quantification of urinary incontinence is required, especially to assess Weak
response to treatment.

3.8 Imaging
Imaging improves our understanding of the anatomical and functional abnormalities that may cause LUTS.
In clinical research, imaging is used to understand the relationship between anatomy and function, between
conditions of the central nervous system (CNS) or of the LUT, and to investigate the relationship between LUT
and pelvic floor imaging and treatment outcomes.
Ultrasound and magnetic resonance imaging (MRI) have largely replaced X-ray imaging in the
evaluation of the pelvic floor. Ultrasound is preferred to MRI because of its ability to produce three-dimensional
and four-dimensional (dynamic) images at lower cost and wider availability.
In general, there is no need for UUT imaging unless a high-pressure bladder, severe POP or
chronic urinary retention is suspected or diagnosed or abnormal renal function tests are observed. In cases of
suspected UI caused by an UUT anomaly or uretero-vaginal fistula, UUT imaging (intravenous urography or CT
scan) may be indicated [102].

3.8.1 Ultrasound
Ultrasonography of the LUT plays a role in the differential diagnosis of women with LUTS, when there is a
suspicion of bladder tumour, stones, etc. and in cases presenting with haematuria.

Ultrasonography has been used in the evaluation of UI and pelvic floor since the 1980s. Different imaging
approaches, such as abdominal, transvaginal, transrectal, perineal and transurethral are described. The bladder
neck and urethra are easily visible and measurements can be done at rest, during straining, coughing and
during pelvic floor contraction. Ultrasonography can be used to assess PFMs and their function. Contraction
of the pelvic floor results in displacement of pelvic structures that can easily be imaged on US. Integrity of
the levator ani muscle can be determined by 3D transperineal US. The specific role of US is discussed in the
condition-specific sections of this guideline.

3.8.2 Detrusor wall thickness


As OAB syndrome is linked to DO, it has been hypothesised that frequent detrusor contractions may increase
detrusor/bladder wall thickness (DWT/BWT). Transvaginal US seems to be more accurate with less inter-
observer variability than transabdominal and transperineal approaches [103]. Several cut-off points have been
suggested, from 4.4 to 6.5 mm. Other studies are contradictory and did not find this correlation. No consensus
exists as to the relationship between OAB and increased BWT/DWT [104], and there is no evidence that BWT/
DWT imaging improves management of OAB in practice. There is no widely accepted, standardised bladder
volume for bladder wall thickness measurement.

In a retrospective study including 227 women with symptoms of voiding difficulty (hesitancy, intermittency
and poor stream), 74 (32.6%) were diagnosed with voiding dysfunction on the basis of free uroflowmetry and
residual urine. While controlling for the effect of DO, the relationships between DWT and different parameters of
voiding function in pressure–flow studies and free uroflowmetry were examined. The results indicated that DWT
was not associated with any urodynamic parameters that may indicate BOO [105].

3.8.3 Magnetic resonance imaging


There is a general consensus that MRI provides good global pelvic floor assessment, including POP, defecatory
function and integrity of the pelvic floor support [106]. However, there is a large variation in MRI interpretation
between observers [107] and little evidence to support its clinical usefulness in the management of LUTS/UI.
There is no conclusive evidence that MRI evaluation of POP is more clinically useful than vaginal examination.
Studies have assessed the use of imaging to assess the mechanism of MUS insertion for SUI. One study
suggested that MUS placement decreased mobility of the mid-urethra but not of the bladder neck [108].
Following MUS, a wider gap between pubic symphysis and sling (assessed by imaging) has been shown to
correlate with a lower chance of cure of SUI [109].

18 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021


3.8.4 Summary of evidence and recommendation for imaging

Summary of evidence LE
There is no consistent evidence that routine urinary tract imaging is useful in the evaluation or 3
management of lower urinary tract symptoms.
There is no consistent evidence that bladder (detrusor) wall thickness measurement is useful in the 3
management of OAB.

Recommendation Strength rating


Do not routinely carry out imaging of the upper or lower urinary tract as part of the Strong
assessment of lower urinary tract symptoms.

4. DISEASE MANAGEMENT
4.1 Overactive bladder
4.1.1 Epidemiology, aetiology, pathophysiology
Overactive bladder is defined by the ICS as ‘urinary urgency, usually accompanied by frequency and nocturia,
with or without UUI, in the absence of urinary tract infection (UTI) or other obvious pathology’ [110]. Overactive
bladder is a chronic condition and can have debilitating effects on QoL. The hallmark urodynamic feature is
DO, although this may not be demonstrated in a large proportion of OAB patients, which may, in part, be due to
failure to reproduce symptoms during urodynamic assessment.

The EPidemiology of InContinence (EPIC) study was one of the largest population-based surveys that studied
the prevalence of LUTS and OAB [111]. Conducted in five countries, including Canada, Germany, Italy,
Sweden, and the UK, it was a cross-sectional telephone survey of adults aged > 18 years. The study included
over 19,000 participants and demonstrated an overall prevalence of OAB symptoms of 11.8% (10.8% in
men and 12.8% in women). Other studies have reported prevalence of up to 30–40%, with rates generally
increasing with age [5].

Various theories have been proposed to explain the pathophysiology of OAB, mainly relating to imbalances
in inhibitory and excitatory neural pathways to the bladder, or the sensitivity of bladder muscle receptors.
However, no definite identifiable cause has been established.

4.1.2 Classification
Overactive bladder is generally classified into ‘wet’ and ‘dry’, based on the presence or absence of associated UI.

4.1.3 Diagnostic evaluation


The evaluation of a patient with symptoms of OAB follows the general pathway of evaluation of the female
LUTS patient.

4.1.3.1 Bladder diaries


Diaries are particularly helpful in establishing and quantifying symptoms of frequency, urgency and UI, and
may be valuable in assessing change over time or response to treatment. A number of observational studies
have demonstrated a close correlation between data obtained from bladder diaries and standard symptom
evaluation [36-39]. The optimum number of days required for bladder diaries appears to be based on a balance
between accuracy and compliance. Diary durations between 3 and 7 days are routinely used in the literature.

4.1.3.2 Urodynamics
Urodynamics is essential in establishing the presence of DO, but its absence does not preclude the diagnosis
of OAB which is based on symptoms alone.

A Cochrane review of seven RCTs showed that use of urodynamic tests increased the likelihood of prescribing
drugs or avoiding surgery. However, there was no evidence that this influence on decision-making altered the
clinical outcome of treatment [112]. A sub-analysis of an RCT comparing fesoterodine to placebo [84, 85] showed
that the urodynamic diagnosis of DO had no predictive value for treatment response.

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 19


4.1.3.3 Summary of evidence and recommendations regarding associated conditions

Summary of evidence LE
Bladder diaries of three to seven days duration may be helpful in quantifying symptoms of OAB, and 3
assessing response to treatment.
Urodynamic diagnosis of DO does not influence treatment outcomes in patients with OAB. 1a

Recommendations Strength rating


Request that patients complete at least a three-day bladder diary at initial valuation and Strong
before each therapeutic intervention for overactive bladder (OAB).
Do not routinely carry out urodynamics when offering first-line treatment to patients with Strong
uncomplicated OAB symptoms.

4.1.4 Disease management


4.1.4.1 Conservative management
In clinical practice, it has long been the convention that non-surgical therapies are recommended first because
they usually carry the lowest risk of harm. While this remains true for non-pharmacological conservative
treatments (e.g. pelvic floor muscle training [PFMT]), increasing concerns regarding the adverse events of some
pharmacological treatments used to treat LUTS (e.g. anticholinergic drugs), particularly regarding cognitive
function have emerged and patients should be fully counselled regarding this potential risk.

4.1.4.1.1 Addressing underlying disease/cognitive impairment


Lower urinary tract symptoms, especially in the elderly, have been associated with multiple comorbid
conditions including:
• cardiac failure;
• chronic renal failure;
• diabetes;
• chronic obstructive pulmonary disease;
• neurological disease;
• general cognitive impairment;
• sleep disturbances, e.g. sleep apnoea;
• depression;
• metabolic syndrome.

It is possible that improvement of associated disease may reduce the severity of urinary symptoms. However,
this is often difficult to assess as patients frequently suffer from more than one condition. In addition,
interventions may be combined and individualised, making it impossible to decide which alteration in an
underlying disease has affected a patient’s symptoms.

One study involving middle-aged women with type 1 diabetes mellitus showed that 10% of these women had
UUI. The study showed no correlation between earlier intensive treatment of type 1 diabetes mellitus and the
prevalence of UI in later life vs. conventional treatment [113].

4.1.4.1.1.1 Summary of evidence and recommendation regarding associated conditions

Summary of evidence LE
There is a lack of evidence that improving any associated comorbid condition improves OAB. 3

Recommendation Strength rating


Review any new medication associated with the development or worsening of UI. Weak

4.1.4.1.2 Adjustment of other medication


Although LUTS are listed as an adverse effect of many drugs in drug compendia, this mainly derives from
uncontrolled individual patient reports and post-marketing surveillance. Few controlled studies have used the
occurrence of LUTS as a primary outcome, or were powered to assess the occurrence of statistically significant
LUTS or worsening rates against placebo. In most cases, it is therefore not possible to be sure that any drug
causes OAB/LUTS.

20 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021


A structured literature review failed to identify any studies addressing whether adjustment of specific
medications could alter existing symptoms of OAB. Also, there is little evidence relating to the occurrence or
worsening of OAB in relation to prescription of any specific drugs.

4.1.4.1.2.1 Summary of evidence and recommendations for adjustment of non-LUTS medication

Summary of evidence LE
There is very little evidence that alteration of non-uroselective medications can cure or improve 3
symptoms of OAB.

Recommendations Strength rating


Take a history of current medication use from all patients with overactive bladder (OAB). Strong
Review any new medication associated with the development or worsening of OAB Weak
symptoms.

4.1.4.1.3 Urinary containment


Urinary containment is important for people with OAB-wet or UUI when active treatment does not cure
the problem, is delayed, or when it is not available or not possible. Some individuals may prefer urinary
containment rather than to undergo active treatment with its associated risks. Containment includes the use
of absorbent pads, urinary catheters, external collection devices and intravaginal devices. Detailed literature
summaries can be found in the current ICUD monograph [114] and in a European Association of Urology
Nurses guidance document [115].

A systematic review of six RCTs comparing different types of pads found that pads filled with superabsorbent
material were better than standard pads, whilst evidence that disposable pads were better than washable
pads was inconsistent [116]. A series of three crossover RCTs examined performance of different pad designs
for differing populations [117, 118]. For women with light UI, disposable insert pads (within washable pouch
pants) were most effective. In adults with moderate/severe UI, disposable pull-up pants were more effective for
women.

A Cochrane review summarised three RCTs comparing different types of long-term indwelling catheters and
found no evidence that one catheter material or type of catheter was superior to another [119]. A systematic
review of non-randomised studies found no differences in UTI outcome or UUT changes between use of
suprapubic or urethral catheter drainage; however, patients with suprapubic catheters were less likely to have
urethral complications [120].

Clean intermittent self-catheterisation (CISC) is the most commonly used therapy to manage high PVR volumes
and urinary retention [115]. It reduces the risk of complications such as UTI, UUT deterioration, bladder stones
and overflow UI etc. It has not yet been established whether the incidence of UTI, other complications and user
satisfaction are affected by either sterile or clean IC, coated or uncoated catheters or by any other strategy
[121]. The use of hydrophilic catheters may be associated with a lower rate of UTI, but further evidence is
needed, as most of it comes from neurogenic patients [122]. The average frequency of catheterisation is four
to six times per day [123] and the catheter sizes most often used are 12-16 Fr. In aseptic IC, an optimum
frequency of five times showed a reduction of UTI [123]. Frequency of catheterisation needs to be based
on individual need and capability, to prevent chronic and repeated over-filling of bladder [124]. Thorough
counselling regarding techniques, frequency, equipment and adverse effects of CISC should be given to all
potential patients in line with good medical practice.

For people using CISC, a Cochrane review found no evidence that one type of catheter or regimen of
catheterisation was better than another [125]. However, there is recent evidence from a narrative review
suggesting that in certain populations using single-use catheters may reduce urethral trauma and UTI [126]. A
Cochrane review summarising five trials comparing bladder washout policies in adults with indwelling urinary
catheters found inconsistent evidence of benefit [127].

A further Cochrane review summarising eight trials testing whether antibiotic prophylaxis was beneficial for
adults using CISC or indwelling catheterisation found it reduced incidence of symptomatic UTI but possible
harms were not assessed [128]. A multicentre RCT from the UK reported that prophylaxis was well-tolerated
but development of antibiotic resistance was a concern [129].

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 21


4.1.4.1.3.1 Summary of evidence and recommendations for urinary containment

Summary of evidence LE
Pads are effective in containing urine. 1b
Antibiotic prophylaxis may help reduce incidence of UTI in patients who self-catheterise or have an 1a
indwelling catheter, but at the cost of increasing antimicrobial resistance.

Recommendations Strength rating


Ensure that women with overactive bladder (OAB) and/or their carers are informed regarding Strong
available treatment options before deciding on urinary containment alone.
Offer incontinence pads and/or containment devices for management of OAB-wet, either for Strong
temporary symptom control or where other treatments are not feasible.
Offer prophylactic antibiotics to patients with recurrent urinary tract infections who perform Strong
clean intermittent self-catheterisation, or have an indwelling catheter, after discussion
regarding the risk of increasing antimicrobial resistance.

4.1.4.1.4 Lifestyle interventions


Examples of lifestyle factors that may be associated with UI include obesity, smoking, level of physical activity
and liquid intake. Modification of these factors may improve symptoms of OAB.

4.1.4.1.4.1 Caffeine intake


Many drinks contain caffeine, particularly coffee, tea and cola. Conflicting epidemiological evidence of urinary
symptoms being aggravated by caffeine intake has focused attention on whether caffeine reduction may
improve LUTS [130, 131]. A recent review of 14 interventional and 12 observational studies reported that
reduction in caffeine intake may reduce symptoms of urgency, but the certainty of evidence was low with
significant heterogeneity [132].

4.1.4.1.4.2 Fluid intake


Modification of fluid intake, particularly restriction, is a strategy commonly used by people with OAB to
relieve symptoms. Advice on fluid intake given by healthcare professionals should be based on 24-hour fluid
intake and urine output measurements. From a general health point of view, it should be advised that fluid
intake should be sufficient to avoid thirst and that an abnormally low or high 24-hour urine output should be
investigated.

The few RCTs that have been published provide inconsistent evidence [133-135]. In most studies, the
instructions for fluid intake were individualised and it is difficult to assess participant adherence to protocol. All
available studies were in women. An RCT showed that a reduction in fluid intake by 25% improved symptoms
in patients with OAB but not UI [135]. Personalised fluid advice compared to generic advice made no difference
to continence outcomes in people receiving anticholinergics for OAB, according to an RCT comparing drug
therapy alone to drug therapy with behavioural advice [136]. Patients should be warned of the possibility of
worsening constipation as a consequence of fluid restriction.

4.1.4.1.4.3 Obesity and weight loss


Being overweight or obese has been identified as a risk factor for LUTS in many epidemiological studies [137,
138].

There is evidence that the prevalence of both UUI and SUI increases proportionately with rising body mass
index (BMI) [139]. However, the evidence base largely relates to obesity and SUI rather than UUI and OAB.
Therefore, no definite inference can be drawn between obesity and the prevalence of OAB.

4.1.4.1.4.4 Smoking
Smoking cessation is a generalised public health measure and has been shown to be weakly associated with
improving urgency, frequency and UI [140, 141].

The effect of smoking cessation on LUTS was described as uncertain in a health technology assessment (HTA)
review [142].

22 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021


4.1.4.1.4.5 Summary of evidence and recommendations for lifestyle interventions

Summary of evidence LE
Reduction of caffeine intake may reduce symptoms of frequency and urgency. 2
Addition of personalised fluid intake advice to pharmacotherapy provided no additional benefit in 2
patients with OAB.
Reduction in fluid intake by 25% may help improve symptoms of OAB but not UI. 1b
Obesity is a risk factor for UI in women, but the relationship to other OAB symptoms remains unclear. 1b
There is weak evidence that smoking cessation will improve the symptoms of OAB. 3

Recommendations Strength rating


Encourage overweight and obese adults with overactive bladder (OAB)/urinary incontinence Strong
to lose weight and maintain weight loss.
Advise adults with OAB that reducing caffeine intake may improve symptoms of urgency Strong
and frequency, but not incontinence.
Review type and amount of fluid intake in patients with OAB. Weak
Provide smoking cessation strategies to patients with OAB who smoke. Strong

4.1.4.1.5 Behavioural and physical therapies


Terminology relating to behavioural and physical therapies remains confusing because of the wide variety of
ways in which treatment regimens and combinations of treatments have been delivered in different studies
[143]. The terms are used to encompass all treatments which require a form of self-motivated personal
retraining by the patient and also include techniques which are used to augment this effect.

Approaches include bladder training (BT) and PFMT, but terms such as bladder drill, bladder discipline,
bladder re-education and behaviour modification are also used. Almost always in clinical practice, these will
be introduced as part of a package of care including lifestyle changes, patient education and possibly some
cognitive therapy as well. The extent to which individual therapists motivate, supervise and monitor these
interventions is likely to vary but it is recognised that these influences are important components of the whole
treatment package.

4.1.4.1.5.1 Prompted voiding and timed voiding


The term ‘prompted voiding’ implies that carers, rather than the patient, initiate the patient going to void with
the aim of preventing or reducing UI. This applies largely to an assisted care setting.

Two systematic reviews (nine RCTs) [144, 145] confirmed a positive effect on continence outcomes for
prompted voiding in comparison to standard care [145]. Timed voiding is defined as fixed, pre-determined,
time intervals between toileting, applicable for those with or without cognitive impairment. A Cochrane review
of timed voiding reviewed two RCTs, finding inconsistent improvement in continence compared with standard
care in cognitively impaired adults [146].

4.1.4.1.5.2 Bladder Training


Bladder training (BT) is a programme of patient education along with a scheduled voiding regimen with
gradually increasing voiding intervals. Specific goals are to correct faulty habit patterns of frequent urination,
improve control over bladder urgency, prolong voiding intervals, increase bladder capacity, reduce incontinent
episodes and restore patient confidence in controlling bladder function. The ideal form or intensity of a BT
programme for OAB/UI is unclear. It is also unclear whether or not BT can prevent the development of OAB/UI.

There have been three systematic reviews on the effect of BT compared to standard care confirming that BT is
more effective than no treatment in improving UUI [67, 142, 147]. The addition of BT to anticholinergic therapy
did not improve UUI compared to anticholinergics alone but it did improve frequency and nocturia [148]. This
review identified seven RCTs in which BT was compared to drug therapy alone and only showed a benefit for
oxybutynin in cure and improvement of UUI [148].

4.1.4.1.5.3 Pelvic floor muscle training


An immediate effect of contracting the PFMs is a simultaneous inhibition of urgency, detrusor contraction and
incontinence [149]. Intensive and regular strength training of the PFMs over time increases PFM strength and
endurance, and changes the morphology of the pelvic floor which may yield a more effective inhibition of the

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 23


detrusor and help to stabilize the proximal urethra and improve urethral function. To date, there is a lack of
basic and mechanistic studies to confirm that change of pelvic floor morphology improves OAB symptoms.

A systematic review of 11 RCTs [150] including women with OAB compared the efficacy of PFMT vs. inactive
control, usual care, other lifestyle modification or other intervention. The descriptive analysis revealed that
PFMT significantly reduced OAB symptoms (frequency and UUI) in five RCTs, while the remaining six reported
no significant difference. Significant heterogeneity in protocols precluded meaningful comparisons.

4.1.4.1.5.4 Electrical stimulation


The details and methods of delivery of electrical stimulation (ES) vary considerably. Electrical stimulation of
the pelvic floor can also be combined with other forms of conservative therapy, e.g. PFMT with and without
biofeedback. Electrical stimulation is often used to assist women who cannot initiate contractions to identify
their PFMs and in patients with OAB and UUI with the aim of inhibiting detrusor contraction. There is, however,
lack of basic and mechanistic studies to confirm this theory.

A systematic review has been done on 51 trials on 3,443 adults with OAB symptoms [151], with quality
of evidence ranging from very low to moderate. Moderate quality evidence suggests ES is more likely to
improve OAB symptoms compared to sham control, no treatment or placebo. Moderate quality evidence also
suggested that ES was more likely to improve OAB symptoms compared to anticholinergic therapy. There was
insufficient evidence for comparisons to PFMT and between different types of ES.

4.1.4.1.5.5 Acupuncture
In a systematic review with meta-analysis of 10 RCTs including 794 patients (590 women), the authors reported
that acupuncture might have an effect in reducing OAB symptoms compared to sham treatment [152].
The studies were of low quality and compared electro-acupuncture (EA) vs. sham acupuncture, or EA plus
tolterodine vs. tolterodine alone.

4.1.4.1.5.6 Posterior tibial nerve stimulation


Electrical stimulation of the posterior tibial nerve (PTNS) delivers electrical stimuli to the sacral micturition
centre via the S2–S4 sacral nerve plexus. Stimulation is percutaneous with a fine (34-G) needle, inserted just
above the medial aspect of the ankle (P-PTNS). Transcutaneous stimulation is also available (T-PTNS) that
delivers stimulation via surface electrodes which do not penetrate skin. Treatment cycles typically consist of
twelve weekly treatments of 30 minutes.

4.1.4.1.5.6.1 Percutaneous posterior tibial nerve stimulation


The reviewed studies included two 12-week RCTs of P-PTNS against sham treatment [153, 154], one
comparing PTNS to tolterodine, and a 3-year extension trial utilising a maintenance protocol in patients with
UUI [155, 156]. The results of studies of PTNS in women with refractory UUI are consistent. Considered
together, these results suggest that PTNS improves UUI in women who did not have adequate improvement
or could not tolerate anti-muscarinic therapy. However, there is no evidence that PTNS cures UUI in women. In
addition, PTNS is no more effective than tolterodine for improvement of UUI in women overall.

4.1.4.1.5.6.2 Transcutaneous posterior tibial nerve stimulation


A small RCT compared transcutaneous PTNS plus standard treatment (PFMT and BT) with PFMT and BT alone
in older women [157]. Women in the T-TPNS group were more likely to achieve improvement at the end of
therapy.

A systematic review of 13 trials (10 RCTs and 3 cohort studies) compared the efficacy of T-PTNS (treatment
period between 4 and 12 weeks) with sham treatment, anticholinergics, and exercise in treatment of adults
with OAB symptoms [158]. Of note, the populations were adult women and men, and some studies included
patients with neurogenic OAB. Meta-analysis was possible in 2 RCTs comparing T-PTNS with sham treatment,
and revealed mean reduction in total ICIQ-UI SF associated with T-PTNS of −3.79 points.

4.1.4.1.5.7 Summary of evidence and recommendations for behavioural and physical therapies

Summary of evidence LE
Bladder training is effective for improvement of UUI in women. 1b
The combination of BT with anticholinergic drugs does not result in greater improvement of UUI, but 1b
may improve frequency and nocturia.

24 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021


Prompted voiding, either alone or as part of a behavioural modification programme, improves 1b
continence in elderly, care-dependent people.
Pelvic floor muscle training may improve symptoms of frequency and incontinence in women. 1b
Electrical stimulation may improve symptoms of OAB in some women, but the type and mode of 1a
delivery of ES remains variable and poorly standardised.
Percutaneous posterior tibial nerve stimulation appears effective for improvement of UUI in women 2b
who have had no benefit from anticholinergic medication.
A maintenance programme of P-PTNS has been shown to be effective up to three years. 1b
Percutaneous-PTNS has comparable effectiveness to tolterodine for improvement of UUI in women. 1b
No serious adverse events have been reported for P-PTNS in UUI. 3
Transcutaneous-PTNS appears to be effective in reducing OAB symptoms compared to sham 1a
treatment.

Recommendations Strength rating


Offer prompted voiding for adults with overactive bladder (OAB) who are cognitively impaired. Strong
Offer bladder training as a first-line therapy to adults with OAB/urgency urinary incontinence Strong
(UUI).
Ensure that pelvic floor muscle training programmes are as intensive as possible. Strong
Consider posterior tibial nerve stimulation as an option for improvement of OAB/UUI in Strong
women who have not benefited from anticholinergic medication.

4.1.4.2 Pharmacological management


4.1.4.2.1 Anticholinergic drugs
Anticholinergic (antimuscarinic) drugs are currently the mainstay of treatment for OAB. They differ in
their pharmacological profiles, e.g. muscarinic receptor affinity and other modes of action and in their
pharmacokinetic properties, e.g. lipid solubility and half-life.

The evaluation of cure or improvement of OAB is made harder by the lack of standard definitions. In general,
systematic reviews note that the overall treatment effect of drugs is usually small but larger than placebo. In
addition, some RCTs have UI as an outcome rather than UUI. Dry mouth is the commonest side effect, though
constipation, blurred vision, fatigue and cognitive dysfunction may occur with anticholinergic drugs [147].

Immediate-release (IR) anticholinergic preparations provide maximum dosage flexibility, including an off-label
‘on-demand’ use. Immediate-release drugs have a greater risk of side effects than extended release (ER)
formulations because of differing pharmacokinetics. A transdermal delivery system (TDS) and gel developed for
oxybutynin gives a further alternative formulation.

Seven systematic reviews of individual anticholinergic drugs vs. placebo were reviewed [147, 159-164]. Most
studies included patients with a mean age of 55–60 years. The evidence reviewed was consistent, indicating
that ER and IR formulations of anticholinergics offer clinically significant short-term improvement rates for OAB
compared to placebo. On balance, IR formulations tend to be associated with more side effects compared to
ER formulations [163].

A network meta-analysis of 128 RCTs comparing anticholinergics with placebo or with other anticholinergics
revealed that all anticholinergics, except imidafenacin, showed significant cure or improvement in OAB
symptoms in women and men [165].

Cure of UUI was deemed to be the most important outcome measure. Table 1 shows a summary of the findings
from systematic reviews [147]. In summary, every drug where cure outcomes for UUI were available showed
superiority compared to placebo, but the absolute size of effect was small. There is limited evidence that
patients who do not respond to a first-line anticholinergic treatment may respond to a higher dose or a different
anticholinergic agent [166, 167]. Risk of adverse events was often represented by trial-withdrawal because of
adverse events, although this does not reflect clinical practice.

The cure rates for darifenacin were not included in the United States (U.S.) Agency for Healthcare Research
and Quality (AHRQ) review. Continence rates were 29–33% for darifenacin compared to 17–18% for placebo
[147]. Transdermal oxybutynin has shown a significant improvement in the number of incontinence episodes
and micturitions per day vs. placebo and other oral formulations but cure of incontinence was not reported as
an outcome [147].

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 25


Oxybutynin topical gel was superior to placebo for improvement of UUI with a higher proportion of
participants being cured [147, 168].

Table 1: S
 ummary of cure rates and discontinuation rates of anticholinergic drugs from RCTs which
reported these outcomes [147]

Drug No. of studies n RR (95% CI) NNT (95% CI)


(of curing UI) (to achieve one cure of UI)
Cure of incontinence
Fesoterodine 2 2,465 1.3 (1.1–1.5) 8 (5–17)
Oxybutynin (includes IR) 4 992 1.7 (1.3–2.1) 9 (6–16)
Propiverine (includes IR) 2 691 1.4 (1.2–1.7) 6 (4–12)
Solifenacin 5 6,304 1.5 (1.4–1.6) 9 (6–17)
Tolterodine (includes IR) 4 3,404 1.2 (1.1–1.4) 12 (8–25)
Trospium (includes IR) 4 2,677 1.7 (1.5–2.0) 9 (7–12)
Discontinuation due to adverse events
RR (95% CI) NNT (95% CI)
(of discontinuation) (for one discontinuation)
Darifenacin 7 3,138 1.2 (0.8–1.8)
Fesoterodine 4 4,433 2.0 (1.3–3.1) 33 (18–102)
Oxybutynin (includes IR) 5 1,483 1.7 (1.1–2.5) 16 (8–86)
Propiverine (includes IR) 2 1,401 2.6 (1.4–5) 29 (16–77)
Solifenacin 7 9,080 1.3 (1.1–1.7) 78 (39–823)
Tolterodine (includes IR) 10 4,466 1.0 (0.6–1.7)
Trospium (includes IR) 6 3,936 1.5 (1.1–1.9) 56 (30–228)
CI = confidence interval; IR = immediate release; n = number of patients; NNT = number to treat;
UI = urinary incontinence; RR = relative risk.

4.1.4.2.1.1 Comparison of different anticholinergic agents


Head-to-head comparison trials of the efficacy and side effects of different anticholinergic agents are of interest
for decision making in practice.

A network meta-analysis revealed no clear best anticholinergic preparation for cure or improvement [165].
Darifenacin (40%), tolterodine IR and oxybutynin ER (13% each) appeared to score highest in indirect
comparisons. Fesoterodine and oxybutynin IR were more effective than both oxybutynin (transdermal) and
tolterodine ER. There were no clinically significant differences between anticholinergics for voiding and UI
outcomes.

Another network meta-analysis of 53 RCTs compared the efficacy and tolerability of solifenacin 5 mg with other
oral anticholinergics in the treatment of adults with OAB symptoms [169]. The analysis revealed that solifenacin
5 mg/day was significantly more effective than tolterodine 4 mg/day for reducing UUI episodes, but significantly
less effective than solifenacin 10 mg/day for micturition episodes. Solifenacin 5 mg/day showed significantly
lower risk of dry mouth compared with other anticholinergics. There were no significant differences for risk of
blurred vision or constipation.

It is notable that nearly all the primary studies in this category were industry-sponsored. Upward dose titration
is often included in the protocol for the experimental arm, but not for the comparator arm. In general, these
studies have been designed to achieve regulatory approval. They have short treatment durations (twelve
weeks) and a primary outcome of a change in OAB symptoms rather than a cure of, or an improvement in, UUI,
which were generally analysed as secondary outcomes. The clinical utility of these trials in real life practice is
debatable. Most trials were of low or moderate quality [161]. The 2012 AHRQ review included a specific section
addressing comparisons of anticholinergic drugs (Table 1).

No single anticholinergic agent improved QoL more than another [161]. Dry mouth is the most prevalent
adverse effect. Good evidence indicates that, in general, higher doses of any drug are likely to be associated
with higher rates of adverse events. Also, ER formulations of short-acting drugs and longer-acting drugs are
generally associated with lower rates of dry mouth than IR preparations [161, 170]. Oxybutynin IR showed
higher rates of dry mouth than tolterodine IR and trospium IR, but lower rates of dry mouth than darifenacin,

26 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021


15 mg daily [161, 170]. Overall, oxybutynin ER has higher rates of dry mouth than tolterodine ER, although
the incidence of moderate or severe dry mouth were similar. Transdermal oxybutynin had a lower rate of dry
mouth than oxybutynin IR and tolterodine ER, but had an overall higher rate of withdrawal due to adverse
skin reactions [161]. Solifenacin 10 mg daily, had higher rates of dry mouth than tolterodine ER [161].
Fesoterodine 8 mg daily, had a higher rate of dry mouth than tolterodine 4 mg daily [171-173]. In general,
similar discontinuation rates were observed, irrespective of differences in the occurrence of dry mouth (doses
have been given where the evidence relates to a specific dose level typically from trials with a dose escalation
element).

4.1.4.2.1.2 Anticholinergic drugs versus conservative treatment


The choice of drug vs. conservative treatment of OAB patients is an important question. More than 100 RCTs
and high-quality reviews are available [148, 161, 162, 174-176]. Most of these were independent studies.
A U.S. HTA [174] found that trials were of a low- or moderate-quality. The main focus of the review was
to compare the different drugs used to treat UUI. In one study, multi-component behavioural modification
produced significantly greater reductions in incontinence episodes compared to oxybutynin and higher patient
satisfaction for behavioural vs. drug treatment.

The combination of BT and solifenacin in female patients with OAB conferred no additional benefit in terms
of continence vs. solifenacin monotherapy [177]. A recent Cochrane review on the benefit of adding PFMT to
other active treatments of UI in women showed insufficient evidence of any benefit in adding PFMT to drug
treatment [178].

One RCT reported a similar improvement in subjective parameters with either transcutaneous electrical nerve
stimulation (T-PTNS) or oxybutynin [179]. One study compared tolterodine ER to transvaginal/anal ES in female
patients with OAB symptoms and/or UUI without differences in UI outcomes [180].

4.1.4.2.1.3 Anticholinergic drugs: adherence and persistence


Most studies on anticholinergic medication are short term (twelve weeks). Adherence in clinical trials is
considered to be much higher than in clinical practice [181]. This topic has been reviewed for the development
of a previous version of these Guidelines [182]. Two open-label extensions of RCTs of fesoterodine 8 mg
showed adherence rates at two years of 49–84% [183, 184]. The main drugs studied were oxybutynin and
tolterodine IR and ER. Non-persistence rates were high for tolterodine at twelve months, and particularly high
(68–95%) for oxybutynin.

Five articles reported ‘median days to discontinuation’, but follow-up periods varied from < 30 days up
to 50 days [185-189]. In a military health system where free medication was provided, the median time to
discontinuation extended to 273 days [186].

Data on adherence/persistence from open-label extension populations are questionable as it could be


argued that these patients are self-selected on the basis of their compliance. A Longitudinal Disease Analyser
database study has indicated an increasing discontinuation rate, following treatment with anticholinergics, from
74.8% at one year to 87% at three years [190].

Several of the RCTs tried to identify the factors associated with low/lower adherence or persistence of
anticholinergics. These were identified as:
• low level of efficacy (41.3%);
• adverse events (22.4%);
• cost (18.7%), although higher adherence rates were observed when drugs were provided at no cost to
the patient [186].

Other reasons for poor adherence included:


• IR vs. ER formulations;
• age (lower persistence among younger adults);
• unrealistic expectations of treatment;
• gender distribution (better adherence/persistence in female patients);
• ethnic group (African-Americans and other ethnic minorities are more likely to discontinue or switch
treatment).

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 27


4.1.4.2.1.4 Summary of evidence and recommendations for anticholinergic drugs

Summary of evidence LE
No anticholinergic drug is clearly superior to another for cure or improvement of OAB/UUI. 1a
Higher doses of anticholinergic drugs are more effective to improve OAB symptoms, but exhibit a 1a
higher risk of side effects.
Once daily (extended release) formulations are associated with lower rates of adverse events compared 1b
to immediate release preparations, although similar discontinuation rates are reported in clinical trials.
Dose escalation of anticholinergic drugs may be appropriate in selected patients to improve treatment 1b
effect although higher rates of adverse events can be expected.
Transdermal oxybutynin (patch) is associated with lower rates of dry mouth than oral anticholinergic 1b
drugs, but has a high rate of withdrawal due to skin reaction.
There is no consistent evidence to show superiority of drug therapy over conservative therapy for 1b
treatment of OAB.
Behavioural treatment may have higher patient satisfaction rates than drug treatment. 1b
There is insufficient evidence as to the benefit of adding PFMT to drug treatment for OAB. 1b
Adherence to anticholinergic treatment is low and decreases over time because of lack of efficacy, 2a
adverse events and/or cost.
Most patients will stop anticholinergic agents within the first three months. 2a

Recommendations Strength rating


Offer anticholineric drugs to adults with overactive bladder (OAB) who fail conservative Strong
treatment.
Consider extended release formulations of anticholinergic drugs, whenever possible. Strong
If an anticholinergic treatment proves ineffective, consider dose escalation or offering an Strong
alternative anticholinergic formulation, or mirabegron, or a combination.
Encourage early review (of efficacy and side effects) of patients on anticholinergic Strong
medication for OAB.

4.1.4.2.2 Beta-3 agonists


Mirabegron was the first clinically available beta-3 agonist. Beta-3 adrenoceptors are the predominant beta
receptors expressed on the smooth muscle cells of the detrusor and their stimulation is thought to induce
detrusor relaxation. Vibegron is another beta-3 agonist commercially available in some countries.

Mirabegron has undergone evaluation in industry-sponsored phase II and phase III trials [191-194]. Three
systematic reviews assessing the clinical effectiveness of mirabegron [191, 192, 195] reported that mirabegron
at doses of 25, 50 and 100 mg, results in significantly greater reduction in UI episodes, urgency episodes and
micturition frequency than placebo, with no difference in the rate of common adverse events [192]. The dry
rates in most of these trials are between 35–40% for placebo, and between 43 and 50% for mirabegron. In
all trials the statistically significant difference is consistent only for improvement but not for cure of UI. Similar
improvements in frequency of UI episodes and micturitions/24 hours was found whether or not patients had
previously tried anticholinergic agents. One systematic review showed that mirabegron is similarly efficacious
as most anticholinergics in reducing UUI episodes [196].

The most common treatment adverse events in the mirabegron groups were hypertension (7.3%),
nasopharyngitis (3.4%) and UTI (3%), with the overall rate similar to placebo [191, 194, 197].

In a twelve-month, active-controlled RCT of mirabegron 50/100 mg vs. tolterodine ER 4 mg, the improvement
in efficacy seen at twelve weeks was sustained at 12-month evaluation in all groups. The reported dry rates at
twelve months were 43%, 45% and 45% for mirabegron 50 mg, 100 mg and tolterodine 4 mg respectively [197].
Post-hoc analyses of RCTs showed that clinical improvement observed in parameters of OAB severity translates
into an improvement in HRQoL and efficacy is maintained in patients with more severe degree of UI [198, 199].

No risk of QTc prolongation on electrocardiogram [200] and no raised intraocular pressure [201] were observed
up to the 100 mg dose; however, patients with uncontrolled hypertension or cardiac arrhythmia were excluded
from these trials. There is no significant difference in rate of side effects at different doses of mirabegron
[197]. Patients on certain concurrent medications (e.g. metoprolol) should be counselled that, due to common
metabolism pathways, their medication dosage may need to be adjusted. In the case of patients taking

28 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021


metoprolol, blood pressure should be monitored after starting mirabegron and, if necessary, the metoprolol
dosing may need to be changed.

Equivalent adherence was observed for tolterodine and mirabegron at twelve months (5.5% and 3.6%),
although the incidence of dry mouth was significantly higher in the tolterodine group [197]. In mirabegron
treated patients, improvement in objective outcome measures correlates directly with clinically relevant PROMs
(Overactive Bladder questionnaire [OAB-q] and Patient Perception of Bladder Condition [PPBC]) [198, 202].
Data from a large Canadian Private Drug Plan database suggest a higher adherence rate for mirabegron
compared to anticholinergics [203].

An RCT in patients who had inadequate response to solifenacin monotherapy 5 mg, demonstrated that
combination treatment with mirabegron 50 mg had a higher chance of achieving clinically meaningful
improvement in UI as compared to dose escalation of solifenacin [204].

4.1.4.2.2.1 Summary of evidence and recommendation for mirabegron

Summary of evidence LE
Mirabegron is better than placebo and as efficacious as anticholinergics for improvement of OAB/UUI 1a
symptoms.
Adverse event rates with mirabegron are similar to placebo. 1a
Patients inadequately treated with solifenacin 5 mg may benefit more from the addition of mirabegron 1b
than dose escalation of solifenacin.

Recommendation Strength rating


Offer mirabegron as an alternative to anticholinergics to women with overactive bladder Strong
who fail conservative treatment.

4.1.4.2.3 Anticholinergics and beta-3 agonists: the elderly and cognition


Trials have been conducted in elderly people with OAB. Considerations in this patient group include the
multifactorial aetiology of OAB in the elderly, comorbidities such as cognitive impairment, the effect of
co-medications and the risk of adverse events. The effects of anticholinergic agents on cognition have been
studied in more detail.

Systematic reviews have included sections on the efficacy and safety of anticholinergics in elderly patients [147,
161]. A 2012 systematic review found inconclusive evidence as to the impact of anticholinergics on cognition [205].

Two recent longitudinal cohort studies in patients using drugs with anticholinergic effect showed a deterioration
in cognitive function, alteration in CNS metabolism and an association with brain atrophy [206, 207]. In general,
the long-term impact of anticholinergic agents specifically approved for OAB treatment on specific patient
cohorts is poorly understood [208-211].

• Oxybutynin: There is evidence that oxybutynin IR may cause/worsen cognitive dysfunction in adults [208,
210, 212, 213]. One RCT with oxybutynin topical gel focused on cognitive and psychomotor function
after one week of treatment showed no clinical meaningful effect on recent memory or other cognitive
functions in healthy old adults [213]. Another retrospective study did not show cognitive impairment
after 4 weeks of treatment with transdermal oxybutynin [210]. Recent evidence has emerged from a
prospective cohort study showing cumulative cognitive deterioration associated with prolonged use of
anticholinergic medication including oxybutynin [206]. More rapid functional deterioration might result
from the combined use of cholinesterase inhibitors and anticholinergic agents in elderly patients with
cognitive dysfunction [214].

• Solifenacin: One pooled analysis [215] has shown that solifenacin does not increase cognitive impairment
in the elderly. No age-related differences in the pharmacokinetics of solifenacin in different age groups was
found, although more frequent adverse events in subjects over 80 years of age were observed. No cognitive
effect on healthy elderly volunteers was shown [216]. In a sub-analysis of a large trial, solifenacin 5–10 mg
improved symptoms and QoL in people ≥ 75 years who had not responded to tolterodine [217]. In patients
with mild cognitive impairment, ≥ 65 years, solifenacin showed no difference in efficacy between age groups
and a lower incidence of most side effects compared to oxybutynin IR [213, 218].

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 29


• Tolterodine: No change in efficacy or side effects related to age has been reported, although a higher
discontinuation rate was found for both tolterodine and placebo in elderly patients [208]. Two RCTs in the
elderly found a similar efficacy and side effect profile to younger patients [219-222]. Post-hoc analysis has
shown little effect on cognition. One non-randomised comparison showed lower rates of depression in
elderly participants treated with tolterodine ER compared to oxybutynin IR [223].

• Darifenacin: Two RCTs in the elderly population (one in patients with UUI and the other in volunteers)
concluded that darifenacin was effective with no risk of cognitive change, measured as memory scanning
tests, compared to placebo [224, 225]. Another study on darifenacin and oxybutynin ER in elderly
subjects concluded that the two agents had a similar efficacy, but that cognitive function was more often
affected in the oxybutynin ER arm [210].

• Trospium chloride: Trospium does not appear to cross the blood brain barrier in healthy individuals due
to its molecular characteristics (quaternary amine structure and hydrophilic properties). Two studies
in healthy volunteers using electro-encephalography (EEG) showed no effect from trospium whilst
tolterodine caused occasional changes and oxybutynin caused consistent changes [226, 227]. No
evidence as to the comparative efficacy and side effect profiles of trospium in different age groups is
available. However, there is some evidence that trospium does not impair cognitive function [211, 228]
and that it is effective compared to placebo in the elderly [229].

• Fesoterodine: Pooled analyses of the RCTs of fesoterodine confirmed the efficacy of the 8 mg but not the
4 mg dose in over 75-year olds [183]. Adherence was lower in the over-75-year-old group but the effect
on mental status was not reported [173, 183, 230]. A more recent RCT showed efficacy of fesoterodine in
the vulnerable elderly with no differences in cognitive function at twelve weeks [231].

• Mirabegron: Analysis of pooled data from three RCTs showed efficacy and safety of mirabegron in elderly
patients [232].

4.1.4.2.3.1 Applicability of evidence to the general elderly population


It is not clear how much the data from pooled analyses and subgroup analyses from large RCTs can be
extrapolated to a general ageing population. Community-based studies of the prevalence of anticholinergic
side effects may be the most helpful [233]. When starting anticholinergics in elderly patients, mental function
should be assessed objectively and monitored [234]. No consensus exists as to the best mental function test to
detect changes in cognition [214, 235].

4.1.4.2.3.2 Anticholinergic burden


A number of medications have anticholinergic effects and if another anticholinergic drug is added the possible
greater cumulative effects on cognition should be considered. Lists of drugs with anticholinergic properties are
available from several sources [236].

No studies were identified specifically in older people with LUTS, but evidence was available from
observational cohort studies relating to the risk in a general population of older people.

Two systematic reviews of largely retrospective cohort studies showed a consistent association between long-
term anticholinergic use and cognitive dysfunction [237, 238]. Longitudinal studies in older people over 2−4
years have found increased rates of decline in cognitive function for patients on anticholinergics or drugs with
anticholinergic effects [206, 207, 239, 240]. It is unclear whether there is a direct correlation between cognitive
dysfunction caused by medication and the long-term risk of the development of dementia.

4.1.4.2.3.3 Summary of evidence and additional recommendations for use of anticholinergic drugs in the elderly

Summary of evidence LE
Anticholinergic drugs are effective in elderly patients suffering from OAB/UUI. 1b
Mirabegron has been shown to be efficacious and safe in elderly women suffering from OAB. 1b
In older women the cognitive impact of drugs which have anticholinergic effects is cumulative and 2
increases with length of exposure.
Oxybutynin may worsen cognitive function in elderly women. 2
Darifenacin, fesoterodine, solifenacin and trospium have not been shown to cause cognitive 1b
dysfunction in elderly women in short-term studies.

30 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021


Recommendations Strength rating
Long-term anticholinergic treatment should be used with caution in elderly women, especially Strong
those who are at risk of, or have pre-existing cognitive dysfunction.
Assess anticholinergic burden and associated co-morbidities in patients being considered Weak
for anticholinergic therapy for overactive bladder syndrome.

4.1.4.2.4 Oestrogens
Oestrogenic drugs including conjugated equine oestrogens, oestradiol, tibolone and raloxifene, are used as
hormone replacement therapy (HRT) for women with natural or therapeutic menopause.

Oestrogen treatment for UI has been tested using oral, transdermal and vaginal routes of administration.
Vaginal (local) treatment is primarily used to treat symptoms of vaginal atrophy in post-menopausal women.
Available evidence related mainly to SUI and although some reviews include participants with UUI, it is difficult
to generalise the results to women with predominantly OAB/UUI.

The association of LUTS with Genitourinary Syndrome of Menopause (GSM) should be considered [241]. GSM
is a relatively new term that describes various menopausal symptoms and signs associated with physical
changes of the vulva, vagina, and LUT. These include mucosal pallor/erythema, loss of vaginal rugae, tissue
fragility/fissures, vaginal petechiae, urethral mucosal prolapse, introital retraction and vaginal dryness. There is
evidence from a systematic review to suggest benefit from vaginal oestrogen therapy in GSM [242]. All vaginal
oestrogens demonstrated superiority in objective endpoints and subjective endpoints of GSM compared to
placebo. Only some trials demonstrated superiority vs. placebo in urogenital symptoms (UI, recurrent UTI,
urgency, frequency). No significant difference was observed between various dosages and dosage forms of
vaginal oestrogen products. Vaginal oestrogen showed superiority over vaginal lubricants and moisturizers for
the improvement of objective clinical endpoints of vulvovaginal atrophy but not for subjective endpoints [242].

Available evidence suggests that vaginal oestrogen treatment with oestradiol and oestriol is not associated
with the increased risk of thromboembolism, endometrial hypertrophy, and breast cancer seen with systemic
administration [243-245].

4.1.4.2.4.1 Summary of evidence and recommendation for oestrogen therapy

Summary of evidence LE
Vaginal oestrogen therapy may improve symptoms associated with GSM, of which OAB may be a 1a
component.

Recommendation Strength rating


Offer vaginal oestrogen therapy to women with lower urinary tract symptoms and Weak
associated symptoms of genito-urinary syndrome of menopause.

4.1.4.3 Surgical management


4.1.4.3.1 Bladder wall injection of botulinum toxin A
Onabotulinum toxin A (onabotA; BOTOX®) 100 U is licenced in Europe to treat OAB with persistent or refractory
UUI in adults of both genders [246, 247]. Surgeons should be aware that other doses of onabotA and other
formulations of botulinum toxin A, abobotulinum toxin A and incobotulinum toxin A, are not licensed for use
in OAB/UUI. Doses for onabotA are not transposable to the other brands of botulinum toxin A. The continued
efficacy of repeat injections is usual, but discontinuation rates may be high [248, 249]. The most important
adverse events related to onabotA 100U injection detected in the regulatory trials were UTI and an increase in
PVR that may require CISC [250].

Following a dose ranging study in which the 100 U of onabotA was established as the ideal dose, a phase
III trial randomised (1:1) the same group of 557 OAB-wet patients whose symptoms were not adequately
managed with anticholinergics to receive bladder wall injections of onabotA (100 U) or saline. At baseline, the
population had on average more than five episodes of UUI, around twelve micturitions per day and a small
PVR. At week twelve, in patients treated with onabotA, UUI episodes/day were halved and the number of
micturitions/day reduced by more than two. A total of 22.9% of the patients in the onabotA arm were fully dry,
against 6.5% in the saline arm [250]. Rates of urinary retention are not reported in systematic reviews, and a
Cochrane review reported no significant difference in PVR between the onabotA and placebo groups [251].

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 31


Quality of life was substantially improved in the onabotA arm, as shown by the > 2.5 times improvement in
I-QOL scores compared to baseline. Cohort studies have shown the effectiveness of bladder wall injections of
onabotA in the elderly and frail elderly [252], though the success rate might be lower and the PVR (> 150 mL)
higher in this group.

The median time to request re-treatment in the pooled analysis of the two RCTs was 24 weeks [247, 250].
Follow-up over 3.5 years showed consistent or increasing duration of effect for each subsequent treatment,
with a median of 7.5 months. Considerable differences between patients have been observed on secondary
analysis [253].

A recent RCT compared onabotA injection 100 U to solifenacin (with dose escalation or switch to trospium
possible in the solifenacin group) and showed similar rates of improvement in UUI over the course of six
months [254]. However, patients receiving onabotA were not only more likely to have cure of UUI (27% vs.
13%, p = 0.003), but also had higher rates of urinary retention during the initial two months (5% vs. 0%) and
of UTIs (33% vs. 13%). Patients taking anticholinergics were more likely to have dry mouth. These results are
further strengthened by a 2017 systematic review and network meta-analysis of onabotulinum toxin A vs.
oral therapies (anticholinergics and mirabegron) for OAB at 12 weeks [255]. This review reported that patients
receiving onabotulinum toxin had the greatest reduction in UUI episodes, urgency episodes, micturition
frequency and the highest odds of achieving dryness as well as ≥ 50% reduction from baseline UI episodes/
day (type not specified). However, adverse events were not reported in this network meta-analysis.

Identification of DO in urodynamics does not appear to influence the outcome of onabotulinum toxin A
injections in patients with UUI [85].

4.1.4.3.1.1 Summary of evidence and recommendations for bladder wall injection of botulinum toxin A

Summary of evidence LE
A single treatment session of onabotulinum toxin A (100 U) injected in the bladder wall is more 1a
effective than placebo at curing and improving UUI/OAB symptoms and QoL.
There is no evidence that repeated injections of onabotulinum toxin A have reduced efficacy but 2a
discontinuation rates are high.
There is a risk of increased PVR and UTI with onabotulinum toxin A injections. 2
The risk of bacteriuria after onabotulinum toxin A (100 U) injection is high but the clinical significance 1b
of this remains uncertain.
Onabotulinum toxin A (100 U) is superior to anticholinergics and mirabegron for cure of UUI and 1a
improvement of symptoms of OAB at twelve weeks.

Recommendations Strength rating


Offer bladder wall injections of onabotulinum toxin A (100 U) to patients with overactive Strong
bladder/urgency urinary incontinence refractory to conservative therapy (such as pelvic floor
muscle training and/or drug treatment).
Warn patients of the limited duration of response, risk of urinary tract infection and the Strong
possible prolonged need for clean intermittent self-catheterisation (ensure that they are
willing and able to do so).

4.1.4.3.2 Sacral nerve stimulation


Sacral nerve stimulation involves placing electrodes adjacent to the sacral nerve roots and delivering an electric
current to the area via an attached battery implanted in the buttock which delivers low-amplitude stimulation
resulting in modulation of neural activity and stabilisation of bladder electrical activity through a mechanism
that is, as yet, not fully understood. In most centres, test stimulation with a temporary or permanent electrode
will be performed to assess response, before undertaking permanent stimulator implantation.

All randomised studies suffer from the limitation that patients cannot be blinded to the treatment allocation since
all recruited subjects had to respond to a test phase before randomisation. A Cochrane review of the literature
until March 2008 [256] identified three RCTs that investigated sacral nerve stimulation in patients with refractory
UUI. The majority of included studies compared a strategy of immediate implantation vs. delayed implantation.

32 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021


One study compared implantation to controls who stayed on medical treatment and received delayed
implantation at six months. Fifty percent of the immediately implanted group had > 90% improvement in UUI
at six months compared to 1.6% of the control group [257]. The effect on generic QoL measured by the SF-36,
was unclear as it differed between the groups in only one of the eight dimensions. The other RCT achieved
similar results, although these patients had already been included in the first report [258].

The results of seventeen case series of patients with UUI, who were treated early with SNS, were reviewed
[259]. After a follow-up duration of between one and three years, approximately 50% of patients with
UUI demonstrated > 90% reduction in UI, 25% demonstrated 50–90% improvement, and another 25%
demonstrated < 50% improvement. Two case series describing the outcome of SNS, with a mean or median
follow-up of at least four years [260, 261] reported continued success (> 50% improvement of original
symptoms) in patients available for follow-up. Cure rates for UUI were 15% [261].

A more recent RCT comparing a strategy of onabotulinum toxinA injection (200 IU), repeated as required,
against a strategy of test and, if indicated, subsequent permanent SNS showed lower cure rates with SNS at
six months: 20% in the onabotulinumtoxinA group and 4% in the SNS group had complete resolution of UUI
(p < 0.001) [262]. Forty-six per cent in the onabotulinumtoxinA group and 26% in the SNS group had at least
a 75% reduction in the number of episodes of UUI (p < 0.001). This 4% cure rate is also lower than the six
month cure rate in another RCT of SNS vs. standard medical therapy which reported a 39% continence rate in
the SNS group at six months; however, the mean (SD) baseline leaks per day (2.4 [± 1.7]) for the SNS group in
this study were lower , reflecting a less severely affected population [263]. Two-year follow-up data from 87%
of participants in this trial suggests no significant deterioration in treatment outcomes over two years, although
satisfaction rates and treatment endorsement remain higher with onabotulinum toxin. Interestingly, the rates of
complete resolution of UI (5% for both) as well as ≥ 75% reduction in UUI episodes (22% onabotulinum toxin
vs. 21% SNM) were equivalent at the 2-year mark [264]. Sacral nerve stimulation revision and removal occurred
in 3% and 9% of this cohort, respectively. Some of these differences in outcome could potentially be explained
by performance bias particularly the difference in type of permanent lead used and potential learning curve
effects within the SNS cohort as well as the use of a relatively high (200 U) dose of onabotulinum toxin A.

A 2018 review of studies including SNS with at least 6 months follow-up reported dry rates of between 43 and
56% [265]. Adverse events occurred in 50% of implanted cases, with surgical revision necessary in 33–41%
[261, 262]. In a sub-analysis of the RCT, the outcomes of UUI patients, with or without pre-implant DO, were
compared. Similar success rates were found in patients with or without urodynamic DO [266].

4.1.4.3.2.1 Summary of evidence and recommendation for sacral nerve stimulation

Summary of evidence LE
Sacral nerve stimulation is more effective than continuation of failed conservative treatment for OAB/ 1b
UUI, but no sham controls have been used.
Sacral nerve stimulation is not more effective than onabotulinumA toxin 200 U injection at 24 months. 1b
In patients who have been implanted 50% improvement of UUI is maintained in at least 50% of 3
patients and 15% may remain cured at four years.
The use of tined, permanent electrodes in a staged approach results in more patients receiving the 4
final implant than occurs with temporary test stimulation.

Recommendation Strength rating


Offer sacral nerve stimulation to patients who have overactive bladder/urgency urinary Strong
incontinence refractory to anticholinergic therapy.

4.1.4.3.3 Cystoplasty/urinary diversion


4.1.4.3.3.1 Augmentation cystoplasty
In augmentation cystoplasty (also known as clam cystoplasty), a detubularised segment of bowel is inserted
into the bivalved bladder wall. The distal ileum is the bowel segment most often used but any bowel segment
can be utilised if it has the appropriate mesenteric length. Most of the evidence pertaining to cystoplasty
comes from patients with neuropathic bladder dysfunction. One study did not find any difference between
bivalving the bladder in the sagittal or in the coronal plane [267, 268]. The procedure can be done, with equal
success by open or robotic techniques, although the robotic consumes considerably more operative time [269].

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 33


There are no RCTs comparing bladder augmentation to other treatments for patients with OAB/UUI. Most often,
bladder augmentation is used to correct neurogenic DO, small capacity or low-compliant bladders caused by
fibrosis, chronic infection such as tuberculosis, radiation or chronic inflammation from interstitial cystitis.

The largest case series of bladder augmentation in a mixed population of idiopathic and neurogenic UUI
included 51 women [270]. At an average follow-up of 74.5 months, only 53% were continent and satisfied with
the surgery, whereas 25% had occasional leaks and 18% continued to have disabling UUI. It seems that the
results for patients with idiopathic DO (58%) appeared to be less satisfactory than for patients with neurogenic
UUI (90%). Malignant transformation was not reported in this series; however, it has been documented in other
series and a systematic review [271-273]. Less than 60 cases have been reported worldwide, and almost all are
exclusively beyond 10 years after the original cystoplasty surgery [274].

Adverse effects were common and have been summarised in a review over five to seventeen years of more
than 267 cases, 61 of whom had non-neurogenic UUI [275]. In addition, many patients may require CISC to
obtain adequate bladder emptying (Table 2). It is unclear if mucolytic agents will effectively reduce mucus
accumulation. The only RCT that was identified comparing various mucolytic agents did not find significant
benefits with the use of N-acetylcysteine, aspirin, or ranitidine. In one small study (n = 40), the use of
subcutaneous octreotide immediately before, and for 15 days after surgery was reported to yield significant
reductions in mucus production, the need for bladder irrigation to clear blockages, and the mean duration of
hospital stay [276]. Before cystoplasty all potential complications should be outlined and both before and after
surgery patients should be well supported by stoma/continence nurses.

Depending on the relative costs of Onabotulinum Toxin A and augmentation cystoplasty, the latter can be
cost effective within five years if the complication rate is low and duration of effect of Onabotulinum Toxin A
< 5 months [277].

Table 2: Complications of bladder augmentation

Short-term complications Affected patients (%)


Bowel obstruction 2
Infection 1.5
Thromboembolism 1
Bleeding 0.75
Fistula 0.4
Long-term complications Affected patients (%)
Clean intermittent self-catheterisation 38
Urinary tract infection/bacteriuria 70% asymptomatic
20% symptomatic
Urinary tract stones 13
Metabolic disturbance 16
Deterioration in renal function 2
Bladder perforation 0.75
Change in bowel symptoms 25

4.1.4.3.3.2 Detrusor myectomy (bladder auto-augmentation)


Detrusor myectomy aims to increase bladder capacity and reduce storage pressures by incising or excising
a portion of the detrusor muscle, to create a bladder mucosal ‘bulge’ or pseudo-diverticulum. It was initially
described as an alternative to bladder augmentation in children [278].

Two case series in adult patients with idiopathic and neurogenic bladder dysfunction, demonstrated poor long-
term results caused by fibrosis of the pseudo-diverticulum [279, 280]. This technique is rarely, if ever, used
nowadays.

4.1.4.3.3.3 Urinary diversion


Urinary diversion remains a reconstructive option for patients with intractable UI after multiple pelvic
procedures, radiotherapy or pelvic pathology leading to irreversible sphincteric incompetence or fistula
formation. These patients may be offered irreversible urinary diversion surgery. Options include ileal conduit
urinary diversion, orthotopic neobladder and heterotopic neobladder with Mitrofanoff continent catheterisable
conduit. There is insufficient evidence to comment on which procedure leads to the most improved QoL.

34 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021


A small study comparing ileal with colonic conduits concluded that there were no differences in the relative
risks of UUT infection and uretero-intestinal stenosis. However, there are no studies that have specifically
examined these techniques in the treatment of intractable OAB/UUI [267]. Therefore, careful consideration on
which operation is undertaken will depend on thorough pre-operative counselling, access to stoma/continence
nurses as well as patient factors to allow for fully informed patient choice.

4.1.4.3.3.4 Summary of evidence and recommendations for cystoplasty/urinary diversion

Summary of evidence LE
There is limited evidence of the effectiveness of augmentation cystoplasty and urinary diversion in 3
treatment of idiopathic OAB.
Augmentation cystoplasty and urinary diversion are associated with high risks of short- and long-term 3
severe complications.
The need to perform CISC following augmentation cystoplasty is very common. 3
There is no evidence comparing the efficacy or adverse effects of augmentation cystoplasty to urinary 3
diversion.
Detrusor myectomy is ineffective in adults with UUI. 3

Recommendations Strength rating


Offer augmentation cystoplasty to patients with overactive bladder (OAB)/urgency urinary Weak
incontinence (UUI) who have failed all other treatment options and have been warned about
the possible small risk of malignancy.
Inform patients undergoing augmentation cystoplasty of the high risk of having to perform Strong
clean intermittent self-catheterisation (ensure they are willing and able to do so) and that
they need life-long surveillance.
Do not offer detrusor myectomy as a treatment for UUI. Weak
Only offer urinary diversion to patients who have failed less invasive therapies for the Weak
treatment of OAB/UUI, who will accept a stoma and have been warned about the possible
small risk of malignancy.

4.1.5 Follow-up
Follow-up for women with OAB is guided by the type of treatment instituted and local service capacity.
Standardisation of follow-up pathways can therefore be difficult. Here we provide recommendations based on
best practice and standards from clinical trials.

4.1.5.1 Recommendations for follow-up of patients with overactive bladder

Recommendations Strength rating


Offer early follow up to women who have been commenced on anticholinergic or beta-3 Strong
agonist therapy.
Offer repeat injections of onabotulinum toxin, as required, to women in whom it has been Strong
effective (refer to the manufacturers guidance regarding the minimum timeframe for repeat
injections).
Offer life-long surveillance to women who have a sacral nerve stimulation implant to monitor Strong
for lead displacement, malfunction and battery wear.
Offer cystoscopic surveillance to women with an augmentation cystoplasty due to the small Weak
risk of malignancy.

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 35


4.2 Stress Urinary Incontinence
4.2.1 Epidemiology, aetiology, pathophysiology
Stress urinary incontinence, defined as ‘involuntary loss of urine on effort or physical exertion’, is a significant
health problem worldwide with social and economic impact on women and society. It is estimated that the
number of women in the U.S. with UI will increase from 18.3 million in 2010 to 28.4 million in 2050 [281]. The
prevalence of SUI appears to peak between the ages of 45–59 years [282].

Data regarding the association of UI with ethnicity are conflicting. In several studies, SUI is more common in
white women than in women of African-American or Asian-American origin [283, 284]. Other factors positively
associated with SUI include parity, obesity, previous hysterectomy or pelvic surgery, diabetes mellitus [285] and
pulmonary disease [286]. Physical activity level is another important factor which also positively correlates with
severity of the problem [287].

Two common, often overlapping, mechanisms for SUI have been described: 1) urethral hypermobility resulting
from loss of support of the bladder neck and urethra and 2) weakness of the urinary sphincter itself (intrinsic
sphincter deficiency) which can result from trauma, radiotherapy, previous pelvic or urogynaecological surgery,
neurological disease or ageing.

The mechanism behind urethral hypermobility as a cause of SUI is based on the “vaginal hammock”
hypothesis [288]. The endopelvic fascia, that is attached to the upper (abdominal) side of the PFMs, links
the muscles to the vagina and represents the ‘hammock’ which can compress the urethra during both rest
and activity. This compression, combined with ‘intrinsic’ urethral sphincter pressure, supports and maintains
the urethra in the correct and closed position preventing involuntary loss of urine, despite any increases
in intravesical pressure. Damage to the supporting tissues (particularly the arcus tendinous fasciae pelvis,
the central part of the fascia) can result in hypermobility of the urethra. Consequently, rather than being
compressed at times of increased intra-abdominal pressure, the urethra moves caudally funnelling the bladder
neck, and is no longer compressed, resulting in SUI [288, 289]. In general, almost all treatments are used
for both subtypes of SUI, but most treatments are more successful in patients with some degree of urethral
hypermobility than for isolated intrinsic weakness of the urinary sphincter [290].

4.2.2 Classification
Patients with SUI can be classified as ‘uncomplicated’ and ‘complicated’ [291]. The Panel reached consensus
on the definition to be used throughout this Guideline document:
• Women with uncomplicated SUI: no history of prior surgery for SUI, no prior extensive pelvic surgery, no
prior pelvic radiation treatment, no neurogenic LUT dysfunction, no bothersome genitourinary prolapse,
absence of voiding symptoms, and no medical conditions that affect the LUT. In cases where additional
significant storage symptoms, especially OAB, are present, consider a possible diagnosis of MUI (see
Section 4.3).
• Women with complicated SUI: previous surgery for incontinence or previous extensive pelvic surgery, a
history of pelvic irradiation, the presence of anterior or apical POP, the presence of voiding symptoms or
the presence of neurogenic LUT dysfunction, and with significant OAB/UUI. Neurogenic LUT dysfunction
is reviewed in the EAU Guidelines on Neuro-Urology and will not be considered further in this guideline
[9]. The treatment of LUTS associated with genitourinary prolapse has been included in this Guideline
(see Section 4.7).

4.2.3 Diagnostic evaluation


4.2.3.1 History and physical examination
There is universal agreement that taking a history should be the first step in the assessment of anyone with
UI. When the history categorises UI as probable SUI the presence of complicated or uncomplicated SUI can
also be determined. Those patients who require rapid referral to an appropriate specialist can also often be
identified from the clinical history.

There is little evidence from clinical trials that carrying out a clinical examination improves clinical outcomes,
but there is widespread consensus that it remains an essential part of the assessment of women with SUI. It
should include abdominal examination, vaginal examination and a careful assessment of any associated pelvic
POP, examination of the perineum and evaluation of PFM strength, as well as a neuro-urological examination.
An attempt to reproduce the SUI should be made. A standing cough test has a higher sensitivity for the
diagnosis of SUI compared to a supine cough test [292]. Despite this, the ICS has proposed a standardisation
of the female cough stress test that includes a supine/lithotomy position with 200-400 mL of fluid in the bladder
and 1–4 coughs [293].

36 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021


4.2.3.1.1 Summary of evidence and recommendation for history and physical examination for SUI

Summary of evidence LE
A standing cough stress test is more sensitive than a supine test. 1b

Recommendation Strength rating


Take a full clinical history and perform a thorough physical examination in all women Strong
presenting with stress urinary incontinence.

4.2.3.2 Patient questionnaires


Although many studies have investigated the validity and reliability of urinary symptom questionnaires and
PROMs, most of these studies did not include homogenous populations of adult women diagnosed with SUI.
This limits the extent to which results and conclusions from these studies can be applied in women with SUI.
Some questionnaires are used for prevalence studies, others are responsive to change and may be used to
measure outcomes, though evidence on their sensitivity is inconsistent [24, 25]. No evidence was found to
indicate whether use of QoL or condition-specific questionnaires have an impact on the outcome of treatment.
To date, there is no one questionnaire that fulfils all requirements for the assessment of women with SUI.

4.2.3.2.1 Summary of evidence and recommendation for patient questionnaires

Summary of evidence LE
Validated condition-specific symptom scores assist in the screening for, and categorisation of UI. 3
Validated symptom scores measure the severity and bother of SUI. 3
Both condition-specific and general health status questionnaires measure current health status, and 3
are responsive to change following treatment.

Recommendation Strength rating


Use a validated and appropriate questionnaire as part of the standardised assessment of Strong
patients with stress urinary incontinence.

4.2.3.3 Post-void residual volume


It is important to evaluate PVR in patients with SUI, in particular in patients who also have voiding symptoms
or POP. The prevalence of a significant PVR in patients with SUI is uncertain, partly because of the lack of a
standard definition of an abnormal PVR volume. Most studies investigating PVR have not included patients with
SUI. In general, the data on PVR can only be applied with caution to adults with non-neurogenic SUI.

In a cohort study of over 900 women with SUI, there was good correlation between PVR estimated by US and
measured by catheterisation. The mean PVR was 39 mL measured by catheterisation and 63 mL estimated by
US, with only 16% of women having a PVR > 100 mL [57].

4.2.3.3.1 Summary of evidence and recommendations for post-void residual volume

Summary of evidence LE
The majority of women with SUI will not have a significant PVR. 3
There is good correlation between PVR estimated using US and measured via catheterisation in 3
women with SUI.

Recommendations Strength rating


Measure post-void residual (PVR) volume, particularly when assessing patients with voiding Strong
symptoms or complicated stress urinary incontinence (SUI).
When measuring PVR, use ultrasound in preference to catheterisation. Strong
Monitor PVR in patients scheduled for treatment which may cause or worsen voiding Strong
dysfunction, including surgery for SUI.

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 37


4.2.3.4 Urodynamics
Urodynamic testing is widely used as an adjunct to clinical diagnosis, based on the assumption that it may help
to provide or confirm diagnosis. The role of urodynamics in SUI evaluation remains poorly defined and is still
under debate.

Invasive urodynamic tests are often performed prior to surgical treatment of SUI. Clinical diagnosis of
incontinence and cystometric findings often do not correlate [67, 68]. The diagnostic accuracy of urethral
pressure profilometry [69] and VLPP measurement in SUI is generally poor [294]. Measurement of MUCP
correlates, although weakly, with incontinence severity [69] and there is conflicting evidence about its
reproducibility [64, 65]. One method of recording MUCP cannot be compared meaningfully to another [66].
Valsalva leak point pressures are not standardised and there is minimal evidence about reproducibility. Valsalva
leak point pressure did not reliably assess incontinence severity in a cohort of women selected for surgical
treatment of SUI [70]. The predictive value of the tests, regarding the outcome of treatment remains unclear.

A Cochrane systematic review including seven RCTs showed that the use of urodynamic tests increased the
likelihood of avoiding surgery for SUI. However, there was no evidence that this influence on decision making
altered the clinical outcome of treatment within trial populations [112].

A high-quality RCT (n = 630) compared office evaluation alone to office evaluation and urodynamics in women
with clinical demonstrable SUI about to undergo surgery for SUI. Whilst urodynamics changed the clinical
diagnosis in 56% of women [295], there was no difference in levels of SUI or any secondary outcome at twelve
months follow-up after SUI surgery [80]. Another similar study also found that the omission of urodynamics
was not inferior in the pre-operative work up of SUI [296]. Patients in whom urodynamics were discordant
with clinical assessment (n = 109) were randomly allocated to receive either immediate surgery or individually
tailored therapy based on the urodynamic findings. In this trial, performing immediate surgery, irrespective of the
result of urodynamics, did not result in inferior outcomes [297]. An RCT, in which 145 women were randomised to
retropubic or transobturator MUS, concluded that when patients were stratified according to pre-operative VLPP
(≤ or > 60 cm H2O), it was not linked to outcome after both synthetic MUS procedures [298].

In another study conflicting evidence was reported. A VLPP or maximum urethral closure pressure in the lowest
quartile was predictive in terms of failure at twelve months [83].

The Panel recognise that it may be valuable to use urodynamic test results to help select the optimum
surgical procedure, but the evidence outlined above would suggest that performing urodynamics in patients
with uncomplicated SUI, which can be diagnosed based on a detailed clinical history and demonstrated at
examination, is not necessary. The role of urodynamics in complicated SUI is still under debate [81, 299].
However in cases of SUI with associated storage symptoms, cases in which the type of incontinence is
unclear, cases where voiding dysfunction is suspected, cases with associated POP or those with a previous
history of SUI surgery, the Panel consensus is that urodynamics should be carefully considered. This is in line
with other guideline documents in this topic area [67].

4.2.3.4.1 Summary of evidence and recommendations for urodynamics

Summary of evidence LE
Pre-operative urodynamics in women with uncomplicated, clinically demonstrable, SUI does not 1b
improve the outcome of surgery for SUI.
There is no consistent correlation between the result of urethral function tests and subsequent 3
success or failure of SUI surgery.
There is no consistent evidence that pre-operative DO is associated with surgical failure of MUS in 3
women.

Recommendations Strength rating


Do not routinely carry out urodynamics when offering treatment for uncomplicated stress Strong
urinary incontinence (SUI).
Perform pre-operative urodynamics in cases of SUI with associated storage symptoms, Weak
cases in which the type of incontinence is unclear, cases where voiding dysfunction is
suspected, cases with associated pelvic organ prolapse or those with a previous history of
SUI surgery.

38 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021


Perform urodynamics if the findings may change the choice of invasive treatment. Weak
Do not use urethral pressure profilometry or leak point pressure to grade severity of Strong
incontinence as they are primarily tests of urethral function.

4.2.3.5 Pad testing


Measurement of urine loss using an absorbent pad worn over a set period of time or during a protocol of
physical exercise can be used to quantify the presence and severity of SUI, as well as a patient’s response to
treatment.

The clinical utility of pad tests for people with UI has been assessed in two systematic reviews [94, 95]. A
one-hour pad test using a standardised exercise protocol and a diagnostic threshold of 1.4 g shows good
specificity but low sensitivity for the diagnosis of SUI and MUI. A 24-hour pad test using a threshold of 4.4 g
is more reproducible but is difficult to standardise with variation according to activity level [96]. A pad test with
a specific short graded exercise protocol also has diagnostic value but a negative test should be repeated or
the degree of provocation increased [97]. The usefulness of pad tests in quantifying severity and predicting
outcome of treatment is uncertain [94, 99]. Pad testing is responsive to change following successful treatment
[100]. Pad testing using a standardised bladder volume (50% of cystometric capacity) has been suggested to
allow for a more reliable assessment of UI in a small study including 25 women [101]. There is no evidence that
one type of pad test is superior to another.

4.2.3.5.1 Summary of evidence and recommendations for pad testing

Summary of evidence LE
A pad test can diagnose UI accurately, but cannot determine the aetiology. 2
Standardisation of bladder volume and degree of provocation improves reproducibility. 2
Twenty-four hours is sufficient duration for home-based testing balancing diagnostic accuracy and 2
adherence.
Change in leaked urine volume on standardised pad tests can be used to measure treatment 2
outcome.

Recommendations Strength rating


Use a pad test of standardised duration and activity protocol. Strong
Use a standardised pad test when quantification of urinary incontinence is required, Weak
especially to assess response to treatment.

4.2.3.6 Imaging
The role of imaging in SUI patients is limited. Many studies have evaluated imaging of bladder neck mobility by
US and MRI, and concluded that SUI cannot be identified by a particular pattern of urethro-vesical movement
[300]. In addition, the generalised increase in urethral mobility after childbirth does not appear to be associated
with de novo SUI [301]. Studies have assessed the use of imaging to investigate the mechanism of action of
MUS inserted for SUI. One study suggested that MUS placement decreased mobility of the mid-urethra but
not mobility of the bladder neck [108]. Following MUS surgery, a wider gap between symphysis and sling
(assessed by imaging) has been shown to correlate with a lower chance of cure of SUI [109]. One study of 72
women post-synthetic sub-urethral MUS surgery has investigated the usefulness of translabial US to assess
tape functionality. In this study different parameters were measured (distance tape to urethra, position and
shape during Valsalva, etc.) and concluded that tape position relative to the patient’s urethra seems to play a
role in treatment outcome [302]. The general role of US in the evaluation and follow-up of women with SUI is
unclear, future research is needed to establish its place in the clinical pathway.

Several imaging studies have investigated the relationship between sphincter volume and function in women
[303] and between sphincter volume and outcome from surgery in women [304]. However, no imaging test has
been shown to predict the outcome of treatment for SUI. Imaging of the pelvic floor can identify levator ani
detachment and hiatus size, although there is little evidence of a relationship to clinical benefit after treatment
of SUI.

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 39


4.2.3.6.1 Summary of evidence and recommendation for imaging

Summary of evidence LE
Imaging can reliably be used to measure bladder neck and urethral mobility, although there is no 2b
evidence of clinical benefit for patients with UI.

Recommendation Strength rating


Do not carry out imaging of the upper or lower urinary tract as part of the routine Strong
assessment of stress urinary incontinence.

4.2.4 Disease management


4.2.4.1 Conservative management
4.2.4.1.1 Obesity and weight loss
Being overweight or obese has been identified as a risk factor for LUTS and UI in many epidemiological studies
[137, 138]. There is evidence that the prevalence of both UUI and SUI increases proportionately with rising BMI
[305]. The proportion of patients who undergo surgery for incontinence who are overweight or obese is higher
than that of the general population [139].

All the available evidence relates to women. Three systematic reviews concluded that weight loss was
beneficial in improving UI [137, 138, 306]. Five further RCTs reported a similar beneficial effect on incontinence
following surgical weight reduction programmes [307-311]. Two large studies in women with diabetes mellitus,
for whom weight loss was the main lifestyle intervention, showed UI did not improve but there was a lower
subsequent incidence of UI among those who lost weight [307, 312]. There have been other cohort studies and
case-control studies suggesting similar effects, including surgery for the morbidly obese [313-317].

In a prospective study in 160 consecutive women who underwent bariatric surgery, surgically-induced weight loss
was associated with a significant improvement in pelvic floor disorders, including UI [318]. Similar results reported
by prospective single-centre studies investigating the effect of bariatric surgery induced weight loss, revealed that
bariatric surgery was associated with substantially reduced UI at 11 months and 3 years [319, 320].

4.2.4.1.1.1 Summary of evidence and recommendation for obesity and weight loss

Summary of evidence LE
Obesity is a risk factor for LUTS and UI in women. 3
Non-surgical weight loss improves UI in overweight and obese women. 1a
Surgical weight loss improves UI in obese women. 1b

Recommendation Strength rating


Encourage overweight and obese women with lower urinary tract symptoms/stress urinary Strong
incontinence to lose weight and maintain weight loss.

4.2.4.1.2 Urinary containment


The evidence for urinary containment derives from the same literature as for containment in OAB-wet. The
readers are therefore referred to Section 4.1.4.1.3.

4.2.4.1.2.1 Summary of evidence and recommendations for urinary containment

Summary of evidence LE
Pads are effective in containing urine. 1b

Recommendations Strength rating


Ensure that women with stress urinary incontinence (SUI) and/or their carers are informed Strong
regarding available treatment options before deciding on urinary containment alone.
Offer incontinence pads and/or containment devices for management of SUI, either for Strong
temporary symptom control or where other treatments are not feasible.

40 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021


4.2.4.1.3 Pelvic floor muscle training
Pelvic floor muscle training is used to improve function of the pelvic floor, thus improving urethral stability.
An immediate effect of a single PFM contraction is narrowing of the levator hiatus area, increase of urethral
closure pressure and lift of the bladder and rectum preventing the occurrence of UI [321-323]. In an RCT
comparing intensive PFMT over a six-month period with no treatment, authors reported increased muscle
strength and endurance, narrowing of the levator hiatus, reduced PFM length, increased muscle volume and
lift of the bladder neck and rectal ampulla [324]. Pelvic floor muscle training may be used to prevent UI, e.g.
in childbearing women before birth, or as part of a planned recovery programme after childbirth. Most often,
PFMT is used to treat existing SUI; sometimes in combination with observation and/or palpation of the muscle
contraction by the therapist or biofeedback (by use of an apparatus measuring the contraction either by
electromyography (EMG), manometry, dynamometry, US or MRI). Electrical stimulation and vaginal cones are
also used in treatment of SUI based on an assumption of the same mechanism of action.

4.2.4.1.3.1 Efficacy of pelvic floor muscle training in stress urinary incontinence


A Cochrane systematic review compared PFMT with no treatment or inactive control treatment and found
that women with SUI in the PFMT groups were eight times more likely to report cure (56% vs. 6%; 4 trials
including 165 women; high-quality evidence) [325]. The review also documented significant improvement in UI
(7 trials, 376 women; moderate-quality evidence), and improvement in UI QoL (6 trials, 348 women; low-quality
evidence). Pelvic floor muscle training reduced leakage by one episode per day in women with SUI (7 trials,
432 women; moderate-quality evidence). Women with SUI in the PFMT groups lost significantly less urine in
short (up to one hour) pad tests. The comparison of short pad tests showed considerable heterogeneity but the
findings still favoured PFMT when using a random-effects model (mean difference 9.71 g in 4 trials including
185 women; moderate-quality evidence). Women in the PFMT group were also more satisfied with treatment
and their sexual outcomes were better. Adverse events were rare and minor.

A Cochrane review concluded that there may be some additional effect of adding biofeedback to PFMT.
However, this was based on RCTs with training dosage and attention favouring biofeedback [326]. In a
recent RCT (61.3% had MUI) comparing the exact same training dosage and attention between groups, use
of biofeedback did not yield any additional effect [327]. Group training is cost effective in treatment of SUI/
UI compared to individual treatment [328]. A Cochrane review concluded that a combination of individual
assessment/education and group training was equally effective compared to individual treatment, but
again the dosage and attention differed between comparison groups [329]. In a more recent RCT with the
exact same training dosage and attention in individual and group training, group training was not inferior to
individual treatment [328]. It is worth noting that all of the PFMT interventions in these reviews follow individual
assessment and teaching before starting PFMT, and most interventions use some sort of measurement tool
(biofeedback) in the assessment.

Both the Cochrane review and the ICI concluded that the use of vaginal cones to train the PFM is more
effective than no treatment, but it is inconclusive whether it is more or less effective than structured PFMT
[325, 330, 331]. Some women are unable to maintain the cone inside, some report discomfort and motivation
problems and adherence may be low [330].

The Cochrane review [325], the ICI [331] and the National Institute for Health and Care Excellence (NICE)
guidelines (2019) [67] all conclude that there is the highest level evidence (1a) to support PFMT in the treatment
of SUI/MUI. All systematic reviews conclude that PFMT is less effective if women with MUI and UUI are
included in the studies and more effective with more intensive and supervised training. According to the NICE
guidelines literature review, PFMT is as effective as surgery for around half of women with SUI, and due to
the risks following surgery and absence of side effects of PFMT, they recommend three months of supervised
PFMT as first-line treatment for SUI and MUI [67].

Pelvic floor muscle training was compared to synthetic MUS surgery in an RCT involving 460 women with
moderate to severe SUI [332]. Crossover between treatment arms was allowed and 49.0% of women in the
physiotherapy group and 11.2% of women in the surgery group crossed over to the alternative treatment.
Subjective improvement was reported by 90.8% of women in the surgery group and 64.4% of women in the
physiotherapy group at 12 months.

4.2.4.1.3.2 Efficacy of electrical stimulation


There is lack of consensus regarding the use of ES to treat SUI. For subjective cure of SUI, a Cochrane review
found moderate quality evidence that ES is probably better than no active treatment (risk ratio [RR] 2.31) [333].
Similar results were found for cure or improvement of SUI (RR: 1.73), but the quality of evidence was low. There

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 41


is uncertainty as to whether there is a difference between ES and sham treatment in terms of subjective cure
alone because of very low quality of evidence (RR: 2.21). For subjective cure or improvement, ES may be better
than sham treatment (RR: 2.03). Any comparison between ES and PFMT and other treatments is hampered by
low quality evidence. Side effects such as pain and discomfort have been reported, and ES is not tolerated by
all women [333].

In an RCT, 132 women assessed by vaginal palpation to have 0-1 on the modified Oxford grading scale (unable
to contract the PFM) were randomly assigned to an 8-week intervention of either learning to contract via
palpation, palpation with pelvic tilt, intravaginal ES or verbal instruction [334]. The results showed that 63.6%,
69.7%, 33.3% and 18.2% in the four groups, respectively, scored 2 after the intervention. Palpation was
significantly more effective that ES, but one third of the ES group had learned a correct PFM contraction [334].
The effect on UI measured by ICIQ-UI-SF was significantly better in the palpation group.

4.2.4.1.3.3 Long-term efficacy of pelvic floor muscle training


In a systematic review including 19 studies, 1,141 women were followed-up long term (between 1 and 15 years)
after PFMT for SUI [335]. Meta-analysis was not performed due to high heterogeneity of outcome measures
and training dosage (frequency, intensity, duration and adherence). Only two studies provided interventions
during the follow-up period. Losses to follow-up during the long-term period ranged between 0% and 39%.
Long-term adherence to PFMT varied between 10% and 70%. Five studies reported that the initial success
rate on SUI and MUI was maintained at long term. Long-term success based on responders in the original trial
varied between 41% and 85%. Surgery rates at long term varied between 4.9% and 58%. It was concluded
that short-term outcome of PFMT can be maintained at long-term follow-up without incentives for continued
training, but there is a high heterogeneity in both interventional and methodological quality in short- and long-
term PFMT studies [335].

4.2.4.1.3.4 Efficacy of pelvic floor muscle training in childbearing women


Pelvic floor muscle training to prevent SUI has been studied during pregnancy and in the postpartum period
and the results are not reported separately for SUI and other subgroups of UI. A Cochrane review concluded
that PFMT in women with and without UI (combined primary and secondary prevention) during pregnancy,
produced a 26% reduced risk of UI during pregnancy and the mid-postnatal period [336]. Furthermore,
pregnant continent women (primary prevention) who exercise the PFM during pregnancy are 62% less likely
to experience UI in late pregnancy and have 29% lower risk of UI 3 to 6 months after giving birth. To date
there is insufficient evidence for a long-term effect of antenatal PFMT beyond 6 to 12 months postpartum. In
treatment studies; compared with “usual” care, there is no evidence that antenatal PFMT in incontinent women
decreases incontinence in late pregnancy (very low-quality evidence), or in the mid- (low-quality evidence), or
late postnatal periods (very low-quality evidence).

There are fewer RCTs in the postpartum period than during pregnancy [336]. No primary prevention studies
were found in women after birth. For PFMT started after delivery, in a mixed group of continent and incontinent
women, there was uncertainty about the effect on UI risk in the late postnatal period (3 trials, 826 women;
moderate-quality evidence), and in postnatal women with persistent UI, there is no evidence that PFMT results
in a difference in UI at more than six to twelve months postpartum (3 trials; 696 women; low-quality evidence).
However, another RCT found that UI was less frequent in the intervention group, with 57% of patients still
symptomatic, compared to 82% of controls (p = 0.03), as was bladder-related bother with a prevalence of
27% in the intervention vs. 60% in the control group (p = 0.005) [337]. Randomised controlled trials of high
interventional and methodological quality are needed in the postpartum period.

4.2.4.1.3.5 Pelvic floor muscle training in the elderly


There are few RCTs on conservative treatment of SUI in the elderly (> 65 years) and many of the studies
combine different modalities e.g. bladder training, lifestyle modifications and PFMT [338]. Some of the studies
on PFMT and SUI in the general population have included women > 65 years and PFMT seems to be equally
effective in elderly women. A systematic review on conservative management included 23 trials, 9 of moderate
to high methodological quality, and concluded that PFMT in combination with physical training was effective
in reducing UI and improving QoL [339]. Prompted voiding and toileting assistance with functional exercise
reduced UI. Other behavioural interventions such a night-time prompted voiding and waking routine had
no effect on UI reduction. The most recent ICI consensus publication stated that although there are limited
studies of PFMT on UI in frail elderly populations, age and frailty alone should not preclude the use of PFMT
in appropriate patients with sufficient cognition to participate [338]. More high-quality RCTs, both in frail and
healthy older women (> 80 years of age) are needed.

42 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021


4.2.4.1.3.6 Summary of evidence and recommendations for pelvic floor muscle training

Summary of evidence LE
Pelvic floor muscle training is better than no treatment for improving SUI and QoL in women with SUI 1a
and MUI across a range of outcomes including cure rate, improvement rate, QoL, number and volume
of urine leaks and treatment satisfaction.
Pelvic floor muscle training exhibits a low rate of adverse events. 1a
Higher-intensity, supervised, treatment regimens confer greater benefit in women receiving PFMT. 1a
There is no extra benefit of combining PFMT with biofeedback. 1b
Short-term benefits of intensive PFMT can be maintained in the long-term. 2a
Pelvic floor muscle training in the antenatal period is associated with a reduced risk of UI in late 1a
pregnancy and in the short-term post-natally.
Postpartum PFMT is effective in women with persistent UI. 1b
There is no benefit of postpartum PFMT in mixed (continent and incontinent) groups of women. 1b
Mid-urethral sling surgery is superior to PFMT for women with moderate to severe SUI. 1b
Pelvic floor muscle training commencing in the early postpartum period improves UI in women for up 1b
to 6 months.
There is conflicting evidence on whether the addition of ES increases the effectiveness of PFMT alone. 2a

Recommendations Strength rating


Offer supervised intensive pelvic floor muscle training (PFMT), lasting at least three months, Strong
as first-line therapy to all women with stress urinary incontinence (SUI) or mixed urinary
incontinence (MUI) (including the elderly and pre- and post-natal).
Ensure that PFMT programmes are as intensive as possible. Strong
Balance the efficacy and lack of adverse events from PFMT against the expected effect and Strong
complications from invasive surgery for SUI.
Do not offer electrical stimulation with surface electrodes (skin, vaginal, anal) alone for the Strong
treatment of SUI.

4.2.4.1.4 Electromagnetic stimulation


Electromagnetic stimulation (EMS) has been evaluated for its role in SUI therapy. In a double-blind RCT of EMS
including 70 women with SUI, no effect of EMS over sham in any outcome was recorded [340].

4.2.4.2 Pharmacological management


4.2.4.2.1 Oestrogen
Oestrogenic drugs including conjugated equine oestrogens, oestradiol, tibolone and raloxifene, are used as
hormone replacement therapy (HRT) for women with natural or therapeutic menopause.

Oestrogen treatment for SUI has been tested using oral, transdermal and vaginal routes of administration.
Available evidence suggests that vaginal oestrogen treatment with oestradiol and oestriol is not associated
with the increased risk of thromboembolism, endometrial hypertrophy, and breast cancer seen with systemic
administration [243-245]. Vaginal (local) treatment is primarily used to treat symptoms of vaginal atrophy in
post-menopausal women.

A Cochrane systematic review looked at the use of oestrogen therapy in post-menopausal women given local
oestrogen therapy and 17 studies are focused on SUI [243]. There is also a narrative review of oestrogen
therapy in urogenital diseases [341]. The Cochrane review found that vaginal oestrogen treatment improved
symptoms of SUI in the short term [243]. The review found small, low-quality trials comparing vaginal
oestrogen treatment with phenylpropanolamine, PFMT, ES and its use as an adjunct to surgery for SUI. Local
oestrogen was less likely to improve UI than PFMT but no differences in UI outcomes were observed for the
other comparisons. A single trial of local oestrogen therapy comparing a ring device to pessaries found no
difference in UI outcomes although more women preferred the ring device. In one trial no significant adverse
effects following vaginal administration of oestradiol for vulvovaginal atrophy over two years were reported [342].

Vaginal oestrogen therapy can be given as conjugated equine oestrogen, oestriol or oestradiol in vaginal
pessaries, vaginal rings or creams. The ideal treatment duration and the long-term effects are uncertain. A
review of local oestrogen treatment showed improvement of UI over placebo with vaginal rings, which were
favoured subjectively over pessaries [343].

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 43


One RCT in post-menopausal women showed benefit in adding intravaginal oestriol to vaginal ES and PFMT in
female SUI [344].

Studies of systemic HRT with non-urogenital primary outcomes have looked for change in urinary continence
in secondary analyses. Large trials using conjugated equine oestrogens showed a higher rate of development
or worsening of UI compared to placebo and no SUI improvement [345-350]. In a single RCT, use of raloxifene
was not associated with development or worsening of UI [351]. Three small RCTs using oral oestriol or
oestradiol as HRT for vulvovaginal atrophy suggested that UI symptoms were improved although the evidence
was unclear [67, 352, 353].

4.2.4.2.1.1 Summary of evidence and recommendations for oestrogens

Summary of evidence LE
Vaginal oestrogen therapy improves SUI for post-menopausal women in the short term. 1a
Neoadjuvant or adjuvant use of local oestrogens are ineffective as an adjunct to surgery for SUI. 2b
Systemic hormone replacement therapy using conjugated equine oestrogens does not improve SUI 1a
and may worsen pre-existing UI.

Recommendations Strength rating


Offer vaginal oestrogen therapy to post-menopausal women with stress urinary Strong
incontinence (SUI) and symptoms of vulvo-vaginal atrophy.
In women taking oral conjugated equine oestrogen as hormone replacement therapy who Strong
develop or experience worsening SUI discuss alternative hormone replacement therapies.

4.2.4.2.2 Duloxetine
Duloxetine inhibits the presynaptic re-uptake of neurotransmitters, serotonin (5-HT) and norepinephrine (NE).
In the sacral spinal cord, an increased concentration of 5-HT and NE in the synaptic cleft increases stimulation
of 5-HT and NE receptors on the pudendal motor neurons, which in turn increases the resting tone and
contraction strength of the urethral striated sphincter.

Duloxetine was evaluated as a treatment for female SUI or MUI in three systematic reviews [162, 354, 355].
Improvement in UI compared to placebo was observed with no clear differences between SUI and MUI. One
study reported cure for UI in about 10% of patients. An improvement in the Urinary Incontinence Quality
of Life questionnaire (I-QoL) was not found in the study, which used this as a primary endpoint. In a further
study comparing duloxetine, 80 mg daily, with PFMT alone, PFMT + duloxetine, and placebo [356], duloxetine
reduced leakage compared to PFMT or no treatment. Global improvement and QoL were better for combined
therapy than no treatment. There was no significant difference between PFMT and no treatment in this trial.

Two open-label studies with a follow-up of one year or more evaluated the long-term effect of duloxetine
in controlling SUI [357, 358]. Both studies had a high patient withdrawal rate, due to lack of efficacy and
a high incidence of adverse events, including nausea and vomiting (40% or more of patients), dry mouth,
constipation, dizziness, insomnia, somnolence and fatigue, amongst other causes.

A systematic review showed significant efficacy for duloxetine compared to placebo in women with SUI, but
with increased risk of adverse events [355]. The reported adverse effects of duloxetine include mental health
problems and suicidal ideation. A meta-analysis of four RCTs including 1,910 women with SUI reported that
“no suicidality, violence or akathisia events were noted”, but suggested that the discontinuation rate due
to adverse events was around 1 in 7 and that the harms of this treatment may outweigh the benefit [359].
Furthermore, a meta-analysis of twelve placebo-controlled trials involving almost 3,000 patients showed that
in patients with major depressive disorders there were no significant differences in the incidence of suicide-
related events with duloxetine vs. placebo [360].

4.2.4.2.2.1 Summary of evidence and recommendations for duloxetine

Summary of evidence LE
Duloxetine improves SUI in women, but the chances of cure are low. 1a

44 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021


Duloxetine may cause significant gastrointestinal and central nervous system side effects leading to a 1a
high rate of treatment discontinuation, although these symptoms may be limited to the first weeks of
treatment.

Recommendations Strength rating


Offer duloxetine (where licensed) to selected patients with SUI unresponsive to other Strong
conservative treatments and who want to avoid invasive treatment, counselling carefully
about the risk of adverse events.
Duloxetine should be initiated and withdrawn using dose titration because of the high risk of Strong
adverse events.

4.2.4.3 Surgical management


4.2.4.3.1 General considerations
The use of polypropylene mesh, synthetic MUS for the treatment of SUI, has recently come under scrutiny
following concerns raised regarding long-term complications. In some European countries such as the United
Kingdom the use of synthetic MUS has been paused and pelvic mesh was the subject of a UK parliamentary
review published in July 2020 [361]. This review has concluded that “For many women mesh surgery is trouble-
free and leads to improvements in their condition. However, this is not the case for all. There is no reliable
information on the true number of women who have suffered complications. While they may be in the minority,
that does not diminish the catastrophic nature of their suffering or the importance of providing support to them
and learning from what has happened to them”.

The range of complications highlighted during the process of this parliamentary review included [361]:
• pain;
• recurrent infections;
• mobility issues;
• recurring or new incontinence/urinary frequency;
• recurring or new prolapse;
• haemorrhage;
• bowel issues;
• erosion of mesh; this can be into the vagina and/or other organs;
• sexual difficulties; including pain on intercourse and a loss of sex life;
• autoimmune issues;
• psychological impacts.

When considering the choice of surgical treatments for SUI the Panel would advise individual clinicians to
abide by any national or local rules that may be in place regarding mesh surgery. Furthermore it is essential
for clinicians to point out the deficiencies in the long-term evidence regarding mesh use in SUI with specific
reference to the complications highlighted above.

In line with the recommendations from NICE [67] and the Scientific Committee on Emerging and Newly
Identified Health Risks (SCENIHR) paper [362] the Panel agreed that surgeons and centres performing surgery
should:
• be trained in the field of incontinence and for each surgical procedure they perform/offer;
• perform sufficient numbers of a procedure to maintain expertise of him/herself and the surgical team;
• be able to offer alternative surgical treatments;
• be able to deal with the complications of surgery;
• provide suitable arrangements for long-term follow-up.

Furthermore, the establishment of accurate and complete databases registering the interventions, patient
profile and surgical complications or all surgical treatments for SUI is recommended to allow the generation of
robust long-term data.

Many surgical procedures are available for uncomplicated SUI patients and the Panel analysed the results of
the different procedures in terms of clinical effectiveness, safety and cost-effectiveness based on the recent
ESTER systematic review and economic evaluation [363] and previous systematic reviews including those from
the Cochrane collaboration [364-368].

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 45


The outcome parameters used to evaluate surgery for SUI have been limited to:
• continence rate;
• patient-reported outcome measures;
• general and procedure-specific complications;
• generic, specific (UI) and associated (sexual and bowel) QoL.

In this context, it has to be taken into account that a number of products may no longer be available and
therefore the recommendations may not be transferable to current devices. The Panel makes a strong
recommendation that new devices are only used as part of a structured research programme and their
outcomes monitored in a registry, until there is adequate evidence of safety and efficacy.

4.2.4.3.1.1 Recommendations for surgical treatment of SUI

Recommendations Strength rating


Offer patients who have explored/failed conservative treatment options a choice of different Strong
surgical procedures, where appropriate, and discuss the advantages and disadvantages of
each approach.
Use new devices for the treatment of SUI only as part of a structured research programme. Strong
Their outcomes must be monitored in a registry or as part of a well-regulated research trial.

4.2.4.3.2 Surgery for women with uncomplicated stress urinary incontinence


The principal procedures evaluated are:
• open and laparoscopic colposuspension;
• autologous “traditional” slings;
• bulking agents;
• synthetic MUS.

4.2.4.3.2.1 Open- and laparoscopic colposuspension surgery


Open colposuspension was previously considered the most appropriate surgical intervention for SUI, and
was used as the comparator in RCTs of newer, less invasive, surgical techniques. These include laparoscopic
techniques, which have enabled colposuspension to be performed with a minimally invasive approach.

Open colposuspension
A number of systematic reviews were found, which covered the subject of open surgery for SUI, with a large
number of RCTs [363, 365-368]. The Cochrane review on open colposuspension [368] included 55 trials
comprising 5,417 women. In most of these trials, open colposuspension was used as the comparator to
an experimental procedure. Within the first year, complete continence rates of approximately 85–90% were
achieved for open colposuspension, while failure rates in terms of recurrent UI were 17% up to five years and
21% at over five years. The risk of re-operation after Burch colposuspension is estimated at 6% within 5 years
[112] and 10.8% within 9 years [113]. The re-operation rate specifically for UI was only 2%. Colposuspension
was associated with a higher rate of development of enterocoele/vault/cervical prolapse (42%) and rectocele
(49%) at five years compared to TVT (23% and 32%, respectively). The rate of cystocoele was similar post-
colposuspension (37%) and after TVT (41%). The Cochrane review concluded that open colposuspension is an
effective treatment for SUI and around 70% of women can expect to be dry at five years after surgery.

Laparoscopic colposuspension
A Cochrane review reported on twelve trials comparing laparoscopic colposuspension to open
colposuspension [366]. Although these procedures had a similar subjective cure rate, there was limited
evidence suggesting the objective outcomes were less good for laparoscopic colposuspension. The ESTER
systematic review [363] showed, based on a network meta-analysis, that at 12 months open colposuspension
was more effective than laparoscopic colposuspension (9 trials, OR: 0.68, p = 0.009) but these findings are
based on low quality evidence. The Surface Under the Cumulative Ranking Scores (SUCRA), which is a
numerical representation of the overall ranking and presents a single number associated with each intervention,
were 76.7% after open colposuspension, and 48.9% after laparoscopic colposuspension (out of a maximum
score of 100%). Laparoscopic colposuspension had a shorter duration of surgery and subsequent hospital stay
and may be slightly more cost-effective when compared with open colposuspension after 24 months follow-up.

Single-port laparoscopic Burch can be an alternative treatment, although data confirming efficacy are limited
[369].

46 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021


Complications
Voiding difficulties appeared to be more common after laparoscopic colposuspension than after retropubic
MUS (7.5% vs. 5.1%) [363]. There was no evidence of a difference for the comparison assessing open
colposuspension vs. retropubic MUS (7.8% vs. 7.5%; OR: 0.87) [363]. The results for the comparisons of
de novo symptoms of urgency or UUI between open colposuspension and retropubic MUS (11% vs. 8%,
OR: 1.49) did not favour either treatment and showed wide confidence intervals [363]. The rate of bladder or
urethral perforation was higher for laparoscopic colposuspension compared with open colposuspension (3.7%
vs. 0.7%; OR: 4.65) [363].

4.2.4.3.2.1.1 Summary of evidence and recommendation for open- and laparoscopic colposuspension surgery
for stress urinary incontinence

Summary of evidence LE
High subjective cure rates are associated with both open- and laparoscopic colposuspension for the 1a
treatment of SUI.
Objective cure rates are higher for open colposuspension compared to laparoscopic colposuspension. 1a
Colposuspension is associated with a higher long-term risk of POP than MUS. 1a
Laparoscopic colposuspension has a shorter hospital stay and may be more cost-effective than open 1a
colposuspension.
Laparoscopic colposuspension is associated with higher rates of intra-operative bladder perforation 1a
and post-operative voiding dysfunction compared to open colposuspension.
The rates of de novo urinary urgency following colposuspension are similar to other surgical 1a
treatments for SUI.

Recommendation Strength rating


Offer colposuspension (open or laparoscopic) to women seeking surgical treatment for Strong
stress urinary incontinence following a thorough discussion of the risks and benefits relative
to other surgical modalities.

4.2.4.3.2.2 Autologous sling


In the past, autologous, cadaveric, xenograft, and synthetic materials have been used for bladder neck

pubovaginal sling (PVS). Nowadays, use of autologous tissue, either rectus sheath or fascia lata, is the most
studied material with the strongest evidence base to support its use [370].

The ESTER systematic review included 3 trials of autologous sling vs. open colposuspension, six trials of
autologous sling vs. retropubic MUS and one trial comparing autologous sling vs. transobturator MUS. The
quality of evidence was overall very low. The pooled estimate showed that fascial sling had a higher cure rate at
one year, than open colposuspension (OR: 1.24), retropubic MUS (OR: 1.06) and transobturator MUS (OR: 1.44)
but without statistical significance. The SUCRA score was 89.4% for women cured after autologous fascial sling.

A sub-analysis from a Cochrane review showed autologous slings had better effectiveness compared to
colposuspension at one to five years follow-up [368]. In an RCT of Burch colposuspension vs. autologous
slings complete continence rates decreased substantially over time in both arms. At five years, the continence
rate of colposuspension was 24.1% compared to 30.8% for fascial slings. Satisfaction remained higher in the
sling group (83% vs. 73%) and was directly related to the continence status [371].

Complications
Adverse events rates were similar for the two treatment groups (Burch 10% and sling 9%) although post-
operative obstruction was found exclusively in the sling group. Voiding difficulties appear to be more
common after autologous sling [15.4% vs. 10.2%; OR: 1.46] than after retropubic MUS. Compared with open
colposuspension, the rate of bladder or urethral perforation was lower for traditional sling [0.6% vs. 3.0%; OR:
0.20] [363].

4.2.4.3.2.2.1 Summary of evidence and recommendation for autologous sling

Summary of evidence LE
High cure rates are associated with autologous sling placement for the treatment of SUI. 1a
Autologous sling is more effective in terms of cure rate than colposuspension. 1a

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 47


Autologous sling has a similar rate of adverse events compared to open colposuspension, with 1a
higher rates of voiding dysfunction and post-operative UTI, but a lower rate of bladder- or urethral
perforation.

Recommendation Strength rating


Offer autologous sling placement to women seeking surgical treatment for stress urinary Strong
incontinence following a thorough discussion of the risks and benefits relative to other
surgical modalities.

4.2.4.3.2.3 Urethral bulking agents


The concept of this procedure originates from the idea that intra- or peri-urethral injection of an agent able to
form artificial cushions under/around the urethra will increase the resistance at the bladder outlet and facilitate
continence.

Two Cochrane reviews (2012 and updated in 2017) identified fourteen RCTs or quasi-RCTs of treatment for
UI in which at least one management arm involved peri-urethral or transurethral injection therapy [372, 373].
Following this review, five additional reviews investigated the effect of injectables for the treatment of female
SUI [374-378], independently of the injected material. One review included results from RCTs only [378]. In the
most recent Cochrane systematic review, 1,814 patients were included from fourteen trials of seven different
types of intra-urethral injection: glutaraldehyde cross-linked collagen (Contigent©), a porcine dermal implant
(Permacol©), solid silicone elastomer (Macroplastique©), autologous fat, pyrolytic carbon (Durasphere©),
calcium hydroxylapatite (Coaptite©), hydrogel (Bulkamid®) and dextran polymer (Zuidex©). The conclusions
state that “the available evidence base remains insufficient to guide practice” [373].

A systematic review of 23 studies using Macroplastique© including 958 patients showed a 75% improvement
with 43% dry rate at less than 6 months and a 64% improvement and 36% cure rate at more than 18 months
[375]. A review of 514 elderly women with SUI treated with various agents showed a reduced pad weight in
73% at 1-year follow-up, independent of the material injected [379]. The heterogeneity of the populations, the
variety of materials used and the lack of long-term follow-up limit guidance of practice. Most of the studies
show a tendency for a short-term improvement in UI, with the exception of one RCT, which could not find a
difference between saline and fat injection [380].

One trial of 30 women showed a weak (but not clinically significant) advantage in terms of patient satisfaction
after mid-urethral injection in comparison to bladder neck injection but with no demonstrable difference in
continence levels [373]. Two trials found a higher risk of urinary retention with intra-urethral injections compared
with transurethral injections, although the latter is associated with a higher risk of temporary urinary retention
[372, 381]. A small RCT found no difference in efficacy between mid-urethral and bladder neck injection of
collagen [382]. One study treated patients who had received radiotherapy with injection of Bulkamid® and
reported around 25% cure at short-term follow-up [383].

Bulking agent injection is generally safe, the most frequent adverse event being UTI. However, autologous
fat or hyaluronic acid should not be used due to the risk of fatal embolism and local abscess formation,
respectively [372, 380].

Comparison with other surgical procedures


Two RCTs compared collagen injection to conventional surgery for SUI (silicon particles vs. autologous sling
and collagen vs. other surgical procedures). The studies reported greater efficacy but higher complication rates
for open surgery [384, 385].

In a recent non-inferiority clinical trial women with primary SUI were randomised to TVT or polyacrylamide
hydrogel urethral bulking agent injection (Bulkamid®) [386]. Mid-urethral TVT slings were associated with better
satisfaction and cure rates than Bulkamid® in primary SUI. In term of objective cure rate the cough stress test
was negative in 95.0% of patients who underwent TVT vs. 66.4% who underwent Bulkamid®.

4.2.4.3.2.3.1 Summary of evidence and recommendations for urethral bulking agents

Summary of evidence LE
Urethral bulking agents may provide short-term improvement and cure, in women with SUI. 1b

48 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021


Bulking agents are less effective than MUS, colposuspension or autologous sling for cure of SUI and 1b
repeat injections may be required in order to achieve sustained benefits.
Autologous fat and hyaluronic acid as bulking agents have a higher risk of adverse events. 1a
Adverse event rates for urethral bulking agents are lower compared to open surgery. 2a
There is no evidence that one type of bulking agent is better than another type. 1b
The peri-urethral route of injection of bulking agents may be associated with a higher risk of urinary 2b
retention compared to the transurethral route.

Recommendations Strength rating


Offer urethral bulking agents to women seeking surgical treatment for stress urinary Strong
incontinence (SUI) following a thorough discussion of the risks and benefits relative to other
surgical modalities.
Offer urethral bulking agents to women with SUI who request a low-risk procedure with the Strong
understanding that efficacy is lower than other surgical procedures, repeat injections are
likely and long-term durability and safety are not established.
Do not offer autologous fat and hyaluronic acid as urethral bulking agents due to the higher Strong
risk of adverse events.

4.2.4.3.2.4 Mid-urethral slings


Early clinical studies identified that non-autologous synthetic slings should be made from monofilament, non-
absorbable material, typically polypropylene, constructed as a 1–2 cm wide mesh with a relatively large pore
size (macroporous) and coloured to facilitate removal [387]. Mid-urethral slings are now the most frequently
used surgical intervention in Europe for women with SUI.

Transobturator route versus retropubic route


A Cochrane meta-analysis of MUS procedures for SUI in women was performed in 2017 spanning January
1947 to June 2014 [388]. Moderate quality evidence from 55 studies showed variable, but comparable,
subjective cure rates between retropubic and transobturator slings (62–98% in the transobturator groups and
71–97% in the retropubic groups) in the short term (up to one year). No difference in the objective cure rate in
the short term was found. However, the ESTER systematic review [363], based on a network meta-analysis
including 36 trials of overall moderate quality, showed that at 12 months retropubic MUS was, on average,
more effective than transobturator MUS (OR: 0.74); SUCRA scores for women cured after retropubic MUS were
89.1% vs. 64.1% after transobturator MUS. However there was no statistically significant difference in these
cure rates between the two approaches (p = 0.4). Similarly, based on 40 moderate quality trials, retropubic
MUS performed better than the transobturator approach in terms of symptom improvement (RR: 0.76) but the
difference was again not statistically significant (p = 0.16).

Analysis of a randomised equivalence trial of retropubic vs. transobturator MUS for the treatment of SUI in
women shows similar findings. This trial confirms equivalence of objective cure rates at 12 months but not at
24 months (77.3% and 72.3% objective cure rate for retropubic and transobturator surgery). For both types
of MUS subjective and objective treatment success decreased over time and equivalence of the retropubic
and the transobturator routes could not be confirmed at 24 and 60 months with retropubic demonstrating
a slightly increased level of benefit, despite satisfaction remaining high in both arms [389]. Five years after
surgical treatment objective success was 7.9% greater in women assigned to retropubic sling compared to
transobturator sling (51.3% vs. 43.4%), not meeting pre-specified criteria for equivalence. Patient satisfaction
decreased over 5 years but remained high and similar between treatment arms (retropubic sling 79% vs.
transobturator sling 85%, p = 0.15) [390].

In terms of long-term complications, data are scant but in one study de novo OAB developed in 14% of
patients at ten years with no significant differences between groups (TOT vs. TVT) [391]. In a multicentre
prospective study of women undergoing TOT a history of failure of previous anti-incontinence procedures was
the only predictor of recurrence of SUI (HR: 5.34, p = 0.009) [391].

A long-term cohort study of retropubic TVT showed an 89.9% objective cure rate and a 76.1% subjective cure
rate at ten years. Overall, 82.6% of patients reported to be highly satisfied with the surgery [392]. A long-term
prospective study on transobturator sling showed that at 145 months the objective and subjective cure rates were
78.9% and 62.6% respectively; with no significant deterioration in SUI cure rates over time [393]. Another long-
term follow-up study of patients treated with TVT showed a sustained response with 95.3%, 97.6%, 97.0% and
87.2% of patients being cured or improved at 5, 7, 11 and 17 years, respectively [394].

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 49


The ESTER network meta-analysis based on cure and improvement suggested that, when comparing surgical
treatments for SUI, retropubic MUS, transobturator MUS and traditional sling had the highest efficacy rates,
but this ranking does not consider the complication profile of these techniques. The short- to medium-term
adverse event data were sparse [363]. The nine procedures compared in ESTER with their associated SUCRA
ratings are shown in Table 3 below.

Table 3: SUCRA curve values for the outcome - number of women cured (indicated by %)*

Procedure* Number of women cured


Traditional sling operations 89.4%
Retropubic MUS operations 89.1%
Open colposuspension 76.7%
Transobturator MUS operations 64.1%
Laparoscopic colposuspension 48.9%
Single-incision sling operations 39.8%
Bladder neck needle suspension 26.9%
Anterior vaginal repair. 12.5%
PFMT 2.6%
*Adapted from ESTER [363].

Several health economic analyses of MUS procedures have been published with conflicting results. In a review
of 26 economic evaluations and on the basis of a cost-utility and value of information analysis over a 10-year
time period, the authors concluded that MUS remains among the most cost-effective approaches [364]. A
primary economic evaluation of retropubic vs. transobturator tapes over a 5-year time period suggested that
the latter may be cost-effective and cost-saving compared to the standard TVT approach [395]. Conversely,
the findings from the ESTER network meta-analysis stated that over a lifetime, retropubic MUS was, on
average, the least costly and most effective surgery but the level of uncertainty in these analyses was high.

Insertion using a skin-to-vagina direction versus a vagina-to-skin direction


The Cochrane review on MUS for female SUI showed no difference in the short- and medium-term subjective
cure rates in vagina-to-skin (inside-out) vs. skin-to-vagina (outside-in) approaches based on moderate quality
evidence [396]. Voiding dysfunction seems to be more frequent in the vagina to skin (inside-out) TOT group but
this approach is associated with a lower frequency of vaginal perforations (RR: 0.25). Due to the low quality of
the evidence it is unclear whether the lower frequency of vaginal perforations of this approach is responsible
for the observed lower rate of vaginal tape erosions.

Likewise, a further meta-analysis of RCTs demonstrated no significant difference in efficacy between outside-
in vs. inside-out approaches, but vaginal perforations were, again, less frequent in the latter group (2.6%
vs. 11.8%, OR: 0.21, p = 0.0002) [397]. The five-year data of a prospective, non-randomised study of the
two techniques showed a very high objective success rate (82.6 vs. 82.5%, respectively) with no difference
between the two approaches [398].

In a secondary analysis of the E-TOT study (a study of transobturator TVTs in the treatment of women with
urodynamic MUI), no difference in the patient-reported success rates was found between the vagina-to-skin
(inside-out) and the skin-to-vagina (outside-in) groups (63.2% and 65.5%, respectively; OR: 1.11) at 9 years
follow-up [399].

Complications of synthetic mid-urethral slings


In the ESTER network meta-analysis it was noted that comparative assessment of adverse events between
different procedures was not always possible due to the lack of available data [363]. Direct comparisons
using head-to-head meta-analyses were mainly carried out for retropubic MUS, transobturator MUS or
single-incision slings. The authors did, however, comment that “For other intervention comparisons, the
number of studies was generally small and the CIs wide. However, there was some evidence to suggest
that bladder perforation was more likely to occur after retropubic MUS than after transobturator MUS, open
colposuspension or traditional sling”. In particular, the retropubic approach for MUS was associated with a
significantly higher rate of bladder perforation than transobturator MUS (5% vs. 0.2%). Regarding de novo
voiding dysfunction, 36 studies compared transobturator MUS with retropubic MUS, favouring transobturator
MUS (OR: 0.51). While for pain, it is worth pointing out that it was defined and measured in many different
ways across individual trials and across Cochrane systematic reviews. Some pain outcomes were categorised

50 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021


by location (e.g. suprapubic) or time (e.g. short- or long-term). These discrepancies made it difficult to combine
data from different studies. Data were available mainly for the comparison between retropubic MUS and
transobturator MUS and other surgical procedures. However, groin pain was more frequent after transobturator
MUS than retropubic MUS [6.3% vs. 1.3%; OR: 3.80]. Converse findings were reported for suprapubic pain
which was higher following TVT (1.2% vs. 4.0%; OR: 0.37). Visceral injury (0.5% vs. 2.4% OR: 0.36), mean
operative time, operative blood loss and hospital stay were lower in the transobturator groups than retropubic
MUS. The overall vaginal erosion risk was low and comparable in both groups [363].

The rate of tape/mesh exposure or extrusion between retropubic and transobturator MUS was similar (2.1%
vs. 2.4%; OR: 1.10). The exact time points at which measurements occurred could not be derived from the
Cochrane systematic reviews but most studies were reported to have a short follow-up period (≤ 12 months),
with only a few studies having a follow-up period of ≥ 2 years [363]. Re-do surgery for UI was more common in
the transobturator group (RR = 8.79); however the data are limited and of low quality.

A population-based study performed in Scotland in over 16,000 women operated on for SUI showed a similar
rate of complications between mesh and non-mesh surgery [400]. However, a recent study of over 92,000
patients followed in the National Health Service (UK) showed a significant (9.8%) rate of complications using a
more broad definition and following patients for a longer period of time [401]. The level of detail regarding the
precise nature of complications in this paper was poor. These findings suggest that, as with any SUI surgery,
MUS surgery can be associated with complications and fully informed consent is mandatory.

In general, the available published evidence would suggest that MUS do not seem to be associated with
significantly higher rates of morbidity and complications compared to other surgeries for SUI such as open
retropubic colposuspension. Pelvic organ prolapse is more common after colposuspension whilst voiding
dysfunction occurs more often after MUS [368]. The ESTER review has commented that the level of detail
regarding short-to-medium adverse event data is poor for all SUI surgeries [363] and the Panel is aware of
the recent findings from the Independent Medicines and Medical Devices Safety Review in the UK which
has raised the possibility that the level of complications from synthetic MUS may be higher than the medical
literature would suggest [361].

The ESTER systematic review included seven studies comparing re-intervention after transobturator MUS and
retropubic MUS [363]. Pooled analysis of these studies showed wide CIs and considerable uncertainty around
the estimated OR (12-month post-surgery: OR: 1.37). At 12 to 60 months after the procedure, rates of repeat
continence surgery were considerably higher in women undergoing transobturator MUS (18.3%) compared
with retropubic MUS (0.5%), although only two studies were available for the analysis. A similar trend was
observed in studies with a longer follow-up period (> 60 months) but the pooled analysis of these studies
showed wide CIs. For retropubic MUS surgery, the bottom-to-top route was 10% more efficacious than top-
to-bottom in terms of subjective cure and it was associated with less voiding dysfunction, bladder perforations
and vaginal erosion [363].

Single-incision mid-urethral slings


Although there have been many studies published on single-incision devices, it should be noted that there
are significant differences in technical design between devices and it may be misleading to make general
statements about them as a class of operation. It should also be noted that some devices have been
withdrawn from the market (e.g. TVT Secur®, Minitape, MiniArc®), and yet evidence relating to these devices
may still be included in current meta-analyses. There was evidence to suggest single-incision slings are quicker
to perform and cause less post-operative thigh pain, but there was no difference in the rate of chronic pain.
There was insufficient evidence for direct comparisons between single-incision slings, and no conclusions have
been reached about differences between devices.

The ESTER systematic review showed, based on low quality evidence, that at 12 months retropubic MUS and
transobturator MUS were, on average, more effective than single-incision sling (TVT, OR: 0.50, p = 0.01 and
TOT, OR: 0.68, p = 0.02). The SUCRA score was 39.8% for women cured after single-incision slings. However,
since not all single-incision devices have been assessed in a comparative RCT, it may be unsafe to assume
that they are collectively technically similar or exhibit the same levels of efficacy.

Complications of single-incision slings


The meta-analysis results for the comparison between single-incision sling and transobturator MUS showed
similar rates of mesh erosion or extrusion between interventions (4.8% vs. 3.7%; OR: 1.23). Rates of ‘post-
operative pain’ were higher after retropubic MUS than after single-incision slings (19.2% vs. 6.8% OR: 0.21).

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 51


The rate of unspecified pain was higher after transobturator MUS than after single-incision sling both at twelve
months (1.0% vs. 5.2%, OR: 0.24) and at 24 months (1.4% vs. 10.4%, OR: 0.16). Single-incision sling was
associated with more repeat surgeries compared with transobturator MUS (5.1% vs. 2.9%, OR: 1.57). At
over 36 months after the procedure, the repeat surgery rate was 10.3% for single-incision slings vs. 7.6% for
transobturator MUS (OR: 1.42) [363].

Sexual function after synthetic mid-urethral sling surgery


A systematic review examining the effect of synthetic MUS on female sexual function suggested different and
contradictory results between studies. Overall, more papers show an improvement, or no change, in sexual
function because of a reduction in coital incontinence, anxiety and avoidance of sex. Dyspareunia was the
most common cause of worsening of sexual function and the precise incidence is difficult to estimate as a lot
of studies do not report it [402]. A meta-analysis of outcome measures in trials of sling procedures suggests
that single-incision slings are associated with a significantly higher improvement in sexual function compared
to standard MUS procedures [403].

4.2.4.3.2.4.1 Summary of evidence and recommendations for mid-urethral slings

Summary of evidence LE
The retropubic MUS appears to provide better patient-reported subjective and objective cure of SUI, 1a
compared with colposuspension.
Mid-urethral synthetic slings inserted by either the transobturator or retropubic route provide 1a
equivalent patient-reported outcomes at one year.
Mid-urethral synthetic slings inserted by the retropubic routes have higher patient-reported cure rates 1b
in the longer term.
Long-term analyses of MUS cohorts showed a sustained response beyond ten years. 2b
The retropubic route of insertion is associated with a higher intra-operative risk of bladder perforation 1a
and a higher rate of voiding dysfunction than the transobturator route.
The transobturator route of insertion is associated with a higher risk of groin pain than the retropubic 1a
route.
Long-term analysis of MUS showed no difference in terms of efficacy for the skin-to-vagina (outside- 2a
in) compared to vagina-to-skin (inside-out) directions up to nine years.
The top-to-bottom (inside-out) direction in the retropubic approach is associated with a higher risk of 1b
post-operative voiding dysfunction.
The comparative efficacy of single-incision slings against conventional MUS is uncertain. 1a
Operation times for insertion of single-incision MUS are shorter than for standard retropubic slings. 1b
Blood loss and immediate post-operative pain are lower for insertion of single-incision slings 1b
compared with conventional MUS.
There is no evidence that other adverse outcomes from surgery are more or less likely with single- 1b
incision slings than with conventional MUS.
In women undergoing surgery for SUI, coital incontinence is likely to improve. 3
Overall, there is conflicting evidence regarding sexual function following SUI surgery. 2a
Improvement in sexual function appears higher with single-incision slings than with standard MUS. 1a
NB: Most evidence on single-incision slings is from studies using the tension-free vaginal tape secure (TVT-S)
device and although this device is no longer available, it is, however, still included in many systematic reviews
and meta-analyses.

Recommendations Strength rating


Offer a mid-urethral sling (MUS) to women seeking surgical treatment for stress urinary Strong
incontinence following a thorough discussion of the risks and benefits relative to other
surgical modalities.
Inform women that long-term outcomes from MUS inserted by the retropubic route are Strong
superior to those inserted via the transobturator route.
Inform women of the complications associated with MUS procedures and discuss all Strong
alternative treatments in the light of recent publicity surrounding surgical mesh.
Inform women who are being offered a single-incision sling that long-term efficacy remains Strong
uncertain.

52 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021


4.2.4.3.2.5 Other treatments for uncomplicated SUI
Intravesical balloon treatment has been explored for women with SUI. The Vesair® gas-filled intravesical
balloon differs from other SUI treatment modalities in that it is not intended to increase outlet resistance or
minimise urethral hypermobility but to attenuate the fluctuation of intravesical pressure when the abdominal
pressure increases [404, 405]. Two sham-controlled randomised trials evaluating the Vesair® intravesical
balloon have been published so far [404, 406]. Both reported significant reductions in incontinence symptoms
and pad weight but QoL was not significantly different between study arms. High levels of adverse events
were reported in both trials as well as significant numbers of withdrawals/device removals. The most common
adverse events were dysuria, urgency, gross haematuria and UTIs.

Mechanical devices have been used to treat SUI for centuries. There are several devices available which act either
by supporting the bladder neck or urethra to address urethral hypermobility, or by occluding the urethral lumen.
A 2014 Cochrane systematic review of eight RCTs that included three small trials comparing
mechanical devices to no treatment found inconclusive evidence of benefit [407]. Another 2014 review of
mechanical devices concluded that there was insufficient evidence to support their use in women [408]. The
place of mechanical devices in the management of SUI remains in question. Currently there is little evidence
from controlled trials on which to judge whether their use is better than no treatment and large well-conducted
trials are required for clarification. There was also insufficient evidence in favour of one device over another and
little evidence to compare mechanical devices with other forms of treatment [407].

Systematic reviews regarding compression devices such as the adjustable compression therapy and
artificial urinary sphincter devices have published evidence to support their use [409, 410]. Although these
procedures are largely reserved for those with recurrent or complicated SUI (see Section 4.2.4.3.3.3 - External
compression devices) these recent additions to the literature include the use of some compression devices for
uncomplicated SUI.

4.2.4.3.2.5.1 Summary of evidence and recommendations for other treatments for uncomplicated SUI

Summary of evidence LE
Vesair® intravesical pressure-attenuating balloon improves SUI compared to sham control at three 1b
months.
Vesair® intravesical pressure-attenuating balloon is associated with significant levels of adverse events. 1b
Implantation of an artificial sphincter can improve or cure incontinence in women with uncomplicated 3
SUI.
Implantation of the adjustable compression therapy (ACT©) device may improve uncomplicated SUI. 3
Complications, mechanical failure and device explantation often occur with both the artificial sphincter 3
and the ACT©.

Recommendations Strength rating


Offer Vesair® intravesical balloon to women with mild-to-moderate stress urinary Weak
incontinence (SUI) who failed conservative treatments only as part of a well-conducted
research trial.
Offer mechanical devices to women with mild-to-moderate SUI who failed conservative Strong
treatments only as part of a well-conducted research trial.
Inform women receiving artificial urinary sphincter or adjustable compression device (ACT©) Strong
that although cure is possible, even in expert centres there is a high risk of complications,
mechanical failure or a need for explantation.

4.2.4.3.3 Surgery for women with complicated stress urinary incontinence


This section will address surgical treatment for women with complicated SUI as defined in Section 4.2.2 -
Classification above. Neurogenic LUT dysfunction is reviewed by the EAU Guidelines on Neuro-Urology [9].
Women with associated genitourinary prolapse are included in Section 4.7 - Pelvic organ prolapse and LUTS.

The principal procedures included are:


• Colposuspension or MUS (synthetic or autologous) following failed primary SUI surgery;
• External compression devices: adjustable compression therapy (ACT©) and artificial urinary sphincter (AUS);
• Adjustable slings.

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 53


4.2.4.3.3.1 Colposuspension or sling (synthetic or autologous) following failed primary SUI surgery
Urinary incontinence following SUI surgery may indicate persistent or recurrent SUI, or the development of
de novo UUI, or both. Careful evaluation including urodynamics is an essential part of the work-up of these
patients.

Most of the data on surgery for SUI refer to primary operations. Even when secondary procedures have been
included, it is unusual for the outcomes in this subgroup to be separately reported. When they are, the numbers
of patients is usually too small to allow meaningful comparisons. This means that no firm recommendations
can be made regarding which modality is best for the treatment of recurrent SUI and previous systematic
reviews have commented that in view of the absence of any evidence, clinicians must rely largely on expert
opinion or personal experience when advising patients about treatment options [411].

The ESTER network meta-analysis revealed that women with transobturator MUS were more likely to undergo
re-do surgery than those who had retropubic MUS and in general fewer repeat surgeries were observed
after retropubic MUS compared with other interventions [363]. A recent update of two Urinary Incontinence
Treatment Network trials [412] compared the re-treatment-free survival rates by initial surgical procedure.
Five-year re-treatment-free survival rates (and standard errors) were 87% (3%), 96% (2%), 97% (1%), and
99% (0.7%) for Burch colposuspension, autologous fascial sling, transobturator, and retropubic MUS groups
respectively (p < 0.0001). Types of surgical re-treatment included autologous fascial sling (19), bulking agent
(18), and synthetic sling (1). This suggests that MUS, may not be preferred in cases of recurrent SUI [412].
In these cohorts, 6% of women after standard anti-incontinence procedures were retreated within five years,
mostly with injection therapy or autologous fascial sling. Not all women with recurrent SUI chose surgical
re-treatment.

A 2019 Cochrane Review attempted to summarise the data regarding different types of MUS procedures for
recurrent SUI after failure of primary surgical therapy [413]. The literature search identified 58 records but all
were excluded from quantitative analysis because they did not meet eligibility criteria. Overall, there were no
data to recommend or refute any of the different management strategies for recurrent or persistent SUI after
failed MUS surgery. Another systematic review looking at the effectiveness of MUS in recurrent SUI included
12 studies and reported an overall subjective cure rate following MUS for recurrent SUI after any previous
surgery of 78.5% at an average 29 months of follow-up [414]. The subjective cure rate following MUS after
previous failed MUS was 73.3% at a follow-up of 16 months. The authors commented that there was a lower
cure rate with transobturator compared to the retropubic tape for recurrent SUI after previous surgery. A further
systematic review aimed to assess the effectiveness and complications of various surgical procedures for
the treatment of female recurrent SUI and reported on data from 350 women in 10 RCTs with a mean follow-
up of 18.1 months [415]. Conversely, the authors found no difference in patient-reported and objective cure/
improvement rates between retropubic and transobturator MUS in the setting of recurrent SUI. There was also
no significant difference between Burch colposuspension and retropubic MUS in terms of patient-reported
improvement or objective cure/improvement.

Systematic review of older trials of open surgery for SUI suggests that the longer-term outcomes of redo open
Burch colposuspension may be poor compared to autologous fascial slings [416]. Similarly, one large non-
randomised comparative series suggested that cure rates after more than two previous operations were 0% for
open colposuspension and 38% for autologous fascial sling [417].

4.2.4.3.3.1.1 Summary of evidence for surgery in those with recurrent SUI following failed primary SUI surgery

Summary of evidence LE
Failure rates of single-incision slings appear higher than with other types of MUS. 1a
The incidence of repeat surgery is higher in those women who underwent primary transobturator MUS 1a
compared to retropubic MUS.
The 5-year failure rate of Burch colposuspension appears higher than for synthetic- or traditional sling 2b
procedures.
Some studies suggest that retropubic synthetic MUS procedures appear to be more effective than 1a
transobturator MUS for the treatment of recurrent SUI, but this is not a consistent finding in the
literature.
Most procedures will be less effective when used as a second-line procedure. 2a
Burch colposuspension has similar short-term patient-reported or objective cure rates when 1b
compared to TVT for the treatment of recurrent SUI.
Autologous sling appears superior to Burch colposuspension for the treatment of recurrent SUI. 2b

54 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021


4.2.4.3.3.2 Adjustable slings
Although adjustable slings are most commonly used as a treatment for complicated SUI, they may also be
considered as a treatment for uncomplicated SUI. There are no RCTs investigating outcome of adjustable
sling insertion for women with SUI. There are limited data from cohort studies on adjustable tension slings with
variable selection criteria and outcome definitions. Few studies include sufficient numbers of patients or have a
long enough follow-up to provide useful evidence.

One adjustable sling is the Remeex system (Neomedic International®, Terrassa, Spain), which was investigated in
a prospective study of 230 women with SUI [418]. After a mean follow-up of 89 months, 165 patients were cured
of SUI (71.7% in the intention-to-treat [ITT] analysis, 80.5% in per protocol analysis [PP]). Forty patients remained
incontinent (17.4% in ITT, 19.5% in PP). Eighty-eight patients required re-adjustment of the sling during the follow-
up. The tension was increased in 82 cases due to recurrence of SUI and reduced in six due to outlet obstruction.

However the currently available adjustable sling devices have differing designs, making it difficult to draw
general conclusions about adjustable slings as a class of procedure.

4.2.4.3.3.2.1 Summary of evidence for adjustable slings

Summary of evidence LE
There is only low level evidence to suggest that adjustable mid-urethral synthetic sling devices may be 3
effective for cure or improvement of SUI in women.
There is no evidence that adjustable slings are superior to standard MUS. 4

4.2.4.3.3.3 External compression devices


External compression devices are usually used in the treatment of recurrent SUI after failure of previous surgery
but can be considered as a primary treatment. Publications in the literature have largely included patients with
profound intrinsic failure of the sphincter mechanism, characterised by very low VLPPs or low urethral closure
pressures [409, 410]. The two intracorporeal external urethral compression devices available are the adjustable
compression therapy (ACT©) device and AUS.

ACT®: Using US or fluoroscopic guidance, the ACT© device is inserted by placement of two inflatable spherical
balloons, one on either side of the bladder neck. The volume of each balloon can be adjusted through a
subcutaneous port placed within the labia majora. A systematic review including eight studies published
between 2007 and 2013 with follow-up ranging from 1-6 years revealed 15-44% of patients considered
that their SUI had been cured and 66–78.4% were satisfied with the result [409]. The explantation rate
ranged between 19 and 31%. In these studies a significant reduction in the number of pads used per day
was consistently observed after ACT® balloon placement and QoL was significantly improved. The authors
concluded that ACT® balloons constitute a reasonable, minimally-invasive alternative for the treatment of
female SUI due to intrinsic sphincter deficiency, especially in patients who have already experienced failure of
standard surgical treatment.

AUS: The major advantage of AUS over other anti-incontinence procedures is the perceived ability to be able
to void normally [407]. There are a few case series of AUS in women, including four series (n = 611), with study
populations ranging from 45 to 215 patients and follow-up ranging from one month to 25 years [419-422].
Case series have been confounded by varying selection criteria, especially the proportion of women who have
neurological dysfunction or who have had previous surgery. Most patients achieved an improvement in SUI, with
reported subjective cure in 59–88%. Common side effects included mechanical failure requiring revision (up to
42% at ten years) and explantation (5.9–15%). In a retrospective series of 215 women followed up for a mean of
six years, the risk factors for failure were older age, previous Burch colposuspension and pelvic radiotherapy [422].

Early reports of laparoscopically implanted AUS do not have sufficient patient populations and/or sufficient
follow-up to be able to draw any conclusions [423, 424].

A more recent systematic review included 17 studies but all were retrospective or prospective non-comparative
case series [410]. Most patients in the included trials had undergone at least one anti-incontinence surgical
procedure prior to AUS implantation (69.1–100%). Outcomes revealed that complete continence rates ranged
from 61 to 100%. The rates of explantation were 0–45%, erosion rates were 0–22% and mechanical failure rates
were 0–44%. The authors concluded that AUS can provide excellent functional outcomes in female patients with
SUI resulting from intrinsic urethral sphincter deficiency but at the cost of a relatively high morbidity.

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 55


4.2.4.3.3.3.1 Summary of evidence for external compression devices

Summary of evidence LE
Implantation of an artificial sphincter improves or cures incontinence in women with SUI caused by 3
sphincter insufficiency.
Implantation of the artificial urinary sphincter (AUS) device may improve complicated SUI. 3
Implantation of the adjustable compression therapy (ACT©) device may improve complicated SUI. 3
Complications, mechanical failure and device explantation often occur with both the artificial sphincter 3
and the ACT©.
Explantation of AUS is more frequent in older women and among those who have had previous Burch 3
colposuspension or pelvic radiotherapy.

4.2.4.3.3.4 Recommendations for complicated stress urinary incontinence

Recommendations Strength rating


Management of complicated stress urinary incontinence (SUI) should only be offered in Strong
centres with appropriate experience (see Section: 4.2.4.3.1).
Base the choice of surgery for recurrent SUI on careful evaluation, including individual Strong
patient factors and considering further investigations such as cystoscopy, multichannel
urodynamics, as appropriate.
Inform women with recurrent SUI that the outcome of a surgical procedure, when used as a Weak
second-line treatment, is generally inferior to its use as a first-line treatment, both in terms
of reduced efficacy and increased risk of complications.
Only offer adjustable mid-urethral sling as a primary surgical treatment for SUI as part of a Strong
structured research programme.
Consider secondary synthetic sling, bulking agents, colposuspension, autologous sling or Weak
artificial urinary sphincter (AUS) as options for women with complicated SUI.
Inform women receiving AUS or adjustable compression device (ACT©) that although cure Strong
is possible, even in expert centres, there is a high risk of complications, mechanical failure
or a need for explantation.

4.2.4.3.4 Surgery for stress urinary incontinence in special patient groups


4.2.4.3.4.1 Stress urinary incontinence surgery in obese women
There is no agreement as to the outcome of incontinence surgery in obese women. Secondary analysis of an
RCT on retropubic and transobturator tapes in the treatment of women with SUI suggests that obese women
experience inferior outcome compared to non-obese women. Stratification of patients according to BMI (< 30
and ≥ 30) shows significant difference in objective dry rates (negative pad test) at one (85.6% vs. 67.8%) and
five years (87.4% vs. 65.9%) and subjective cure (absence of SUI symptoms) at one (85.8% vs. 70.7%) and
five years (76.7% vs. 53.6%, respectively). Between one and five years, 6.7% and 16.3% of patients initially dry
(negative pad test) after surgery developed a positive pad test, respectively [425, 426].
Conversely, short-term outcome of single-incision MiniArc® sling showed comparable objective cure
rates (negative cough stress test) at two years (86% and 81% in non-obese and obese women, respectively);
similar improvement of the Urinary Distress Inventory 6 and Incontinence Impact questionnaire 7 was observed
in non-obese and obese women [427].

4.2.4.3.4.2 Stress urinary incontinence surgery in elderly women


Age appears to be a significant factor in outcome from SUI surgery but there is conflicting evidence in the
literature. An RCT of 537 women comparing retropubic to transobturator tape, showed that increasing age
was an independent risk factor for failure of surgery over the age of 50 [428]. An RCT assessing risk factors
for the failure of TVT vs. transobturator tension-free vaginal tape (TVT-O) in 162 women also found that age
is a specific risk factor (adjusted OR: 1.7 per decade) for recurrence at one year [429]. In addition, based on
a sub-analysis of a trial cohort of 655 women at 2 years follow-up, it was shown that elderly women were
more likely to have a positive stress test at follow-up (OR: 3.7) were less likely to report objective or subjective
improvement in stress and UUI, and were more likely to undergo re-treatment for SUI (OR: 3.9). There was no
difference in time to post-operative normal voiding [430].
Another RCT comparing immediate TVT vs. no surgery (or delayed TVT) in older women, confirmed
efficacy of surgery in terms of QoL and satisfaction, but with more complications in the surgical arm [431].

56 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021


Conversely, a cohort study evaluating 181 women undergoing TVT-O surgery, found that women over 70 years
had similar outcomes when compared to women under 70 years old in terms of cure rates (92.5% vs. 88.3%,
p = 0.40), voiding dysfunction, vaginal erosion and groin pain at a median follow-up of 24 months [432].

Furthermore, a systematic review of the efficacy of treatments of UI in older patients suggests that MUS are
successful in older patients (≥ 65 years) with 5.2-17.6% reporting persistent SUI after surgery. No difference in
the frequency of de novo UUI, persistent UUI and persistent SUI was found in older patients [365].

A cohort study of 256 women undergoing vagina-to-skin (inside-out) TOT reported similar efficacy in older vs.
younger women, but there was a higher risk of de novo urgency in older patients [433].

4.2.4.3.4.3 Summary of evidence and recommendations for SUI surgery in special patient groups

Summary of evidence LE
Incontinence surgery may be safely performed in obese women, however, outcomes may be inferior. 1
The risk of failure from surgical repair of SUI, and the risk of suffering adverse events, appears to 2b
increase with age.
There is no evidence that any surgical procedure has greater efficacy or safety in older women than 4
another procedure.

Recommendations Strength rating


Inform obese women with stress urinary incontinence (SUI) about the increased risks Weak
associated with surgery, together with the lower probability of benefit.
Inform older women with SUI about the increased risks associated with surgery, together Weak
with the likelihood of lower probability of benefit.

4.2.5 Follow-up
The follow-up of patients with SUI will be dependent on the treatment given. For conservative and physical
therapies sufficient time should be allowed for the demonstration of treatment effect. For pharmacological
treatment early follow-up is recommended. For most surgical interventions short-term follow-up should be
arranged to assess efficacy and identify any surgical complications in the early post-operative phase.

The Panel is supportive of long-term outcome assessment via registries and recognises the paucity of high-
quality long-term data specifically regarding complications from surgery.

4.3 Mixed urinary incontinence


The term ‘mixed urinary incontinence’ is extremely broad because it may refer to equal stress and urgency
symptoms, stress-predominant symptoms, urgency-predominant symptoms, urodynamic SUI (USUI or USI)
with DO or USUI with clinical urgency symptoms, but no DO [434]. The challenge of this broad definition is that
it leads to inconsistencies when evaluating treatment options and outcomes.

4.3.1 Epidemiology, aetiology and pathophysiology


The prevalence rates of MUI vary widely in the literature. A large majority of epidemiological studies have either
not considered subtypes of UI, or only reported on SUI, UUI and MUI. The current literature is unclear regarding
the population prevalence and risks for the different UI subtypes [8]. There are a large number of urinary
symptom questionnaires employed in epidemiological research all with varying evidence of validity. Caution is
needed when comparing epidemiological studies that do or do not report a separate MUI subgroup, and when
generalising from population level data to clinical practice. The problems arise from significant heterogeneity
in terms of types of questionnaires/surveys used, population parameters, variable response rates, varying
definitions of MUI, and outcome measures.

It seems apparent, however, that MUI is the second most common form of UI, after SUI, with most studies
reporting a 7.5–25% prevalence [8]. Furthermore, one can extrapolate that of women with UI, approximately
one-third have MUI [435]. In a secondary analysis of a large clinical trial 655 women were evaluated for the
presence of incontinence and their response to treatment [436]. They found that 50−90% of women fell into
the category of MUI based on patient-reported answers to the Medical Epidemiologic and Social Aspects of
Aging (MESA) and Urinary Distress Inventory (UDI) questionnaires. However, when objective criteria such as
urodynamic findings were used, only 8% of women were categorised as having MUI.

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 57


Mixed incontinence is usually caused by a combination of the same factors that cause stress and urgency
incontinence. Several factors may be responsible for its development, including oestrogen deficiency,
abnormalities in histomorphology, and microstructural changes [437]. One report postulates that an
incompetent sphincter and bladder neck allows urine to enter the proximal urethra during stress, causing a
urethro-detrusor reflex that triggers an involuntary detrusor contraction, which then causes urgency and UUI
[438]. Another paper also showed that urine flow across urethral mucosa led to an increase in the excitability of
the micturition reflex [439]. Ultimately, it is likely that one theory or risk factor does not explain the development
of MUI and its symptoms; it is more probable that disturbances in several elements and the inability of the
bladder to compensate results in the development of MUI [437].

4.3.2 Diagnostic evaluation


Assessment of patients with MUI begins with a thorough history of the patient’s urinary symptoms and follows
the recommendations set out in the general evaluation and diagnosis of LUTS section. It is conventional to try
and categorise MUI as either stress or urge predominant.

Mixed UI is difficult to diagnose, as the condition comprises many phenotypes. Some women exhibit detrusor
contractions provoked by physical stressors, some have unprovoked detrusor contractions, and many have no
abnormal detrusor contractions, but still report urine leakage with the sensation of urgency. Some women with
urgency symptoms do not manifest UUI because their urethral sphincter is strong and often able to prevent
urine leakage [440].

The role of urodynamics in MUI is unclear, but establishing objective degrees of SUI and DO incontinence may
help in counselling patients about the most appropriate initial treatment option.

4.3.2.1 Summary of evidence and recommendations for the diagnosis of mixed urinary incontinence

Summary of evidence LE
There is no evidence that urodynamics affects outcomes of treatment for MUI. 3

Recommendations Strength rating


Complete a thorough history and examination as part of the assessment of mixed urinary Strong
incontinence (MUI).
Characterise MUI as either stress-predominant or urgency-predominant where possible. Weak
Use bladder diaries and urodynamics as part of the multi-modal assessment of patients Strong
with MUI to help inform the most appropriate management strategy.

4.3.3 Disease Management


4.3.3.1 Conservative management
Women with MUI generally have more severe symptoms and respond less well to treatment than women with
only SUI or UUI [441]. Clinicians are encouraged to begin treatment for MUI with conservative management
directed toward the most bothersome component of the woman’s symptom spectrum and to reserve surgery
as a last resort [440].

4.3.3.1.1 Pelvic floor muscle training in mixed urinary incontinence


An RCT comparing PFMT with and without an audiotape for 71 women with UI did not find any difference
between the two treatment arms [442]. Mean number of incontinent episodes per day decreased from 3.9 overall
to 3.2 for participants with MUI. Six months after completing the course of exercises approximately one third of
all enrolees reported that they continued to note good or excellent improvement and desired no further treatment.

A small RCT including 34 women with SUI and MUI compared 8 weeks of PFMT with no treatment and found
that PFM training significantly increased PFM strength, improved QoL, and reduced the frequency of UI
episodes compared to no treatment [443]. Another RCT including SUI and MUI confirmed these results [444].

A multicentre randomised controlled non-inferiority trial on 467 women with MUI was conducted in 10
hospitals. Participants were randomised 1:1 to receive electro-acupuncture (36 sessions over 12 weeks with
24 weeks of follow-up) or PFMT-solifenacin (5 mg/day) over 36 weeks. In women with moderate-to-severe MUI,
electro-acupuncture was not inferior to PFMT-solifenacin in decreasing the 72-hour incontinence episodes
(between-group difference, -1.34%) [445].

58 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021


In a comparative study of the effectiveness of behavioural and PFMT (combined with MUS vs. sling alone
in women with MUI, 416 (86.7%) had post-baseline outcome data and were included in primary 12-month
analyses [446]. The UDI score in both groups significantly decreased (178.0 to 30.7 points in the combined
group, 176.8 to 34.5 points in the sling-only group). The model-estimated between-group difference did not
meet the minimal clinically important difference threshold. Adherence to the behavioural and PFMT regimes,
which is a prerequisite for achieving effect, was not reported in the study.

A Cochrane review comparing PFMT with no or sham treatment included 31 RCTs from 14 countries, but there
was only one study including women with MUI and one with UUI and none of them reported data on cure,
improvement, or number of episodes of these subgroups [325].

Another Cochrane review comparing different approaches to delivery of PFMT (21 RCTs) concluded that
increased intensity of delivery of the therapy improves response and that there is no consistent difference
between group therapy and individualised treatment sessions [329]. This concurs with the latest ICI publication
[331]. No other consistent differences between techniques were found.

The effect of combining biofeedback with PFMT has already been fully addressed in SUI Section 4.2.4.1.3
- Pelvic floor muscle training and there was no evidence of any additional benefit in a population with
predominantly MUI.

4.3.3.1.2 Bladder training


Details on bladder training programs are elucidated in Section 4.2.4 (SUI). The ICI 2017 [331] concluded that for
women with UUI or MUI, PFMT and BT are effective first-line conservative therapies. One RCT assigned 108
women with diagnoses of SUI (n = 50), UUI (n = 16), or MUI (n = 42) to six weeks of BT and PFMT or BT alone
[447]. The results showed that overall, and in the SUI and MUI subgroups, significantly more patients in the BT
and PFMT group reported cure and improved symptoms.

4.3.3.1.3 Electrical stimulation


A Cochrane review on ES for SUI included participants with SUI or stress-predominant MUI. Twenty-five
percent of the included trials were deemed to have a high risk of bias due to a variety of factors including
baseline differences between groups and industry funding. For subjective cure or improvement of SUI, low-
quality evidence indicated that ES was better than no active treatment (RR: 1.73), or sham treatment (RR: 2.03).
Electrical stimulation for OAB and SUI is covered in Sections 4.1.4.1.5.4 and 4.2.4.1.3.2, of the respective
topics.

4.3.3.2 Summary of evidence and recommendations for conservative management in MUI

Summary of evidence LE
Pelvic floor muscle training appears less effective for MUI than for SUI alone. 2
Pelvic floor muscle training is better than no treatment for improving UI and QoL in women with MUI. 1a
Bladder training, combined with PFMT, may be beneficial in the treatment of MUI. 1b

Recommendations Strength rating


Treat the most bothersome symptom first in patients with mixed urinary incontinence (MUI). Weak
Offer bladder training as a first-line therapy to adults with MUI. Strong
Offer supervised intensive pelvic floor muscle training, lasting at least three months, as a Strong
first-line therapy to all women with MUI (including elderly and postnatal women).

4.3.3.3 Pharmacological management


Many RCTs include patients with MUI with predominant symptoms of either SUI or UUI but few report
outcomes separately for those with MUI compared to pure SUI or UUI groups.

4.3.3.3.1 Tolterodine
In an RCT of 854 women with MUI, tolterodine ER was effective for improvement of UUI, but not SUI
suggesting that the efficacy of tolterodine for UUI was not altered by the presence of SUI [448]. In another
study (n = 1,380) tolterodine was equally effective in reducing urgency and UUI symptoms, regardless of
whether there was associated SUI [449]. Similar results were found for solifenacin [450, 451].

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 59


4.3.3.3.2 Duloxetine
In one RCT of duloxetine vs. placebo in 588 women, subjects were stratified into either stress-predominant,
urgency-predominant or balanced MUI groups. Duloxetine was effective for improvement of incontinence and
QoL in all subgroups, although results in stress-predominant groups were better [452]. Treatment-emergent
adverse event rate in the duloxetine group was high at 61.3% with discontinuation rates of 15.7%. Adverse
event rates were higher in those participants taking other concomitant anti-depressant agents.

Duloxetine was also found to have equal efficacy for SUI and MUI from an RCT (n = 553) following secondary
analysis of respective subpopulations [453]. No data on adverse events was reported in this study.

4.3.3.3.3 Summary of evidence and recommendations for pharmacological management of MUI

Summary of evidence LE
Limited evidence suggests that anticholinergic drugs are effective for improvement of the UUI 2
component in patients with MUI.
Duloxetine is effective for improvement of both SUI and MUI symptoms, but adverse event rates are 1b
high.

Recommendations Strength rating


Treat the most bothersome symptom first in patients with mixed urinary incontinence (MUI). Weak
Offer anticholinergic drugs or beta-3 agonists to patients with urgency-predominant MUI. Strong
Offer duloxetine (where licensed) to selected patients with stress-predominant MUI Weak
unresponsive to other conservative treatments and who want to avoid invasive treatment,
counselling carefully about the risk of adverse events.

4.3.3.4 Surgical management


The surgical treatment options for MUI include all the anti-incontinence procedures outlined in the SUI chapter.

Many RCTs include both patients with pure SUI or pure UUI as well as patients with MUI. However, very few
RCTs report separate outcomes for MUI subgroups.

Post-hoc analysis of a large RCT showed that in women undergoing either autologous fascial sling or Burch
colposuspension, the outcomes were poorer for women with a concomitant complaint of pre-operative urgency
[430]. A similar post-hoc review of another RCT comparing transobturator and retropubic MUS showed that the
greater the severity of pre-operative urgency, the more likely that treatment would fail [99]. However, an earlier
study had found that surgery provided similar outcomes, whether or not urgency was present prior to surgery
(this study included only a few patients with urodynamic DO). Another RCT including 93 patients with MUI
showed a statistical improvement in continence and QoL in the group that had TVT and Botox® rather than with
either treatment alone [454].

Case series tend to show poorer results in patients with MUI compared with those with pure SUI. In a case series
of 192 women undergoing MUS insertion, overall satisfaction rates were lower for women with mixed symptoms
and DO on pre-operative urodynamics compared to those with pure SUI and normal urodynamics (75% vs. 98%,
respectively) [455]. A comparison of two parallel cohorts of patients undergoing Burch colposuspension for SUI,
with and without DO, found inferior outcomes in women with MUI [456].

One cohort of 450 women, showed that in urgency-predominant MUI, the success rate of TVT fell to 52%
compared to 80% in stress-predominant MUI [457]. In a study with 1,113 women treated with TVT-O, SUI
was cured equally in stress-predominant MUI or urgency-predominant MUI. However, women with stress-
predominant MUI were found to have significantly better overall outcomes than women with urgency
predominant MUI [458].

In contrast to studies examining older surgical methods, more recent studies (generally small case series)
reported that UUI symptoms improve in 30% to 85% of women with MUI after MUS surgery [459].

In a prospective, multicentre, comparative trial 42 women who had a TVT for MUI had a greater improvement in
urgency and QoL scores than 90 women who had a TOT. There were no significant differences in the cure and
satisfaction rates between the two groups [460].

60 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021


In a single-centre prospective study, 86 consecutive women underwent TOT for MUI. At a mean follow-up
of 59 months, SUI was cured objectively in 83.7% and subjectively in 87.2% of the patients. The continence
rates were 74.4% for UUI and 66.3% for MUI (cure of both components). The patient-reported success rate
was 87.2% (‘much better’ or ‘very much better’ on Patient Global Impression of Improvement scale). There
were statistically significant improvements in all domains except general health. The univariate analysis found
no significant risk factor for persistence of SUI. Median age > 60 years and menopause were predictive for
persistence of UUI. Median and mean age > 60 years were predictive of persistence of overall incontinence
[461]. Overall, the outcome for women with pre-existing UUI remains uncertain.

In a secondary analysis of a study of transobturator TVTs in the treatment of women with urodynamic MUI, no
difference in patient-reported success rates was found between the vagina-to-skin (inside-out) and the skin-to-
vagina (outside-in) groups (63.2% and 65.5%, respectively; OR: 1.11, 95% CI: 0.33–3.70, p > 0.999) at 9 years
follow-up [399].

Analysis of the trial populations included in the meta-analysis on single-incision slings suggests that the
evidence is generalisable to women who have predominantly SUI, and no other clinically severe LUT
dysfunction. The evidence is not adequate to guide choice of surgical treatment for those women with MUI,
severe POP, or a history of previous surgery for SUI.

In general research trials should define accurately what is meant by MUI. There is a need for well-designed
trials comparing treatments in populations with MUI, and in which the type of MUI has been accurately defined.

4.3.3.4.1 Summary of evidence and recommendations for surgery in patients with MUI

Summary of evidence LE
Women with MUI are less likely to be cured of their UI by SUI surgery than women with SUI alone. 1b
The response of pre-existing urgency symptoms to SUI surgery is unpredictable. 3

Recommendations Strength rating


Treat the most bothersome symptom first in patients with mixed urinary incontinence (MUI). Weak
Warn women that surgery for MUI is less likely to be successful than surgery for stress Strong
urinary incontinence alone.
Inform women with MUI that one single treatment may not cure urinary incontinence; it may Strong
be necessary to treat other components of the incontinence problem as well as the most
bothersome symptom.

4.4 Underactive bladder


Underactive bladder is a common clinical entity, defined by the ICS as ‘a symptom complex characterised by a
slow urinary stream, hesitancy, and straining to void, with or without a feeling of incomplete bladder emptying
sometimes with storage symptoms’ [462]. Diagnosis of UAB is made based on clinical symptoms and can have
a highly variable presentation and aetiology.

This differs from DU, which is a diagnosis based on urodynamic studies. Detrusor underactivity is defined by
the ICS as ‘‘a detrusor contraction of reduced strength and/or duration, resulting in prolonged bladder emptying
and/or failure to achieve complete bladder emptying within a normal time span’’[1]. Acontractile detrusor is
specified when there is no detrusor contraction.

Female voiding dysfunction is defined by the ICS as a diagnosis based on symptoms and urodynamic
investigations characterised by abnormally slow and/or incomplete micturition, based on abnormally slow urine
flow rates and/or abnormally high PVR, ideally on repeated measurement to confirm abnormality. Pressure-flow
studies can be required to determine the cause of the voiding dysfunction [28].

4.4.1 Epidemiology, aetiology, pathophysiology


4.4.1.1 Epidemiology
Underactive bladder as an entity remains difficult to study in part because its corresponding urodynamic
correlate, “detrusor underactivity” remains loosely defined, leading to significant variability in diagnostic criteria
across research studies and significant overlap of symptoms with other conditions. As a consequence of the
variability in definition, reported prevalence also varies and ranges from 12% to 45% in females with increased
prevalence seen with age [76] and in institutionalised elderly women [463].

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 61


Several studies have demonstrated similar prevalence rates for DU in the ambulatory setting of around
12%–19.4% [464-466]. As would be expected, voiding symptoms consistent with UAB are slightly higher. A
Detroit population study surveyed 291 women with 20% reporting difficulty with emptying their bladder [467].
In a large cross-sectional, population-based internet survey conducted in the USA, UK and Sweden including
15,861 women ≥ 40 years, 20.1% referred to weak flow, 27.4% to incomplete bladder emptying and 38.3% to
terminal dribbling [5].

Some studies have identified the coexistence of DO during filling and DU in the voiding phase of urodynamic
studies (formerly known as DHIC, “detrusor hyperactivity with impaired contractility”) as a common finding
in elderly women. Up to 38.1% of incontinent institutionalised women showed DHIC in urodynamic studies
[468, 469].

4.4.1.2 Aetiology
The presence of DU in diverse clinical groups suggests a multifactorial aetiology [470]. Idiopathic DU is
probably in part an age-dependent decrease in detrusor contractility with no other identifiable causes,
but young women can also be identified as having DU. There are many secondary causes of DU including
neurogenic (multiple sclerosis, multiple systemic atrophy, spinal cord injury, spina bifida, Parkinson,
hydrocephalus, transverse myelitis, stroke, Guillain-Barré syndrome, diabetes mellitus, pelvic nerve injury, etc.),
myogenic (acute prolonged bladder overdistension, diabetes mellitus, BOO) and iatrogenic (pelvic surgery)
[471].

4.4.1.3 Pathophysiology
There are many pathways involved in normal detrusor contraction, and there are different possible sites of
dysfunction [76] with a variety of mechanisms involved in UAB:
• Central circuits and centres (prefrontal cortex, PAG, PMC, hypothalamus): failure of integration or
processing;
• Efferent pathways (sacral cord, sacral nerves, pelvic nerves, postganglionic neurons): impaired activation
of detrusor;
• Afferent pathways (peripheral afferent nerves, anterolateral white column, posterior column): early
termination of voiding reflex;
• Muscle (detrusor myocyte, extracellular matrix): loss of intrinsic contractility.

Different aetiologies can share common pathophysiological mechanisms: for example, diabetes mellitus affects
mainly afferent pathways and the detrusor muscle; and neurogenic diseases affect central circuits and efferent-
afferent pathways.

One study suggests that in patients with DU, there is significant urothelial dysfunction, increased sub-urothelial
inflammation and apoptosis, and altered sensory protein expression [472]. Impaired urothelial signalling and
sensory transduction pathways may reflect part of the pathophysiology of DU. Pelvic ischaemia is another
proposed mechanism of DU in ageing patients [472]. See Figure 1.

62 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021


Figure 1: Management and treatment of women presenting with urinary incontinence

Site of dysfunction Major aetiological factors Mechanisms

Brain circuits
Pontine micturition center
Periaqueductal gray Failure of integration
Limbus or processing
Hypothalamus
Prefrontal cortex

Efferent pathways
Sacral cord
Sacral nerves Neurological Impaired activation
Pelvic nerves disease or injury of detrusor
Postganglionic neurons
Normative
ageing
Afferent pathways
Peripheral afferents Bladder outlet Early termination of
Anterolateral white column obstruction voiding reflex
Posterior column

Diabetes
mellitus

Detrusor muscle
Loss of intrinsic
Detrusor myocite
contractility
Extracellular matrix

*Figure reproduced with permission from the publisher, from Osman N. et al., [473].

4.4.2 Classification
There is no current classification system of UAB. Patients can be classified according to presumed aetiology or
pathogenic mechanism, but without sufficient longitudinal data or high-level evidence to establish prognostic
factors, the classification of UAB patients in terms of relevant clinical characteristics or risk of complications is
not possible.

4.4.3 Diagnostic evaluation


4.4.3.1 Symptoms associated with detrusor underactivity
A retrospective study correlated LUTS with urodynamic findings in 1,788 patients (1,281 women). In women
with DU (defined as PdetQmax < 20, Qmax <15, BVE% < 90 and excluding obstruction on VUDS), the authors
found a statistically higher occurrence of reduced and/or interrupted stream, hesitancy, feeling of incomplete
bladder emptying, palpable bladder, and absent and/or decreased sensation compared with women with a
normal pressure flow study [474]. A qualitative study on a small sample of male and female patients diagnosed
with DU reported a variety of LUT symptoms and associated impact on QoL. Storage symptoms of nocturia,
increased daytime frequency, and urgency, and the voiding symptoms of slow stream, hesitancy, and straining
were reported by over half of the patients. A sensation of incomplete emptying and post-micturition dribble
were also frequently described. The impact of their symptoms on QoL was highly variable, but in general
storage symptoms were more bothersome [475].

Based on current data, it is not possible to find a pivotal symptom or collection of symptoms to identify DU
patients. The ICI Questionnaire‐underactive bladder (ICIQ‐UAB) has been developed as a research PROM tool,
that needs further validation before use in common clinical practice [476].

4.4.3.2 Urodynamic studies


Non-invasive studies like uroflowmetry, PVR measurement and BVE determination are potentially useful to identify
women who might have DU. There is considerable symptomatic overlap with BOO and uroflowmetry and PVR
findings may also be similar. Only invasive urodynamics with pressure-flow studies can reliably distinguish DU
from BOO and these urodynamic diagnoses can co-exist. There is no consensus on the best method to diagnose
DU in women. In addition, diagnosis in women is particularly difficult as women can void by relaxing the pelvic
floor, that is, without a detectable detrusor contraction during the pressure-flow study and without increased
abdominal pressure [477]. The simplest methods to define and diagnose DU are based on the use of cut-off
values of Qmax and PdetQmax, possibly combined with cut-off values of PVR and BVE. There is no consensus

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 63


on which cut-off values should be used [478]. It is obvious that the prevalence of DU depends on the criteria
used. In a retrospective study on 1,015 women, DU was found in 14.9% when using Qmax < 12 mL/s or PVR
> 150 mL; in 9.6% when using PdetQmax < 30 cm H2O and Qmax < 10 mL/s; and in 6.4% when using PdetQmax
< 20 cm H2O, Qmax < 15 mL/s and BVE < 90% [479].

More elaborate methods combine urodynamic data into an index or a physical quantity that reflects bladder
contraction strength. A value below a certain threshold would thus diagnose DU. Again, there is no consensus
regarding what is normal and what is abnormal. Table 4 provides an overview of the best-known parameters,
their background and typical values. The parameter Watt’s factor (WF) estimates the power generated by the
detrusor per unit area of bladder surface [480]. Its value varies during voiding. Usually, its maximum value
WFmax is considered. Alternatively, its value at Qmax can be used. Projected isovolumetric pressure (PIP) is
a gross simplification of the bladder output relation and estimates the maximum detrusor pressure that can
be generated by the bladder when the outlet is closed, the isovolumetric detrusor pressure. The bladder
contractility index (BCI) is simply a reduction of PIP to an index [46]. The population in which PIP and BCI were
developed mainly consisted of males. Projected isovolumetric pressure (PIP1) also estimates the isovolumetric
detrusor pressure, but was developed in an entirely female population via an experimental method [481].

A third method of quantifying bladder contraction strength involves “stop tests”. One study compared 3 types
of direct measurement of the isovolumetric pressure: (i) the voluntary stop test, in which the patient voluntarily
interrupts flow, (ii) the mechanical stop test, in which flow is interrupted by a balloon catheter, and (iii) the
continuous occlusion test, in which the subject tries to void against a blocked outlet. The latter had the best
reliability and best detected drug-induced changes. The results of the mechanical stop test were however very
similar [482].

All parameters discussed above give some information on the strength of the detrusor contraction in a given
void. They do not necessarily reflect what the detrusor might potentially achieve under optimum conditions
[483]. Also, they give no information on another important aspect of a voiding contraction, namely its duration.
No parameters for this are available. Finally, in a given patient an abnormally low bladder contraction strength
does not necessarily imply an insufficient bladder contraction strength to achieve optimal voiding. Table 4
summarises different parameters to measure detrusor contraction in female patients.

Table 4: Most used parameters to measure detrusor contraction in female patients

Parameter Basis Population Values


Watt’s factor [480] Hill equation of muscle 8 asymptomatic femaleIdeal voiding (bell-shaped flow
contraction in a spherical volunteers aged 28-45curves): WFmax 11-24 W/m2
organ, with fixed years (median 34 years)
Non-ideal voiding: WFmax
constants obtained from 5-10 W/m2
experimental and clinical Normally WFmax > 7 W/m2
studies (expert opinion, unspecified
population) [484]
Projected isovolumetric Bladder Output Relation, Unspecified population, Classification based on expert
detrusor pressure (PIP, simplified to a straight line mainly men with BPO opinion:
cm H2O) and Bladder with fixed slope of 5 cm > 150: strong contraction
Contractility Index (BCI, H2O/mL/sec (Formula: 100-150: normal contraction
using PIP as an index) PdetQmax + 5xQmax) 50-100: weak contraction
[46, 485] < 50: very weak contraction
Projected isovolumetric Comparison of Qmax and 100 women with UUI 5th-95th percentile: 29-78 cm
detrusor pressure PdetQmax values with stop aged 53-89 (mean: 70) H2O
(PIP1, cm H2O) [481] test results years Mean: 49 cm H2O
(Formula: pdetQmax + Qmax) Median: 48 cm H2O
Proposed typical values:
30-75 cm H2O
Continuous occlusion Direct measurement of 70 women with UUI Mean ± SD: 48.7 ± 24.4 cm
test [482] isovolumetric voiding aged 53-89 (mean: 70) H2O
contraction years
BPO = benign prostatic obstruction; PdetQmax = detrusor pressure at maximum flow rate; PIP = projected
isovolumetric pressure; Qmax = maximum flow rate; SD = standard deviation; UUI = urgency urinary
incontinence; WF = Watt’s factor.

64 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021


4.4.4 Disease management
As there are so many different possible causes and pathogenic mechanisms involved in female UAB,
preventive and therapeutic strategies are difficult to define. Among preventive strategies, early recognition after
major surgery or labour might prevent long-term problems associated with prolonged bladder over-distension.
Nerve-sparing techniques for radical pelvic surgery are more favourable in terms of early recovery of bladder
function [486, 487].

Treatment of female DU includes strategies to ensure bladder drainage, increase bladder contraction, decrease
urethral resistance or a combination [484]. The goals of management of UAB are to improve symptoms and
QoL, to reduce the risk of complications for impaired bladder emptying, but also to identify situations where
interventions may not be appropriate.

4.4.4.1 Conservative management


4.4.4.1.1 Behavioural interventions
Regular or timed voiding to avoid bladder over-distension should be encouraged in women with impaired
bladder sensations. Assisted voiding by abdominal straining with adequate relaxation of the PFM has been
recommended, as well as double or triple voiding in an attempt to improve bladder emptying. None of these
manoeuvres have proven their efficacy in a randomised study. There is a possible association between voiding
by excessive abdominal straining and the risk of POP development [488]. A small retrospective study in women
with neurogenic acontractile detrusor secondary to spina bifida showed that Valsalva voiding may increase the
risk of rectal prolapse compared with CISC [489].

4.4.4.1.2 Pelvic floor muscle relaxation training with biofeedback


There are no RCTs on PFM relaxation training in female adults with UAB. Contradictory to common beliefs, one
study found significant relaxation of the PFM after a contraction [490] and a second study found that PFMT
over time increased the speed of PFM relaxation after a single contraction [491]. However, muscle contraction
is known to be followed by relaxation. There is, however, some evidence from the paediatric literature including
one randomised study that compared efficacy of PFM relaxation with biofeedback plus a combined therapy
(including hydration, scheduled voiding, toilet training and diet) vs. combined therapy alone in children with
non-neuropathic UAB and voiding dysfunction. Mean number of voiding episodes was significantly increased in
the relaxation training group compared with the group with only combined treatment (6.6 ± 1.6 vs. 4.5 ± 1 times
a day, p < 0.000). Post-void residual volume and voiding time decreased considerably, whereas maximum urine
flow increased significantly in the relaxation group compared with the combined treatment group (17.2 ± 4.7 vs.
12.9 ± 4.6 mL/s, p < 0.01) [492].

4.4.4.1.3 Clean intermittent self-catheterisation


See Section 4.1.4.1.3 for details on CISC.

4.4.4.1.4 Indwelling catheter


Indwelling urinary catheter may be an option for some women who have failed all other treatments and are
unable to perform CISC. Complications include UTI, stone formation and urethral damage. Suprapubic
catheterisation may be preferable over urethral catheterisation to minimise the risk of urethral trauma and pain
[493].

4.4.4.1.5 Intravesical electrical stimulation


Intravesical electrical stimulation (IVES) can be used to improve bladder dysfunction by stimulating A-delta
mechanoreceptor afferents, but requires preservation of afferent circuit and healthy detrusor muscle. One
retrospective study in 16 patients (11 female) found that two-thirds of patients with a weak detrusor after
prolonged bladder overdistension regained balanced voiding after IVES due to detrusor reinforcement [494].

4.4.4.1.6 Intraurethral insert


The intraurethral insert is a short silicone catheter containing an internal valve and pump mechanism positioned
in the female urethra. See BOO Section 4.5 for more information.

4.4.4.2 Pharmacology management


4.4.4.2.1 Parasympathomimetics
Theoretical approaches to UAB pharmacological treatment include direct stimulation of detrusor cells
muscarinic receptors using agonists like carbachol or bethanechol, or inhibiting acetylcholinesterase (enzyme
that inhibits the endogenous muscarinic agonist acetylcholine) using agents such as distigmine, pyridostigmine
or neostigmine.

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 65


A systematic review on the use of parasympathomimetics in patients with UAB included 10 RCTs (controls
typically received placebo or no treatment). While three studies reported statistically significant improvements
relative to control group, six did not and one even reported significant worsening of symptoms. There was no
evidence for differences between individual drugs, specific uses of such drugs, or in outcome measures [495].
The review concluded that the available studies do not support the use of parasympathomimetics for treating
UAB, especially when frequent and/or serious possible side-effects (gastrointestinal upset, blurred vision,
bronchospasm and bradycardia) are taken into account.

4.4.4.2.2 Alpha-blockers
In order to improve bladder emptying, decreasing outlet resistance through sympathetic blockade at
the bladder neck/urethra has been investigated. One prospective study with tamsulosin showed similar
improvement in terms of uroflowmetry parameters (specifically in the percentage of patients who had a good
therapeutic response) in both women with BOO and women with DU (39.4% and 32.7% respectively) [496].
Another longitudinal study including 14 women with DU showed clinical and urodynamic improvements after
tamsulosin [497]. A prospective single-blind randomised study in female patients with DU compared efficacy of
alpha-blocker, cholinergic drugs or combination therapy, with the latter exhibiting the best results [498].

4.4.4.2.3 Prostaglandins
Prostaglandins are prokinetic agents that promote smooth muscle contraction. Prostaglandin E2 and F2 have
been used intravesically to treat urinary retention after surgery in several studies. A Cochrane systematic review
showed a statistically significant association between intravesically administered prostaglandin and successful
voiding among post-operative patients with urinary retention (RR: 3.07). However, the success rate is relatively
low (32%) compared to placebo. It should also be noted that the 95% CI was very wide, RCTs included in the
pooled analysis were underpowered with methodological limitations, and the event rate was very low indicating
a very low certainty of the evidence [499]. Intravesical prostaglandin treatment is rarely used and further
research is necessary before it can be taken up more widely.

4.4.4.3 Surgical management


4.4.4.3.1 Sacral nerve stimulation
Sacral nerve stimulation is an FDA approved therapy for non-obstructive urinary retention. The mechanism
of action has not been fully elucidated, but activation of afferent sensory pathways, modulation-activation
of central nervous system and inhibition of inappropriate activation of the guarding reflex are some of the
mechanisms proposed.

An RCT included 37 patients in the implantation arm and 31 in the standard medical therapy arm, showing a
mean decrease in PVR volume in the implanted group compared with control of 270 mL and a mean increase
in voided volume of 104 mL [500]. A meta-analysis of 7 studies showed a mean difference in PVR reduction of
236 mL and a mean voided volume increase of 299 mL [501]. The response rate during the trial phase ranged
from 33–90% (mean 54.2) and the success rate of permanent implant ranged from 55–100% (mean 73.9%),
highlighting that patient selection is crucial [502]. A subgroup of women with idiopathic urinary retention
(Fowler’s syndrome) seem to have a higher response rate of 68-77% [503].

In conclusion, SNS is a valid option for female patients with DU, with proper patient selection. Women
should have preserved bladder contractility on urodynamic tests and mechanical/anatomical BOO should
be excluded. Patients with evidence of anatomical bladder outflow obstruction, suspected loss of intrinsic
detrusor contractility or neurogenic bladder dysfunction show lower response rates [504].

4.4.4.3.2 OnabotulinumtoxinA
OnabotulinumtoxinA injections in external striated urethral sphincter may improve voiding in patients with
DU by reducing outlet resistance and possibly reducing the guarding reflex. Some retrospective case studies
show improvement in voiding symptoms, recovery of spontaneous voiding and improvement in urodynamic
parameters (reduction of voiding pressure and/or urethral closure pressures, PVR) [505, 506]. The duration of
symptomatic relief is short, typically three months.

4.4.4.3.3 Transurethral incision of the bladder neck


Transurethral incision of the bladder neck has been described in short series of women with refractory DU. In a
retrospective case study up to 40/82 (48.8%) of women achieved satisfactory outcomes (spontaneous voiding
with voiding efficiency ≥ 50%), but 5 (6.1%) patients developed SUI and 2 (2.4%) developed a vesico-vaginal
fistula [507].

66 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021


4.4.4.3.4 Reduction cystoplasty
This is a very uncommon procedure with a few case reports described only in men [508].

4.4.4.3.5 Myoplasty
One retrospective multicentre study reported the long-term results of latissimus dorsi detrusor myoplasty in
patients with bladder acontractility, with 71% recovering complete spontaneous voiding, with a mean PVR of
25 mL [509]. No other groups have published their experience to reproduce these findings.

4.4.4.4 Summary of therapeutic evidence on detrusor underactivity


The level of evidence of most of therapeutic interventions on DU is low. Only CISC remains as a gold standard
to reduce the adverse consequences of a high PVR and incomplete voiding, in spite of the low level of
evidence that supports this statement.

4.4.4.4.1 Summary of evidence and recommendations for underactive bladder

Summary of evidence LE
Clean intermittent self-catheterisation has proven efficacy in patients who are unable to empty their 3
bladder.
Indwelling transurethral catheterisation and suprapubic cystostomy are associated with a range of 3
complications as well as an enhanced risk of UTI.
Intravesical electrical stimulation may be useful in some patients after prolonged bladder overdistension. 3
Parasympathomimetics do not improve clinical and urodynamic parameters of UAB patients and 1b
frequent and/or serious side-effects may arise.
There is limited evidence about effectiveness of alpha-blockers in women with UAB. 2b
Very low certainty evidence indicates that intravesically administered prostaglandins may promote 1a
successful voiding in patients with urinary retention after surgery.
Sacral nerve stimulation improves voided volume and decreases PVR in women with DU. 1b
There is limited evidence for the effectiveness of OnabotulinumtoxinA external urethral sphincter 3
injections to improve voiding in women with UAB.
Transurethral bladder neck incision may improve voiding in women with DU, but complications (SUI, 3
vesico-vaginal fistulae) may appear.
There is very limited evidence for effectiveness of detrusor myoplasty. 3

Recommendations Strength rating


Encourage double voiding in those women who are unable to completely empty their bladder. Weak
Warn women with underactive bladder (UAB) who use abdominal straining to improve Weak
emptying about pelvic organ prolapse risk.
Use clean intermittent self-catheterisation (CISC) as a standard treatment in patients who Strong
are unable to empty their bladder.
Thoroughly instruct patients in the technique and risks of CISC. Strong
Offer indwelling transurethral catheterisation and suprapubic cystostomy only when other Weak
modalities for urinary drainage have failed or are unsuitable.
Do not routinely recommend intravesical electrical stimulation in women with UAB. Weak
Do not routinely recommend parasympathomimetics in the treatment of women with UAB. Strong
Offer alpha-blockers before more invasive techniques. Weak
Offer intravesical prostaglandins to women with urinary retention after surgery only in the Weak
context of well-regulated clinical trials.
Offer onabotulinumtoxinA external sphincter injections before more invasive techniques as Weak
long as the patient is informed that the evidence to support this treatment is of low quality.
Offer sacral nerve stimulation to women with UAB refractory to conservative measures Strong
Do not routinely offer detrusor myoplasty as a treatment for detrusor underactivity. Weak

4.4.5 Follow-up
Natural history and clinical evolution at long-term follow-up of women with DU is not well known. No
longitudinal cohort studies with long-term follow-up are described in the literature. The interval between follow-
up visits will depend on patient characteristics, treatments given and the frequency of urinary complications.

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 67


4.5 Bladder outlet obstruction
4.5.1 Introduction
Bladder outlet obstruction is defined by the ICS as obstruction during voiding, characterised by increased
detrusor pressure and reduced urine flow rate [1]. Its precise diagnosis requires urodynamic evaluation
including an assessment of both pressure and flow.

Voiding dysfunction is a broad term which is distinct from BOO and is defined by the ICS as ‘a diagnosis
made by symptoms and urodynamic investigations characterised by abnormally slow and/or incomplete
micturition, based on abnormally slow urine flow rates and or raised PVRs, ideally on repeated measurement
to confirm abnormality’ [110]. Pressure-flow studies are required to determine the precise cause of the voiding
dysfunction. Bladder outlet obstruction is one possible cause of voiding dysfunction but there are also
non-obstructive causes and the two terms should not be used interchangeably. Another term that is to be
differentiated from BOO and voiding dysfunction is dysfunctional voiding, which is a specific and discrete form
of voiding dysfunction with an intermittent and/or fluctuating flow rate [110].

4.5.2 Epidemiology, aetiology, pathophysiology


4.5.2.1 Epidemiology
Estimates of prevalence of BOO among women vary, with figures of 2.7% to 29% reported in the literature
[510]. One large series of women undergoing urodynamic evaluation for LUTS found that around 20% are
diagnosed with outlet obstruction. The wide variance between studies is due to several factors, including
differences in definitions and diagnostic criteria for female BOO, differences in study populations, and variation
in study methodologies. The estimated prevalence rates of LUTS due to BOO in women are lower than those
reported in men (18.7–18.9% vs. 24.3–24.7%), respectively [511].

Prevalence of voiding LUTS was found to be associated with age [53, 512, 513], parity [53, 514], prolapse [53,
514] and prior continence surgery [53, 514]. Bladder outlet obstruction has long been postulated to cause mainly
voiding symptoms [515] but recent data from a series of 1,142 consecutive women referred for evaluation of LUT
symptoms suggest that storage symptoms may be predominant in female patients diagnosed with BOO and
excess daytime urinary frequency was the most common symptom reported by 69% [510].

4.5.2.2 Pathophysiology
Bladder outlet obstruction is one of multiple causes of voiding dysfunction in women. The obstruction can be
either anatomical (mechanical) or functional. In anatomic BOO, there is a physical or mechanical obstruction
to the outflow of urine, whereas in functional BOO there is a non-anatomic, non-neurogenic obstruction of the
outlet usually resulting from non-relaxation of bladder neck, sphincter or PFM or increased urethral sphincter
tone or PFM contraction during the void, as is observed in patients with dysfunctional voiding.

Mechanisms for anatomic (mechanical) obstruction include external compression, fibrosis, stricture or injury to
the urethra and kinking of the urethra due to POP. Progressive fibroblastic reaction around the urethra induced
by mesh tapes or slings used in UI surgery may also bring about anatomic (mechanical) obstruction [471]. In a
retrospective review of 192 females diagnosed with BOO, 64% had mechanical obstruction [510].

Functional obstruction, on the other hand, may be caused by failure of relaxation, or contraction, of the bladder
neck and/or urethral sphincter complex or the PFMs during a sustained detrusor contraction [515]. The exact
causes of this lack of relaxation, or contraction, is often elusive but might be due to sympathetic hyperactivity
or hypertrophy of the bladder neck smooth muscle for primary bladder neck obstruction [516] or may be mostly
behavioural as in dysfunctional voiding [517].

4.5.2.3 Aetiology
Conditions associated with anatomic BOO include POP, incontinence surgery, urethral stricture, urethral stenosis,
urethral diverticulum, urethral caruncle, urethral malignancies and para-urethral masses.

Conditions associated with functional BOO include primary bladder neck obstruction, dysfunctional voiding,
and idiopathic urinary retention (Fowler’s syndrome).

In primary bladder neck obstruction the bladder neck fails to open adequately during voiding in the absence
of an anatomical obstruction [518]. It is estimated that 4.6–16% of women presenting with voiding symptoms
have primary bladder neck outlet obstruction [516].

68 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021


Dysfunctional voiding is due to involuntary intermittent contractions of the peri-urethral striated or levator
muscles during voiding in a neurologically normal woman, thought to be caused by faulty learned toileting
behaviour [471]. There is also some evidence of a link between dysfunctional voiding and a history of sexual
abuse [519].

Idiopathic urinary retention, also known as Fowler’s syndrome, is a primary disorder of the external urethral
sphincter with hypertrophy of the muscle fibres, which fail to relax during micturition. It is associated with
decreased detrusor contractility via enhancement of the guarding reflex. It is seen most often, but not
exclusively, in young women with urinary retention and is characterised by increased urinary sphincter volume
and activity/tone, which may be hormonally triggered [520].

Alpha-adrenergic agonists, such as pseudo-ephedrine commonly contained in decongestants, could lead to


some form of functional obstruction due to their stimulatory effects, which may contract the bladder neck and
lead to urinary retention [521].

Neurologic conditions can also bring about functional BOO in females. These will not be considered in this
Guideline document and are covered elsewhere [9].

4.5.3 Classification
4.5.3.1 Anatomic bladder outlet obstruction
Anatomic BOO involves a physical or mechanical obstruction of the outflow of urine.

4.5.3.2 Functional bladder outlet obstruction


Functional BOO involves a non-anatomic, non-neurogenic obstruction of the outflow of urine resulting from
non-relaxation or increased tone in the bladder neck and/or urethral sphincter complex or the PFMs (Table 5).
Neurologic causes of functional BOO will not be considered in this Guideline document and are covered
elsewhere [9].

Table 5: Main causes of female bladder outlet obstruction

Functional BOO Anatomical BOO


• Primary bladder neck obstruction • Urethral stricture
• Dysfunctional voiding • Anti-incontinence surgery
• Idiopathic urinary retention (Fowler’s • Pelvic organ prolapse
syndrome) • Urethral diverticulum
• Urethral caruncle
• Urethral malignancies
• Para-urethral masses

4.5.3.3 Recommendation for the classification of bladder outlet obstruction

Recommendation Strength rating


Use standardised classification of bladder outlet obstruction in women (anatomical or Strong
functional) and research populations should be fully characterised using such classification.

4.5.4 Diagnostic evaluation


The diagnosis of BOO in women, although dependent on formal pressure flow studies, may be suggested by a
number of clinical and other non-invasive assessments.

4.5.4.1 Clinical history


In terms of the clinical history a range of LUTS may be elicited and these may not be confined to voiding LUTS.
Women may not present until they are suffering with the possible complications of BOO such as recurrent
UTI, chronic urinary retention or acute/chronic kidney disease [510]. The evidence regarding clinical utility
of symptoms for the diagnosis of BOO is inconclusive. In a single-centre retrospective study involving 587
women, 38 of whom were diagnosed with BOO, the authors concluded that symptom assessment alone was
insufficient for the diagnosis and a full urodynamic evaluation was essential [522]. A smaller retrospective
study of 57 premenopausal women with bothersome LUTS found a significantly higher proportion of women
with bladder dysfunction presenting with the symptom of UUI. Patients with voiding phase dysfunction had

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 69


higher total scores and voiding symptom subscores in the American Urological Association Symptom Index
(AUASI) [523]. Perhaps some of the difficulty in evaluating the diagnostic accuracy of urinary symptoms comes
from the observation that a significant proportion of female patients presenting with obstruction will also
have concomitant urodynamic abnormalities. In a large study of over 5,000 women with urinary symptoms
identifying 163 with BOO additional urodynamic diagnoses were noted in 54% [524]. Similarly, in a study
involving 101 women with a primary diagnosis of SUI, the prevalence of BOO (based on maximum urine flow
[Qmax] of less than 12 mL per second and maximum detrusor pressure at maximum flow of more than 25 cm
H2O) was 16% [525]. Symptoms alone were not sufficient to discriminate between the various different diagnostic
groups of women in these 2 studies. Lower urinary tract symptoms appear to be fairly sensitive to change
following intervention for BOO. A prospective study in 53 women with clinically suspected voiding dysfunction
describes significant symptom improvement in 12 of 16 patients who underwent surgical intervention [526].

4.5.4.2 Clinical examination


There are no studies evaluating the clinical utility of physical examination in women with suspected BOO.
Despite this it is widely considered as a key part of the medical assessment. It allows for visual inspection of
the urethra and vagina for possible causes of mechanical obstruction as well as an assessment of the pelvic
floor, which may be the cause of functional obstruction.

4.5.4.3 Uroflowmetry and post-void residual volume


Reduced Qmax and/or incomplete bladder emptying can result from both a weakness in the contractile
strength of the detrusor muscle or the presence of increased outlet resistance due to functional or anatomical/
mechanical BOO. The use of a uroflowmetry measurements to differentiate between anatomical and functional
BOO was explored in a retrospective study of 157 women [517] which concluded that Qmax was statistically
significantly lower in those patients with anatomical obstruction but a large degree of overlap was noted.
The largest evaluation of the diagnostic utility of urine flow studies and PVR volume estimation derives
from a retrospective analysis of over 1,900 patients with symptoms of voiding dysfunction of whom over
800 were diagnosed with BOO based on urodynamic assessment [527]. In this series functional BOO was
over 6 times more common than anatomical/mechanical obstruction which is discordant with most of the
other epidemiological literature for female BOO. The authors found that although urine flow rate alone was
not accurate enough to diagnose BOO, a PVR of 200 mL or more was able to differentiate bladder neck
dysfunction from the other causes of BOO, with a receiver-operator characteristics (ROC) AUC of 0.69.
Conversely, in a retrospective study involving 101 women primarily presenting with SUI, a good correlation
between abnormal uroflowmetry and urodynamic obstruction (phi = 0.718, p < 0.0001) was found [525]. In a
prospective study of over 50 women with a clinical diagnosis of voiding dysfunction abnormal uroflow curves
were observed in around 40% of women, but BOO based on pressure-flow results was confirmed in only 52%
of these women [526].

4.5.4.4 Ultrasound
The major utility of US scanning in women with BOO is to detect possible complications such as bladder
wall thickening or upper tract dilatation/hydronephrosis. However, the diagnostic capabilities of US have
been investigated in a prospective case control study involving 27 patients with cystoscopically confirmed
bladder neck obstruction [528]. The value of shear wave elastography (SWE) and acoustic radiation force
impulse imaging (ARFI) in the diagnosis of female BOO was compared and the authors concluded that ARFI
was more accurate than SWE, but a combination of the two techniques was superior to both in this small
study. Ultrasound scanning was further evaluated in a small study of just 15 women with BOO diagnosed
urodynamically [529]. The authors proposed that trans-vaginal ultrasonography was able to demonstrate
a closed bladder neck during attempts at micturition and concluded that this modality was useful in the
evaluation of the possible causal factors of female BOO such as primary bladder neck obstruction.

4.5.4.5 Magnetic resonance imaging


The role of MRI in the diagnostic evaluation of female patients with suspected BOO is poorly defined. Although
it allows for the precise anatomical evaluation of pelvic structures there are no reports describing its clinical
utility in the diagnosis of female BOO. Magnetic resonance imaging in patients with urethral stricture disease
can determine the degree of peri-urethral fibrosis, although the prognostic and clinical significance of such
finding has not been established [530].

4.5.4.6 Electromyography
Electromyography (EMG) has been most extensively studied in the subgroup of women with BOO due to
idiopathic urinary retention caused by a high-tone non-relaxing sphincter (Fowler’s syndrome). Abnormal EMG
activity may be associated with non-relaxation of the striated sphincter, abnormally high urethral pressure, and,

70 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021


through an exaggerated guarding reflex, poor bladder sensation and reduced detrusor contractile strength
[519, 531]. Complex repetitive discharges and decelerating bursts are specific EMG abnormalities that have
been described in patients with high-tone non-relaxing sphincter although these abnormalities have also
been noted in asymptomatic volunteers [532, 533]. A review on the subject of voiding dysfunction in women
included 65 studies with only a small number addressing the diagnostic utility of electromyography [471]. The
authors commented that increased EMG activity of the PFM can be seen during voiding or non-relaxation and
when this is coupled with pressure-flow information from urodynamics may be useful to differentiate between
functional and anatomical obstruction. Further evidence for this comes from a retrospective study of 157
women with roughly equal numbers of women with functional and anatomical obstruction concluding that a
low level of EMG activity is characteristic of anatomical obstruction [517]. Additional neurophysiological tests
such as anal sphincter EMG, bulbocavernosus reflex, and pudendal sensory evoked potentials can assess
the integrity of the somatic S2, 3, 4 nerve roots; however, their clinical utility in the context of non-neurogenic
female BOO needs to be better defined [519].

4.5.4.7 Cystourethroscopy
Cystourethroscopy can be useful to visualise any anatomical/mechanical obstruction and provide information
regarding its nature, location and calibre. Given that pelvic malignancy may cause anatomical BOO,
cystourethroscopy is considered an essential part of the diagnostic pathway. Formal urethral calibration may
be useful for women with BOO secondary to urethral stricture disease and various different urethral calibre
thresholds have been used, from 14Fr to 20Fr [534].

4.5.4.8 Urodynamics and video-urodynamics


Pressure flow studies are the mainstay of BOO diagnosis and the characteristic abnormalities are a
combination of low flow and concomitant high detrusor pressure [518]. However, while the general definition of
BOO is well-established with some data supporting its clinical validity in male patients [535], the urodynamic
definition of female BOO remains a matter of controversy [515]. Several urodynamic criteria have been
introduced during the past 20 years but none have been established as a standard due to lack of clinical
validation [515, 536]. The Blaivas and Groutz nomogram which plots free Qmax and maximum detrusor pressure
(Pdet,max) measured during urodynamic studies is one of the most popular [537] but has been suggested
to overestimate obstruction [74]. The addition of fluoroscopic imaging suggested by Nitti and colleagues
introduces a video-urodynamic criterion for obstruction and has found popularity [77]. However, both methods
lack data supporting their clinical validity, especially regarding their predictive value for therapeutic intervention
outcomes [75].

In a large retrospective study of 1,914 patients, 810 of whom were diagnosed with BOO, several urodynamic
cut-off values were determined by ROC curve analysis to optimise the diagnostic accuracy of video-
urodynamic studies [527]:
• PdetQmax of 30 cm H2O or greater for differentiating BOO from bladder dysfunction and normal studies
(ROC AUC = 0.78);
• the Abrams-Griffiths number greater than 30 for differentiating anatomic from functional BOO (ROC AUC
= 0.66);
• PdetQmax of 30 cm H2O or greater for differentiating dysfunctional voiding from poor relaxation of the
external sphincter (ROC AUC = 0.93).

Other smaller studies with a similar methodology of utilising ROC curve analysis have concluded that neither
pressure flow data only nor clinical symptoms alone may be sufficient for diagnosing obstruction in women
[538], therefore independent validation of any suggested thresholds is necessary.

More recently, Solomon and Greenwell devised a female BOO nomogram which parallels the ICS nomogram
used for male BOO [539]. It allows the calculation of an alternative BOO female index (BOOIf), using a formula
closely aligned to its male counterpart: BOOIf = PdetQmax-2.2Qmax.

It is interpreted with a different algorithm however:


• BOOIf < 0: less than 10% probability of obstruction;
• 5 < BOOIf < 18: equivocal, at least 50% likelihood of obstruction;
• BOOIf >18: 90% likelihood of obstruction.

The Solomon-Greenwell nomogram is the first which has been tested for clinical validity. In a recent series of 21
unselected consecutive women treated for BOO, the authors observed significant improvement of all urodynamic
parameters (Qmax, PdetQmax, BOOIf) in female patients who became asymptomatic post-operatively [540].

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 71


An alternative urodynamic parameter of area under the detrusor pressure curve during voiding (corrected for
voided volume) has been proposed following a prospective study involving 103 women [541]. The authors
concluded that this variable appears to be the most discriminating urodynamic parameter for the diagnosis of
female BOO. This suggested diagnostic method has not been independently validated.

Voiding cystourethrography alone or in conjunction with concomitant pressure flow studies may be useful in
delineating the site of obstruction. Characteristic features include:
• radiographic evidence of obstruction between the bladder neck and distal urethra in the presence of a
sustained detrusor contraction [77];
• lack of funnelling appearance of the bladder neck/tight bladder neck in primary bladder neck obstruction;
• proximal dilatation of the urethra with distal narrowing in women with urethral stricture disease or pelvic
floor hypertonicity.

4.5.4.9 Summary of evidence and recommendations for the diagnosis of bladder outlet obstruction

Summary of evidence LE
The evaluation of LUTS by history and examination alone is insufficient to accurately diagnose female 3
BOO.
Urine flow studies cannot diagnose BOO in women with high levels of accuracy. 3
Ultrasound scanning is unable to diagnose BOO in women with high levels of accuracy. 2b
Electromyography alone is unable to diagnose BOO in women with high levels of accuracy although it 3
may be of use both in combination with pressure flow studies and in the differentiation of anatomical
vs. functional obstruction.
Urodynamics (often combined with videofluoroscopy) is the standard test for evaluating female BOO. 3

Recommendations Strength rating


Take a full clinical history and perform a thorough clinical examination in women with Strong
suspected bladder outlet obstruction (BOO).
Do not rely on measurements from urine flow studies alone to diagnose female BOO. Strong
Perform cystourethroscopy in women with suspected anatomical BOO. Strong
Perform urodynamic evaluation in women with suspected BOO. Strong

4.5.5 Disease management


Therapeutic interventions for BOO aim to decrease outlet resistance in order to increase urinary flow, improve
bladder emptying and thus reduce voiding and storage LUT symptoms [75, 515, 536]. Treatment choice is
commonly dictated by the nature of the underlying cause of the obstruction.

4.5.5.1 Conservative management


4.5.5.1.1 Behavioural modification
Behavioural modification aims to improve or correct maladaptive voiding patterns through the analysis
and alteration of the relationship between the patient’s symptoms and her environment, lifestyle and
habits. Behavioural modification interventions are often tailored to individual patients’ needs, symptoms and
circumstances and can include elements such as education regarding normal voiding function, self-monitoring
of symptoms, changes in lifestyle factors that may affect symptoms, avoidance of constipation and alteration
of voiding technique. Ultimately, techniques aim to improve the coordination between the detrusor and the
sphincter resulting in their synergistic action [75, 515, 536].

The vast majority of individual components of self-management have not been critically evaluated and most
recommendations are traditionally derived from consensus methodology. General interventions such as those
listed above may help with symptoms resulting from BOO but no quantification of their effect is possible.

4.5.5.1.2 Pelvic floor muscle training +/- biofeedback


Pelvic floor muscle training aims to improve pelvic floor function and urethral stability. In the context of BOO,
physiotherapy aims to teach patients to relax their PFMs and striated urethral sphincter during voiding. Pelvic
floor muscle contraction, particularly in women with pelvic floor dysfunction, has been shown to result in a
significant reduction in vaginal resting pressure and surface EMG activity [490]. A 12-week PFMT program
in post-menopausal women demonstrated significant improvement in the speed of relaxation after PFM
contraction and a decrease in the PFM tone [491].

72 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021


As mentioned in the section discussing UAB (see Section 4.4.4.1.2 - PFM relaxation training with biofeedback),
most of the evidence supporting PFMT in dysfunctional voiding are from studies involving children.

A case-series reported improved PFM relaxation and voiding function following PFMT with biofeedback in 15
women with dysfunctional voiding based on a dilated proximal urethra on voiding cystourethrography (VCUG)
and hyperactivity of the pelvic muscles or external urethral sphincter on EMG during voiding. No clinical
outcomes were reported by this series [542].

4.5.5.1.3 Electrical stimulation


Application of electrodes that allow for controlled contraction and relaxation of the PFM may theoretically
facilitate the relaxation of the external sphincter and pelvic floor but no critical evaluation of this intervention in
women with BOO has ever been published.

4.5.5.1.4 Use of vaginal pessary


Intravaginal devices such as pessaries aim to relieve voiding symptoms and improve bladder emptying by the
physical correction of the obstruction caused by a prolapsed pelvic organ. In a prospective study of 18 women
with grade 3 to 4 cystoceles and diagnosed with BOO by urodynamics (defined as PdetQmax > 25 cm H20,
Qmax < 15 mL/sec), normal voiding was noted in 17 (94%) immediately after placement of a vaginal pessary.
No other outcomes were available in this series [543]. No long-term data are available on the use of vaginal
pessary for BOO.

4.5.5.1.5 Urinary containment devices


Urinary containment devices include body-worn absorbent products. Their use in BOO is to achieve social
continence in patients with urinary retention and associated overflow UI and they are often only a temporary
measure. There are no published studies on the outcomes or adverse events associated with the use of urinary
containment devices for the management of female BOO. While there may be no studies exclusively involving
women with BOO, there are many involving women with UI who may have BOO as an underlying cause.

4.5.5.1.6 Urinary catheterisation


Significant urinary retention from BOO may be addressed by actively bypassing the obstruction and draining
the residual urine. Catheterisation may be used as a treatment itself or as an adjunct to an initial treatment of
urethral dilatation or urethrotomy or bladder neck incision. There are two ways of using a catheter: CISC or
indwelling catheterisation [115].

Post-UI surgery BOO may be managed by short-term catheterisation for the majority of those who will suffer
from transient post-operative voiding difficulty. For a few women who develop chronic urinary retention, CISC
or indwelling catheterisation may be offered [471].

A small RCT investigated the effectiveness of CISC to prevent recurrence after internal optical urethrotomy for
urethral stricture disease. In the treatment group, CISC was done twice a day for one week, and once a day
for 4 weeks, then once weekly for 7 weeks post-urethrotomy. Freedom from stricture recurrence, determined
by a urethrogram and uroflowmetry performed 12 weeks post-surgery, was higher in the catheterisation group
compared to no catheterisation (78.5% vs. 55.4%) [544]. This finding mirrors the Cochrane systematic review
on self-dilatation for urethral stricture among men that showed less recurrence with the performance of self-
dilatation [545].

In a series of 20 patients with voiding dysfunction after TVT who were put on a CISC programme, overall cure
rate was 59%, with cure defined as consistent residual volume of less than 100 mL. Half of these patients were
voiding normally within 12 weeks [546].

A patient satisfaction survey involving 188 patients on CISC/self-dilatation, which included 38 patients with
urethral stricture, showed positive (pleased or satisfied) outcomes in 54.3% of all patients, while 28.2% had
mixed feelings and 9.6% were unhappy. No rates were given specifically for the BOO group [547].

4.5.5.1.7 Intra-urethral inserts


An intra-urethral insert is a short silicone catheter containing an internal valve and pump mechanism positioned
in the female urethra. The valve-pump mechanism is operated by an external control unit, which activates to
open the valve and the pump to draw urine from the bladder and allow voiding. At the end of urination the
pump ceases to rotate and the valve closes to regain continence. The insert is routinely replaced once a month.

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 73


Only one study reported the use of this device in 92 women with voiding dysfunction of various aetiologies
including multiple sclerosis, prior pelvic surgery, pelvic radiation, diabetes mellitus, spinal stenosis and injury.
The device was removed within 7 days of insertion in 60% of the cases due to discomfort, pericatheter leakage
and technical difficulty. An additional 20% of patients had late discontinuations. All those who continued to use
the device were satisfied, with PVR volumes remaining at less than 100 mL. Adverse events included migration
into the bladder in 6 cases and symptomatic UTI in 4 cases [548, 549]. Extended, long-term data on the use of
urethral inserts are not available.

4.5.5.1.8 Extracorporeal magnetic stimulation


Extracorporeal magnetic stimulation involves the patient sitting on a device that induces consistent contraction
and relaxation of PFM by repeated magnetic stimulation of motor nerve fibres. Extracorporeal magnetic
stimulation contracts and then relaxes the PFM following a set frequency and interval. It is postulated that
patients could therefore learn to spontaneously contract or relax the PFM which may enhance their ability to
relax their pelvic floor while voiding [550].

In a small (n = 60) prospective non-randomised trial, alfuzosin was compared to EMS and to the combination
of alfuzosin + EMS in women with functional BOO. They observed significant increase of Qmax and significant
decrease of IPSS in all groups and significantly greater improvement in the QoL question of the IPSS in the
combination therapy group [550].

4.5.5.1.9 S
 ummary of evidence and recommendations for conservative treatment of bladder outlet
obstruction

Summary of evidence LE
Pelvic floor muscle relaxation training with biofeedback may result in relaxation of the pelvic muscles 3
and external urethra in women with dysfunctional voiding.
There is no available evidence in the published literature on the clinical effect of ES for the NA
management of female BOO.
In women with large (grade 3 to 4) cystoceles causing BOO, placement of a vaginal pessary may 3
improve voiding efficiency.
Regular CISC after urethrotomy is better than no catheterisation to prevent recurrence of urethral 1b
strictures.
A CISC program in women with voiding dysfunction after TVT has a cure rate of 59%. 3
Women who use the intra-urethral device had lower PVR volume but the majority required its removal 3
due to complications.
Extracorporeal magnetic stimulation combined with alfuzosin may be more effective than either of 2a
these therapies used alone in female patients with functional BOO.

Recommendations Strength rating


Offer pelvic floor muscle training (PFMT) aimed at pelvic floor muscle relaxation to women Weak
with functional bladder outlet obstruction (BOO).
Prioritise research that will investigate and advance the understanding of the mechanisms Strong
and impact of PFMT on the coordinated relaxation of the pelvic floor during voiding.
Offer the use of a vaginal pessary to women with grade 3 to 4 cystocoeles and BOO who Weak
are not eligible/inclined towards other treatment options.
Offer urinary containment devices to women with BOO to address urinary leakage as a Weak
result of BOO, but not as a treatment to correct the condition.
Offer clean intermittent self-catheterisation to women with urethral strictures or post-urinary Weak
incontinence surgery for BOO.
Do not offer an intraurethral device to women with BOO. Strong

4.5.6 Pharmacologic management


4.5.6.1 Alpha-adrenergic blockers
Alpha-adrenergic blockers are postulated to relieve LUTS caused by BOO in females via smooth muscle
relaxation in the bladder neck thus decreasing bladder outlet resistance [551].

Systematic reviews on the use of alpha-blockers in women generally involve studies with a population that
includes females complaining of LUTS and voiding dysfunction. Confirmation of BOO is often not required

74 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021


in the trials included in these systematic reviews [552, 553]. These reviews showed significant improvements
in symptoms and urodynamic parameters associated with their use [552-554]. A meta-analysis of 14 RCTs
comparing alpha-blockers and placebo in women with LUTS showed statistically significant symptom relief
after alpha-blocker treatment relative to placebo (MD: -1.60, p = 0.004), but no significant difference in
Qmax, PVR and adverse event rates [552]. This is in contrast with prospective non-comparative trials which
consistently show improvements in voiding and storage symptoms, bother scores, and urodynamic parameters
(Qmax, PVR, PdetQmax, MUCP) after alpha-blocker use compared to baseline [496, 497, 555-557].

A systematic review performed by the Panel of studies on alpha-blocker used specifically for women with BOO
included one placebo-controlled RCT, one RCT comparing two types of alpha-blockers, and 6 prospective
non-comparative studies. In the only placebo-controlled RCT reporting subgroup analyses in women with
urodynamically proven BOO (based on the Bladder and Groutz nomogram) no statistically significant difference
was observed in the changes of IPSS, IPSS sub scores, Qmax, PVR and bladder diary after eight weeks of
alfuzosin (n = 58) vs. placebo (n = 59). Of note, no EMG and/or voiding cystourethrography was used to
distinguish between dysfunctional voiding and primary bladder neck obstruction [558].

Information on the comparative effectiveness of the different types of alpha-blockers is limited to one RCT.
A small trial on 37 women with IPSS > 8, Qmax < 12 mL/sec and PVR > 50 mL, compared tamsulosin and
prazosin over a 3-month treatment period. More patients on tamsulosin were completely satisfied with their
treatment (16/20 vs. 9/20, p < 0.05). Both treatment groups showed significant improvement in symptom
scores from baseline but no further statistical comparison between the groups was done. However, a larger
decrease in AUA symptom score was seen with the tamsulosin group compared to the prazosin group. More
adverse events were reported with prazosin group (13 cases vs. 1 case) [559].

A small three-arm non-RCT in women with functional BOO compared alfuzosin monotherapy and EMS. The
combination of alfuzosin and EMS showed greater improvement in storage symptoms and QoL with EMS with
or without alfuzosin than with alfuzosin monotherapy alone [550].

4.5.6.2 Striated muscle relaxants


Baclofen is a gamma-aminobutyric acid (GABA) agonist that exerts its effect on the GABAergic interneurons
in the sacral intermediolateral cell column responsible for the relaxation of the striated urinary sphincter during
voiding. Intrathecal administration has been shown to improve voiding in a trial among spinal cord injured
patients. Oral baclofen has also been widely studied [536].

A randomised placebo-controlled crossover trial investigated the efficacy and safety of a 4-week course of
oral baclofen 10 mg 3 times/day in 60 women diagnosed with BOO, based on an increased EMG activity with
a sustained detrusor contraction during voiding. It showed lower number of voids, significant improvements in
Qmax and PdetQmax with baclofen compared with placebo. Post-void residual, maximum cystometric capacity
(MCC) and MUCP parameters were not significantly different between groups. Adverse event rates were also
similar, with the most common complaints including somnolence, dizziness and nausea. An important limitation
of this study was the lack of patient-reported outcomes to assess symptoms and QoL [560].

A small case series reported the outcomes of 20 women with functional BOO who were given oral baclofen
5 mg 3 times/day for 12 weeks. There was significant improvement in the mean voided volume and BVE of
the patients. However, Qmax, PdetQmax, PVR and urethral profile pressures did not significantly change. No
significant adverse events were noted [561].

4.5.6.3 Oestrogens
The relative reduction in urethral wall compliance seen in atrophic urethritis due to oestrogen deprivation
may be responsible for the obstruction in the urethra in post-menopausal women. Oestrogen therapy is thus
theoretically expected to improve the condition. There are no published studies on the use of oestrogens
specifically for the management of female BOO.

4.5.6.4 Sildenafil
Sildenafil, by inhibiting PDE5, increases the levels of nitric oxide in the female urethral sphincter, thereby
promoting urethral relaxation.

A placebo-controlled randomised crossover trial which included 20 females with partial or complete retention
or obstructive voiding, with high MUCP and elevated US-estimated sphincter volume (> 1.6 cm) showed that

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 75


while there was a significant improvement in symptom scores and urodynamic parameters from baseline with
sildenafil treatment, this difference was not significant when compared with placebo [562].

4.5.6.5 Thyrotropin-releasing hormone


Intravenous thyrotropin-releasing hormone has been postulated as a neurotransmitter that induces urethral
relaxation [563]. The exact mechanism is unclear.

In a small RCT of 16 women with voiding difficulty, 8 women (3 with BOO) were randomised to receive 200 ug
intravenous bolus of TRH, and 8 (3 with BOO) received saline. No difference in the decline in functional profile
lengths and maximum urethral closure pressures were noted between treatment groups, despite a significant
decline noted from baseline in the treatment group. No subgroup analysis of women with BOO was reported
[563].

4.5.6.6 Summary of evidence and recommendations for pharmacologic treatment

Summary of evidence LE
Alpha-blocker use is associated with significant improvement in symptom scores from baseline, but 1a
not urodynamic parameters compared with placebo.
Tamsulosin is associated with greater improvement in symptoms score compared with prazosin. 1b
Non-specific alpha-blockers are associated with higher rates of adverse events. 1b
Oral baclofen is better than placebo in improving Qmax and PdetQmax, but not other urodynamic 1b
parameters. Its effects on symptoms are not well reported.
Current evidence does not show that sildenafil is superior to placebo in improving symptoms or 1b
urodynamic parameters of female patients with BOO.
Trials including women with voiding problems of mixed aetiologies showed no difference in 1b
urodynamic outcomes between intravenous thyrotropin-releasing hormone and placebo.

Recommendations Strength rating


Offer uroselective alpha-blockers, as an off-label option, to women with functional bladder Weak
outlet obstruction (BOO) following discussion of the potential benefits and adverse events.
Offer oral baclofen to women with BOO particularly those with increased electromyography Weak
activity and a sustained detrusor contraction during voiding.
Only offer sildenafil to women with BOO as part of a well-regulated clinical trial. Strong
Do not offer thyrotropin-releasing hormone to women with BOO. Strong

4.5.7 Surgical treatment


4.5.7.1 Intra-sphincteric botulinum toxin injection
Botulinum toxin inhibits the presynaptic release of acetylcholine, which results in the reduction of the urethral
sphincter tone. It is also believed to cause a decrease in the release of norepinephrine in the urethra to
counteract external urethral sphincter overactivity [564].

Evidence on the use of botulinum toxin for female BOO is limited to small case series. Most studies
included mixed populations without subgroup analyses, or the diagnosis of voiding dysfunction could not be
ascertained as solely resulting from BOO. No comparative trial exclusively involving female BOO patients using
botulinum toxin has been identified in the literature.

A systematic review including several reports of small case series using variable doses of botulinum toxin A
injected peri-urethrally in females with dysfunctional voiding showed improvements in symptoms, reduction in
residual volumes and reduction in voiding detrusor pressures. Larger series in adults (both males and females,
n > 100) showed success rates of 86–100% [564].

In a randomised, double-blind, placebo-controlled study (n = 73) 100 U onabotulinumtoxinA vs. saline resulted
in significantly lower IPSS scores and larger voided volumes compared with placebo in 31 adults (men and
women with voiding dysfunction [defined by a spinning top appearance on real-time fluoroscopy, poorly
relaxed urethral sphincter on EMG, and a normal-to-high voiding pressure with a low and/or intermittent
urinary flow rate, a PVR volume > 300 mL, and a low voiding efficiency]). Other urodynamic parameters were
comparable between the groups [565]. A subgroup analysis on the female population of this study was not
available.

76 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021


Two small case series on women with BOO reported the effects of 100 U intra-sphincteric injection of BTA.
Both showed improvement in symptom and bother scores and significant reduction in PVR [519, 566]. One
study reported increased Qmax and improved static urethral pressure profile (UPP) [519]. The average symptom-
free duration was noted to be 16.8 weeks in another study [566]. Adverse events included UTI and temporary
need for CISC. No SUI was reported.

4.5.7.2 Sacral nerve stimulation


Sacral nerve stimulation is a type of neuromodulation that allows continuous ES from an electrode placed
alongside a sacral nerve via a surgically implanted pulse generator. The ES is postulated to decrease the
urethral tone. In addition, SNS is also postulated to work by blockage of the inhibitory urethral afferent
impulses, which cause inhibition of normal bladder contraction.

No comparative trial has been identified in the literature on the use of neuromodulation for female BOO.

The majority of the publications on neuromodulation for voiding dysfunction are retrospective reviews of cases,
involving a mix of patient populations who underwent the procedure for different indications. In studies that
indicated a subgroup of patients with urinary retention, there was either no urodynamic confirmation of the
nature of the retention or separate outcomes were not reported for participants with retention.

A review of 60 women who underwent sacral nerve stimulation for urinary retention associated with outlet
obstruction (defined as UPP > 100 cm H2O, increased urethral sphincter volume > 1.8 mL, and abnormal EMG
with repetitive discharges and decelerating bursts) showed an overall spontaneous voiding rate of 72% over
a mean follow up of 4 years. Of those who continued to require CISC up to twice/day post-operatively, the
frequency was less than prior to surgery (degree not specified). There were 99 adverse events and 63 surgical
revisions. In this series, half of the patients underwent a one-stage SNS procedure and the other half a two-
stage procedure. The proportion of patients who required CISC-assisted voiding was higher in the two-stage
group (27% vs. 17%). More serious adverse events (defined as ‘events requiring admission or surgical revision
to resolve issues such as loss of response, lead migration and surgical revisions’) were associated with the
one-stage procedure [562].

A single-centre series in a subgroup of 32 patients diagnosed with idiopathic urinary retention (Fowler’s
syndrome) who underwent SNS, 62.5% achieved a > 50% reduction in the CISC rate [567].

4.5.7.3 Pelvic organ prolapse surgery


Pelvic organ prolapse surgery may relieve BOO by correcting the urethral kinking caused by the prolapse or
by relieving the urethral compression brought about by the prolapsing organ [75, 515, 536]. No comparative
studies on prolapse surgery for female BOO have been published.

Bladder outlet obstruction due to POP may be addressed by corrective surgery. Based on reviews, the majority
of patients who had BOO caused by POP who had a repair of their cystocoele demonstrated improvement of
their voiding difficulties [471, 568].

A multicentre prospective study involving 277 women with at least grade 2 symptomatic POP who underwent
surgery demonstrated a significant reduction in voiding symptoms and PVR volume one year post-operatively
[569].

A retrospective study of 50 women who underwent laparoscopic sacrocolpopexy for POP showed a significant
increase in the mean post-operative Qmax and a decrease in the PdetQmax and PVR in those ≥ 65 years old. The
OAB symptom score (OABSS) improved but there was no significant difference in the ICIQ-SF score post-
operatively [570].

In a case series of 35 females with stage 3-4 POP presenting with a pre-operative PVR > 100 mL (mean
226 mL), 89% had PVR volumes of < 100 mL post-surgery [571]. In another case series of 39 patients with
cystocoeles who complained of voiding symptoms pre-operatively, 30 (79%) achieved normal voiding,
defined as no obstructive symptoms and a PVR below 50 mL, after bladder neck suspension with anterior
colporrhaphy [572].

4.5.7.4 Urethral dilatation


Urethral dilation involves the passage of sequentially greater diameter dilators into the urethra, causing
the obstructing fibrotic tissue to break open and thereby widening the lumen. It is considered the primary

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 77


procedure of choice for women suspected of urethral stricture disease [534]. Dilation of up to 30-40Fr has been
done. There is no standard dilatation technique; dilatation of up to 43F has been described, although other
authors suggest dilating to 30Fr or 35Fr.

A systematic review on female urethral stricture management included 3 trials involving urethral dilatation.
Pooled analysis from these studies with data from 93 females showed a mean success rate of 49% after
urethral dilation to 41Fr with a mean follow up of 46 months. Mean time to failure was 12 months. In treatment-
naïve patients, success rate (as defined by trialists) was 58% while in patients who had undergone previous
dilatation, success rate was 27.2% [530].

An RCT of 50 women with OAB syndrome and associated urodynamically-confirmed BOO (defined as a Qmax
of less than 15 mL/sec with a voided volume of 100 mL or above and/or PVR volume over 200 mL, not due
to a urethral stricture) compared the effect of cystoscopy and bladder distension with urethral dilatation (22)
and cystoscopy only (28) after 6 week follow-up. Significantly more patients who had cystoscopy only had
persistent urgency at 6 weeks and 6 months post-operatively. Urodynamic parameters did not significantly
change pre- and post-operatively in both groups. The greater improvement in QoL scores based on the King’s
Health Questionnaire (KHQ) domain scores seen in the non-urethral dilatation group in this trial should be
interpreted cautiously because of the higher baseline scores in this group. Of note, there were no significant
changes in Qmax, PVR, voided volume or PdetQmax in any of the two groups at 6 weeks questioning the role of
any of these two options for the therapeutic management of BOO. Also, six patients (12%) developed post-
operative SUI [573].

A prospective trial of 86 women with primary urethral stricture compared on-demand vs. intermittent urethral
dilatation to 24Fr (dilate every 2 months). It showed an overall increase in Qmax and decrease in PVR post-
dilatation. Significantly greater improvements were seen in the intermittent urethral dilatation group [574].

Three small case series showed improvements in symptoms with relief of urgency and/or UUI but inconsistent
results in terms of significant improvement in Qmax, PVR and PdetQmax. Benefits were poorly sustained, with the
majority of patients requiring additional or repeat intervention in the long-term [575-577].

Worsening or new-onset SUI is a concern with urethral dilatation but it is less of a concern than after urethrotomy
or surgical reconstruction. Patients have also reported frequency and urgency post-dilatation [577].

4.5.7.5 Urethrotomy
Urethrotomy involves the incision of the urethra endoscopically or using a urethrotome. It addresses the
urethral narrowing by cutting open the scar tissue which is causing the obstruction [75, 515, 536]. No
comparative study has investigated the effectiveness of urethrotomy in female BOO.

A prospective study of 10 females with urethral strictures investigated the effect of Otis urethrotomy to 40Fr
followed by 6 weekly dilatations. There was significant improvement in IPSS, QoL, voided volume, Qmax and
PVR at 6 months. Only the improvements in PVR and QoL were maintained on long-term follow-up (mean 82
months) [575].

4.5.7.6 Bladder neck incision/resection


Transurethral bladder neck incision decreases the resistance at the bladder neck by cutting open the hypertrophic
bladder neck smooth muscle in patients with primary bladder neck obstruction. Transurethral incision of the
bladder neck may be performed with a unilateral incision at 12 o’clock or with bilateral incisions placed at the 5
and 7 o’clock or at 2 and 10 o’clock or at 3 and 9 o’clock positions, or 4 incisions at 3, 6, 9 and 12 o’clock. This
may be done using a resectoscope with a Collin’s knife, a cold knife, or using laser energy. Some authors report
additional resection of the bladder neck between the 5 and 7 o’clock positions during the procedure.

Evidence on bladder neck incision or resection for female BOO is limited to non-comparative trials. A review of
studies on bladder neck incision for the treatment of bladder neck obstruction in women reports success rates
between 76–100% [518].

Bladder neck incision was compared with V-Y-reconstruction using Nesbit’s technique in a retrospective study
of 17 females with BOO, diagnosed by various uroradiological, endoscopic and urodynamic investigations. The
results showed similar symptomatic improvement rates and post-operative PVRs between the two groups. V-Y
plasty was noted to have a longer operative- and catheter time, lower uroradiological improvement rate, higher
transfusion rate, and higher adverse event rate [578].

78 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021


Several prospective case series consistently reported significant improvements in IPSS, QoL, Qmax, PdetQmax
and PVR after treatment compared to baseline, regardless of the site of the incision, type of energy used or the
length of follow up [579-582].

The largest case series with 84 patients diagnosed with primary bladder neck obstruction (based on the lack of
funnel shape of the bladder neck during voiding on voiding cystourethrogram, a Pdet > 20 cm H20 and a Qmax
< 12 mL/sec) showed success in 84.5% of patients with improvement in IPSS, QoL, Qmax and PdetQmax after a
mean follow up of 27.4 months (6–78 months). Complications included vesico-vaginal fistula (VVF) (3.6%), SUI
(4.7%) and urethral stricture (3.6%) [579].

No comparisons have been made between the different incision techniques (location of incision, length, depth,
implement used – cold knife vs. hot knife vs. laser, with or without resection). However, in a case series of 84
patients, complications of VVF and SUI were noted in the cohort of patients who had their incisions at 5 & 7
o’clock, and not in those who had their incisions at 2 & 10 o’clock [579].

Adverse events include SUI, requirement for re-operation and recurrence. Post-operative SUI was reported in
3-33% [518].

4.5.7.7 Urethroplasty/urethral reconstruction


The surgical reconstruction of the female urethra has been used in the management of extensive female
urethral stricture. Several urethroplasty techniques have been reported including the use of vaginal or labial
flaps, as well as vaginal and buccal grafts after cutting open the fibrotic tissue causing the urethral obstruction
[583]. The use of bladder flaps has also been reported [584]. Recently laboratory-engineered tissue grafts have
also been used [585].

The surgical approaches have been described based on the position relative to the urethra; dorsal, ventral or
circumferential. The dorsal approach is believed to provide better mechanical support and a more vascularised bed
for a graft or flap. However, there is greater risk of damage to the sphincter and clitoral bodies with this approach.
The ventral approach is more familiar to most surgeons and requires less urethral mobilisation. However, it is
reported as being more prone to urethra-vaginal fistulae, although it is not clear to what extent [534].

Reviews of studies reporting outcomes of urethroplasties state success rates ranging from 57–100% [534,
586]. Pooled analysis from 6 studies using vaginal or labial flaps showed a mean success rate of 91% with a
mean follow-up of 32 months. Vaginal or labial graft urethroplasty was reported to have an 80% success rate
with a mean follow-up of 22 months.

Oral mucosal grafts, reported in 7 studies, had a mean success of 94% after a mean 15-month follow up [534].
A later review of studies on dorsal buccal mucosal reported graft success rates of 62-100%, pooled success
rate 86% [587]. Stricture recurrence rate in a long-term study with a mean follow-up of 32 months showed a
stricture recurrence rate of 23.1% [586].

A retrospective comparative study on 10 females who underwent urethral dilatation and 12 who underwent
dorsal onlay pedicled labium flap urethroplasty, reported both groups with significant improvements from
baseline in terms of QoL, AUA symptom score, PVR and Qmax. The urethroplasty group had significantly better
QoL scores and Qmax (17.0 vs. 12) on follow-up as compared to the dilatation group [588]. Adverse events
associated with urethroplasties include new-onset SUI and urgency and worsening of UUI.

4.5.7.8 Urethrolysis
Bladder outlet obstruction in females occurring as a complication of surgical procedures for SUI may be
managed surgically by lysis of the urethra aiming to regain urethral mobility. Urethrolysis may involve removal of
peri-urethral anti-incontinence sutures, scar tissue and fibrosis.

No comparative trials have been published on urethrolysis. Case series showed success rates measured as
improved voiding and lower residual volumes, improvement or resolution of symptoms and QoL and improvement
of urodynamic parameters post treatment [589-591]. De novo SUI was reported in 39% in one study [591].

A study on 21 patients who underwent urethrolysis suggested an association of persistent post-operative


bladder symptoms with greater delay in doing urethrolysis. Patients who presented with post-operative storage
and voiding symptoms after a mean of 17 months follow-up had a longer time to urethrolysis compared to
those who had no complaints (31 vs. 9 months) [592].

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 79


4.5.7.9 Removal/excision/section/loosening of mid-urethral sling
In women who develop BOO after placement of a mid-urethral sling, surgical management may include tape
loosening, incision or division, and excision and/or removal of the tape [471].

Several small retrospective reviews of cases using different techniques of sling revision (incision, partial
excision, excision) showed good success rates in terms of symptom reduction, resumption of voiding with
significant reduction in PVRs and improvement of urodynamic parameters. Stress UI recurs in a very small
proportion of patients and often to a lesser degree than prior to the sling procedure. Studies show long-term
efficacy, including preservation of continence.

In a series of 63 females who presented with voiding dysfunction and persistent PVR > 100 mL after tape
surgery for UI, different techniques were compared. Comparisons involved sling revision (sling division (n = 46)
vs. partial sling excision (n = 13) vs. sling revision (division or excision) with an additional anti-SUI procedure
(n = 4). The authors reported an overall success rate of 87% (success defined as PVR < 150 mL). No
statistically significant difference in success rates was demonstrated across the different revision techniques.
There was a higher need for surgery for recurrent SUI in the partial sling excision group without an anti-SUI
procedure (23% vs. 2.2 and 0) [593].

4.5.7.9.1 Timing of sling revision


One study showed that patients who underwent surgical release more than 180 days after initial anti-UI surgery
had significantly less recurrent SUI as compared to patients who underwent the release sooner (15% vs. 46%,
p = 0.0008) [594].

4.5.7.10 Summary of evidence and recommendations for surgical management of BOO

Summary of evidence LE
Intrasphincteric injection of botulinum toxin results in the improvement of symptoms and urodynamic 2
parameters.
Sacral nerve stimulation results in spontaneous voiding and a reduction in CISC rate in the majority of 3
female BOO patients in idiopathic urinary retention.
More serious adverse events and surgical revisions were associated with the one-stage neuromodulator 3
implantation procedure.
Repair of pelvic organ prolapse improved PVR volume and voiding symptoms. 3
Urethral dilatation in women with BOO results in significant improvement in OAB symptoms, but 1b
improvements in urodynamic parameters of voiding are inconsistent.
Programmed intermittent urethral dilatation results in better outcomes compared with on demand 3
dilatation.
Effects of urethral dilation are poorly sustained, requiring repeat intervention in the long term. 3
Internal urethrotomy followed by regular dilatations resulted in significant improvement in symptoms 3
and urodynamic parameters in women with BOO.
Bladder neck incision in females with BOO results in improvements in symptoms and urodynamic 3
parameters.
Complications of bladder neck incision are not common, but include vesico-vaginal fistula, SUI, and 3
urethral stricture.
Urethroplasty using grafts or flaps in women with BOO due to urethral stricture have good success 3
rates with significant improvements of symptoms, QoL scores and urodynamic parameters compared
to baseline.
Urethroplasty results in better QoL and Qmax compared to urethral dilatation. 2
Long-term results showed significant stricture recurrence rate after urethroplasty. 3
Urethrolysis performed on women with voiding problems after anti-UI surgery resulted in 3
improvements in symptoms, QoL and urodynamic parameters post-operatively.
Delayed urethrolysis was associated with persistent post-operative bladder symptoms. 3
Sling revision in women who presented with urinary retention or voiding problems and significant 3
PVRs after sling surgery for UI resulted in improvements in symptoms and urodynamic parameters,
resumption of voiding and reductions in PVRs.
Sling revision is associated with the risk of recurrent SUI. 3

80 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021


Recommendations Strength rating
Offer intrasphincteric injection of botulinum toxin to women with functional bladder outlet Weak
obstruction (BOO).
Offer sacral nerve stimulation to women with functional BOO. Weak
Advise women with voiding symptoms associated with pelvic organ prolapse (POP) that Weak
symptoms may improve after POP surgery.
Offer urethral dilatation to women with urethral stenosis causing BOO, but advise on the Weak
likely need for repeated intervention.
Offer internal urethrotomy with post-operative urethral self-dilatation to women with BOO Weak
due to urethral stricture disease but advise on its limited long-term improvement and the
risk of post-operative urinary incontinence (UI).
Do not offer urethral dilatation or urethrotomy as a treatment for BOO to women who have Weak
previously undergone mid-urethral synthetic tape insertion due to the theoretical risk of
causing urethral mesh extrusion.
Inform women of limited long-term improvement (only in terms of post-void residual and Weak
quality of life) after internal urethrotomy.
Offer bladder neck incision to women with BOO secondary to primary bladder neck Weak
obstruction.
Advise women who will undergo bladder neck incision on the small risk of developing stress Strong
urinary incontinence (SUI), vesico-vaginal fistula or urethral stricture post-operatively.
Offer urethroplasty to females with BOO due to recurrent urethral stricture after failed Weak
primary treatment.
Caution women on the possible recurrence of strictures on long-term follow-up after Weak
urethroplasty.
Offer urethrolysis to women who have voiding difficulties after anti-UI surgery. Weak
Offer sling revision (release, incision, partial excision, excision) to women who develop Strong
urinary retention or significant voiding difficulty post tape surgery for UI.
Caution women about the risk for recurrent SUI and the need for a repeat/concurrent anti-UI Strong
surgery after sling revision.

4.5.8 Follow up
Women with BOO should be followed up and monitored regularly due to the risk of further deterioration of
voiding or renal function in case of persistence and progression of the obstruction. For those who received
treatment, monitoring must be done for the recurrence of the BOO. In particular, women who underwent
urethral dilation, urethrotomy or urethroplasty for urethral stricture need to be monitored for the recurrence of
the stricture.

4.6 Nocturia
Nocturia was defined by the ICS in 2002 as ‘the complaint that the individual has to wake at night one or more
times to void’ and quantified in an updated document in 2019 as ‘the number of times an individual passes
urine during their main sleep period, from the time they have fallen asleep up to the intention to rise from that
period’ [595]. A systematic review of the literature in this topic area has been conducted by the EAU Guidelines
Panel on Urinary Incontinence covering publications up to and including 2017 [596]. This was supplemented
with a scoping search in 2020 covering more recent publications.

4.6.1 Epidemiology, aetiology, pathophysiology


The prevalence of nocturia varies according to age with around 4–18% of women aged 20-40 experiencing
2 or more episodes per night, compared to 28–62% of women aged 70 or older [597]. In a study of 1,000
community-dwelling older adults, nocturia in females was associated with older age, African-American race,
a history of UI, swelling of the lower limbs and hypertension [598]. A report on over 5,000 adults aged 30–79
years identified around 28% with nocturia and found additional correlates with increased BMI, cardiac disease,
type 2 diabetes and diuretic use [599]. A recent systematic review and meta-analysis has concluded that
nocturia is probably associated with an approximate 1.3 fold increased risk of death [600].

The aetiology of nocturia is multifactorial and can include both urological and non-urological causes. Urological
conditions which may exhibit nocturia as a significant symptom include OAB syndrome, BOO, DU and
dysfunctional voiding. Non-urological causes include 24 hour polyuria (which includes nocturnal polyuria),
congestive heart failure, sleep apnoea, restless leg syndrome, peripheral vascular disease, sleep disorders,
night-time food ingestion, dependent oedema and excessive fluid intake [601]. Given the varied aetiology of

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 81


this symptom there are a range of possible pathophysiological mechanisms. These include: (1) 24 hour polyuria
(e.g. diabetes mellitus, primary polydipsia, and diabetes insipidus); (2) nocturnal polyuria (e.g. behavioural,
peripheral oedema, obstructive sleep apnoea, glycosuria, hormonal abnormalities and cardiac dysfunction);
(3) diminished bladder capacity (e.g. OAB syndrome/DO, pelvic floor dysfunction, BOO, pharmaceuticals, LUT
calculi or tumours and neurological bladder dysfunction); and (4) primary or secondary sleep disorders [602].

4.6.2 Classification
Classification of nocturia is dependent on bladder diary analysis and several parameters have been defined as
important [603]:
• Nocturnal urine volume - total volume of urine passed during the night (this includes the 1st morning void
but does not include the last void prior to sleep);
• Maximum voided volume - the largest single voided volume in 24 hours;
• Nocturia index - the nocturnal urine volume divided by the maximum voided volume;
• Nocturnal polyuria index - the nocturnal urine volume divided by the 24 hour urine volume;
• Nocturnal urine production - the nocturnal urine volume divided by the duration of sleep in hours.

The analysis of these parameters will allow the clinical classification of nocturia based on the physiological
abnormalities that can cause nocturia:
• 24 hour polyuria;
• nocturnal polyuria;
• diminished bladder capacity;
• sleep disorders.

4.6.3 Diagnostic evaluation


The evaluation of nocturia should include a thorough medical history and physical examination with
particular reference made to a history of sleep disorders, fluid balance, associated LUTS, cardiovascular and
endocrine comorbidities, renal disease, current medications and previous urological history [603]. Several
nocturia-specific symptom scores exist such as the ICI Questionnaire-nocturia, the Nocturia Quality of Life
Questionnaire (N-QoL) and the Nocturia Impact Diary, some of which were developed in men. A further
screening tool is also available which aims to identify causes of nocturia; the Targeting the individual’s
Aetiology of Nocturia to Guide Outcomes (TANGO) assessment tool [604-606].

A bladder diary is a vital initial investigation in any patient that is complaining of nocturia and further
supplementary investigations are guided by any abnormalities identified. Bladder diary analysis can allow for
the calculations of the parameters detailed in Section 4.6.2. A low nocturnal bladder capacity or global bladder
capacity will be highlighted by reduced voided volumes either during nocturnal hours or both night and day.
This may suggest an underlying urological condition such as OAB syndrome, BOO or DU. Global polyuria
is defined as a 24-hour urine production of more than 40 mL/kg [607] and may be present in conditions
such as diabetes mellitus or diabetes insipidus. The definition of nocturnal polyuria is age dependent and
the thresholds for this diagnosis range from 20% (in younger persons) to 33% (in those over 65) of the 24
hour urine volume being produced during sleeping hours. This may also be observed in patients with loss of
circadian rhythm, cardiovascular disease, sleep apnoea or sleep disorders [603]. A large study conducted
across European and American centres involving almost 2,000 patients has identified nocturnal polyuria as a
contributory cause of nocturia in 89% of patients who were being treated for LUT abnormalities such as OAB
syndrome or benign prostatic enlargement [608].

As an alternative to a three, or more, day bladder diary a nocturnal-only diary has been investigated in
men [609]. Overall, results showed acceptable sensitivity and specificity from the nocturnal bladder diary
in comparison with the standard bladder diary for the majority of parameters. The nocturnal-only diary was
obviously not able to diagnose 24 hour polyuria and has not yet been validated for use in women.

4.6.3.1 Summary of evidence and recommendations for the diagnosis of nocturia

Summary of evidence LE
A thorough medical history is an integral part of the evaluation of women presenting with nocturia. 4
Nocturia-specific questionnaires are sensitive to symptom changes. 3
A bladder diary allows for calculation of important indices and can identify potential causes of nocturia. 3
Nocturnal-only bladder diaries have been evaluated in men only. 3

82 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021


Recommendations Strength rating
Take a complete medical history from women with nocturia. Strong
Use a validated questionnaire during the assessment of women with nocturia and for Weak
re-evaluation during and/or after treatment.
Use a three-day bladder diary to assess nocturia in women. Strong
Do not use nocturnal-only bladder diaries to evaluate nocturia in women. Weak

4.6.4 Disease management


When evaluating the results of trials involving treatment strategies for nocturia it is vital to examine for clinical
significance as statistical significance can be achieved with very small reductions in nocturia episodes.

4.6.4.1 Conservative management


The individual components of self-management have not been critically evaluated and most recommendations
are traditionally derived from consensus methodology. Interventions such as those listed below may help with
nocturia but, for the majority, no quantification of their effect is possible:
• reduction of fluid intake at specific times;
• avoidance/moderation of intake of caffeine or alcohol;
• distraction techniques;
• bladder retraining;
• PFMT;
• reviewing medication;
• treatment of constipation.

The available data for conservative treatment of nocturia exhibit significant heterogeneity. In the EAU
systematic review [596], three studies [610-612] were favourable for conservative treatment with PFMT, with
another failing to confirm benefit [613].

The highest level of evidence comes from a study of 131 patients (as a secondary analysis from a prospective
RCT which had urgency-predominant UI as the primary inclusion criterion) and found that training in PFM
contraction, which included 4 sessions of biofeedback-assisted PFMT reduced nocturia by a median 0.50
episodes per night and was significantly more effective than anticholinergic drug treatment or placebo [610].
The certainty of evidence associated with this treatment is moderate.

A smaller RCT of 50 women with “urinary complaints”, randomised 1:1 to bladder training and PFMT compared
to a control group receiving no treatment, showed a significant decrease in patients’ complaints of nocturia
[611]. Another RCT in only 24 women compared PFMT only to transcutaneous electrical nerve stimulation
therapy (TENS) plus PFMT [612]. Although the authors did not find significant differences between the groups,
the change in nocturia episodes before and after treatment was statistically significant in both groups. This
study was underpowered by the authors’ own admission. The level of certainty of the evidence from these two
trials is low.

In a secondary analysis from a prospective RCT, 210 women with UUI were evaluated for change from baseline
in the number of episodes of nocturia and nocturnal incontinence between groups allocated to medical
treatment (tolterodine ER 4 mg) alone vs. medical treatment plus PFMT [613]. No significant difference between
the groups was found and the actual difference in nocturia episodes in either treatment arm was small. The
level of certainty of the evidence from this trial is low.

A recent RCT has explored both individual and group PFMT with a specific secondary outcome of number of
patients with two or more nocturia episodes per night [328]. The authors reported similar reductions with over
30% of patients who had two or more episodes of nocturia at baseline no longer experiencing this level of
symptoms at one year following PFMT.

One small, single-centre RCT in which functional magnetic stimulation (FMS) was compared to no treatment
in 39 women reported a significant decrease in nocturia (together with voiding frequency and pad use) in the
treatment group compared to the control group [614].

In patients with obstructive sleep apnoea who complain of nocturia, treatment with continuous positive airway
pressure has been shown to be effective in a systematic review and meta-analysis of 5 RCTs involving both
male and female patients [615]. This treatment was associated with an average numerical reduction in nocturia
of over two episodes per night.

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 83


4.6.4.1.1 Summary of evidence and recommendations for the conservative management of nocturia

Summary of evidence LE
Individual or group PFMT appear to be equally effective in terms of reduction in nocturia episodes. 1b
The majority of studies evaluating PFMT for nocturia in women with additional urinary symptoms have 1b
shown positive results both in comparison to placebo and to anticholinergic drugs.
Treatment of nocturia secondary to obstructive sleep apnoea with continuous positive airway pressure 1a
results in reductions in nocturia episodes.

Recommendations Strength rating


Offer women with lower urinary tract symptoms (LUTS) lifestyle advice prior to, or concurrent Strong
with, treatment.
Offer pelvic floor muscle training for nocturia (either individually or in the group setting) to Strong
women with urinary incontinence or other storage LUTS.
Offer women with nocturia and a history suggestive of obstructive sleep apnoea a referral Strong
to a sleep clinic for an assessment of suitability for continuous positive airway pressure
treatment.

4.6.4.2 Pharmacology management


4.6.4.2.1 Desmopressin
Desmopressin is a synthetic analogue of the hormone vasopressin and is most often used for management
of nocturia due to nocturnal polyuria. In the recent systematic review [596] 3 trials specifically conducted in
women were found but more additional data could be extracted from studies in mixed populations. The earliest
evidence comes from a 1982 single-site crossover trial involving 25 women treated with either 20 mcg of
desmopressin or placebo revealed a significant decrease in nocturnal urine output at 6 weeks [616]. A more
recent multicentre, multinational double-blind RCT involving 141 women used desmopressin in doses 0.1, 0.2,
0.4 mg orally at bedtime after a dose-titration period [617]. This increases the likelihood of a positive outcome
because non-responders were excluded at this stage. At 3 weeks significant reductions in nocturnal urinary
frequency and nocturnal diuresis were reported. In another multicentre double-blind RCT a total of 58 women
were randomised into 5 groups (12 receiving placebo, 12 receiving desmopressin 10 μg, 11 receiving 25 μg, 11
receiving 50 μg and 12 receiving 100 μg) for 4 weeks [618]. A dose-response relationship was observed and
female patients appeared more sensitive to desmopressin. Statistically significant changes in nocturnal urine
volumes were reported in favour of the higher desmopressin dose. Differences in the nocturnal polyuria index
also tended to favour desmopressin over placebo and favoured the higher desmopressin dose. The level of
certainty of the evidence from these 3 trials is low.

Desmopressin can be safely combined with anticholinergics with significant benefit in women with OAB and
nocturnal polyuria, as shown by a multicentre RCT of 97 patients [619]. A post-hoc analysis of data comparing
3-month once-daily combination (desmopressin 25 μg/tolterodine 4 mg, n = 49) or monotherapy (tolterodine
4 mg/placebo, n = 57) revealed a significant reduction in nocturnal void volume and time to first nocturnal void
in favour of combination therapy. The level of certainty of the evidence from this trial is moderate.

Pooled data from three RCTs were used to examine the adverse event profile of desmopressin, specifically
hyponatraemia [620]. The authors reported that the majority tolerate desmopressin treatment without clinically
significant hyponatremia, but risk increased with age and lower baseline serum sodium concentration. They
advised that desmopressin treatment in elderly patients should include careful monitoring of the serum sodium
concentration and should be avoided in patients with a baseline serum sodium concentration below normal
range [620].

4.6.4.2.2 Anticholinergics
The systematic review [596] identified 3 RCTs involving anticholinergics such as oxybutynin 2.5 mg/day [610]
and tolterodine 4 mg/day [613, 619]. A secondary analysis from a prospective RCT involving 131 women with
nocturia followed up for 8 weeks found that women receiving 2.5 mg once a day immediate-release oxybutynin
(with the possibility of self-titration and dose escalation to 5 mg three times a day) had less nocturia episodes
than women receiving placebo [610]. Women receiving oxybutynin plus behavioural therapy also exhibited
a statistically significant decrease in nocturia episodes compared to both placebo and oxybutynin alone. A
multicentre RCT which included 305 women followed up for 8 weeks examined the efficacy of tolterodine
tartrate 4 mg alone or in combination with behavioural training [613]. Statistically significant differences

84 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021


compared to baseline were observed in mean nocturia episodes and nocturnal incontinence episodes in
both groups, but no difference was reported between the two treatment groups. The level of certainty of the
evidence from this trial is moderate.

In an RCT including 97 women with nocturnal polyuria and OAB syndrome, comparing three months of once-
daily combination (desmopressin 25 μg/tolterodine 4 mg, n = 49) or monotherapy (tolterodine 4 mg/placebo,
n = 57) a significant reduction in mean number of nocturnal voids compared to baseline was reported in both
groups [619]. The level of certainty of the evidence from this trial is moderate.

A well-designed large comparative study followed 407 women with OAB and nocturia for 4 weeks [621]. The
patients were given tolterodine as monotherapy in one group, and tolterodine combined with estazolam (a
benzodiazepine) in the other group for four weeks. Significant changes from baseline in both groups for the
main outcome of number of nocturia episodes were reported. The combination showed a significant benefit
for women with OAB and nocturia compared to monotherapy in terms of differences in number of nocturia
episodes per night, urgency episodes in 24 hours, UUI episodes in 24 hours, voided volume per micturition.
The level of certainty of the evidence from this trial is very low.

4.6.4.2.3 Oestrogens
In the recent systematic review [596] only a single RCT investigating the efficacy of oestrogen for nocturia was
identified [622]. This trial compared an oestradiol-releasing vaginal ring with an oestriol vaginal pessary in 251
women followed up for 6 months. There was no difference between the treatment groups in the number of
women reporting nocturia though they reported significant change from baseline in both treatment arms with
over 50% of subjects responding in each arm. The certainty of evidence for this outcome was low.

4.6.4.2.4 Diuretic treatment


In a randomised placebo-controlled study an afternoon dose of 40 mg of furosemide (taken 6 hours before
bedtime) in an attempt to establish and complete a diuresis before bedtime was given to elderly men [623]. In
the 43 men who completed the study, night-time frequency in the furosemide group fell by 0.5 compared to
placebo, and percentage night-time voided volume fell by 18%. No such study has been carried out in female
patients.

4.6.4.3 Surgical management


Surgical treatment is in general reserved for those with underlying correctable LUT disorders. The effect of
surgical treatments on symptom of nocturia can be found in the relevant condition-specific sections of this
guideline.

4.6.4.4 Summary of evidence and recommendations for the pharmacological management of nocturia

Summary of evidence LE
Desmopressin treatment for nocturia shows significant reductions in nocturnal urine output, nocturnal 1b
urinary frequency and nocturnal polyuria index.
The majority of nocturia patients tolerate desmopressin treatment without clinically significant 1a
hyponatremia; however, the risk increases with increasing age and decreasing baseline serum sodium
concentration.
Treatment of nocturia in OAB patients with anticholinergic drugs shows reduction in nocturia episodes. 1b
Combination of PFMT and pharmacological treatment with anticholinergics does not appear to confer 1b
additional benefit over anticholinergics alone.
Combination of anticholinergic and desmopressin treatment appears to reduce nocturnal voided volume 1b
and time to 1st nocturnal void in women with nocturnal polyuria.
Vaginal oestrogen may be beneficial in the treatment of nocturia in around 50% of women. 1b
Afternoon (timed) diuretic treatment with furosemide reduces nocturia episodes and nocturnal voided 1b
volume in men but no similar studies have been conducted in women.
Examination for clinical significance is important when evaluating the results of trials involving treatment 3
strategies for nocturia as statistical significance can be achieved with very small reductions in nocturia
episodes.

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 85


Recommendations Strength rating
Offer desmopressin treatment for nocturia secondary to nocturnal polyuria to women Strong
following appropriate counselling regarding the potential benefits and associated risks
(including hyponatremia).
Carefully monitor serum sodium concentration in elderly patients treated with Strong
desmopressin. Avoid prescribing desmopressin to patients with a baseline serum sodium
concentration below normal range.
Offer a anticholinergic treatment for nocturia to women with urgency incontinence or other Strong
storage lower urinary tract symptoms following appropriate counselling regarding the
potential benefits and associated risks.
Inform women with nocturia that the combination treatment with behavioural therapy and Weak
anticholinergic drugs is unlikely to provide increased efficacy compared with either modality
alone.
Offer combination treatment with anticholinergics and desmopressin to women with OAB Weak
and nocturia secondary to nocturnal polyuria following appropriate counselling regarding
the potential benefits and associated risks.
Offer vaginal oestrogen treatment to women with nocturia following appropriate counselling Weak
regarding the potential benefits and associated risks.
Offer timed diuretic treatment to women with nocturia secondary to polyuria following Weak
appropriate counselling regarding the potential benefits and associated risks.

4.6.5 Follow-up
The follow-up of patients with nocturia will be dependent on both the underlying aetiology of this symptom and
the treatment given.

4.7 Pelvic organ prolapse and LUTS


4.7.1 Epidemiology, aetiology, pathophysiology
Pelvic organ prolapse is a common condition in adult women. The prevalence of POP ranges from 3-6% when
bothersome symptoms are used to characterise the condition and goes up to as high as 50% when a purely
anatomical definition is used [624].

The lifetime risk for POP surgery is an estimated 12.6% [625]. Parity, vaginal delivery, ageing and obesity are
the most commonly recognised risk factors [626].

Although the aetiology of POP is not fully understood, birth trauma to the levator ani complex is recognised as
central to its development. In normal physiology an intact levator ani complex functionally closes the genital
hiatus surrounding the vagina limiting the pressure gradient between the intra-abdominal and intravaginal
areas. During physical activities this reduces the stress on the endopelvic fascia and its condensations
(e.g. ligaments), which are crucial in securing the bladder, uterus and rectum to their surroundings. Current
aetiological concepts include widening of the levator hiatus due to birth trauma, which will create a low-
pressure area in the vagina and consequently an increased stress on the ligaments, fascial elements and PFMs
during physical activity. When the supporting function of the muscles and connective tissues fail, POP may
develop [627]. This concept also explains the time lapse between birth trauma and the occurrence of POP.

Pelvic organ prolapse and LUTS often occur simultaneously in women. Although in isolation both POP and
LUTS are already prevalent conditions in women, the prevalence of LUTS in women with POP exceeds that
of LUTS in women without POP symptoms [624]. The observation that LUTS symptoms may improve, or
worsen, after POP treatment also suggests a link between these two entities [624]. Clinical examples include
the occurrence of BOO symptoms in the case of a severe POP, and the disappearance of SUI symptoms with
progression of POP (and conversely the occurrence of SUI after treatment of POP) [628].

4.7.2 Classification
Since 1996 POP has been classified according to the Pelvic Organ Prolapse-Quantification (POP-Q) system
[629]. For specifics on how to perform the POP-Q measurement and the 9 standard points to be measured, as
shown in Figures 2 and 3, we refer to the original publications [629, 630].

The vagina is divided into anterior (bladder), posterior (rectum) and apical (cervix or vaginal vault)
compartments. After scoring the position of the 9 POP-Q points, a prolapse of each compartment is graded
numerically from stage 0 to 4, with stage 0 being no prolapse and stage 4 being a complete eversion of
the uterus/vaginal vault. A crucial landmark in staging of the POP is the hymenal remnant. Any POP with

86 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021


a maximum descent that is still 1 cm above the hymen (e.g. in the vagina) is considered a stage 1 POP. A
maximum descent between 1 cm above and 1 cm below (outside the vagina) the hymen is a stage 2 POP. Any
descent beyond 1 cm below the hymen is a stage 3 POP.

The figures below show the POP-Q staging in comparison to the Baden Walker system (and others) used
before the international consensus on the POP-Q staging was introduced as the new standard.

Figure 2: Prolapse classification system*

*Figure reproduced with permission from the publisher, from Theofratus JP, et al. [630].

Figure: 3: Pelvic Organ Prolapse-Quantification staging*

*Figure reproduced with permission from the publisher, from Bump RC, et al, [629]. The standardisation of
terminology of female pelvic organ prolapse and pelvic floor dysfunction.

4.7.3 Diagnostic evaluation


Pelvic organ prolapse is a clinical diagnosis and is staged according to the POP-Q system. In general, a
POP that is above the hymen should only produce mild symptoms at most [631]. In cases where there is a
discrepancy between the clinical symptoms and POP-Q staging, it is advised to consider performing the
POP-Q measurement in a standing position rather than supine, or re-evaluate at a later time in the day.
Magnetic resonance imaging assessment demonstrated a marked difference in POP staging between supine
and standing position [632]. Additional diagnostic tests for POP are mainly indicated if there are accompanying
symptoms like LUTS or bowel dysfunction. Imaging techniques are not advised for the routine diagnostic work-
up of patients presenting with POP [67]. The role of urodynamics in the diagnostic work up of SUI has been
discussed in the SUI Section of this guideline.

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 87


The use of techniques to reduce the POP during urodynamic evaluation in order to diagnose occult SUI is
common practice. This information may be used to decide if additional anti-UI surgery should be offered at the
time of POP surgery or to counsel patients on the possible after-effects of POP treatment.

There are several POP reduction methods that may be used during physical examination or urodynamic
evaluation. In a multicentre observational study, five different cough/stress tests were compared for their
ability to detect SUI in women with POP [633]. Stress urinary incontinence during at least one of the five
tests occurred in 60/205 (29.2%) of women without SUI symptoms. Looking at single test performance, the
detection rate of occult SUI in women without SUI symptoms increases from 4.4% in case of no reduction, to
22% in case of reduction with a pessary.

A large randomised trial which included women with POP without symptoms of SUI, who were randomised to
sacrocolpopexy with or without Burch colposuspension [634]. Three hundred and twenty two stress-continent
women with stages 2–4 prolapse underwent standardised urodynamic testing, and the protocol included
five prolapse reduction methods. Pre-operatively, 12 of 313 (3.7%) women demonstrated urodynamic SUI
without prolapse reduction. Pre-operative detection of urodynamic SUI with prolapse reduction at 300 mL
was by pessary, 6% (5 of 88); manual, 16% (19 of 122); forceps, 21% (21 of 98); swab, 20% (32 of 158); and
speculum, 30% (35 of 118). Another large trial included women with POP without SUI symptoms randomised
to vaginal POP surgery with or without (sham incision) MUS [635]. Before surgery, 33.5% (111/331) of women
demonstrated SUI at a prolapse-reduction cough stress test. In an observational study of 172 women with POP
without SUI, 19% of women were diagnosed with occult SUI on basic office evaluation (with prolapse reduction
with swab on forceps) and 29% on urodynamic evaluation [636].

In summary, SUI can be demonstrated in women with POP without symptoms of SUI after POP reduction in up
to 30% of cases. There is no consensus on the best reduction technique.

Although the detection rate of occult SUI increases after reduction of POP in SUI-symptom-negative women,
its clinical value is under debate.

In one trial, pre-operative stress continent women were evaluated during urodynamic testing with prolapse
reduction to determine if they were more likely to report post-operative SUI, regardless of concomitant
colposuspension (controls 58% vs. 38% [p = 0.04] and Burch 32% vs. 21% [p = 0.19]) [634]. In another trial,
women with SUI during the cough stress test after POP reduction reported UI at 3 months in 29.6% in the sling
group, compared with 71.9% in the sham group (adjusted OR: 0.13) [635]. Women with a positive prolapse-
reduction stress test before surgery appeared to receive more benefit from a sling at 3 months, but not at 12
months, than did those with a negative test.

In a large observational study women did not receive additional anti-UI surgery even if they had SUI after POP
reduction pre-operatively. In this scenario 9% (16/172) of all women developed post-operative SUI and six
women (4%) underwent surgery for de novo SUI [636]. Women with demonstrable pre-operative SUI were more
at risk of post-operative SUI: 28% vs. 5% (diagnostic OR: 7). Based on urodynamic evaluation only, one more
woman was predicted to have post-operative SUI, but all 6 women who underwent treatment for de novo SUI
showed SUI during basic office evaluation.

In a model developed to predict the risk of de novo SUI in women undergoing POP surgery based on findings
from two trials, a total of 12 pre-operative predictors were tested [637]. A positive finding of SUI during a pre-
operative prolapse reduction test was included in this model, but it failed to be of significant predictive value as
a single item. In addition, pre-operative POP stage was not associated with the risk of de novo SUI.

4.7.3.1 Summary of evidence and recommendation for the detection of SUI in women with POP

Summary of evidence LE
POP reduction during cough stress test, in office or during UDS, will detect SUI in approximately 30% 2a
of continent women.
Women with SUI after POP reduction pre-operatively (occult SUI) are likely to be at increased risk of 2a
developing SUI symptoms after POP surgery.

88 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021


Recommendation Strength rating
Perform a pelvic organ prolapse (POP) reduction test in continent women to identify those Strong
with occult stress urinary incontinence and counsel them about the pros and cons of
additional anti-incontinence surgery at the time of POP surgery.

4.7.3.2 Urodynamics in women with POP and LUTS (without stress urinary incontinence).
The role of urodynamics is less clear in women presenting with POP and concurrent LUTS, other than SUI.
Pelvic organ prolapse is a complex condition incorporating not only different compartments of the vagina,
but also presenting at different stages in terms of severity. Information about detrusor activity, as assessed
with urodynamics, may provide information about the risk of developing DO after surgery, but also on the
risk of urinary retention due to DU. An observational study assessed predictors for DO following POP surgery
for POP-Q stage 3 or higher in 1,503 women and the authors concluded that a pre-operative maximum
urethral closure pressure of ≥ 60 cm H2O, a Qmax of < 15 mL/sec, a maximum detrusor voiding pressure (Dmax)
≥ 20 cm H2O and a PVR volume of ≥ 200 mL were independent risk factors for post-operative DO [638]. A
small observational study (n = 49) evaluated those with pre-operative DU (detrusor pressure at maximum
flow was ≤ 10 cm H2O and Qmax of ≤ 12 mL/s) after POP surgery. Surgery objectively “cured” DU in 47% of
women and urodynamic findings normalised after surgery [639]. The 2019 NICE Guidelines do not include a
recommendation to perform urodynamics as part of the diagnostic workup of POP, except for the combination
with symptomatic SUI [67].

4.7.4 Disease management


Pelvic organ prolapse symptoms can be treated with PFMT, vaginal pessary use, surgery or a combination of
these treatments. The scope of this guideline is to focus on LUTS in women and therefore only data on the
effect of treatment of urinary symptoms are presented.

4.7.4.1 Conservative treatment of pelvic organ prolapse


The 2013 NICE guideline on Urinary Incontinence and Pelvic Organ Prolapse in Women updated its
management section in 2019, including a full evidence review [67]. The overall conclusion with respect to
conservative treatment for POP was that the evidence is of low quality. A total of 13 RCT’s were identified.
Seven studies presented data on changes in urinary symptoms [640-646]. An additional search identified 4
RCT’s that addressed the addition of PFMT to POP surgery [647-650], and one that compared combined
PFMT/Pilates therapy with lifestyle advice by leaflet [651].

Five studies [641, 643-645, 651] compared PFMT to lifestyle advice/leaflet, one study [642] compared PFMT to
PFMT with pessary, one study [646] PFMT to pessary therapy, and five studies compared surgery for POP with
or without the addition of PFMT [640, 647-650].

4.7.4.1.1 Pelvic floor muscle training versus lifestyle advice


An RCT (n = 109) reported that at 6 months follow-up the ICIQ-UI-SF scores improved in favour of the PFMT
group compared with a control group receiving lifestyle advice only (difference from baseline PFMT 2.40
points and control 0.2 points, p = 0.002) [641]. However, the difference of 2.4 points from baseline, in favour of
PFMT, has to be viewed with caution as the mean baseline score in the PFMT group was higher compared to
the control group (7.4 vs. 5.9, p = 0.05). Likewise, it has to be noted that the absolute ICIQ-UI-SF values at 6
months follow-up were not significantly different between groups (PFMT 4.8 vs. control 5.2).

Two publications from one RCT reported on the 3, 6 and 12 month results of lifestyle advice only vs. lifestyle
advice combined with group PFMT [643, 644]. The UDI-6 and UIQ-7 questionnaires were used to assess
urinary symptoms. At 3 months follow-up both groups (53 women in the lifestyle group and 56 in the lifestyle
+ PFMT cohort) reported significantly improved UDI-6 scores whilst the lifestyle-only group also reported
significantly greater improvement on the UIQ-7. Between-group comparison showed no differences in UDI-6
and UIQ-7 scores at 6 months. At 12 months follow-up, the majority of women had sought additional treatment
(70% in the lifestyle-only group and 48% in the lifestyle/PFMT group, p = 0.05). The number of patients
remaining on the original therapy was too small to reach strong conclusions.

One RCT reported on the 6 and 12 months follow-up of 225 women with POP-Q stage 1–3 randomised to
individualised PFMT and 222 women randomised to lifestyle leaflet information only (control) [645]. Urinary
symptoms were assessed with a single question on the existence of UI, a single question regarding the need to
strain to void and a single question regarding incomplete bladder emptying which were supplemented with the
ICIQ-SF questionnaire score. At 6 months, significantly more women in the control group reported UI, the need
to strain to empty bladder and the feeling of incomplete emptying compared to the PFMT group. The score on

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 89


the ICIQ-SF was also significantly worse in the control group as compared to the PFMT group. However, at
12 months there was no significant difference on these items between groups. It has to be noted that 50% in
the control group received additional treatment within the 12-month study period. Twenty-seven percent had
additional PFMT, which may have had an effect on the 12-month data.

Another RCT reported on the 24-month follow-up of 414 women with stage 1–3 POP (207 assigned to PFMT/
Pilates and 207 to lifestyle advice) [651]. Urinary symptoms were assessed with the ICIQ-UI-SF and a question
about UI and difficulty emptying the bladder. At 24 months the ICIQ-UI-SF score was significantly better in the
intervention group (mean difference -0.83, p = 0.008). However, the proportion of women reporting any UI did
not differ between groups, nor did the number of pads used weekly.

4.7.4.1.2 Pelvic floor muscle training versus pelvic floor muscle training with pessary
One RCT compared PFMT alone to PFMT and pessary therapy for symptomatic POP [642]. Urinary tract
symptom changes were assessed using the Urogenital Distress Inventory-6 (UDI-6) and the Urinary Impact
Questionnaire (UIQ) at 6 and 12 months follow-up. At 12 months follow-up there was no difference in the
between-group comparison. With respect to the UIQ, women in the pessary/PFMT showed a significant
improvement from baseline, but the PFMT-only group did not. Women in the pessary/PFMT group reported
significantly more frequent de novo SUI (48% vs. 22%), and also more improvement of pre-existing voiding
difficulty (62.5% vs. 35.5%).

4.7.4.1.3 Pelvic floor muscle training versus pessary only


One RCT reported on the 24-month follow-up of 82 women with symptomatic POP randomised to pessary
therapy and 80 women randomised to PFMT [652]. The UDI-6 was used as the outcome measure for urinary
symptoms. Both in the ITT and per protocol analyses the UDI score did not differ significantly between groups
at 24 months of follow-up.

4.7.4.1.4 Surgery versus surgery with pelvic floor muscle training


An assessor-blinded RCT compared surgery for POP with or without additional pre-and post-operative PFMT.
At 12 months after surgery there were no significant differences between groups on the change in scores of the
UDI nor the IIQ scores [640].

Another RCT reported on the 6-month follow-up of 57 women (28 surgery/29 surgery with PFMT). The UDI-6
was used to assess urinary symptoms. There was a statistically significant improvement on the UDI-6 score for
both groups, but not between groups [648].

Another RCT reported on the results of a 2x2 factorial design in which women were first randomised between
two surgical techniques for POP and in addition between additional PFMT (n = 188) or not (n = 186) [649]. The
UDI was used to assess urinary symptoms up to 24 months. No significant differences were found between the
addition of PFMT to surgery or not. Another study of the same population reported on SUI in particular [650].
and no significant differences were found between women who had additional PFMT and those who had not.

Finally, in 2020 an RCT reported on the 40 and 90 days follow-up of 48 women randomised to supervised
PFMT before and after surgery and 40 women having surgery only [647]. The UDI-6 was used to assess urinary
symptoms. No statistically significant differences in UDI-6 scores were identified at 40 and 90 days.

The NICE guideline on the management of POP advocates considering supervised PFMT for at least 16
weeks as initial treatment for symptomatic prolapse [67]. The use of pessary is also to be considered, alone or
combined with PFMT. It is important to recognise that the benefit is expected on typical POP symptoms, like
feeling or seeing a bulge out of the vagina, and not on LUTS, as the reported RCTs show. From a urological
perspective, initiating conservative treatment for asymptomatic POP in order to treat UI or bladder emptying
problems is not supported by the data.

4.7.4.1.5 S
 ummary of evidence and guidelines for the conservative treatment of pelvic organ prolapse and
lower urinary tract symptoms

Summary of evidence LE
Pelvic floor muscle therapy improves LUTS for up to six months in POP patients who do not have 2a
additional pessary or surgical treatment.
If pessary therapy or surgical intervention is used for POP, PFMT does not show an additional benefit. 2a

90 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021


Recommendations Strength rating
Inform women with pelvic organ prolapse (POP), who do not need a vaginal pessary or Strong
surgical intervention, about the potential relief from lower urinary tract symptoms (LUTS) from
pelvic floor muscle therapy (PFMT).
Do not offer pre-operative PFMT in order to improve outcome of LUTS if pessary therapy or Strong
surgical intervention is indicated for POP.

4.7.4.2 Pelvic organ prolapse surgery and overactive bladder


Only a few studies specifically address the effect of POP surgery on OAB symptoms. A systematic review of
12 studies, excluding women with SUI, evaluated OAB symptoms before and after surgery [653]. All but one
study reported an improvement of OAB symptoms. The same authors performed a prospective analysis of 505
women who had POP surgery with or without mesh [654]. Symptoms were assessed with UDI questions and
each symptom was dichotomised into not bothersome or bothersome. Mean follow-up was 12.7 months. The
incidence of bothersome urinary frequency reduced from 36.6% to 14.6%, with de novo symptoms occurring
in 6.1%. Bothersome urgency symptoms reduced in 36.8% to 12.9% of women, with 5.0% developing de novo
symptoms. Urgency UI symptoms reduced from 21.2% to 6.1% of women, with 5.3% developing de novo
symptoms.

One observational study evaluated frequency and urgency symptoms without consideration of bother in 87
women undergoing POP surgery and showed an improvement in frequency by 75%, and in urgency in 83% [655].
The effect of the POP-Q stage did not seem to influence the effect of surgery on OAB symptoms [654, 655].

Another observational study (n = 43) evaluated the effect of posterior repair on OAB/DO and showed a 70–75%
improvement rate in both parameters after surgery [656].

4.7.4.3 Pelvic organ prolapse surgery and bladder outlet obstruction


The criteria for BOO are based on urodynamic assessment. Pelvic organ prolapse can be categorised as
anatomical BOO which is addressed in Sections 4.5.2.2 and 4.5.3.1.

4.7.4.4 Pelvic organ prolapse surgery and stress urinary incontinence


The aim of this section is to address the options available to women who require surgery for POP and who
have associated SUI (either before or after reduction of prolapse), and to assess the value of prophylactic anti-
UI surgery in women with no evidence of SUI.

A systematic review and meta-analysis of ten trials on prolapse surgery with or without an anti-incontinence
procedure was reported in 2018 [657]. In addition, a Cochrane review including nineteen trials (n = 2,717)
evaluating bladder function after surgery for POP presented analyses of women with POP and SUI, women
with POP and occult SUI and women with POP who were continent [658].

4.7.4.4.1 Vaginal pelvic organ prolapse surgery in women with stress urinary incontinence
Two trials addressed post-operative SUI in patients who had been diagnosed with SUI pre-operatively and had
vaginal POP surgery [659, 660]. Two trials (n = 185 and n = 134) compared the use of MUS at initial POP surgery
to POP surgery alone. The RR for post-operative SUI was 0.30 in favour of the combined POP surgery and
MUS group. One of these two trials also compared the use of MUS at initial POP surgery and at 3 months if SUI
persisted [659]. At 12 months follow-up there was no difference between the groups regarding post-operative UI
(RR: 0.41); however, 44% of the women without initial MUS never required surgery and 29% were dry.

4.7.4.4.2 Abdominal pelvic organ prolapse surgery in women with stress urinary incontinence
One RCT randomised 47 women with POP and SUI to an abdominal POP surgical procedure; e.g.
sacrocolpopexy with or without Burch colposuspension. Additional SUI surgery did not improve post-operative
SUI as compared to sacrocolpopexy alone (RR: 1.38) [661]. This finding remained consistent over 5 years
follow-up [662]. Another RCT compared the addition of a MUS or Burch colposuspension to an abdominal
sacrocolpopexy in 113 women with POP and SUI [663]. At 2 years follow-up the RR for post-operative SUI was
0.54 in favour of the MUS group.

4.7.4.4.3 Vaginal POP surgery in women with prolapse and occult stress urinary incontinence
Five RCTs including a total of 194 women who had vaginal POP repair alone and 174 women who had an
additional MUS at the time of primary surgery were identified [635, 664-667]. The RR of post-operative SUI was
0.38 in favour of the MUS group.

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 91


4.7.4.5 Vaginal pelvic organ prolapse surgery in continent women
One RCT comparing vaginal POP surgery alone with concomitant MUS in 220 women found that post-
operative SUI occurred in 46/113 (40.7%) women who had POP surgery alone, compared to 30/107 (28.0%)
who had additional MUS (RR: 0.69) [658].

4.7.4.5.1 Abdominal pelvic organ prolapse surgery in continent women


Two RCTs, compared abdominal sacrocolpopexy with (n = 189) or without (n = 190) Burch colposuspension
with an outcome favouring the addition of Burch colposuspension (RR for de novo SUI 0.69) [668, 669].

4.7.4.6 Adverse events associated with combined pelvic organ prolapse and stress urinary incontinence
surgery
Data from 6 RCT’s on vaginal POP surgery with MUS were pooled to assess adverse events [635, 659, 660,
665-667]. Urgency UI was less frequent after combination surgery as compared to POP surgery alone (28 vs.
42%, RR: 0.7), but there was a tendency towards more voiding problems. Adverse events directly related to
surgery occurred more often in the combination group (28% vs. 15%, RR: 1.8), as did serious adverse events
such as bladder perforation, urethral injuries, tape exposure (14% vs. 8%, RR: 1.7) [657].

In summary, it is difficult to generalise the results of trials using different procedures to treat both POP and UI.
It seems that with a combined procedure the rate of SUI post-operatively is lower but post-operative voiding
symptoms and complication rates are higher. Studies using MUS have generally shown more significant
differences in UI outcomes with combined procedures than when other types of anti-UI procedure have been
used. It must be taken into account that although more women are dry after combined surgery for POP with
MUS, there are potential adverse events that should be balanced against potential benefits.

4.7.5 Summary of evidence and recommendations for surgery in women with both pelvic organ
prolapse and stress urinary incontinence

Summary of evidence LE
Women with pelvic organ prolapse and urinary incontinence
Surgery for POP and SUI shows a higher rate of cure of UI in the short-term than POP surgery alone. 1a
There is conflicting evidence on the relative long-term benefit of surgery for POP and SUI vs. POP 1a
surgery alone.
Combined surgery for POP + SUI carries a higher risk of adverse events than POP surgery alone. 1a
Continent women with pelvic organ prolapse
Continent women with pelvic organ prolapse are at risk of developing SUI post-operatively. 1a
The addition of a prophylactic anti-UI procedure reduces the risk of post-operative UI but increases 1a
the risk of adverse events.
Women with pelvic organ prolapse and overactive bladder
There is some low-level inconsistent evidence to suggest that surgical repair of POP can improve 2b
symptoms of overactive bladder.

Recommendations for women requiring surgery for bothersome pelvic organ Strength rating
prolapse (POP) who have symptomatic or occult stress urinary incontinence (SUI)
Offer simultaneous surgery for POP and SUI only after a full discussion of the potential risks Strong
and benefits of combined surgery vs. POP surgery alone.
Inform women of the increased risk of adverse events with combined prolapse and anti- Strong
urinary incontinence surgery compared to prolapse surgery alone.
Recommendations for women requiring surgery for bothersome POP who do not have
symptomatic or occult SUI
Inform women that there is a risk of developing de novo SUI after prolapse surgery. Strong
Warn women that the benefit of combined surgery for POP and SUI may be outweighed by Strong
the increased risk of adverse events compared to prolapse surgery alone.

4.8 Urinary fistula


The evidence relating to diagnosis and treatment of urinary fistulae is generally of low-level and is largely
composed of case series and other consensus statements. In particular, the epidemiology, aetiology, diagnosis,
treatment and prevention of obstetric and non-obstetric fistulae have been described in detail during the recent
ICI conference [670]. Most non-obstetric fistulae are iatrogenic in origin, with the majority caused by pelvic

92 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021


surgery (e.g. hysterectomy for benign or malignant conditions, bowel resection, and urological surgery). The
risks during pelvic surgery increase relative to the complexity of the resection, the extent of primary disease
and the existence of prior radiotherapy (especially for recurrent disease). When a fistula occurs following
radiotherapy for primary treatment, this may be an indication of tumour recurrence.

4.8.1 Epidemiology, aetiology and pathophysiology


4.8.1.1 Obstetric fistula
According to the WHO, fistulae affect more than two million women, mostly from sub-Saharan African and
Asian countries. The pooled prevalence of fistula from population studies is at 0.29/1000 pregnancies [671].
Poor quality obstetric care, staff unaccountability, late referral and poor nursing standards have been identified
as health system causes [671]. However, obstructed labours are poorly documented. The main individual risk
factors include age at first marriage, short stature, pregnancy with a male child, failure to attend ante-natal
care, low socio-economic status, low social class, lack of employment and illiteracy [672-674]. Obstetric
fistulae have detrimental consequences on global and individual health and are associated with malnutrition,
sexual dysfunction, anxiety, depression, insomnia, social isolation, worsening poverty and suicide [675, 676].

4.8.1.2 Iatrogenic fistula


Poor obstetric care is usually responsible for VVF in the developing world. By contrast, in the developed world,
gynaecological or pelvic surgeries are the main causes of VVF.

4.8.1.2.1 Post-gynaecological surgery


An injury to the urinary tract during hysterectomy for benign conditions (60–75%), hysterectomy for malignant
conditions (30%) and caesarean section (6%) are the main causes of post-operative VVF in the developed
world [677, 678]. The risk of pelvic organ fistula following hysterectomy ranges from 0.1 to 4% [679].
Furthermore, fistulae may also occur as a result of primary or recurrent malignancy, or as a consequence of
cancer treatment by surgery, radiotherapy, and/or chemotherapy.

In a study including 536 women undergoing a radical hysterectomy for invasive cervical cancer, bladder injury
occurred in 1.5% with VVF forming in 2.6% and uretero-vaginal fistula (UVF) in 2.4% of cases [680]. Overall,
the rate of urogenital fistula appears to be approximately 9 times higher following radical hysterectomy for
malignant disease as compared to that following simple hysterectomy (abdominal or vaginal for benign
conditions) [681]. Bladder sparing techniques during pelvic exenteration can increase the risk of fistula
formation [682].

4.8.1.2.2 Radiation fistula


The risk of fistula seems to be higher for post-operative external radiation (1.9%) compared to intravaginal
brachytherapy (0.8%) [683], without any predictive factor being identified [684]. This is most likely due to
the heterogeneity of data regarding the tumour type and stage, the form of radiation and the site and dose
delivered.

4.8.1.2.3 Rare causes of vesico-vaginal fistula


Foreign bodies such as pessaries, sex toys, cups etc. can be a cause of delayed presentation of VVF [685-
687]. Ketamine abuse has also been shown to be responsible for fistula formation [688].

4.8.1.3 Summary of evidence for epidemiology, aetiology and pathophysiology of urinary fistula

Summary of evidence LE
The risk of injury to the urinary tract and subsequent fistula formation is higher in women with 2
malignant disease undergoing radical surgery than in women with benign disease undergoing simple
surgical procedures.
The rate of fistula formation following radiotherapy for gynaecological cancer appears to be of the 4
same order as that following surgical treatment.

4.8.2 Classification
Due to the plethora of VVF classification systems, a consensual classification system needs to be adopted. The
Waaldijk and Goh classifications are widely used for diagnosis and follow-up [689-691]. They were originally
designed for obstetric fistulae and their use in iatrogenic fistulae is less relevant [692]. Waaldijk’s classification
is based on the size and site of the fistula and divides them into 3 main categories - type 1 are VVF with no
urethral involvement, type 2 fistulae are those which involve the urethra (and are sub-classified into those with

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 93


circumferential and non-circumferential urethral involvement) and type 3 are fistula involving other parts of the
urinary tract. Goh’s classification also uses the presence or absence of urethral involvement to sub-categorise
VVF and also takes into account the degrees of fibrosis present. The WHO classification (Table 6) was originally
developed for obstetric fistulae and separates fistulae into simple and complex.

Table 6: Adapted WHO Classification of fistulae [671]*

Simple fistula with good prognosis Complex fistula with uncertain prognosis
• Single fistula < 4 cm • Fistula > 4 cm
• Vesico-vaginal fistula • Multiple fistula
• Closing mechanism not involved • Recto-vaginal mixed fistula, cervical fistula
• No circumferential defect • Closing mechanism involved
• Minimal tissue loss • Scarring
• Ureters not involved • Circumferential defect
• First attempt to repair • Extensive tissue loss
• Intravaginal ureters
• Failed previous repair
• Radiation fistula
*Although this classification was developed for obstetric fistula initially, it could be relevant for iatrogenic fistula
as well.

4.8.2.1 Recommendation for the classification of urinary fistula

Recommendation Strength rating


Use a classification system for urinary tract fistulae to try to standardise terminology in this Strong
subject area.

4.8.3 Diagnostic evaluation


Leakage of urine is the hallmark sign of a urogenital fistula. The leakage is usually painless, may be intermittent
if it is position dependent, but more usually is constant. Unfortunately, intra-operative diagnosis of a GU or GI
injury is made in only about half of all cases [693]. The diagnosis of VVF usually requires clinical assessment
often in combination with appropriate imaging or laboratory studies. Direct visual inspection, cystoscopy,
retrograde bladder filling with a coloured fluid or placement of a tampon into the vagina to identify staining
may facilitate the diagnosis of a VVF. A double-dye test to differentiate between a UVF and VVF may be useful
in some cases [678]. Testing the creatinine level in either the extravasated or collected fluid will confirm fluid
leakage as urine. Contrast-enhanced CT with late excretory phase reliably diagnoses urinary fistulae and provides
information about ureteric integrity and the possible presence of associated urinoma. Magnetic resonance
imaging, in particular with T2 weighting, also provides diagnostic information regarding fistulae [694].

4.8.4 Management of fistula


4.8.4.1 Management of vesico-vaginal fistula
4.8.4.1.1 Conservative management
4.8.4.1.1.1 Spontaneous closure
The reported spontaneous closure rate is 13% ± 23% [695], although this applies largely to small fistulae (size
< 1 cm) [670, 696]. Hence, immediate management is usually by urinary catheterisation or diversion; however,
within the first two weeks following fistula occurrence, surgical exploration and repair can be considered.

4.8.4.1.1.2 Pharmacotherapies
Several case reports describe a successful fistula closure rate following the induction of amenorrhoea by
oestrogen, oestrogen/progesterone combinations or luteinising hormone releasing hormone (LHRH) analogues
specifically for small (< 7 mm), uretero- or vesico-uterine fistula following caesarean section [697-703]. One
RCT comparing the efficacy of using fibrin glue compared to Martius flap inter-positioning (n = 14; < 4 cm and
n = 5; > 5 cm) did not report statistically different outcomes between the two types of treatment [704].

4.8.4.1.1.3 Palliation and skin care


During the intervening period between diagnosis and repair, UI pads with the aim of prevention of skin
complications related to chronic urinary leakage can be provided and the use of a barrier cream or local
oestrogen can also be considered [705, 706].

94 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021


4.8.4.1.1.4 Nutrition
Nutritional support is essential in patients with fistulae induced by malignant disease or radiotherapy [707], or
following diversion surgery [707-709].

4.8.4.1.1.5 Physiotherapy
Early involvement of the physiotherapist in pre-operative management and rehabilitation of fistula patients
suffering from limb weakness, foot drop and limb contracture is essential [710, 711].

4.8.4.1.1.6 Antimicrobial therapy


Active infection in the genital or urinary tracts should be treated prior to surgical repair [712].

4.8.4.1.1.7 Counselling
Confident and realistic counselling by the surgeon is essential and the involvement of nursing staff or
counsellors with experience of fistula patients is also highly desirable.

4.8.4.1.2 Surgical management


4.8.4.1.2.1 Timing of surgery
Findings from uncontrolled case series suggest no difference in success rates for early (within 3 weeks) or
delayed (after 3 months) closure of VVF.

4.8.4.1.2.2 Surgical approaches


Vaginal procedures
There are two main types of closure techniques applied to the repair of urinary fistulae, the classical
saucerisation/partial colpocleisis [695] and the more commonly used dissection and repair in layers or ‘flap-
splitting’ technique [713]. There are no data comparing their outcomes.

Abdominal procedures
Repair by the abdominal route is indicated when high fistulae are fixed at the vaginal vault and are inaccessible
via a vaginal approach. A transvesical repair has the advantage of being entirely extraperitoneal. A simple
transperitoneal repair is used less often although it is favoured by some using the laparoscopic approach. A
combined transperitoneal and transvesical procedure may be utilised for fistula repair following Caesarean
section. There are no RCTs comparing abdominal and vaginal approaches. Results of secondary and
subsequent repairs are not as successful as the initial repair [714].

A single RCT compared trimming of the fistula edge with no trimming and found no difference in success rates but
failed repairs in trimmed cases ended up with larger recurrences than untrimmed cases, which were smaller [715].

Laparoscopic and robotic procedures


Very small series (single figures) have been reported using these techniques, but whilst laparoscopic repair is
feasible with and without robotic assistance, it is not possible to compare outcomes with alternative surgical
approaches.

Tissue interposition
Tissue flaps are often added as an additional layer of repair during VVF surgery. Most commonly, such flaps
are utilised in the setting of recurrence after a prior attempt at repair, for VVF related to previous radiotherapy
(described later), ischaemic or obstetric fistulae, large fistulae, and finally those associated with a difficult or
tenuous closure due to poor tissue quality. However, there is no high-level evidence that the use of such flaps
improves outcomes for either complicated or uncomplicated VVF.

Post-operative management
There is no high-level evidence to support any particular practice in post-operative management but most
reported series used catheter drainage for at least ten days and longer periods in complex or radiation-
associated fistulae (up to three weeks). The performance of post-operative cystogram prior to catheter removal
can miss a persistent fistula if not done with a micturition phase or if the fistula is located at the bladder neck.

4.8.4.1.3 Management of complications of vesico-vaginal fistulae


The complications of VVF repair are varied and can include:

• Persistence or recurrence of fistula;


• Persistence or recurrence of UI;

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 95


• Persistence of LUT symptoms or occurrence of new LUT symptoms, including de novo OAB symptoms
and/or SUI;
• Infections: wound and UTIs/urosepsis;
• Ureteric obstruction (ligation, fibrosis, injury);
• Bladder outlet obstruction (meatal stenosis, urethral stricture, bladder neck obstruction);
• Bladder contracture;
• Vaginal stenosis;
• Sexual dysfunction (vaginismus/dyspareunia);
• Rare complications (granulomas/diverticulum formation);
• Neurological complications (foot drop/neurogenic bladder);
• Psychological trauma (social isolation/divorce/mental illness);
• Infertility.

The literature on the treatment and management of complications of fistula repairs is extremely scarce and is
mostly experience-based. It is impossible to give any specific evidence-based guidance.

4.8.4.2 Management of radiation fistulae


Modified surgical techniques are often required, and indeed, where the same techniques have been applied to
both surgical and post-radiation fistulae, the results from the latter have been consistently poorer [716]. Due
to the wide field abnormality surrounding many radiotherapy-associated fistulae, approaches include, on the
one hand, permanent urinary and/or faecal diversion [716, 717] or alternatively preliminary urinary and faecal
diversion, with later undiversion in selected cases following reconstruction. In cases where life expectancy is
deemed to be very short, ureteric occlusion might be more appropriate.

4.8.4.3 Management of ureteric fistulae


4.8.4.3.1 General principles
Patients at higher risk of ureteric injury require experienced surgeons who can identify and protect the ureter
and its blood supply to prevent injury and also recognise injury promptly when it occurs. Immediate repair of
any intra-operative injury should be performed observing the principles of debridement, adequate blood supply
and tension-free anastomosis with internal drainage using stents [718]. Delayed presentation of upper tract
injury should be suspected in patients whose recovery after relevant abdominal or pelvic surgery is slower
than expected, if there is any fluid leak, and if there is any unexpected dilatation of the pelvicalyceal system.
Whilst there is no evidence to support the use of one surgical approach over another, there is consensus that
repair should adhere to the standard principles of tissue repair and safe anastomosis, and be undertaken by
an experienced team. Conservative management is possible with internal or external drainage, endoluminal
management using nephrostomy and stenting where available, and early (< two weeks) or delayed (> three
months) surgical repair when required [719]. Functional and anatomical imaging should be used to follow up
patients after repair to guard against development of ureteric stricture and deterioration in renal function.

4.8.4.3.2 Uretero-vaginal fistulae


Uretero-vaginal fistula occurring in the early post-operative phase predominantly after hysterectomy is the
most frequent presentation of UUT fistulae in urological practice. An RCT in 3,141 women undergoing open- or
laparoscopic gynaecological surgery found that prophylactic insertion of ureteric stents made no difference to
the low risk (1%) of ureteric injury [720].

Endoscopic management is sometimes possible by retrograde stenting, percutaneous nephrostomy and


antegrade stenting if there is pelvicalyceal dilatation, or ureteroscopic realignment [721]. However, the long-
term success rate is unknown. If endoluminal techniques fail or result in secondary stricture, the abdominal
approach to repair is standard and may require end-to-end anastomosis, re-implantation into the bladder using
psoas hitch or Boari flap, or replacement with bowel segments with or without reconfiguration. As a last resort,
nephrectomy may be considered, particularly in the context of a poorly functioning kidney and an otherwise
normal contralateral kidney [722-726].

4.8.4.3.3 Management of urethrovaginal fistulae


4.8.4.3.3.1 Aetiology
Whilst they are rare, most urethrovaginal fistulae in adults have an iatrogenic aetiology. Causes include
surgical treatment of SUI with bulking agents or synthetic slings, surgery for urethral diverticulum and genital
reconstruction in adults. Irradiation and even conservative treatment of prolapse with pessaries can lead to the
formation of fistulae.

96 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021


4.8.4.3.3.2 Diagnosis
Clinical vaginal examination, including the three-swab test, is often sufficient to diagnose the presence of a
urethrovaginal fistula. Urethroscopy and cystoscopy can be performed to assess the extent and location of
the fistulae. In cases of difficult diagnosis, VCUG or US can be useful. 3D-MRI or CT scan is becoming utilised
more widely to clarify anatomy [727, 728].

4.8.4.3.3.3 Surgical management


Choice of surgery will depend on the size, localisation and aetiology of the fistula and the amount of tissue
loss. Principles of reconstruction include identifying the fistula, creation of a plane between vaginal wall and
urethra, watertight closure of urethral wall, eventual interposition of tissue, and closure of the vaginal wall.

One case series reports that a vaginal approach yielded a success rate of 70% at first attempt and 92% at
second attempt, and that an abdominal approach only leads to a successful closure in 58% of cases [729]. A
vaginal approach required less operating time, had less blood loss and a shorter hospitalisation time.

Most authors describe surgical principles that are identical to those of vesico-vaginal fistula repair: primary
closure rates of 53-95.4% have been described. A series of 71 women, treated for urethra-vaginal fistula
reports 90.1% of fistulae were closed at the first vaginal intervention. Additionally, 7.4% were closed during a
second vaginal intervention. Despite successful closure, SUI developed in 52%. The stress incontinent patients
were treated with synthetic or autologous slings and nearly 60% became dry and an additional 32% improved.
Urethral obstruction occurred in 5.6% and was managed by urethral dilation or urethrotomy [730].

4.8.4.3.3.4 Flaps and neo-urethra


The simplest flap is a vaginal advancement flap to cover the urethral suture line. Labial tissue can be harvested
as a pedicled skin flap. This labial skin can be used as a patch to cover the urethral defect, but can also be
used to create a tubular neo-urethra [731, 732]. The construction of a neo-urethra has mostly been described
in traumatic aetiologies. In some cases a transpubic approach has been used [733]. The numbers of patients
reported are small and there are no data on the long-term outcome of fistula closure and continence rates.
The underlying bulbocavernosus tissue can be incorporated in the pedicled flap and probably offers a better
vascularisation and more bulking to the repair. This could allow a safer placement of a sling afterwards, in those
cases where bothersome SUI would occur post-operatively [734, 735].

4.8.4.3.3.5 Martius flap


While in obstetrical fistula repair it was not found to have any benefit, in a large retrospective study in 440
women the labial bulbocavernosus muscle/fat flap by Martius is still considered by some to be an important
adjunctive measure in the treatment of GU fistulae where additional bulking with well-vascularised tissue is
needed [736]. The series of non-obstetrical aetiology are small and all of them are retrospective. There are no
prospective data, nor randomised studies [737]. The indications for Martius flap in the repair of urethro-vaginal
fistulae remain unclear.

4.8.4.3.3.6 Rectus muscle flap


Rectus abdominis muscle flaps have been described by some authors [738, 739].

4.8.4.3.3.7 Alternative approaches


An alternative retropubic retro-urethral technique has been described by Koriatim [740]. This approach allows a
urethro-vesical flap tube to be fashioned to form a continent neo-urethra.

4.8.4.4 Summary of evidence and recommendations for the management of urinary fistula

Summary of evidence LE
Spontaneous closure of surgical fistulae does occur, and appears more likely for small fistulae 3
although it is not possible to establish the rate with any certainty.
There is no evidence that the timing of repair makes a difference to the chances of successful closure 3
of a fistula.
There is no high-quality evidence of differing success rates for repair of vesico-vaginal fistulae by 3
vaginal, abdominal, transvesical and transperitoneal approaches.
A period of continuous bladder drainage may be crucial to successful fistula repair but there is no 3
high-level evidence to support one regimen over another.
A variety of interpositional grafts can be used in either abdominal or vaginal procedures, although 3
there is little evidence to support their use in any specific setting.

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 97


Post-radiation fistula
Successful repair of irradiated fistulae may require prior urinary diversion and the use of non-irradiated 3
tissues to effect repair.
Ureteric fistula
Prophylactic ureteric stent insertion does not reduce risk of ureteric injury during gynaecological 2
surgery.
Antegrade endoluminal distal ureteric occlusion combined with nephrostomy tube diversion often 4
palliates urinary leakage due to malignant fistula in the terminal phase.
Urethro-vaginal fistula
Urethro-vaginal fistula repair may be complicated by SUI, urethral stricture and urethral shortening 3
which may necessitate long-term follow-up.

Recommendations Strength rating


General
When reporting on outcomes after fistula repair, authors should make a clear distinction Strong
between fistula closure rates and post-operative urinary incontinence (UI) rates and the time
at which the follow-up was organised.
Do not routinely use ureteric stents as prophylaxis against injury during routine Strong
gynaecological surgery.
Suspect ureteric injury or fistula in patients following pelvic surgery if a fluid leak or Strong
pelvicalyceal dilatation occurs post-operatively, or if drainage fluid contains high levels of
creatinine.
Use three-dimensional imaging techniques to diagnose and localise urinary fistulae Weak
particularly in cases with negative direct visual inspection or cystoscopy.
Manage upper urinary tract fistulae initially by conservative or endoluminal techniques Weak
where such expertise and facilities exist.
Surgical principles
Surgeons involved in fistula surgery should have appropriate training, skills, and experience Weak
to select an appropriate procedure for each patient.
Attention should be given as appropriate to skin care, nutrition, rehabilitation, counselling Weak
and support prior to, and following, fistula repair.
Tailor the timing of fistula repair to the individual patient and surgeon requirements once any Weak
oedema, inflammation, tissue necrosis, or infection, are resolved.
Ensure that the bladder is continuously drained following fistula repair until healing is Weak
confirmed (expert opinion suggests: 10–14 days for simple and/or postsurgical fistulae;
14–21 days for complex and/or post-radiation fistulae).
Where urinary and/or faecal diversions are required, avoid using irradiated tissue for repair. Weak
Use interposition graft when repair of radiation-associated fistulae is undertaken. Weak
Repair persistent uretero-vaginal fistulae by an abdominal approach using open, Weak
laparoscopic or robotic techniques according to availability and competence.
Urethro-vaginal fistulae should preferably be repaired by a vaginal approach. Weak

4.9 Urethral diverticulum


A female urethral diverticulum is a sac-like protrusion composed of the entire urethral wall or only by the
urethral mucosa, situated between the peri-urethral tissues and the anterior vaginal wall.

4.9.1 Epidemiology, aetiology, pathophysiology


Urethral diverticulum is an uncommon condition with a prevalence estimated to range between 1 and 6%.
Amongst women with LUTS attending a tertiary referral centre one study reported a prevalence of up to 10%
[741]. However, as many patients are asymptomatic or misdiagnosed, the true incidence is unknown [742-744].
Given the rarity of the condition, most published series are small and single institutional. Urethral diverticulum
are thought to arise from repeated obstruction, infection and subsequent rupture of periurethral glands into the
urethral lumen, resulting in an epithelialised cavity that communicates with the urethra [742].

Iatrogenic damage to the urethra may also play a role, as up to 20% of women with urethral diverticula are
noted to have a history of prior urethral surgery, dilation, or traumatic delivery [742, 745]. Iatrogenic urethral
diverticula formation associated with synthetic suburethral sling has also been reported [746-748].

98 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021


4.9.2 Classification

Table 7: Classification system for female urethral diverticula based on characteristics*

Localisation Mid-urethral
Distal
Proximal
Full length
Configuration Single
Multiloculated
Saddle shaped
Communication Mid-urethral
No communication visualised
Distal
Proximal
Continence Stress urinary incontinence
Continent
Post-void dribble
Mixed incontinence
*Limited LNS C3 classification of urethral diverticula [745, 749, 750].

4.9.3 Diagnosis
The commonly encountered symptoms for urethral diverticulum such as pain, urgency, frequency, recurrent
UTIs, vaginal discharge, dyspareunia, voiding difficulties or UI [751], are common to many other LUT
dysfunctions. Consequently, there is no pathognomonic cluster of symptoms to identify urethral diverticula.
Many patients with urethral diverticula are asymptomatic. However, urethral diverticulum often presents with a
palpable urethral mass. It may be possible to express a purulent exudate from the urethra. Occasionally a stone
may develop within the diverticulum.

The diagnosis of a urethral diverticulum may be achieved by physical examination, VCUG and MRI. Other
investigations include urethrocystoscopy, endocavitary (often transvaginal or sometimes transurethral) pelvic
floor US and double balloon urethrography.

No robust diagnostic accuracy studies address the question of the best test to confirm the diagnosis in a
woman with a clinical suspicion of urethral diverticulum. However, a case series of 27 patients concluded
that endoluminal (vaginal or rectal) MRI has better diagnostic accuracy than VCUG [752] and determines the
size and extent of urethral diverticula more accurately. In a case series of 60 subjects, it was reported that the
sensitivity, specificity, positive predictive value and negative predictive value of MRI is 100%, 83%, 92% and
100%, respectively [753]. Another case series reported 100% specificity and sensitivity of MRI in 60 patients
[754]. However, in a case series of 41 patients, authors reported a 25% discrepancy between MRI and surgical
findings [755]. Endoluminal MRI with either a vaginal or rectal coil may provide even better image quality than
simple MRI [756].

Magnetic resonance imaging is the gold standard for the diagnosis and planning surgical repair. Magnetic
resonance imaging also proved to be useful in diagnosing inflammation or tumour in the diverticulum [757, 758].

Urethrocystoscopy can be used to visualise the ostia of the diverticulum. Knowledge of the ostia’s location and
number can assist with surgical planning since each of these ostia need to be closed after diverticulectomy.
However, given the challenges of urethroscopy in females the ostia is only seen in 42% of cases [751].

If a VCUG is performed antero-posterior and lateral images are required to optimally characterise the
configuration of the diverticulum. There is a high risk of false negatives since the ostia of the diverticulum
must be patent and the patient must be able to void during the study. In more complex diverticulum where
there is septation, the entire diverticula may not be visualised underestimating its complexity or size [759]. The
sensitivity of VCUG is 73.5% which is significantly worse than MRI [751].

Ultrasound can be performed transabdominally, transvaginally or transperineally to identify the diverticulum. In


particular the transvaginal approach US allows imaging of the urethra from the meatus to the bladder neck in
several planes and can identify the number, size, location and contents of the diverticulum. This technique is

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 99


challenging and requires a skilled ultrasonographer. In addition, the probe can compress the urethra, causing
distortion [759]. A meta-analysis reported that US of any kind had a sensitivity of 82.0% which was inferior to
MRI [759]. However a recent publication on translabial US reported a sensitivity of 95% [760]; therefore, this
approach may be explored further by researchers in the future.

For patients who cannot have an MRI and the ostia is not seen on cystoscopy, double balloon urethrography
can be an option. Modern series have reported a 94.7% sensitivity which is comparable to that of MRI. The
technique uses positive pressure to force contrast into the diverticular sac between two balloons; one placed in
the bladder and one outside the ostia of the diverticulum. It is technically difficult to achieve a seal sufficient to
create a closed urethral space and avoid contrast leaking around the catheter. The procedure can be painful for
the patient and carries a risk of UTI. An experienced radiologist is required as well as specialised equipment.
Given the current popularity of other imaging modalities many units may not have access to this technique
[759].

4.9.3.1 Associated voiding dysfunction


Although the presentation of urethral diverticulum is often non-specific and variable, urethral diverticula can be
associated with voiding dysfunction and SUI or UUI.

One recent series reported SUI occurring in 60% of patients with urethral diverticulum [761]. A urethral
diverticulum is most often located at the level of the mid-urethra. This location often overlaps with the external
sphincter. However urethral diverticulum may also extend proximally toward the bladder neck in the vicinity
of the proximal sphincter mechanism. This morphology may, in part, explain the association between urethral
diverticulum and SUI with potentially more proximal lesions at risk for post-operative SUI [762].

Urethral diverticulum may also be associated with BOO due to the mass effect of the urethral diverticulum,
urinary retention, or urgency and UUI [763]. Pain and dysuria associated with urethral diverticulum may also
result in acquired voiding dysfunction.

Pressure flow studies may have a role in the pre-operative assessment of patients with urethral diverticula
and coexisting voiding dysfunction or SUI [744, 764-766]. Indeed, urodynamics may evaluate for coexisting
detrusor dysfunction or document the presence or absence of SUI or obstruction prior to repair [767, 768].
Urethral pressure profilometry has also been used in the assessment or diagnosis of urethral diverticulum
noting a biphasic pattern, or pressure drop at the level of the lesion during the study [764, 766, 769]. Video-
urodynamics may be helpful in differentiating SUI from paradoxical UI due to fluid accumulation in the urethral
diverticulum. In addition, resting and straining images obtained during fluoroscopic imaging may document an
open bladder neck at rest. This may be a consideration in some patients with an extensive urethral diverticulum
at the level of the mid-urethra and potential implications for post-operative UI due to compromise of both
sphincter mechanisms.

4.9.4 Disease management


For women with minimal symptoms who would prefer to avoid invasive treatment, conservative management
can be considered. Patients should be warned of the small risk of cancer (1–6%) within the diverticulum [770,
771].

4.9.4.1 Surgical treatment


No RCTs were found to inform regarding the relative effectiveness of available surgical treatments in a woman
who has a bothersome urethral diverticulum. Thorough evaluation of the anatomy of the diverticulum is
essential in planning reconstructive surgery.

There are three surgical approaches to diverticulum: marsupialization, endoscopic incision, and curative
treatment with diverticulectomy.

Surgical removal is the most commonly reported treatment in contemporary case series. The principles
of successful transvaginal diverticulectomy are to: dissect a well vascularised vaginal flap, preserve the
periurethral fascia for closure, remove all the diverticular wall, excise the ostia and close the urethra in
a watertight fashion, close the incision in a multi-layered fashion with no overlapping suture lines, and
preservation or creation of continence.

The decision to use a labial fat pad flap, commonly known as a Martius flap, is variable and used more
frequently in the following situations: recurrent cases, large urethral defects or for deficient vaginal flaps for

100 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021
closure [745, 749] transection of the urethra required for access to a circumferential diverticulum [758] or in the
case of complex configuration [763], and if there is a planned future sling procedure required for UI to facilitate
the dissection at that time [745].

Marsupialisation involves incision into the mass on the vaginal side to drain the infected contents. The wall is
then sutured open with absorbable suture to allow drainage and prevent reaccumulation of infectious materials.
This approach leaves the cystic structure in place and can theoretically cause a urethra-vaginal fistula since
there is communication with the diverticular ostia, but it is a rapid procedure with little dissection required. This
approach has been advocated in the pregnant patient to decompress the diverticulum to allow safe vaginal
delivery. A small case series suggests that 75% pregnant women with urethral diverticula managed expectantly
end up requiring postpartum surgery [772].

Endoscopic incision is a rarely reported treatment option [773, 774]. This procedure involves finding the narrow
neck of the ostia and incising it with a resectoscope. This unroofing of the diverticulum transforms the narrow
communication with the urethra that causes symptoms when it gets obstructed into a wide mouthed sac that
drains freely.

4.9.4.2 Management of concomitant stress urinary incontinence


Many women present with concomitant SUI and urethral diverticula, and may request both conditions to be
simultaneously treated. A meta-analysis reports that diverticulectomy actually cured the SUI even without a
concomitant anti-incontinence procedure, but no data regarding symptom severity was given and it could be
assumed that many of these cured patients had more minimal UI before surgery [751]. Therefore, additional
surgical corrections may be required [762, 774]. However, there is no consensus on appropriate timing of
surgical management of these two conditions. Thus, patients with symptomatic, bothersome SUI in association
with urethral diverticulum may be offered simultaneous anti-UI surgery. Although historical series have shown
good results with concomitant bladder neck suspension [768], more contemporary series have utilised
pubovaginal fascial slings in patients with satisfactory outcomes [775-778]. Mid-urethral synthetic slings are
not recommended as a concomitant anti-UI procedure at the time of urethral diverticulectomy [779]. Synthetic
material adjacent to a fresh suture line following diverticulectomy in the setting of potentially infected urine may
place the patient at higher risk for subsequent urethral erosion and vaginal extrusion of the sling material, as
well as urethrovaginal fistula formation and foreign body granuloma formation.

Transvaginal urethral diverticulectomy has a high success rate (defined by being dry) of between 84% and
98%, with a re-operation rate of 2–13% after primary repair during a mean follow-up of 12–50 months [742,
745, 762, 780]. The resolution of symptoms after surgery has been reported to reach 68.8% but less than half
of studies comment on symptom improvement [781]. One case series reported rates of storage symptoms
decreased significantly post-operatively from 60% to 16% following urethral diverticulum surgery [762]. Other
series with long-term follow-up, however, have demonstrated rates of post-operative urgency of 54% [782],
and de novo UUI in 36% of patients [774]. Such symptoms post-operatively may indicate urethral diverticulum
persistence, urethral diverticulum recurrence, or de novo OAB syndrome or urethral obstruction.

Early common post-operative complications include: UTI (0–39%), de novo SUI (3.8–33%), and de novo urinary
retention (0–9%), especially in the setting of concomitant placement of an autologous pubovaginal sling [742,
745, 762, 780]. Delayed complications such as urethral stricture are reported in 0–5.2% of cases [742, 745,
774, 780]. Urethrovaginal fistula is a devastating complication presenting in 0.9–8.3% of cases [783]. A distal
fistula located beyond the sphincteric mechanism can present with split urinary stream or vaginal voiding and
may not require repair. However, a fistula located anywhere from the mid-urethra to the bladder neck may result
in UI. These patients should undergo repair with consideration of an adjuvant tissue flap, such as a Martius
flap, to aid in closure. The timing of the fistula repair is not well defined, with a delay of 3–6 months after
the initial repair generally being a good balance between patient discomfort and optimal tissue quality. Rare
complications include: distal urethral necrosis, bladder injury, urethral injury, ureteric injury, and vaginal scarring
or narrowing with consequent dyspareunia [783].

One case series reported a recurrence rate of 33% in U-shaped and of 60% in circumferential diverticulum
within one year [749], Ingber et al. found a 10.7% recurrence rate in 122 women undergoing diverticulectomy,
with a higher risk of recurrence in those with proximal or multiple diverticula or after previous pelvic surgery
[782] or radiation. A recurrent urethral diverticulum following initial successful urethral diverticulectomy may
occur as a result of a new infection or traumatic insult such as childbirth, a new urethral diverticulum, or
recurrence of the original lesion. Recurrence of urethral diverticulum may be due to incomplete removal of the
urethral diverticulum, inadequate closure of the urethra or residual dead space (circumferential diverticula), or

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 101
other technical factors. Repeat urethral diverticulectomy surgery represents a unique challenge due to altered
anatomy, scarring, and difficulty identifying proper anatomical planes.

Stress UI can be worsened or occur de novo after diverticulectomy. This is most likely due to sphincteric
damage from the dissection or scarification preventing urethral closure. De novo SUI (10.6% of women) seems
to be more common in proximal and in large size (> 30 mm) diverticula [762]. However, Lee et al. noted at least
some de novo SUI in 49% of patients following urethral diverticulectomy, the majority of which was minor and
did not require additional therapy [784]. Only 10% of these individuals underwent a subsequent SUI operation.
Treatment for SUI after a diverticulectomy is not well described in the literature. The most commonly reported
surgery is an autologous pubovaginal sling [773] followed by retropubic suspension [774] however there are
two cases reported of synthetic mesh sling to treat the SUI without mesh complications [749, 762], but this is
certainly controversial.

4.9.4.3 Pathological findings


Most urethral diverticula are lined with squamous cells, urothelium or columnar epithelium [745, 785, 786]. In
this meta-analysis there is a high prevalence of chronic or acute inflammation (68.6%) and the most commonly
reported lesions are nephrogenic metaplasia which occurs in 8%. Diverticula may undergo neoplastic
alterations (6%) including invasive adenocarcinomas [787], followed by squamous cell carcinoma in 0.7%. It
is unknown if the diverticulum forms first and then transforms into a malignancy or if the malignancy develops
first. These malignancies are treated in a similar fashion to urethral cancer in the female.

4.9.5 Summary of evidence and recommendations for urethral diverticulum

Summary of evidence LE
Magnetic resonance imaging has the best sensitivity and specificity for the diagnosis of urethral 3
diverticula.
Surgical removal of symptomatic urethral diverticula provides good long-term results; however, 3
women should be counselled of the risk of recurrence and de novo SUI.

Recommendations Strength rating


Offer surgical removal of symptomatic urethral diverticula. Weak
If conservative treatment is adopted, warn patients of the small (1–6%) risk of cancer Weak
developing within the diverticulum.
Carefully question and investigate patients for co-existing voiding dysfunction and urinary Strong
incontinence.
Following appropriate counselling, address bothersome stress urinary incontinence at the Weak
time of urethral diverticulectomy with concomitant non-synthetic sling.
Counsel patients regarding the possibility of de novo or persistent lower urinary tract Strong
symptoms including urinary incontinence despite technically successful urethral
diverticulectomy.

5. REFERENCES
1. Abrams, P., et al. The standardisation of terminology of lower urinary tract function: report from the
Standardisation Sub-committee of the International Continence Society. Neurourol Urodyn, 2002.
21: 167.
https://pubmed.ncbi.nlm.nih.gov/11857671
2. Milsom, I., et al. Effect of bothersome overactive bladder symptoms on health-related quality of life,
anxiety, depression, and treatment seeking in the United States: results from EpiLUTS. Urology,
2012. 80: 90.
https://pubmed.ncbi.nlm.nih.gov/22748867
3. Felde, G., et al. Anxiety and depression associated with urinary incontinence. A 10-year follow-up
study from the Norwegian HUNT study (EPINCONT). Neurourol Urodyn, 2017. 36: 322.
https://pubmed.ncbi.nlm.nih.gov/26584597

102 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021
4. Eapen, R.S., et al. Review of the epidemiology of overactive bladder. Res Rep Urol, 2016. 8: 71.
https://pubmed.ncbi.nlm.nih.gov/27350947
5. Coyne, K.S., et al. The prevalence of lower urinary tract symptoms (LUTS) in the USA, the UK and
Sweden: results from the Epidemiology of LUTS (EpiLUTS) study. BJU Int, 2009. 104: 352.
https://pubmed.ncbi.nlm.nih.gov/19281467
6. Coyne, K.S., et al. The burden of lower urinary tract symptoms: evaluating the effect of LUTS on
health-related quality of life, anxiety and depression: EpiLUTS. BJU Int, 2009. 103 Suppl 3: 4.
https://pubmed.ncbi.nlm.nih.gov/19302497
7. Irwin, D.E., et al. The economic impact of overactive bladder syndrome in six Western countries.
BJU Int, 2009. 103: 202.
https://pubmed.ncbi.nlm.nih.gov/19278532
8. Incontinence, 6th Edn. 2017, P. Abrams, Cardozo, L, Wagg, A, Wein, A. (Eds). Bristol, UK.
https://www.ics.org/publications/ici_6/Incontinence_6th_Edition_2017_eBook_v2.pdf
9. Blok, B., et al. EAU Guidelines on Neuro-urology. Edn presented at the 36th Annual Congress,
Milan, 2021. EAU Guidelines Office, Arnhem, The Netherlands.
https://uroweb.org/guideline/neuro-urology/
10. Radmayr, C., et al. EAU Guidelines on Paediatric Urology. Edn presented at the 36th Annual
Congress, Milan, 2021. EAU Guidelines Office, Arnhem, The Netherlands.
https://uroweb.org/guideline/paediatric-urology/
11. Peyronnet, B., et al. What are the benefits and harms of conservative, pharmacological, and surgical
management options for women with bladder outlet obstruction? CRD42020183839 2020.
https://www.crd.york.ac.uk/prospero/display_record.php?ID=CRD42020183839
12. Farag, F., et al. What are the benefits and potential harms of the surgical and non-surgical treatment
options for the management of women with overactive bladder syndrome? 2020.
https://www.crd.york.ac.uk/prospero/display_record.php?RecordID=192207
13. Guyatt, G.H., et al. GRADE: an emerging consensus on rating quality of evidence and strength of
recommendations. BMJ, 2008. 336: 924.
https://pubmed.ncbi.nlm.nih.gov/18436948
14. Guyatt, G.H., et al. What is “quality of evidence” and why is it important to clinicians? BMJ, 2008.
336: 995.
https://pubmed.ncbi.nlm.nih.gov/18456631
15. Phillips B, et al. Oxford Centre for Evidence-based Medicine Levels of Evidence. 1998. Updated by
Jeremy Howick March 2009 [accessed March 2021].
https://www.cebm.net/2009/06/oxford-centre-evidence-based-medicine-levels-evidence-
march-2009/
16. Guyatt, G.H., et al. Going from evidence to recommendations. BMJ, 2008. 336: 1049.
https://pubmed.ncbi.nlm.nih.gov/18467413
17. D’Ancona, C., et al. The International Continence Society (ICS) report on the terminology for adult
male lower urinary tract and pelvic floor symptoms and dysfunction. Neurourol Urodyn, 2019.
38: 433.
https://pubmed.ncbi.nlm.nih.gov/30681183
18. Bonkat, G., et al. EAU Guidelines on Urological Infections. Edn presented at the 36th Annual
Congress, Milan, 2021. EAU Guidelines Office, Arnhem, The Netherlands.
https://uroweb.org/guideline/urological-infections/
19. Engeler, D., et al. EAU Guidelines on Chronic Pelvic Pain. Edn presented at the 36th Annual
Congress, Milan, 2021. EAU Guidelines Office, Arnhem, The Netherlands.
https://uroweb.org/guideline/chronic-pelvic-pain/
20. U.S. Department of Health and Human Services, F.D.A. Guidance for Industry - Patient-Reported
Outcome Measures: Use in Medical Product Development to Support Labeling Claims. 2009.
https://www.fda.gov/media/77832/download
21. Farrell, S.A., et al. Women’s ability to assess their urinary incontinence type using the QUID as an
educational tool. Int Urogynecol J, 2013. 24: 759.
https://pubmed.ncbi.nlm.nih.gov/22940842
22. Hess, R., et al. Long-term efficacy and safety of questionnaire-based initiation of urgency urinary
incontinence treatment. Am J Obstet Gynecol, 2013. 209: 244 e1.
https://pubmed.ncbi.nlm.nih.gov/23659987
23. ICIQ Development Group. International Consultation on Incontinence Questionnaire Female Lower
Urinary Tract Symptoms Modules (ICIQ-FLUTS) ICIQ Project, 2014.
https://iciq.net/iciq-fluts

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 103
24. Chan, S.S., et al. Responsiveness of the Pelvic Floor Distress Inventory and Pelvic Floor Impact
Questionnaire in women undergoing treatment for pelvic floor disorders. Int Urogynecol J, 2013.
24: 213.
https://pubmed.ncbi.nlm.nih.gov/22669425
25. Kim, J., et al. Is there a relationship between incontinence impact questionnaire 7 score after
surgery for stress urinary incontiennce and patient-perceived satisfaction and improvement? J Urol,
2013. 189: e647.
https://www.researchgate.net/publication/288303017
26. Rosier, P.F.W.M., et al. Committee 6. Urodynamic testing, In: Incontinence, 6th Edn. 2017, Abrams,
P., Cardozo, L., Wagg, A., Wein, A. (Eds). Bristol, UK.
https://www.ics.org/publications/ici_6/Incontinence_6th_Edition_2017_eBook_v2.pdf
27. Shy, M., et al. Objective Evaluation of Overactive Bladder: Which Surveys Should I Use? Curr
Bladder Dysfunct Rep, 2013. 8: 45.
https://pubmed.ncbi.nlm.nih.gov/23439804
28. Haylen, B.T., et al. An International Urogynecological Association (IUGA)/International Continence
Society (ICS) joint terminology and classification of the complications related directly to the insertion
of prostheses (meshes, implants, tapes) and grafts in female pelvic floor surgery. Neurourol Urodyn,
2011. 30: 2.
https://pubmed.ncbi.nlm.nih.gov/21181958
29. Brown, J.S., et al. Measurement characteristics of a voiding diary for use by men and women with
overactive bladder. Urology, 2003. 61: 802.
https://pubmed.ncbi.nlm.nih.gov/12670569
30. Nygaard, I., et al. Reproducibility of the seven-day voiding diary in women with stress urinary
incontinence. Int Urogynecol J Pelvic Floor Dysfunct, 2000. 11: 15.
https://pubmed.ncbi.nlm.nih.gov/10738929
31. Bright, E., et al. Developing and validating the International Consultation on Incontinence
Questionnaire bladder diary. Eur Urol, 2014. 66: 294.
https://pubmed.ncbi.nlm.nih.gov/24647230
32. Jimenez-Cidre, M.A., et al. The 3-day bladder diary is a feasible, reliable and valid tool to evaluate
the lower urinary tract symptoms in women. Neurourol Urodyn, 2015. 34: 128.
https://pubmed.ncbi.nlm.nih.gov/24264859
33. Ertberg, P., et al. A comparison of three methods to evaluate maximum bladder capacity:
cystometry, uroflowmetry and a 24-h voiding diary in women with urinary incontinence. Acta Obstet
Gynecol Scand, 2003. 82: 374.
https://pubmed.ncbi.nlm.nih.gov/12716323
34. Fitzgerald, M.P., et al. Variability of 24-hour voiding diary variables among asymptomatic women.
J Urol, 2003. 169: 207.
https://pubmed.ncbi.nlm.nih.gov/12478137
35. Burgio, K.L., et al. Behavioral vs drug treatment for urge urinary incontinence in older women: a
randomized controlled trial. JAMA, 1998. 280: 1995.
https://pubmed.ncbi.nlm.nih.gov/9863850
36. Fayyad, A.M., et al. Urine production and bladder diary measurements in women with type 2
diabetes mellitus and their relation to lower urinary tract symptoms and voiding dysfunction.
Neurourol Urodyn, 2010. 29: 354.
https://pubmed.ncbi.nlm.nih.gov/19760759
37. Homma, Y., et al. Assessment of overactive bladder symptoms: comparison of 3-day bladder diary
and the overactive bladder symptoms score. Urology, 2011. 77: 60.
https://pubmed.ncbi.nlm.nih.gov/20951412
38. Stav, K., et al. Women overestimate daytime urinary frequency: the importance of the bladder diary.
J Urol, 2009. 181: 2176.
https://pubmed.ncbi.nlm.nih.gov/19296975
39. van Brummen, H.J., et al. The association between overactive bladder symptoms and objective
parameters from bladder diary and filling cystometry. Neurourol Urodyn, 2004. 23: 38.
https://pubmed.ncbi.nlm.nih.gov/14694455
40. Yap, T.L., et al. A systematic review of the reliability of frequency-volume charts in urological
research and its implications for the optimum chart duration. BJU Int, 2007. 99: 9.
https://pubmed.ncbi.nlm.nih.gov/16956355
41. Bright, E., et al. Urinary diaries: evidence for the development and validation of diary content,
format, and duration. Neurourol Urodyn, 2011. 30: 348.
https://pubmed.ncbi.nlm.nih.gov/21284023

104 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021
42. Buchsbaum, G.M., et al. Utility of urine reagent strip in screening women with incontinence for
urinary tract infection. Int Urogynecol J Pelvic Floor Dysfunct, 2004. 15: 391.
https://pubmed.ncbi.nlm.nih.gov/15278254
43. Arinzon, Z., et al. Clinical presentation of urinary tract infection (UTI) differs with aging in women.
Arch Gerontol Geriatr, 2012. 55: 145.
https://pubmed.ncbi.nlm.nih.gov/21963175
44. Moore, E.E., et al. Urinary incontinence and urinary tract infection: temporal relationships in
postmenopausal women. Obstet Gynecol, 2008. 111: 317.
https://pubmed.ncbi.nlm.nih.gov/18238968
45. Ouslander, J.G., et al. Does eradicating bacteriuria affect the severity of chronic urinary incontinence
in nursing home residents? Ann Intern Med, 1995. 122: 749.
https://pubmed.ncbi.nlm.nih.gov/7717597
46. Abrams, P. Bladder outlet obstruction index, bladder contractility index and bladder voiding
efficiency: three simple indices to define bladder voiding function. BJU Int, 1999. 84: 14.
https://pubmed.ncbi.nlm.nih.gov/10444116
47. Griffiths, D.J., et al. Variability of post-void residual urine volume in the elderly. Urol Res, 1996. 24: 23.
https://pubmed.ncbi.nlm.nih.gov/8966837
48. Marks, L.S., et al. Three-dimensional ultrasound device for rapid determination of bladder volume.
Urology, 1997. 50: 341.
https://pubmed.ncbi.nlm.nih.gov/9301695
49. Nygaard, I.E. Postvoid residual volume cannot be accurately estimated by bimanual examination. Int
Urogynecol J Pelvic Floor Dysfunct, 1996. 7: 74.
https://pubmed.ncbi.nlm.nih.gov/8798090
50. Ouslander, J.G., et al. Use of a portable ultrasound device to measure post-void residual volume
among incontinent nursing home residents. J Am Geriatr Soc, 1994. 42: 1189.
https://pubmed.ncbi.nlm.nih.gov/7963206
51. Stoller, M.L., et al. The accuracy of a catheterized residual urine. J Urol, 1989. 141: 15.
https://pubmed.ncbi.nlm.nih.gov/2908944
52. Gehrich, A., et al. Establishing a mean postvoid residual volume in asymptomatic perimenopausal
and postmenopausal women. Obstet Gynecol, 2007. 110: 827.
https://pubmed.ncbi.nlm.nih.gov/17906016
53. Robinson, D., et al. Defining female voiding dysfunction: ICI-RS 2011. Neurourol Urodyn, 2012.
31: 313.
https://pubmed.ncbi.nlm.nih.gov/22415792
54. Haylen, B.T., et al. Immediate postvoid residual volumes in women with symptoms of pelvic floor
dysfunction. Obstet Gynecol, 2008. 111: 1305.
https://pubmed.ncbi.nlm.nih.gov/18515513
55. Lukacz, E.S., et al. Elevated postvoid residual in women with pelvic floor disorders: prevalence and
associated risk factors. Int Urogynecol J Pelvic Floor Dysfunct, 2007. 18: 397.
https://pubmed.ncbi.nlm.nih.gov/16804634
56. Milleman, M., et al. Post-void residual urine volume in women with overactive bladder symptoms.
J Urol, 2004. 172: 1911.
https://pubmed.ncbi.nlm.nih.gov/15540753
57. Tseng, L.H., et al. Postvoid residual urine in women with stress incontinence. Neurourol Urodyn,
2008. 27: 48.
https://pubmed.ncbi.nlm.nih.gov/17563112
58. Siroky, M.B., et al. The flow rate nomogram: I. Development. J Urol, 1979. 122: 665.
https://pubmed.ncbi.nlm.nih.gov/159366
59. Haylen, B.T., et al. Maximum and average urine flow rates in normal male and female populations--
the Liverpool nomograms. Br J Urol, 1989. 64: 30.
https://pubmed.ncbi.nlm.nih.gov/2765766
60. Rosier, P., et al. International Continence Society Good Urodynamic Practices and Terms 2016:
Urodynamics, uroflowmetry, cystometry, and pressure-flow study. Neurourol Urodyn, 2017. 36: 1243.
https://pubmed.ncbi.nlm.nih.gov/27917521
61. Abrams, P., et al. United Kingdom Continence Society: Minimum standards for urodynamic studies,
2018. Neurourol Urodyn, 2019. 38: 838.
https://pubmed.ncbi.nlm.nih.gov/30648750
62. Brostrom, S., et al. Short-term reproducibility of cystometry and pressure-flow micturition studies in
healthy women. Neurourol Urodyn, 2002. 21: 457.
https://pubmed.ncbi.nlm.nih.gov/12232880

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 105
63. Broekhuis, S.R., et al. Reproducibility of same session repeated cystometry and pressure-flow
studies in women with symptoms of urinary incontinence. Neurourol Urodyn, 2010. 29: 428.
https://pubmed.ncbi.nlm.nih.gov/19618451
64. Dorflinger, A., et al. Urethral pressure profile: is it affected by position? Neurourol Urodyn, 2002.
21: 553.
https://pubmed.ncbi.nlm.nih.gov/12382246
65. Wang, A.C., et al. A comparison of urethral pressure profilometry using microtip and double-lumen
perfusion catheters in women with genuine stress incontinence. BJOG, 2002. 109: 322.
https://pubmed.ncbi.nlm.nih.gov/11950188
66. Zehnder, P., et al. Air charged and microtip catheters cannot be used interchangeably for urethral
pressure measurement: a prospective, single-blind, randomized trial. J Urol, 2008. 180: 1013.
https://pubmed.ncbi.nlm.nih.gov/18639301
67. NICE Guidance – Urinary incontinence and pelvic organ prolapse in women: management. BJU Int,
2019. 123: 777.
https://pubmed.ncbi.nlm.nih.gov/31008559
68. van Leijsen, S.A., et al. The correlation between clinical and urodynamic diagnosis in classifying
the type of urinary incontinence in women. A systematic review of the literature. Neurourol Urodyn,
2011. 30: 495.
https://pubmed.ncbi.nlm.nih.gov/21298721
69. Schick, E., et al. Predictive value of maximum urethral closure pressure, urethral hypermobility and
urethral incompetence in the diagnosis of clinically significant female genuine stress incontinence.
J Urol, 2004. 171: 1871.
https://pubmed.ncbi.nlm.nih.gov/15076296
70. Albo, M.E., et al. Burch colposuspension versus fascial sling to reduce urinary stress incontinence.
N Engl J Med, 2007. 356: 2143.
https://pubmed.ncbi.nlm.nih.gov/17517855
71. Klarskov, N. Urethral pressure reflectometry. A method for simultaneous measurements of pressure
and cross-sectional area in the female urethra. Dan Med J, 2012. 59: B4412.
https://pubmed.ncbi.nlm.nih.gov/22381095
72. Dokmeci, F., et al. Comparison of ambulatory versus conventional urodynamics in females with
urinary incontinence. Neurourol Urodyn, 2010. 29: 518.
https://pubmed.ncbi.nlm.nih.gov/19731314
73. Radley, S.C., et al. Conventional and ambulatory urodynamic findings in women with symptoms
suggestive of bladder overactivity. J Urol, 2001. 166: 2253.
https://pubmed.ncbi.nlm.nih.gov/11696746
74. Akikwala, T.V., et al. Comparison of diagnostic criteria for female bladder outlet obstruction. J Urol,
2006. 176: 2093.
https://pubmed.ncbi.nlm.nih.gov/17070266
75. Rademakers, K., et al. Recommendations for future development of contractility and obstruction
nomograms for women. ICI-RS 2014. Neurourol Urodyn, 2016. 35: 307.
https://pubmed.ncbi.nlm.nih.gov/26872573
76. Osman, N.I., et al. Detrusor underactivity and the underactive bladder: a new clinical entity? A
review of current terminology, definitions, epidemiology, aetiology, and diagnosis. Eur Urol, 2014.
65: 389.
https://pubmed.ncbi.nlm.nih.gov/24184024
77. Nitti, V.W., et al. Diagnosing bladder outlet obstruction in women. J Urol, 1999. 161: 1535.
https://pubmed.ncbi.nlm.nih.gov/10210391
78. Clement, K.D., et al. Urodynamic studies for management of urinary incontinence in children and
adults. Cochrane Database Syst Rev, 2013. 2013: CD003195.
https://pubmed.ncbi.nlm.nih.gov/24166676
79. Rachaneni, S., et al. Does preoperative urodynamics improve outcomes for women undergoing
surgery for stress urinary incontinence? A systematic review and meta-analysis. BJOG, 2015. 122:
8.
https://pubmed.ncbi.nlm.nih.gov/25041381
80. Nager, C.W., et al. A randomized trial of urodynamic testing before stress-incontinence surgery.
N Engl J Med, 2012. 366: 1987.
https://pubmed.ncbi.nlm.nih.gov/22551104
81. Serati, M., et al. Urodynamics useless before surgery for female stress urinary incontinence: Are you
sure? Results from a multicenter single nation database. Neurourol Urodyn, 2016. 35: 809.
https://pubmed.ncbi.nlm.nih.gov/26061435

106 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021
82. Gravas, S., et al., EAU Guidelines on the management of Non-Neurogenice Male LUTS. Edn
presented at the 36th Annual Congress, Milan, 2021. EAU Guidelines Office, Arnhem, The
Netherlands.
https://uroweb.org/guideline/treatment-of-non-neurogenic-male-luts/
83. Nager, C.W., et al. Baseline urodynamic predictors of treatment failure 1 year after mid urethral sling
surgery. J Urol, 2011. 186: 597.
https://pubmed.ncbi.nlm.nih.gov/21683412
84. Nitti, V.W., et al. Response to fesoterodine in patients with an overactive bladder and urgency
urinary incontinence is independent of the urodynamic finding of detrusor overactivity. BJU Int,
2010. 105: 1268.
https://pubmed.ncbi.nlm.nih.gov/19889062
85. Rovner, E., et al. Urodynamic results and clinical outcomes with intradetrusor injections of
onabotulinumtoxinA in a randomized, placebo-controlled dose-finding study in idiopathic overactive
bladder. Neurourol Urodyn, 2011. 30: 556.
https://pubmed.ncbi.nlm.nih.gov/21351127
86. Koldewijn, E.L., et al. Predictors of success with neuromodulation in lower urinary tract dysfunction:
results of trial stimulation in 100 patients. J Urol, 1994. 152: 2071.
https://pubmed.ncbi.nlm.nih.gov/7966677
87. South, M.M., et al. Detrusor overactivity does not predict outcome of sacral neuromodulation test
stimulation. Int Urogynecol J Pelvic Floor Dysfunct, 2007. 18: 1395.
https://pubmed.ncbi.nlm.nih.gov/17364132
88. Dawson, T., et al. Factors predictive of post-TVT voiding dysfunction. Int Urogynecol J Pelvic Floor
Dysfunct, 2007. 18: 1297.
https://pubmed.ncbi.nlm.nih.gov/17347790
89. Hong, B., et al. Factors predictive of urinary retention after a tension-free vaginal tape procedure for
female stress urinary incontinence. J Urol, 2003. 170: 852.
https://pubmed.ncbi.nlm.nih.gov/12913715
90. Miller, E.A., et al. Preoperative urodynamic evaluation may predict voiding dysfunction in women
undergoing pubovaginal sling. J Urol, 2003. 169: 2234.
https://pubmed.ncbi.nlm.nih.gov/12771757
91. Groen, J., et al. Bladder contraction strength parameters poorly predict the necessity of long-term
catheterization after a pubovaginal rectus fascial sling procedure. J Urol, 2004. 172: 1006.
https://pubmed.ncbi.nlm.nih.gov/15311024
92. Abdel-Fattah, M., et al. Pelvicol pubovaginal sling versus tension-free vaginal tape for treatment of
urodynamic stress incontinence: a prospective randomized three-year follow-up study. Eur Urol,
2004. 46: 629.
https://pubmed.ncbi.nlm.nih.gov/15474274
93. Lemack, G.E., et al. Normal preoperative urodynamic testing does not predict voiding dysfunction
after Burch colposuspension versus pubovaginal sling. J Urol, 2008. 180: 2076.
https://pubmed.ncbi.nlm.nih.gov/18804239
94. Al Afraa, T., et al. Normal lower urinary tract assessment in women: I. Uroflowmetry and post-void
residual, pad tests, and bladder diaries. Int Urogynecol J, 2012. 23: 681.
https://pubmed.ncbi.nlm.nih.gov/21935667
95. Krhut, J., et al. Pad weight testing in the evaluation of urinary incontinence. Neurourol Urodyn, 2014.
33: 507.
https://pubmed.ncbi.nlm.nih.gov/23797972
96. Painter, V., et al. Does patient activity level affect 24-hr pad test results in stress-incontinent
women? Neurourol Urodyn, 2012. 31: 143.
https://pubmed.ncbi.nlm.nih.gov/21780173
97. Rimstad, L., et al. Pad stress tests with increasing load for the diagnosis of stress urinary
incontinence. Neurourol Urodyn, 2014. 33: 1135.
https://pubmed.ncbi.nlm.nih.gov/23913797
98. Costantini, E., et al. Sensitivity and specificity of one-hour pad test as a predictive value for female
urinary incontinence. Urol Int, 2008. 81: 153.
https://pubmed.ncbi.nlm.nih.gov/18758212
99. Richter, H.E., et al. Demographic and clinical predictors of treatment failure one year after
midurethral sling surgery. Obstet Gynecol, 2011. 117: 913.
https://pubmed.ncbi.nlm.nih.gov/21422865

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 107
100. Ward, K.L., et al. A prospective multicenter randomized trial of tension-free vaginal tape and
colposuspension for primary urodynamic stress incontinence: two-year follow-up. Am J Obstet
Gynecol, 2004. 190: 324.
https://pubmed.ncbi.nlm.nih.gov/14981369
101. Lose, G., et al. Pad-weighing test performed with standardized bladder volume. Urology, 1988. 32: 78.
https://pubmed.ncbi.nlm.nih.gov/3388665
102. Khullar, V., et al., Committee 7 Imaging, neurophysiological testing and other tests, In: Incontinence,
6th Edn. 2017, Abrams, P., Cardozo, L., Wagg, A., Wein, A. (Eds). Bristol, UK.
https://www.ics.org/publications/ici_6/Incontinence_6th_Edition_2017_eBook_v2.pdf
103. Panayi, D.C., et al. Transvaginal ultrasound measurement of bladder wall thickness: a more reliable
approach than transperineal and transabdominal approaches. BJU Int, 2010. 106: 1519.
https://pubmed.ncbi.nlm.nih.gov/20438565
104. Antunes-Lopes, T., et al. Biomarkers in lower urinary tract symptoms/overactive bladder: a critical
overview. Curr Opin Urol, 2014. 24: 352.
https://pubmed.ncbi.nlm.nih.gov/24841379
105. Lekskulchai, O., et al. Is detrusor hypertrophy in women associated with voiding dysfunction? Aust
N Z J Obstet Gynaecol, 2009. 49: 653.
https://pubmed.ncbi.nlm.nih.gov/20070717
106. Woodfield, C.A., et al. Imaging pelvic floor disorders: trend toward comprehensive MRI. AJR Am
J Roentgenol, 2010. 194: 1640.
https://pubmed.ncbi.nlm.nih.gov/20489108
107. Lockhart, M.E., et al. Reproducibility of dynamic MR imaging pelvic measurements: a multi-
institutional study. Radiology, 2008. 249: 534.
https://pubmed.ncbi.nlm.nih.gov/18796659
108. Shek, K.L., et al. The urethral motion profile before and after suburethral sling placement. J Urol,
2010. 183: 1450.
https://pubmed.ncbi.nlm.nih.gov/20171657
109. Chantarasorn, V., et al. Sonographic appearance of transobturator slings: implications for function
and dysfunction. Int Urogynecol J, 2011. 22: 493.
https://pubmed.ncbi.nlm.nih.gov/20967418
110. Haylen, B.T., et al. An International Urogynecological Association (IUGA)/International Continence
Society (ICS) joint report on the terminology for female pelvic floor dysfunction. Neurourol Urodyn,
2010. 29: 4.
https://pubmed.ncbi.nlm.nih.gov/19941278
111. Irwin, D.E., et al. Population-based survey of urinary incontinence, overactive bladder, and other
lower urinary tract symptoms in five countries: results of the EPIC study. Eur Urol, 2006. 50: 1306.
https://pubmed.ncbi.nlm.nih.gov/17049716
112. Glazener, C.M., et al. Urodynamic studies for management of urinary incontinence in children and
adults. Cochrane Database Syst Rev, 2012. 1: CD003195.
https://pubmed.ncbi.nlm.nih.gov/22258952
113. Sarma, A.V., et al. Risk factors for urinary incontinence among women with type 1 diabetes: findings
from the epidemiology of diabetes interventions and complications study. Urology, 2009. 73: 1203.
https://pubmed.ncbi.nlm.nih.gov/19362350
114. Abrams, P., et al. 5th International Consultation on Incontinence, Paris, February 2012.
http://www.icud.info/incontinence.html
115. Geng, V., et al., Catheterisation Indwelling catheters in adults – Urethral and Suprapubic - Evidence-
based Guidelines for Best Practice in Urological Health Care. 2012. European Asssociation of
Urology Nurses, Arnhem, the Netherlands.
https://nurses.uroweb.org/guideline/catheterisation-indwelling-catheters-in-adults-urethral-and-
suprapubic/
116. Brazzelli, M., et al. Absorbent products for containing urinary and/or fecal incontinence in adults.
J Wound Ostomy Continence Nurs, 2002. 29: 45.
https://pubmed.ncbi.nlm.nih.gov/11810074
117. Fader, M., et al. A multi-centre evaluation of absorbent products for men with light urinary
incontinence. Neurourol Urodyn, 2006. 25: 689.
https://pubmed.ncbi.nlm.nih.gov/17009303
118. Fader, M., et al. Absorbent products for urinary/faecal incontinence: a comparative evaluation of key
product designs. Health Technol Assess, 2008. 12: iii.
https://pubmed.ncbi.nlm.nih.gov/18547500

108 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021
119. Jahn, P., et al. Types of indwelling urinary catheters for long-term bladder drainage in adults.
Cochrane Database Syst Rev, 2012. 10: CD004997.
https://pubmed.ncbi.nlm.nih.gov/23076911
120. Hunter, K.F., et al. Long-term bladder drainage: Suprapubic catheter versus other methods: a
scoping review. Neurourol Urodyn, 2013. 32: 944.
https://pubmed.ncbi.nlm.nih.gov/23192860
121. Prieto, J., et al. Intermittent catheterisation for long-term bladder management. Cochrane Database
Syst Rev, 2014: CD006008.
https://pubmed.ncbi.nlm.nih.gov/25208303
122. Tradewell, M., et al. Systematic review and practice policy statements on urinary tract infection
prevention in adults with spina bifida. Transl Androl Urol, 2018. 7: S205.
https://pubmed.ncbi.nlm.nih.gov/29928619
123. Woodbury, M.G., et al. Intermittent catheterization practices following spinal cord injury: a national
survey. Can J Urol, 2008. 15: 4065.
https://pubmed.ncbi.nlm.nih.gov/18570710
124. Cottenden A, et al. Management using continence products. In: Incontinence, 6th Edn. 2017,
Abrams, P., Cardozo, L., Wagg, A., Wein, A. (Eds). Bristol, UK.
https://www.ics.org/publications/ici_6/Incontinence_6th_Edition_2017_eBook_v2.pdf
125. Prieto, J., et al. Catheter designs, techniques and strategies for intermittent catheterisation: What
is the evidence for preventing symptomatic UTI and other complications? A Cochrane systematic
review. Eur Urol Suppl, 2014. 13: e762.
https://pubmed.ncbi.nlm.nih.gov/17943874
126. Hakansson, M.A. Reuse versus single-use catheters for intermittent catheterization: what is safe
and preferred? Review of current status. Spinal Cord, 2014. 52: 511.
https://pubmed.ncbi.nlm.nih.gov/24861702
127. Hagen, S., et al. Washout policies in long-term indwelling urinary catheterisation in adults. Cochrane
Database Syst Rev, 2010: CD004012.
https://pubmed.ncbi.nlm.nih.gov/20238325
128. Niel-Weise, B.S., et al. Urinary catheter policies for long-term bladder drainage. Cochrane Database
Syst Rev, 2012: CD004201.
https://pubmed.ncbi.nlm.nih.gov/22895939
129. Fisher, H., et al. Continuous low-dose antibiotic prophylaxis for adults with repeated urinary tract
infections (AnTIC): a randomised, open-label trial. Lancet Infect Dis. , 2018. 18: 957.
https://pubmed.ncbi.nlm.nih.gov/30037647
130. Arya, L.A., et al. Dietary caffeine intake and the risk for detrusor instability: a case-control study.
Obstet Gynecol, 2000. 96: 85.
https://pubmed.ncbi.nlm.nih.gov/10862848
131. Bryant, C.M., et al. Caffeine reduction education to improve urinary symptoms. Br J Nurs, 2002.
11: 560.
https://pubmed.ncbi.nlm.nih.gov/11979209
132. Le Berre, M., et al. What do we really know about the role of caffeine on urinary tract symptoms?
A scoping review on caffeine consumption and lower urinary tract symptoms in adults. Neurourol
Urodyn, 2020. 39: 1217.
https://pubmed.ncbi.nlm.nih.gov/32270903
133. Swithinbank, L., et al. The effect of fluid intake on urinary symptoms in women. J Urol, 2005.
174: 187.
https://pubmed.ncbi.nlm.nih.gov/15947624
134. Dowd, T.T., et al. Fluid intake and urinary incontinence in older community-dwelling women.
J Community Health Nurs, 1996. 13: 179.
https://pubmed.ncbi.nlm.nih.gov/8916607
135. Hashim, H., et al. How should patients with an overactive bladder manipulate their fluid intake? BJU
Int, 2008. 102: 62.
https://pubmed.ncbi.nlm.nih.gov/18284414
136. Zimmern, P., et al. Effect of fluid management on fluid intake and urge incontinence in a trial for
overactive bladder in women. BJU Int, 2010. 105: 1680.
https://pubmed.ncbi.nlm.nih.gov/19912207
137. Hunskaar, S. A systematic review of overweight and obesity as risk factors and targets for clinical
intervention for urinary incontinence in women. Neurourol Urodyn, 2008. 27: 749.
https://pubmed.ncbi.nlm.nih.gov/18951445

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 109
138. Subak, L.L., et al. Weight loss to treat urinary incontinence in overweight and obese women. N Engl
J Med, 2009. 360: 481.
https://pubmed.ncbi.nlm.nih.gov/19179316
139. Chen, C.C., et al. Obesity is associated with increased prevalence and severity of pelvic floor
disorders in women considering bariatric surgery. Surg Obes Relat Dis, 2009. 5: 411.
https://pubmed.ncbi.nlm.nih.gov/19136310
140. Hannestad, Y.S., et al. Are smoking and other lifestyle factors associated with female urinary
incontinence? The Norwegian EPINCONT Study. BJOG, 2003. 110: 247.
https://pubmed.ncbi.nlm.nih.gov/12628262
141. Danforth, K.N., et al. Risk factors for urinary incontinence among middle-aged women. Am J Obstet
Gynecol, 2006. 194: 339.
https://pubmed.ncbi.nlm.nih.gov/16458626
142. Imamura, M., et al. Systematic review and economic modelling of the effectiveness and cost-
effectiveness of non-surgical treatments for women with stress urinary incontinence. Health Technol
Assess, 2010. 14: 1.
https://pubmed.ncbi.nlm.nih.gov/20738930
143. Bo, K., et al. An International Urogynecological Association (IUGA)/International Continence Society
(ICS) joint report on the terminology for the conservative and nonpharmacological management of
female pelvic floor dysfunction. Int Urogynecol J, 2017. 28: 191.
https://pubmed.ncbi.nlm.nih.gov/27921161
144. Eustice, S., et al. Prompted voiding for the management of urinary incontinence in adults. Cochrane
Database Syst Rev, 2000: CD002113.
https://pubmed.ncbi.nlm.nih.gov/10796861
145. Flanagan, L., et al. Systematic review of care intervention studies for the management of
incontinence and promotion of continence in older people in care homes with urinary incontinence
as the primary focus (1966-2010). Geriatr Gerontol Int, 2012. 12: 600.
https://pubmed.ncbi.nlm.nih.gov/22672329
146. Ostaszkiewicz, J., et al. Habit retraining for the management of urinary incontinence in adults.
Cochrane Database Syst Rev, 2004: CD002801.
https://pubmed.ncbi.nlm.nih.gov/15106179
147. Shamliyan, T., et al., Nonsurgical Treatments for Urinary Incontinence in Adult Women: Diagnosis
and Comparative Effectiveness. 2012, IUGA-ICS Conservative Management for Female Pelvic Floor
Dysfunction: Rockville (MD), U.S.A.
https://pubmed.ncbi.nlm.nih.gov/22624162
148. Rai, B.P., et al. Anticholinergic drugs versus non-drug active therapies for non-neurogenic overactive
bladder syndrome in adults. Cochrane Database Syst Rev, 2012. 12: CD003193.
https://pubmed.ncbi.nlm.nih.gov/23235594
149. Shafik, A., et al. Overactive bladder inhibition in response to pelvic floor muscle exercises. World
J Urol, 2003. 20: 374.
https://pubmed.ncbi.nlm.nih.gov/12682771
150. Bo, K., et al. Is pelvic floor muscle training effective for symptoms of overactive bladder in women?
A systematic review. Physiotherapy, 2020. 106: 65.
https://pubmed.ncbi.nlm.nih.gov/32026847
151. Stewart, F., et al. Electrical stimulation with non-implanted electrodes for overactive bladder in
adults. Cochrane Database Syst Rev, 2016. 12: CD010098.
https://pubmed.ncbi.nlm.nih.gov/27935011
152. Zhao, Y., et al. Acupuncture for adults with overactive bladder: A systematic review and meta-
analysis of randomized controlled trials. Medicine (Baltimore), 2018. 97: e9838.
https://pubmed.ncbi.nlm.nih.gov/29465566
153. Finazzi-Agro, E., et al. Percutaneous tibial nerve stimulation effects on detrusor overactivity
incontinence are not due to a placebo effect: a randomized, double-blind, placebo controlled trial.
J Urol, 2010. 184: 2001.
https://pubmed.ncbi.nlm.nih.gov/20850833
154. Peters, K.M., et al. Randomized trial of percutaneous tibial nerve stimulation versus Sham efficacy in
the treatment of overactive bladder syndrome: results from the SUmiT trial. J Urol, 2010. 183: 1438.
https://pubmed.ncbi.nlm.nih.gov/20171677
155. Peters, K.M., et al. Randomized trial of percutaneous tibial nerve stimulation versus extended-
release tolterodine: results from the overactive bladder innovative therapy trial. J Urol, 2009.
182: 1055.
https://pubmed.ncbi.nlm.nih.gov/19616802

110 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021
156. Peters, K.M., et al. Percutaneous tibial nerve stimulation for the long-term treatment of overactive
bladder: 3-year results of the STEP study. J Urol, 2013. 189: 2194.
https://pubmed.ncbi.nlm.nih.gov/23219541
157. Schreiner, L., et al. Randomized trial of transcutaneous tibial nerve stimulation to treat urge urinary
incontinence in older women. Int Urogynecol J, 2010. 21: 1065.
https://pubmed.ncbi.nlm.nih.gov/20458465
158. Booth, J., et al. The effectiveness of transcutaneous tibial nerve stimulation (TTNS) for adults with
overactive bladder syndrome: A systematic review. Neurourol Urodyn, 2018. 37: 528.
https://pubmed.ncbi.nlm.nih.gov/28731583
159. Chapple, C., et al. The effects of antimuscarinic treatments in overactive bladder: a systematic
review and meta-analysis. Eur Urol, 2005. 48: 5.
https://pubmed.ncbi.nlm.nih.gov/15885877
160. Chapple, C.R., et al. The effects of antimuscarinic treatments in overactive bladder: an update of a
systematic review and meta-analysis. Eur Urol, 2008. 54: 543.
https://pubmed.ncbi.nlm.nih.gov/18599186
161. McDonagh, M.S., et al. Drug Class Review: Agents for Overactive Bladder: Final Report Update 4.
2009: Portland (OR), U.S.A.
https://pubmed.ncbi.nlm.nih.gov/21089246
162. Shamliyan, T.A., et al. Systematic review: randomized, controlled trials of nonsurgical treatments for
urinary incontinence in women. Ann Intern Med, 2008. 148: 459.
https://pubmed.ncbi.nlm.nih.gov/18268288
163. Buser, N., et al. Efficacy and adverse events of antimuscarinics for treating overactive bladder:
network meta-analyses. Eur Urol, 2012. 62: 1040.
https://pubmed.ncbi.nlm.nih.gov/22999811
164. Reynolds, W.S., et al. Comparative Effectiveness of Anticholinergic Therapy for Overactive Bladder
in Women: A Systematic Review and Meta-analysis. Obstet Gynecol, 2015. 125: 1423.
https://pubmed.ncbi.nlm.nih.gov/26000514
165. Herbison, P., et al. Which anticholinergic is best for people with overactive bladders? A network
meta-analysis. Neurourol Urodyn, 2019. 38: 525.
https://pubmed.ncbi.nlm.nih.gov/30575999
166. Chapple, C., et al. Superiority of fesoterodine 8 mg vs 4 mg in reducing urgency urinary
incontinence episodes in patients with overactive bladder: results of the randomised, double-blind,
placebo-controlled EIGHT trial. BJU Int, 2014. 114: 418.
https://pubmed.ncbi.nlm.nih.gov/24552358
167. Kaplan, S.A., et al. Efficacy and safety of fesoterodine 8 mg in subjects with overactive bladder after
a suboptimal response to tolterodine ER. Int J Clin Pract, 2014. 68: 1065.
https://pubmed.ncbi.nlm.nih.gov/24898471
168. Goldfischer, E.R., et al. Efficacy and safety of oxybutynin topical gel 3% in patients with urgency
and/or mixed urinary incontinence: a randomized, double-blind, placebo-controlled study. Neurourol
Urodyn, 2015. 34: 37.
https://pubmed.ncbi.nlm.nih.gov/24133005
169. Nazir, J., et al. Comparative efficacy and tolerability of solifenacin 5 mg/day versus other oral
antimuscarinic agents in overactive bladder: A systematic literature review and network meta-
analysis. Neurourol Urodyn, 2018. 37: 986.
https://pubmed.ncbi.nlm.nih.gov/29140559
170. Novara, G., et al. A systematic review and meta-analysis of randomized controlled trials with
antimuscarinic drugs for overactive bladder. Eur Urol, 2008. 54: 740.
https://pubmed.ncbi.nlm.nih.gov/18632201
171. Chapple, C., et al. Clinical efficacy, safety, and tolerability of once-daily fesoterodine in subjects with
overactive bladder. Eur Urol, 2007. 52: 1204.
https://pubmed.ncbi.nlm.nih.gov/17651893
172. Herschorn, S., et al. Comparison of fesoterodine and tolterodine extended release for the treatment
of overactive bladder: a head-to-head placebo-controlled trial. BJU Int, 2010. 105: 58.
https://pubmed.ncbi.nlm.nih.gov/20132103
173. DuBeau, C.E., et al. Efficacy and tolerability of fesoterodine versus tolterodine in older and younger
subjects with overactive bladder: a post hoc, pooled analysis from two placebo-controlled trials.
Neurourol Urodyn, 2012. 31: 1258.
https://pubmed.ncbi.nlm.nih.gov/22907761

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 111
174. Hartmann, K.E., et al. Treatment of overactive bladder in women. Evid Rep Technol Assess (Full
Rep), 2009: 1.
https://pubmed.ncbi.nlm.nih.gov/19947666
175. Goode, P.S., et al. Incontinence in older women. JAMA, 2010. 303: 2172.
https://pubmed.ncbi.nlm.nih.gov/20516418
176. Gormley, E.A., et al. Diagnosis and treatment of overactive bladder (non-neurogenic) in adults: AUA/
SUFU guideline. J Urol, 2012. 188: 2455.
https://pubmed.ncbi.nlm.nih.gov/23098785
177. Mattiasson, A., et al. Efficacy of simplified bladder training in patients with overactive bladder
receiving a solifenacin flexible-dose regimen: results from a randomized study. BJU Int, 2010.
105: 1126.
https://pubmed.ncbi.nlm.nih.gov/19818077
178. Ayeleke, R.O., et al. Pelvic floor muscle training added to another active treatment versus the same
active treatment alone for urinary incontinence in women. Cochrane Database Syst Rev, 2015:
CD010551.
https://pubmed.ncbi.nlm.nih.gov/26526663
179. Manriquez, V., et al. Transcutaneous posterior tibial nerve stimulation versus extended release
oxybutynin in overactive bladder patients. A prospective randomized trial. Eur J Obstet Gynecol
Reprod Biol, 2016. 196: 6.
https://pubmed.ncbi.nlm.nih.gov/26645117
180. Franzen, K., et al. Electrical stimulation compared with tolterodine for treatment of urge/urge
incontinence amongst women--a randomized controlled trial. Int Urogynecol J, 2010. 21: 1517.
https://pubmed.ncbi.nlm.nih.gov/20585755
181. Kosilov, K.V., et al. Randomized controlled trial of cyclic and continuous therapy with trospium
and solifenacin combination for severe overactive bladder in elderly patients with regard to patient
compliance. Ther Adv Urol, 2014. 6: 215.
https://pubmed.ncbi.nlm.nih.gov/25435915
182. Nambiar, A.K., et al. EAU Guidelines on Assessment and Nonsurgical Management of Urinary
Incontinence. Eur Urol, 2018. 73: 596.
https://pubmed.ncbi.nlm.nih.gov/29398262
183. Sand, P.K., et al. Long-term safety, tolerability and efficacy of fesoterodine in subjects with
overactive bladder symptoms stratified by age: pooled analysis of two open-label extension studies.
Drugs Aging, 2012. 29: 119.
https://pubmed.ncbi.nlm.nih.gov/22276958
184. Scarpero, H., et al. Long-term safety, tolerability, and efficacy of fesoterodine treatment in men and
women with overactive bladder symptoms. Curr Med Res Opin, 2011. 27: 921.
https://pubmed.ncbi.nlm.nih.gov/21355814
185. D’Souza, A.O., et al. Persistence, adherence, and switch rates among extended-release and
immediate-release overactive bladder medications in a regional managed care plan. J Manag Care
Pharm, 2008. 14: 291.
https://pubmed.ncbi.nlm.nih.gov/18439051
186. Sears, C.L., et al. Overactive bladder medication adherence when medication is free to patients.
J Urol, 2010. 183: 1077.
https://pubmed.ncbi.nlm.nih.gov/20092838
187. Shaya, F.T., et al. Persistence with overactive bladder pharmacotherapy in a Medicaid population.
Am J Manag Care, 2005. 11: S121.
https://pubmed.ncbi.nlm.nih.gov/16161385
188. Yeaw, J., et al. Comparing adherence and persistence across 6 chronic medication classes.
J Manag Care Pharm, 2009. 15: 728.
https://pubmed.ncbi.nlm.nih.gov/19954264
189. Yu, Y.F., et al. Persistence and adherence of medications for chronic overactive bladder/urinary
incontinence in the california medicaid program. Value Health, 2005. 8: 495.
https://pubmed.ncbi.nlm.nih.gov/16091027
190. Kalder, M., et al. Discontinuation of treatment using anticholinergic medications in patients with
urinary incontinence. Obstet Gynecol, 2014. 124: 794.
https://pubmed.ncbi.nlm.nih.gov/25198276
191. Chapple, C.R., et al. Mirabegron in overactive bladder: a review of efficacy, safety, and tolerability.
Neurourol Urodyn, 2014. 33: 17.
https://pubmed.ncbi.nlm.nih.gov/24127366

112 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021
192. Cui, Y., et al. The efficacy and safety of mirabegron in treating OAB: a systematic review and meta-
analysis of phase III trials. Int Urol Nephrol, 2014. 46: 275.
https://pubmed.ncbi.nlm.nih.gov/23896942
193. Herschorn, S., et al. A phase III, randomized, double-blind, parallel-group, placebo-controlled,
multicentre study to assess the efficacy and safety of the beta(3) adrenoceptor agonist, mirabegron,
in patients with symptoms of overactive bladder. Urology, 2013. 82: 313.
https://pubmed.ncbi.nlm.nih.gov/23769122
194. Yamaguchi, O., et al. Phase III, randomised, double-blind, placebo-controlled study of the beta3-
adrenoceptor agonist mirabegron, 50 mg once daily, in Japanese patients with overactive bladder.
BJU Int, 2014. 113: 951.
https://pubmed.ncbi.nlm.nih.gov/24471907
195. Wu, T., et al. The role of mirabegron in overactive bladder: a systematic review and meta-analysis.
Urol Int, 2014. 93: 326.
https://pubmed.ncbi.nlm.nih.gov/25115445
196. Maman, K., et al. Comparative efficacy and safety of medical treatments for the management of
overactive bladder: a systematic literature review and mixed treatment comparison. Eur Urol, 2014.
65: 755.
https://pubmed.ncbi.nlm.nih.gov/24275310
197. Chapple, C.R., et al. Randomized double-blind, active-controlled phase 3 study to assess 12-month
safety and efficacy of mirabegron, a beta(3)-adrenoceptor agonist, in overactive bladder. Eur Urol,
2013. 63: 296.
https://pubmed.ncbi.nlm.nih.gov/23195283
198. Castro-Diaz, D., et al. The effect of mirabegron on patient-related outcomes in patients with
overactive bladder: the results of post hoc correlation and responder analyses using pooled data
from three randomized Phase III trials. Qual Life Res, 2015. 24: 1719.
https://pubmed.ncbi.nlm.nih.gov/25688038
199. Chapple, C., et al. Efficacy of the beta3-adrenoceptor agonist mirabegron for the treatment of
overactive bladder by severity of incontinence at baseline: a post hoc analysis of pooled data from
three randomised phase 3 trials. Eur Urol, 2015. 67: 11.
https://pubmed.ncbi.nlm.nih.gov/25092537
200. Malik, M., et al. Proarrhythmic safety of repeat doses of mirabegron in healthy subjects: a
randomized, double-blind, placebo-, and active-controlled thorough QT study. Clin Pharmacol Ther,
2012. 92: 696.
https://pubmed.ncbi.nlm.nih.gov/23149929
201. Martin, N., et al. Randomised, double-blind, placebo-controlled study to assess the ocular safety of
mirabegron in normotensive IOP research subjects. Eur Urol Suppl, 2014. 13: e686.
https://www.sciencedirect.com/science/article/abs/pii/S1569905612606836
202. Kelleher, C., et al. A post-HOC analysis of pooled data from 3 randomised phase 3 trials
ofmirabegronin patients with overactive bladder (OAB): Correlations between objective and
subjective outcome measures. Int Urogynecol J Pelvic Floor Dysfunct, 2013. 24: S119. [No abstract
available].
203. Wagg, A., et al. Persistence and adherence with the new beta-3 receptor agonist, mirabegron, versus
antimuscarinics in overactive bladder: Early experience in Canada. Can Urol Assoc J, 2015. 9: 343.
https://pubmed.ncbi.nlm.nih.gov/26644809
204. MacDiarmid, S., et al. Mirabegron as Add-On Treatment to Solifenacin in Patients with Incontinent
Overactive Bladder and an Inadequate Response to Solifenacin Monotherapy. J Urol, 2016. 196: 809.
https://pubmed.ncbi.nlm.nih.gov/27063854
205. Tannenbaum, C., et al. A systematic review of amnestic and non-amnestic mild cognitive
impairment induced by anticholinergic, antihistamine, GABAergic and opioid drugs. Drugs Aging,
2012. 29: 639.
https://pubmed.ncbi.nlm.nih.gov/22812538
206. Gray, S.L., et al. Cumulative use of strong anticholinergics and incident dementia: a prospective
cohort study. JAMA Intern Med, 2015. 175: 401.
https://pubmed.ncbi.nlm.nih.gov/25621434
207. Risacher, S.L., et al. Association between anticholinergic medication use and cognition, brain
metabolism, and brain atrophy in cognitively normal older adults. JAMA Neurology, 2016. 73: 721.
https://jamanetwork.com/journals/jamaneurology/article-abstract/2514553
208. Kessler, T.M., et al. Adverse event assessment of antimuscarinics for treating overactive bladder: a
network meta-analytic approach. PLoS One, 2011. 6: e16718.
https://pubmed.ncbi.nlm.nih.gov/21373193

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 113
209. Paquette, A., et al. Systematic review and meta-analysis: do clinical trials testing antimuscarinic
agents for overactive bladder adequately measure central nervous system adverse events? J Am
Geriatr Soc, 2011. 59: 1332.
https://pubmed.ncbi.nlm.nih.gov/21718264
210. Kay, G., et al. Differential effects of the antimuscarinic agents darifenacin and oxybutynin ER on
memory in older subjects. Eur Urol, 2006. 50: 317.
https://pubmed.ncbi.nlm.nih.gov/16687205
211. Isik, A.T., et al. Trospium and cognition in patients with late onset Alzheimer disease. J Nutr Health
Aging, 2009. 13: 672.
https://pubmed.ncbi.nlm.nih.gov/19657549
212. Lackner, T.E., et al. Randomized, placebo-controlled trial of the cognitive effect, safety, and
tolerability of oral extended-release oxybutynin in cognitively impaired nursing home residents with
urge urinary incontinence. J Am Geriatr Soc, 2008. 56: 862.
https://pubmed.ncbi.nlm.nih.gov/18410326
213. Wagg, A., et al. Randomised, multicentre, placebo-controlled, double-blind crossover study
investigating the effect of solifenacin and oxybutynin in elderly people with mild cognitive
impairment: the SENIOR study. Eur Urol, 2013. 64: 74.
https://pubmed.ncbi.nlm.nih.gov/23332882
214. Sink, K.M., et al. Dual use of bladder anticholinergics and cholinesterase inhibitors: long-term
functional and cognitive outcomes. J Am Geriatr Soc, 2008. 56: 847.
https://pubmed.ncbi.nlm.nih.gov/18384584
215. Wagg, A., et al. Efficacy and tolerability of solifenacin in elderly subjects with overactive bladder
syndrome: a pooled analysis. Am J Geriatr Pharmacother, 2006. 4: 14.
https://pubmed.ncbi.nlm.nih.gov/16730617
216. Wesnes, K.A., et al. Exploratory pilot study assessing the risk of cognitive impairment or sedation in
the elderly following single doses of solifenacin 10 mg. Expert Opin Drug Saf, 2009. 8: 615.
https://pubmed.ncbi.nlm.nih.gov/19747069
217. Zinner, N., et al. Impact of solifenacin on quality of life, medical care use, work productivity, and
health utility in the elderly: an exploratory subgroup analysis. Am J Geriatr Pharmacother, 2009.
7: 373.
https://pubmed.ncbi.nlm.nih.gov/20129258
218. Herschorn, S., et al. Tolerability of solifenacin and oxybutynin immediate release in older (> 65
years) and younger (</= 65 years) patients with overactive bladder: sub-analysis from a Canadian,
randomized, double-blind study. Curr Med Res Opin, 2011. 27: 375.
https://pubmed.ncbi.nlm.nih.gov/21175373
219. Drutz, H.P., et al. Clinical efficacy and safety of tolterodine compared to oxybutynin and placebo in
patients with overactive bladder. Int Urogynecol J Pelvic Floor Dysfunct, 1999. 10: 283.
https://pubmed.ncbi.nlm.nih.gov/10543335
220. Michel, M.C., et al. Does gender or age affect the efficacy and safety of tolterodine? J Urol, 2002.
168: 1027.
https://pubmed.ncbi.nlm.nih.gov/12187215
221. Millard, R., et al. Clinical efficacy and safety of tolterodine compared to placebo in detrusor
overactivity. J Urol, 1999. 161: 1551.
https://pubmed.ncbi.nlm.nih.gov/10210394
222. Zinner, N.R., et al. Efficacy, safety, and tolerability of extended-release once-daily tolterodine
treatment for overactive bladder in older versus younger patients. J Am Geriatr Soc, 2002. 50: 799.
https://pubmed.ncbi.nlm.nih.gov/12028164
223. Jumadilova, Z., et al. Retrospective evaluation of outcomes in patients with overactive bladder
receiving tolterodine versus oxybutynin. Am J Health Syst Pharm, 2006. 63: 2357.
https://pubmed.ncbi.nlm.nih.gov/17106009
224. Chapple, C., et al. Darifenacin treatment of patients >or= 65 years with overactive bladder: results of
a randomized, controlled, 12-week trial. Curr Med Res Opin, 2007. 23: 2347.
https://pubmed.ncbi.nlm.nih.gov/17706004
225. Lipton, R.B., et al. Assessment of cognitive function of the elderly population: effects of darifenacin.
J Urol, 2005. 173: 493.
https://pubmed.ncbi.nlm.nih.gov/15643227
226. Pietzko, A., et al. Influences of trospium chloride and oxybutynin on quantitative EEG in healthy
volunteers. Eur J Clin Pharmacol, 1994. 47: 337.
https://pubmed.ncbi.nlm.nih.gov/7875185

114 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021
227. Todorova, A., et al. Effects of tolterodine, trospium chloride, and oxybutynin on the central nervous
system. J Clin Pharmacol, 2001. 41: 636.
https://pubmed.ncbi.nlm.nih.gov/11402632
228. Staskin, D.R., et al. Trospium chloride once-daily extended release is effective and well tolerated for
the treatment of overactive bladder syndrome: an integrated analysis of two randomised, phase III
trials. Int J Clin Pract, 2009. 63: 1715.
https://pubmed.ncbi.nlm.nih.gov/19930332
229. Sand, P.K., et al. Trospium chloride once-daily extended release is efficacious and tolerated in
elderly subjects (aged >/= 75 years) with overactive bladder syndrome. BJU Int, 2011. 107: 612.
https://pubmed.ncbi.nlm.nih.gov/20707790
230. Kraus, S.R., et al. Efficacy and tolerability of fesoterodine in older and younger subjects with
overactive bladder. Urology, 2010. 76: 1350.
https://pubmed.ncbi.nlm.nih.gov/20974482
231. Dubeau, C.E., et al. Effect of fesoterodine in vulnerable elderly subjects with urgency incontinence: a
double-blind, placebo controlled trial. J Urol, 2014. 191: 395.
https://pubmed.ncbi.nlm.nih.gov/23973522
232. Wagg, A., et al. Review of the efficacy and safety of fesoterodine for treating overactive bladder and
urgency urinary incontinence in elderly patients. Drugs Aging, 2015. 32: 103.
https://pubmed.ncbi.nlm.nih.gov/25673122
233. Ancelin, M.L., et al. Non-degenerative mild cognitive impairment in elderly people and use of
anticholinergic drugs: longitudinal cohort study. BMJ, 2006. 332: 455.
https://pubmed.ncbi.nlm.nih.gov/16452102
234. Wagg, A., et al. Review of cognitive impairment with antimuscarinic agents in elderly patients with
overactive bladder. Int J Clin Pract, 2010. 64: 1279.
https://pubmed.ncbi.nlm.nih.gov/20529135
235. Wagg, A., et al. Long-term safety, tolerability and efficacy of flexible-dose fesoterodine in elderly
patients with overactive bladder: open-label extension of the SOFIA trial. Neurourol Urodyn, 2014.
33: 106.
https://pubmed.ncbi.nlm.nih.gov/23460503
236. Boustani, M., et al. Impact of anticholinergics on the aging brain: a review and practical application.
Aging Health, 2008. 4: 311.
https://www.futuremedicine.com/doi/abs/10.2217/1745509X.4.3.311
237. Cai, X., et al. Long-term anticholinergic use and the aging brain. Alzheimers Dement, 2013. 9: 377.
https://pubmed.ncbi.nlm.nih.gov/23183138
238. Campbell, N., et al. The cognitive impact of anticholinergics: a clinical review. Clin Interv Aging,
2009. 4: 225.
https://pubmed.ncbi.nlm.nih.gov/19554093
239. Carriere, I., et al. Drugs with anticholinergic properties, cognitive decline, and dementia in an elderly
general population: the 3-city study. Arch Intern Med, 2009. 169: 1317.
https://pubmed.ncbi.nlm.nih.gov/19636034
240. Fox, C., et al. Anticholinergic medication use and cognitive impairment in the older population: the
medical research council cognitive function and ageing study. J Am Geriatr Soc, 2011. 59: 1477.
https://pubmed.ncbi.nlm.nih.gov/21707557
241. Rogers, R.G., et al. An International Urogynecological Association (IUGA)/International Continence
Society (ICS) joint report on the terminology for the assessment of sexual health of women with
pelvic floor dysfunction. Neurourol Urodyn, 2018. 37: 1220.
https://pubmed.ncbi.nlm.nih.gov/29441607
242. Biehl, C., et al. A systematic review of the efficacy and safety of vaginal estrogen products for the
treatment of genitourinary syndrome of menopause. Menopause, 2019. 26: 431.
https://pubmed.ncbi.nlm.nih.gov/30363010
243. Cody, J.D., et al. Oestrogen therapy for urinary incontinence in post-menopausal women. Cochrane
Database Syst Rev, 2012. 10: CD001405.
https://pubmed.ncbi.nlm.nih.gov/23076892
244. Lyytinen, H., et al. Breast cancer risk in postmenopausal women using estrogen-only therapy.
Obstet Gynecol, 2006. 108: 1354.
https://pubmed.ncbi.nlm.nih.gov/17138766
245. Yumru, A.E., et al. The use of local 17beta-oestradiol treatment for improving vaginal symptoms
associated with post-menopausal oestrogen deficiency. J Int Med Res, 2009. 37: 198.
https://pubmed.ncbi.nlm.nih.gov/19215691

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 115
246. Mangera, A., et al. Contemporary management of lower urinary tract disease with botulinum toxin
A: a systematic review of botox (onabotulinumtoxinA) and dysport (abobotulinumtoxinA). Eur Urol,
2011. 60: 784.
https://pubmed.ncbi.nlm.nih.gov/21782318
247. Chapple, C., et al. OnabotulinumtoxinA 100 U significantly improves all idiopathic overactive
bladder symptoms and quality of life in patients with overactive bladder and urinary incontinence: a
randomised, double-blind, placebo-controlled trial. Eur Urol, 2013. 64: 249.
https://pubmed.ncbi.nlm.nih.gov/23608668
248. Veeratterapillay, R., et al. Discontinuation rates and inter-injection interval for repeated intravesical
botulinum toxin type A injections for detrusor overactivity. Int J Urol, 2014. 21: 175.
https://pubmed.ncbi.nlm.nih.gov/23819724
249. Mohee, A., et al. Long-term outcome of the use of intravesical botulinum toxin for the treatment of
overactive bladder (OAB). BJU Int, 2013. 111: 106.
https://pubmed.ncbi.nlm.nih.gov/22672569
250. Nitti, V.W., et al. OnabotulinumtoxinA for the treatment of patients with overactive bladder and
urinary incontinence: results of a phase 3, randomized, placebo controlled trial. J Urol, 2013.
189: 2186.
https://pubmed.ncbi.nlm.nih.gov/23246476
251. Duthie, J.B., et al. Botulinum toxin injections for adults with overactive bladder syndrome. Cochrane
Database Syst Rev, 2011: CD005493.
https://pubmed.ncbi.nlm.nih.gov/22161392
252. White, W.M., et al. Short-term efficacy of botulinum toxin a for refractory overactive bladder in the
elderly population. J Urol, 2008. 180: 2522.
https://pubmed.ncbi.nlm.nih.gov/18930481
253. Nitti, V.W., et al. Durable Efficacy and Safety of Long-Term OnabotulinumtoxinA Treatment in
Patients with Overactive Bladder Syndrome: Final Results of a 3.5-Year Study. J Urol, 2016.
196: 791.
https://pubmed.ncbi.nlm.nih.gov/27038769
254. Visco, A.G., et al. Anticholinergic therapy vs. onabotulinumtoxina for urgency urinary incontinence.
N Engl J Med, 2012. 367: 1803.
https://pubmed.ncbi.nlm.nih.gov/23036134
255. Drake, M.J., et al. Comparative assessment of the efficacy of onabotulinumtoxinA and oral therapies
(anticholinergics and mirabegron) for overactive bladder: a systematic review and network meta-
analysis. BJU Int, 2017. 120: 611.
https://pubmed.ncbi.nlm.nih.gov/28670786
256. Herbison, G.P., et al. Sacral neuromodulation with implanted devices for urinary storage and voiding
dysfunction in adults. Cochrane Database Syst Rev, 2009: CD004202.
https://pubmed.ncbi.nlm.nih.gov/19370596
257. Weil, E.H., et al. Sacral root neuromodulation in the treatment of refractory urinary urge
incontinence: a prospective randomized clinical trial. Eur Urol, 2000. 37: 161.
https://pubmed.ncbi.nlm.nih.gov/10705194
258. Schmidt, R.A., et al. Sacral nerve stimulation for treatment of refractory urinary urge incontinence.
Sacral Nerve Stimulation Study Group. J Urol, 1999. 162: 352.
https://pubmed.ncbi.nlm.nih.gov/10411037
259. Brazzelli, M., et al. Efficacy and safety of sacral nerve stimulation for urinary urge incontinence: a
systematic review. J Urol, 2006. 175: 835.
https://pubmed.ncbi.nlm.nih.gov/16469561
260. Groen, J., et al. Sacral neuromodulation as treatment for refractory idiopathic urge urinary
incontinence: 5-year results of a longitudinal study in 60 women. J Urol, 2011. 186: 954.
https://pubmed.ncbi.nlm.nih.gov/21791355
261. van Kerrebroeck, P.E., et al. Results of sacral neuromodulation therapy for urinary voiding
dysfunction: outcomes of a prospective, worldwide clinical study. J Urol, 2007. 178: 2029.
https://pubmed.ncbi.nlm.nih.gov/17869298
262. Amundsen, C.L., et al. OnabotulinumtoxinA vs Sacral Neuromodulation on Refractory Urgency
Urinary Incontinence in Women: A Randomized Clinical Trial. JAMA, 2016. 316: 1366.
https://pubmed.ncbi.nlm.nih.gov/27701661
263. Siegel, S., et al. Results of a prospective, randomized, multicenter study evaluating sacral
neuromodulation with InterStim therapy compared to standard medical therapy at 6-months in
subjects with mild symptoms of overactive 24415559 bladder. Neurourol Urodyn, 2015. 34: 224.
https://pubmed.ncbi.nlm.nih.gov/24415559

116 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021
264. Amundsen, C.L., et al. Two-Year Outcomes of Sacral Neuromodulation Versus OnabotulinumtoxinA
for Refractory Urgency Urinary Incontinence: A Randomized Trial. Eur Urol, 2018. 74: 66.
https://pubmed.ncbi.nlm.nih.gov/29482936
265. Tutolo, M., et al. Efficacy and Safety of Sacral and Percutaneous Tibial Neuromodulation in Non-
neurogenic Lower Urinary Tract Dysfunction and Chronic Pelvic Pain: A Systematic Review of the
Literature. Eur Urol, 2018. 73: 406.
https://pubmed.ncbi.nlm.nih.gov/29336927
266. Groenendijk, P.M., et al. Urodynamic evaluation of sacral neuromodulation for urge urinary
incontinence. BJU Int, 2008. 101: 325.
https://pubmed.ncbi.nlm.nih.gov/18070199
267. Cody, J.D., et al. Urinary diversion and bladder reconstruction/replacement using intestinal
segments for intractable incontinence or following cystectomy. Cochrane Database Syst Rev, 2012:
CD003306.
https://pubmed.ncbi.nlm.nih.gov/22336788
268. Kockelbergh, R.C., et al. Clam enterocystoplasty in general urological practice. Br J Urol, 1991.
68: 38.
https://pubmed.ncbi.nlm.nih.gov/1873689
269. Cohen, A.J., et al. Comparative Outcomes and Perioperative Complications of Robotic Vs Open
Cystoplasty and Complex Reconstructions. Urology, 2016. 97: 172.
https://pubmed.ncbi.nlm.nih.gov/27443464
270. Awad, S.A., et al. Long-term results and complications of augmentation ileocystoplasty for
idiopathic urge incontinence in women. Br J Urol, 1998. 81: 569.
https://pubmed.ncbi.nlm.nih.gov/9598629
271. Lane, T., et al. Carcinoma following augmentation ileocystoplasty. Urol Int, 2000. 64: 31.
https://pubmed.ncbi.nlm.nih.gov/10782030
272. Stone, A.R., et al. Carcinoma associated with augmentation cystoplasty. Br J Urol, 1987. 60: 236.
https://pubmed.ncbi.nlm.nih.gov/3676669
273. Leedham, P.W., et al. Adenocarcinoma developing in an ileocystoplasty. Br J Surg, 1973. 60: 158.
https://pubmed.ncbi.nlm.nih.gov/4685940
274. Biardeau, X., et al. Risk of malignancy after augmentation cystoplasty: A systematic review.
Neurourol Urodyn, 2016. 35: 675.
https://pubmed.ncbi.nlm.nih.gov/25867054
275. Greenwell, T.J., et al. Augmentation cystoplasty. BJU Int, 2001. 88: 511.
https://pubmed.ncbi.nlm.nih.gov/11678743
276. Covert, W.M., et al. The role of mucoregulatory agents after continence-preserving urinary diversion
surgery. Am J Health Syst Pharm, 2012. 69: 483.
https://pubmed.ncbi.nlm.nih.gov/22382478
277. Padmanabhan, P., et al. Five-year cost analysis of intra-detrusor injection of botulinum toxin type A
and augmentation cystoplasty for refractory neurogenic detrusor overactivity. World J Urol, 2011.
29: 51.
https://pubmed.ncbi.nlm.nih.gov/21110030
278. Cartwright, P.C., et al. Bladder autoaugmentation: partial detrusor excision to augment the bladder
without use of bowel. J Urol, 1989. 142: 1050.
https://pubmed.ncbi.nlm.nih.gov/2795729
279. Leng, W.W., et al. Enterocystoplasty or detrusor myectomy? Comparison of indications and
outcomes for bladder augmentation. J Urol, 1999. 161: 758.
https://pubmed.ncbi.nlm.nih.gov/10022679
280. ter Meulen, P.H., et al. A study on the feasibility of vesicomyotomy in patients with motor urge
incontinence. Eur Urol, 1997. 32: 166.
https://pubmed.ncbi.nlm.nih.gov/9286647
281. Wu, J.M., et al. Forecasting the prevalence of pelvic floor disorders in U.S. Women: 2010 to 2050.
Obstet Gynecol, 2009. 114: 1278.
https://pubmed.ncbi.nlm.nih.gov/19935030
282. Hunskaar, S., et al. The prevalence of urinary incontinence in women in four European countries.
BJU Int, 2004. 93: 324.
https://pubmed.ncbi.nlm.nih.gov/14764130
283. Thom, D.H., et al. Differences in prevalence of urinary incontinence by race/ethnicity. J Urol, 2006.
175: 259.
https://pubmed.ncbi.nlm.nih.gov/16406923

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 117
284. Mitchell, E.S., et al. Correlates of urinary incontinence during the menopausal transition and early
postmenopause: observations from the Seattle Midlife Women’s Health Study. Climacteric, 2013.
16: 653.
https://pubmed.ncbi.nlm.nih.gov/23560943
285. Minassian, V.A., et al. Urinary incontinence in women: variation in prevalence estimates and risk
factors. Obstet Gynecol, 2008. 111: 324.
https://pubmed.ncbi.nlm.nih.gov/18238969
286. Brown, J.S., et al. Urinary incontinence in older women: who is at risk? Study of Osteoporotic
Fractures Research Group. Obstet Gynecol, 1996. 87: 715.
https://pubmed.ncbi.nlm.nih.gov/8677073
287. Bø, K., et al. Is Physical Activity Good or Bad for the Female Pelvic Floor? A Narrative Review.
Sports Med, 2020. 50: 471.
https://pubmed.ncbi.nlm.nih.gov/31820378
288. DeLancey, J.O. Structural support of the urethra as it relates to stress urinary incontinence: the
hammock hypothesis. Am J Obstet Gynecol, 1994. 170: 1713.
https://pubmed.ncbi.nlm.nih.gov/8203431
289. Aoki, Y., et al. Urinary incontinence in women. Nat Rev Dis Primers, 2017. 3: 17042.
https://pubmed.ncbi.nlm.nih.gov/28681849
290. Hillary, C.J., et al. Considerations in the modern management of stress urinary incontinence
resulting from intrinsic sphincter deficiency. World J Urol, 2015. 33: 1251.
https://pubmed.ncbi.nlm.nih.gov/26060138
291. Medina, C.A., et al. Evaluation and surgery for stress urinary incontinence: A FIGO working group
report. Neurourol Urodyn, 2017. 36: 518.
https://pubmed.ncbi.nlm.nih.gov/26950893
292. Patnam, R., et al. Standing Vs Supine; Does it Matter in Cough Stress Testing? Female Pelvic Med
Reconstr Surg, 2017. 23: 315.
https://pubmed.ncbi.nlm.nih.gov/28079569
293. Guralnick, M.L., et al. ICS Educational Module: Cough stress test in the evaluation of female urinary
incontinence: Introducing the ICS-Uniform Cough Stress Test. Neurourol Urodyn, 2018. 37: 1849.
https://pubmed.ncbi.nlm.nih.gov/29926966
294. Rosier, P., et al., Committee 6: Urodynamic Testing, in 5th International Consultation on
Incontinence, Paris February, 2012, In: Incontinence, 5th Edn. 2013, Abrams, P., Cardozo, L.,
Khoury, S., Wein, A. (Eds). Bristol, UK.
https://www.ics.org/Publications/ICI_5/INCONTINENCE.pdf
295. Sirls, L.T., et al. The effect of urodynamic testing on clinical diagnosis, treatment plan and outcomes
in women undergoing stress urinary incontinence surgery. J Urol, 2013. 189: 204.
https://pubmed.ncbi.nlm.nih.gov/22982425
296. van Leijsen, S.A., et al. Can preoperative urodynamic investigation be omitted in women with stress
urinary incontinence? A non-inferiority randomized controlled trial. Neurourol Urodyn, 2012. 31: 1118.
https://pubmed.ncbi.nlm.nih.gov/30576004
297. van Leijsen, S.A., et al. Value of urodynamics before stress urinary incontinence surgery: a
randomized controlled trial. Obstet Gynecol, 2013. 121: 999.
https://pubmed.ncbi.nlm.nih.gov/23635736
298. Costantini, E., et al. Preoperative Valsalva leak point pressure may not predict outcome of mid-
urethral slings. Analysis from a randomized controlled trial of retropubic versus transobturator mid-
urethral slings. Int Braz J Urol, 2008. 34: 73.
https://pubmed.ncbi.nlm.nih.gov/18341724
299. Finazzi-Agrò, E., et al. Comments on “A randomized trial of urodynamic testing before stress-
incontinence surgery” (N Engl J Med. 2012 May 24;366(21):1987-1997) From the Italian Society of
Urodynamics. Neurourol Urodyn, 2013. 32: 301.
https://pubmed.ncbi.nlm.nih.gov/23023985
300. Lewicky-Gaupp, C., et al. “The cough game”: are there characteristic urethrovesical movement
patterns associated with stress incontinence? Int Urogynecol J Pelvic Floor Dysfunct, 2009. 20: 171.
https://pubmed.ncbi.nlm.nih.gov/18850057
301. Shek, K.L., et al. The effect of childbirth on urethral mobility: a prospective observational study.
J Urol, 2010. 184: 629.
https://pubmed.ncbi.nlm.nih.gov/20639028
302. Kociszewski, J., et al. Tape functionality: sonographic tape characteristics and outcome after TVT
incontinence surgery. Neurourol Urodyn, 2008. 27: 485.
https://pubmed.ncbi.nlm.nih.gov/18288705

118 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021
303. Morgan, D.M., et al. Urethral sphincter morphology and function with and without stress
incontinence. J Urol, 2009. 182: 203.
https://pubmed.ncbi.nlm.nih.gov/19450822
304. Nguyen, L., et al. Surgical technique to overcome anatomical shortcoming: balancing post-
prostatectomy continence outcomes of urethral sphincter lengths on preoperative magnetic
resonance imaging. J Urol, 2008. 179: 1907.
https://pubmed.ncbi.nlm.nih.gov/18353395
305. Nygaard, I., et al. Prevalence of symptomatic pelvic floor disorders in US women. JAMA, 2008. 300:
1311.
https://pubmed.ncbi.nlm.nih.gov/18799443
306. Gozukara, Y.M., et al. The improvement in pelvic floor symptoms with weight loss in obese women
does not correlate with the changes in pelvic anatomy. Int Urogynecol J, 2014. 25: 1219.
https://pubmed.ncbi.nlm.nih.gov/24711149
307. Brown, J.S., et al. Lifestyle intervention is associated with lower prevalence of urinary incontinence:
the Diabetes Prevention Program. Diabetes Care, 2006. 29: 385.
https://pubmed.ncbi.nlm.nih.gov/16443892
308. Bump, R.C., et al. Obesity and lower urinary tract function in women: effect of surgically induced
weight loss. Am J Obstet Gynecol, 1992. 167: 392.
https://pubmed.ncbi.nlm.nih.gov/1497041
309. Subak, L.L., et al. Does weight loss improve incontinence in moderately obese women? Int
Urogynecol J Pelvic Floor Dysfunct, 2002. 13: 40.
https://pubmed.ncbi.nlm.nih.gov/11999205
310. Wing, R.R., et al. Improving urinary incontinence in overweight and obese women through modest
weight loss. Obstet Gynecol, 2010. 116: 284.
https://pubmed.ncbi.nlm.nih.gov/20664387
311. Subak, L.L., et al. Weight loss: a novel and effective treatment for urinary incontinence. J Urol, 2005.
174: 190.
https://pubmed.ncbi.nlm.nih.gov/15947625
312. Phelan, S., et al. Weight loss prevents urinary incontinence in women with type 2 diabetes: results
from the Look AHEAD trial. J Urol, 2012. 187: 939.
https://pubmed.ncbi.nlm.nih.gov/22264468
313. Burgio, K.L., et al. Changes in urinary and fecal incontinence symptoms with weight loss surgery in
morbidly obese women. Obstet Gynecol, 2007. 110: 1034.
https://pubmed.ncbi.nlm.nih.gov/17978117
314. Deitel, M., et al. Gynecologic-obstetric changes after loss of massive excess weight following
bariatric surgery. J Am Coll Nutr, 1988. 7: 147.
https://pubmed.ncbi.nlm.nih.gov/3361039
315. Laungani, R.G., et al. Effect of laparoscopic gastric bypass surgery on urinary incontinence in
morbidly obese women. Surg Obes Relat Dis, 2009. 5: 334.
https://pubmed.ncbi.nlm.nih.gov/19342304
316. Mishra, G.D., et al. Body weight through adult life and risk of urinary incontinence in middle-aged
women: results from a British prospective cohort. Int J Obes (Lond), 2008. 32: 1415.
https://pubmed.ncbi.nlm.nih.gov/18626483
317. Richter, H.E., et al. The impact of obesity on urinary incontinence symptoms, severity, urodynamic
characteristics and quality of life. J Urol, 2010. 183: 622.
https://pubmed.ncbi.nlm.nih.gov/20018326
318. Leshem, A., et al. Effects of Bariatric Surgery on Female Pelvic Floor Disorders. Urology, 2017. 105: 42.
https://pubmed.ncbi.nlm.nih.gov/28315786
319. Knepfler, T., et al. Bariatric surgery improves female pelvic floor disorders. J Visc Surg, 2016.
153: 95.
https://pubmed.ncbi.nlm.nih.gov/26678846
320. Subak, L.L., et al. Urinary Incontinence Before and After Bariatric Surgery. JAMA Intern Med, 2015.
175: 1378.
https://pubmed.ncbi.nlm.nih.gov/26098620
321. Miller, J.M., et al. A pelvic muscle precontraction can reduce cough-related urine loss in selected
women with mild SUI. J Am Geriatr Soc, 1998. 46: 870.
https://pubmed.ncbi.nlm.nih.gov/9670874
322. Bø, K. Pelvic floor muscle training is effective in treatment of female stress urinary incontinence, but
how does it work? Int Urogynecol J Pelvic Floor Dysfunct, 2004. 15: 76.
https://pubmed.ncbi.nlm.nih.gov/15014933

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 119
323. Zubieta, M., et al. Influence of voluntary pelvic floor muscle contraction and pelvic floor muscle
training on urethral closure pressures: a systematic literature review. Int Urogynecol J, 2016.
27: 687.
https://pubmed.ncbi.nlm.nih.gov/26407561
324. Braekken, I.H., et al. Morphological changes after pelvic floor muscle training measured by
3-dimensional ultrasonography: a randomized controlled trial. Obstet Gynecol, 2010. 115: 317.
https://pubmed.ncbi.nlm.nih.gov/20093905
325. Dumoulin, C., et al. Pelvic floor muscle training versus no treatment, or inactive control treatments,
for urinary incontinence in women. Cochrane Database Syst Rev, 2018. 10: CD005654.
https://pubmed.ncbi.nlm.nih.gov/30288727
326. Herderschee, R., et al. Feedback or biofeedback to augment pelvic floor muscle training for urinary
incontinence in women. Cochrane Database Syst Rev, 2011: CD009252.
https://pubmed.ncbi.nlm.nih.gov/21735442
327. Hagen, S., et al., Effectiveness and cost-effectiveness of biofeedback-assisted pelvic floor muscle
training for female urinary incontinence: a multicentre randomised controlled trial (abstract 485),
ICS2019. ICS: Gothenburg, Sweden.
https://www.ics.org/2019/abstract/489
328. Dumoulin, C., et al. Group-Based vs Individual Pelvic Floor Muscle Training to Treat Urinary
Incontinence in Older Women: A Randomized Clinical Trial. JAMA Intern Med, 2020. 180: 1284.
https://pubmed.ncbi.nlm.nih.gov/32744599
329. Hay-Smith, E.J., et al. Comparisons of approaches to pelvic floor muscle training for urinary
incontinence in women. Cochrane Database Syst Rev, 2011: CD009508.
https://pubmed.ncbi.nlm.nih.gov/22161451
330. Herbison, G.P., et al. Weighted vaginal cones for urinary incontinence. Cochrane Database Syst Rev,
2013. 2013: CD002114.
https://pubmed.ncbi.nlm.nih.gov/23836411
331. Dumoulin, C., et. al. Committee 12, Conservative management, In: Incontinence, 6th Edn. 2017,
Abrams, P., Cardozo, L., Wagg, A., Wein, A. (Eds). Bristol, UK.
https://www.ics.org/publications/ici_6/Incontinence_6th_Edition_2017_eBook_v2.pdf
332. Labrie, J., et al. Surgery versus physiotherapy for stress urinary incontinence. N Engl J Med, 2013.
369: 1124.
https://pubmed.ncbi.nlm.nih.gov/24047061
333. Stewart, F., et al. Electrical stimulation with non-implanted devices for stress urinary incontinence in
women. Cochrane Database Syst Rev, 2017. 12: CD012390.
https://pubmed.ncbi.nlm.nih.gov/29271482
334. Mateus-Vasconcelos, E.C.L., et al. Effects of three interventions in facilitating voluntary pelvic floor
muscle contraction in women: a randomized controlled trial. Braz J Phys Ther, 2018. 22: 391.
https://pubmed.ncbi.nlm.nih.gov/29429823
335. Bø, K., et al. Does it work in the long term?--A systematic review on pelvic floor muscle training for
female stress urinary incontinence. Neurourol Urodyn, 2013. 32: 215.
https://pubmed.ncbi.nlm.nih.gov/22847318
336. Woodley, S.J., et al. Pelvic floor muscle training for preventing and treating urinary and faecal
incontinence in antenatal and postnatal women. Cochrane Database Syst Rev, 2020. 5: CD007471.
https://pubmed.ncbi.nlm.nih.gov/32378735
337. Sigurdardottir, T., et al. Can postpartum pelvic floor muscle training reduce urinary and anal
incontinence?: An assessor-blinded randomized controlled trial. Am J Obstet Gynecol, 2020.
222: 247 e1.
https://pubmed.ncbi.nlm.nih.gov/31526791
338. Wagg, A., et al. Committee 11, Incontinence in frail older persons, In: Incontinence, 6th Edn. 2017,
Abrams, P., Cardozo, L., Wagg, A., Wein, A. (Eds). Bristol, UK.
339. Stenzelius, K., et al. The effect of conservative treatment of urinary incontinence among older and
frail older people: a systematic review. Age Ageing, 2015. 44: 736.
https://pubmed.ncbi.nlm.nih.gov/26112402
340. Gilling, P.J., et al. A double-blind randomized controlled trial of electromagnetic stimulation of the
pelvic floor vs sham therapy in the treatment of women with stress urinary incontinence. BJU Int,
2009. 103: 1386.
https://pubmed.ncbi.nlm.nih.gov/19154474
341. Robinson, D., et al. Estrogens and the lower urinary tract. Neurourol Urodyn, 2011. 30: 754.
https://pubmed.ncbi.nlm.nih.gov/21661025

120 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021
342. Mettler, L., et al. Long-term treatment of atrophic vaginitis with low-dose oestradiol vaginal tablets.
Maturitas, 1991. 14: 23.
https://pubmed.ncbi.nlm.nih.gov/1791769
343. Weber, M.A., et al. Local Oestrogen for Pelvic Floor Disorders: A Systematic Review. PLoS One,
2015. 10: e0136265.
https://pubmed.ncbi.nlm.nih.gov/26383760
344. Castellani, D., et al. Low-Dose Intravaginal Estriol and Pelvic Floor Rehabilitation in Post-
Menopausal Stress Urinary Incontinence. Urol Int, 2015. 95: 417.
https://pubmed.ncbi.nlm.nih.gov/26043913
345. Jackson, S., et al. The effect of oestrogen supplementation on post-menopausal urinary stress
incontinence: a double-blind placebo-controlled trial. Br J Obstet Gynaecol, 1999. 106: 711.
https://pubmed.ncbi.nlm.nih.gov/10428529
346. Fantl, J.A., et al. Efficacy of estrogen supplementation in the treatment of urinary incontinence. The
Continence Program for Women Research Group. Obstet Gynecol, 1996. 88: 745.
https://pubmed.ncbi.nlm.nih.gov/8885906
347. Grady, D., et al. Postmenopausal hormones and incontinence: the Heart and Estrogen/Progestin
Replacement Study. Obstet Gynecol, 2001. 97: 116.
https://pubmed.ncbi.nlm.nih.gov/11152919
348. Hendrix, S.L., et al. Effects of estrogen with and without progestin on urinary incontinence. JAMA,
2005. 293: 935.
https://pubmed.ncbi.nlm.nih.gov/15728164
349. Rossouw, J.E., et al. Risks and benefits of estrogen plus progestin in healthy postmenopausal
women: principal results From the Women’s Health Initiative randomized controlled trial. JAMA,
2002. 288: 321.
https://pubmed.ncbi.nlm.nih.gov/12117397
350. Steinauer, J.E., et al. Postmenopausal hormone therapy: does it cause incontinence? Obstet
Gynecol, 2005. 106: 940.
https://pubmed.ncbi.nlm.nih.gov/16260510
351. Goldstein, S.R., et al. Incidence of urinary incontinence in postmenopausal women treated with
raloxifene or estrogen. Menopause, 2005. 12: 160.
https://pubmed.ncbi.nlm.nih.gov/15772563
352. Molander, U., et al. Effect of oral oestriol on vaginal flora and cytology and urogenital symptoms in
the post-menopause. Maturitas, 1990. 12: 113.
https://pubmed.ncbi.nlm.nih.gov/2255263
353. Wang, C.J., et al. Low dose oral desmopressin for nocturnal polyuria in patients with benign
prostatic hyperplasia: a double-blind, placebo controlled, randomized study. J Urol, 2011. 185: 219.
https://pubmed.ncbi.nlm.nih.gov/21074790
354. Mariappan, P., et al. Duloxetine, a serotonin and noradrenaline reuptake inhibitor (SNRI) for the
treatment of stress urinary incontinence: a systematic review. Eur Urol, 2007. 51: 67.
https://pubmed.ncbi.nlm.nih.gov/17014950
355. Li, J., et al. The role of duloxetine in stress urinary incontinence: a systematic review and meta-
analysis. Int Urol Nephrol, 2013. 45: 679.
https://pubmed.ncbi.nlm.nih.gov/23504618
356. Ghoniem, G.M., et al. A randomized controlled trial of duloxetine alone, pelvic floor muscle training
alone, combined treatment and no active treatment in women with stress urinary incontinence.
J Urol, 2005. 173: 1647.
https://pubmed.ncbi.nlm.nih.gov/15821528
357. Bump, R.C., et al. Long-term efficacy of duloxetine in women with stress urinary incontinence. BJU
Int, 2008. 102: 214.
https://pubmed.ncbi.nlm.nih.gov/18422764
358. Vella, M., et al. Duloxetine 1 year on: the long-term outcome of a cohort of women prescribed
duloxetine. Int Urogynecol J Pelvic Floor Dysfunct, 2008. 19: 961.
https://pubmed.ncbi.nlm.nih.gov/18231697
359. Maund, E., et al. Considering benefits and harms of duloxetine for treatment of stress urinary
incontinence: a meta-analysis of clinical study reports. CMAJ, 2017. 189: E194.
https://pubmed.ncbi.nlm.nih.gov/28246265
360. Acharya, N., et al. Duloxetine: meta-analyses of suicidal behaviors and ideation in clinical trials for
major depressive disorder. J Clin Psychopharmacol, 2006. 26: 587.
https://pubmed.ncbi.nlm.nih.gov/17110815

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 121
361. Cumberledge, J. First Do No Harm. The report of the Independent Medicines and Medical Devices
Safety Review. 2020.
https://psnet.ahrq.gov/issue/first-do-no-harm-report-independent-medicines-and-medical-devices-
safety-review
362. Chapple, C.R., et al. Consensus Statement of the European Urology Association and the European
Urogynaecological Association on the Use of Implanted Materials for Treating Pelvic Organ Prolapse
and Stress Urinary Incontinence. Eur Urol, 2017. 72: 424.
https://pubmed.ncbi.nlm.nih.gov/28413126
363. Brazzelli, M., et al. Surgical treatments for women with stress urinary incontinence: the ESTER
systematic review and economic evaluation. Health Technol Assess, 2019. 23: 1.
https://pubmed.ncbi.nlm.nih.gov/30929658
364. Javanbakht, M., et al. Surgical treatments for women with stress urinary incontinence: a systematic
review of economic evidence. Syst Rev, 2020. 9: 85.
https://pubmed.ncbi.nlm.nih.gov/32312310
365. Franzen, K., et al. Surgery for urinary incontinence in women 65 years and older: a systematic
review. Int Urogynecol J, 2015. 26: 1095.
https://pubmed.ncbi.nlm.nih.gov/25477140
366. Freites, J., et al. Laparoscopic colposuspension for urinary incontinence in women. Cochrane
Database Syst Rev, 2019. 12: CD002239.
https://pubmed.ncbi.nlm.nih.gov/31821550
367. Glazener, C.M., et al. Anterior vaginal repair for urinary incontinence in women. Cochrane Database
Syst Rev, 2017. 7: CD001755.
https://pubmed.ncbi.nlm.nih.gov/28759116
368. Lapitan, M.C., et al. Open retropubic colposuspension for urinary incontinence in women. Cochrane
Database Syst Rev, 2016. 2: CD002912.
https://pubmed.ncbi.nlm.nih.gov/26878400
369. Gumus, II, et al. Laparoscopic single-port Burch colposuspension with an extraperitoneal approach
and standard instruments for stress urinary incontinence: early results from a series of 15 patients.
Minim Invasive Ther Allied Technol, 2013. 22: 116.
https://pubmed.ncbi.nlm.nih.gov/22909022
370. Guerrero, K.L., et al. A randomised controlled trial comparing TVT, Pelvicol and autologous fascial
slings for the treatment of stress urinary incontinence in women. BJOG, 2010. 117: 1493.
https://pubmed.ncbi.nlm.nih.gov/20939862
371. Brubaker, L., et al. 5-year continence rates, satisfaction and adverse events of burch urethropexy
and fascial sling surgery for urinary incontinence. J Urol, 2012. 187: 1324.
https://pubmed.ncbi.nlm.nih.gov/22341290
372. Kirchin, V., et al. Urethral injection therapy for urinary incontinence in women. Cochrane Database
Syst Rev, 2012. 5: CD003881.
https://pubmed.ncbi.nlm.nih.gov/22336797
373. Kirchin, V., et al. Urethral injection therapy for urinary incontinence in women. Cochrane Database
Syst Rev, 2017. 7: CD003881.
https://pubmed.ncbi.nlm.nih.gov/28738443
374. Davis, N.F., et al. Injectable biomaterials for the treatment of stress urinary incontinence: their
potential and pitfalls as urethral bulking agents. Int Urogynecol J, 2013. 24: 913.
https://pubmed.ncbi.nlm.nih.gov/23224022
375. Ghoniem, G.M., et al. A systematic review and meta-analysis of Macroplastique for treating female
stress urinary incontinence. Int Urogynecol J, 2013. 24: 27.
https://pubmed.ncbi.nlm.nih.gov/22699885
376. Kasi, A.D., et al. Polyacrylamide hydrogel (Bulkamid(R)) for stress urinary incontinence in women: a
systematic review of the literature. Int Urogynecol J, 2016. 27: 367.
https://pubmed.ncbi.nlm.nih.gov/26209952
377. Siddiqui, Z.A., et al. Intraurethral bulking agents for the management of female stress urinary
incontinence: a systematic review. Int Urogynecol J, 2017. 28: 1275.
https://pubmed.ncbi.nlm.nih.gov/28220200
378. Matsuoka, P.K., et al. The efficacy and safety of urethral injection therapy for urinary incontinence in
women: a systematic review. Clinics (Sao Paulo), 2016. 71: 94.
https://pubmed.ncbi.nlm.nih.gov/26934239
379. Zhao, Y., et al. Bulking agents - an analysis of 500 cases and review of the literature. Clin Exp
Obstet Gynecol, 2016. 43: 666.
https://pubmed.ncbi.nlm.nih.gov/30074316

122 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021
380. Lee, P.E., et al. Periurethral autologous fat injection as treatment for female stress urinary
incontinence: a randomized double-blind controlled trial. J Urol, 2001. 165: 153.
https://pubmed.ncbi.nlm.nih.gov/11125386
381. Schulz, J.A., et al. Bulking agents for stress urinary incontinence: short-term results and
complications in a randomized comparison of periurethral and transurethral injections. Int
Urogynecol J Pelvic Floor Dysfunct, 2004. 15: 261.
https://pubmed.ncbi.nlm.nih.gov/15517671
382. Kuhn, A., et al. Where should bulking agents for female urodynamic stress incontinence be injected?
Int Urogynecol J Pelvic Floor Dysfunct, 2008. 19: 817.
https://pubmed.ncbi.nlm.nih.gov/18157642
383. Krhut, J., et al. Treatment of stress urinary incontinence using polyacrylamide hydrogel in women
after radiotherapy: 1-year follow-up. Int Urogynecol J, 2016. 27: 301.
https://pubmed.ncbi.nlm.nih.gov/26342812
384. Carr, L.K., et al. Autologous muscle derived cell therapy for stress urinary incontinence: a
prospective, dose ranging study. J Urol, 2013. 189: 595.
https://pubmed.ncbi.nlm.nih.gov/23260547
385. Maher, C.F., et al. Pubovaginal sling versus transurethral Macroplastique for stress urinary
incontinence and intrinsic sphincter deficiency: a prospective randomised controlled trial. BJOG,
2005. 112: 797.
https://pubmed.ncbi.nlm.nih.gov/15924540
386. Itkonen Freitas, A.M., et al. Tension-Free Vaginal Tape Surgery versus Polyacrylamide Hydrogel
Injection for Primary Stress Urinary Incontinence: A Randomized Clinical Trial. J Urol, 2020.
203: 372.
https://pubmed.ncbi.nlm.nih.gov/31479396
387. Abrams, P., et al. Synthetic vaginal tapes for stress incontinence: proposals for improved regulation
of new devices in Europe. Eur Urol, 2011. 60: 1207.
https://pubmed.ncbi.nlm.nih.gov/21855204
388. Ford, A.A., et al. Mid-urethral sling operations for stress urinary incontinence in women. Cochrane
Database Syst Rev, 2017. 7: CD006375.
https://pubmed.ncbi.nlm.nih.gov/28756647
389. Albo, M.E., et al. Treatment success of retropubic and transobturator mid urethral slings at 24
months. J Urol, 2012. 188: 2281.
https://pubmed.ncbi.nlm.nih.gov/23083653
390. Kenton, K., et al. 5-year longitudinal followup after retropubic and transobturator mid urethral slings.
J Urol, 2015. 193: 203.
https://pubmed.ncbi.nlm.nih.gov/25158274
391. Serati, M., et al. Tension-free Vaginal Tape-Obturator for Treatment of Pure Urodynamic Stress
Urinary Incontinence: Efficacy and Adverse Effects at 10-year Follow-up. Eur Urol, 2017. 71: 674.
https://pubmed.ncbi.nlm.nih.gov/27597239
392. Svenningsen, R., et al. Long-term follow-up of the retropubic tension-free vaginal tape procedure.
Int Urogynecol J, 2013. 24: 1271.
https://pubmed.ncbi.nlm.nih.gov/23417313
393. Natale, F., et al. Transobturator Tape: Over 10 Years Follow-up. Urology, 2019. 129: 48.
https://pubmed.ncbi.nlm.nih.gov/30890420
394. Nilsson, C.G., et al. Seventeen years’ follow-up of the tension-free vaginal tape procedure for female
stress urinary incontinence. Int Urogynecol J, 2013. 24: 1265.
https://pubmed.ncbi.nlm.nih.gov/23563892
395. Lier, D., et al. Surgical treatment of stress urinary incontinence-trans-obturator tape compared with
tension-free vaginal tape-5-year follow up: an economic evaluation. BJOG, 2017. 124: 1431.
https://pubmed.ncbi.nlm.nih.gov/27506185
396. Ford, A.A., et al. Mid-urethral sling operations for stress urinary incontinence in women. Cochrane
Database Syst Rev, 2015: CD006375.
https://pubmed.ncbi.nlm.nih.gov/26130017
397. Fusco, F., et al. Updated Systematic Review and Meta-analysis of the Comparative Data on
Colposuspensions, Pubovaginal Slings, and Midurethral Tapes in the Surgical Treatment of Female
Stress Urinary Incontinence. Eur Urol, 2017. 72: 567.
https://pubmed.ncbi.nlm.nih.gov/28479203
398. Cheung, R.Y., et al. Inside-out versus outside-in transobturator tension-free vaginal tape: a 5-year
prospective comparative study. Int J Urol, 2014. 21: 74.
https://pubmed.ncbi.nlm.nih.gov/23675961

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 123
399. Abdel-Fattah, M., et al. Long-term outcomes for transobturator tension-free vaginal tapes in women
with urodynamic mixed urinary incontinence. Neurourol Urodyn, 2017. 36: 902.
https://pubmed.ncbi.nlm.nih.gov/28028822
400. Morling, J.R., et al. Adverse events after first, single, mesh and non-mesh surgical procedures for
stress urinary incontinence and pelvic organ prolapse in Scotland, 1997-2016: a population-based
cohort study. Lancet, 2017. 389: 629.
https://pubmed.ncbi.nlm.nih.gov/28010993
401. Keltie, K., et al. Complications following vaginal mesh procedures for stress urinary incontinence: an
8 year study of 92,246 women. Sci Rep, 2017. 7: 12015.
https://pubmed.ncbi.nlm.nih.gov/28931856
402. Alwaal, A., et al. Female sexual function following mid-urethral slings for the treatment of stress
urinary incontinence. Int J Impot Res, 2016. 28: 121.
https://pubmed.ncbi.nlm.nih.gov/27146350
403. Fan, Y., et al. Incontinence-specific quality of life measures used in trials of sling procedures for
female stress urinary incontinence: a meta-analysis. Int Urol Nephrol, 2015. 47: 1277.
https://pubmed.ncbi.nlm.nih.gov/26093584
404. Wyndaele, J.J., et al. A randomized, controlled clinical trial of an intravesical pressure-attenuation
balloon system for the treatment of stress urinary incontinence in females. Neurourol Urodyn, 2016.
35: 252.
https://pubmed.ncbi.nlm.nih.gov/25598453
405. McCammon, K., et al. Three-month primary efficacy data for the SUCCESS Trial; a phase III, multi-
center, prospective, randomized, controlled study treating female stress urinary incontinence with
the vesair intravesical balloon. Neurourol Urodyn, 2018. 37: 440.
https://pubmed.ncbi.nlm.nih.gov/29095516
406. Winkler, H., et al. Twelve-Month Efficacy and Safety Data for the “Stress Incontinence Control,
Efficacy and Safety Study”: A Phase III, Multicenter, Prospective, Randomized, Controlled Study
Treating Female Stress Urinary Incontinence Using the Vesair Intravesical Balloon. Female Pelvic
Med Reconstr Surg, 2018. 24: 222.
https://pubmed.ncbi.nlm.nih.gov/28953076
407. Lipp, A., et al. Mechanical devices for urinary incontinence in women. Cochrane Database Syst Rev,
2014. 2014: CD001756.
https://pubmed.ncbi.nlm.nih.gov/25517397
408. Shaikh, S., et al. Mechanical devices for urinary incontinence in women. Cochrane Database Syst
Rev, 2006: CD001756.
https://pubmed.ncbi.nlm.nih.gov/16855977
409. Phé, V., et al. A systematic review of the treatment for female stress urinary incontinence by ACT®
balloon placement (Uromedica, Irvine, CA, USA). World J Urol, 2014. 32: 495.
https://pubmed.ncbi.nlm.nih.gov/23783882
410. Peyronnet, B., et al. AMS-800 Artificial urinary sphincter in female patients with stress urinary
incontinence: A systematic review. Neurourol Urodyn, 2019. 38 Suppl 4: S28.
https://pubmed.ncbi.nlm.nih.gov/30298943
411. Bakali, E., et al. Treatment of recurrent stress urinary incontinence after failed minimally invasive
synthetic suburethral tape surgery in women. Cochrane Database Syst Rev, 2013: CD009407.
https://pubmed.ncbi.nlm.nih.gov/23450602
412. Zimmern, P.E., et al. Management of recurrent stress urinary incontinence after burch and sling
procedures. Neurourol Urodyn, 2016. 35: 344.
https://pubmed.ncbi.nlm.nih.gov/25598512
413. Bakali, E., et al. Interventions for treating recurrent stress urinary incontinence after failed minimally
invasive synthetic midurethral tape surgery in women. Cochrane Database Syst Rev, 2019. 9:
CD009407.
https://pubmed.ncbi.nlm.nih.gov/31482580
414. Pradhan, A., et al. Effectiveness of midurethral slings in recurrent stress urinary incontinence: a
systematic review and meta-analysis. Int Urogynecol J, 2012. 23: 831.
https://pubmed.ncbi.nlm.nih.gov/22576328
415. Agur, W., et al. Surgical treatment of recurrent stress urinary incontinence in women: a systematic
review and meta-analysis of randomised controlled trials. Eur Urol, 2013. 64: 323.
https://pubmed.ncbi.nlm.nih.gov/23680414
416. Nikolopoulos, K.I., et al. The surgical management of recurrent stress urinary incontinence: a
systematic review. Acta Obstet Gynecol Scand, 2015. 94: 568.
https://pubmed.ncbi.nlm.nih.gov/25737292

124 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021
417. Amaye-Obu, F.A., et al. Surgical management of recurrent stress urinary incontinence: A 12-year
experience. Am J Obstet Gynecol, 1999. 181: 1296.
https://pubmed.ncbi.nlm.nih.gov/10601904
418. Errando-Smet, C., et al. A re-adjustable sling for female recurrent stress incontinence and intrinsic
sphincteric deficiency: Long-term results in 205 patients using the Remeex sling system. Neurourol
Urodyn, 2018. 37: 1349.
https://pubmed.ncbi.nlm.nih.gov/29130569
419. Chung, E., et al. 25-year experience in the outcome of artificial urinary sphincter in the treatment of
female urinary incontinence. BJU Int, 2010. 106: 1664.
https://pubmed.ncbi.nlm.nih.gov/20500509
420. Costa, P., et al. The use of an artificial urinary sphincter in women with type III incontinence and a
negative Marshall test. J Urol, 2001. 165: 1172.
https://pubmed.ncbi.nlm.nih.gov/11257664
421. Heitz, M., et al. [Therapy of female urinary incontinence with the AMS 800 artificial sphincter.
Indications, outcome, complications and risk factors]. Urologe A, 1997. 36: 426.
https://pubmed.ncbi.nlm.nih.gov/9424794
422. Vayleux, B., et al. Female urinary incontinence and artificial urinary sphincter: study of efficacy and
risk factors for failure and complications. Eur Urol, 2011. 59: 1048.
https://pubmed.ncbi.nlm.nih.gov/21420781
423. Mandron, E., et al. Laparoscopic artificial urinary sphincter implantation for female genuine stress
urinary incontinence: technique and 4-year experience in 25 patients. BJU Int, 2010. 106: 1194.
https://pubmed.ncbi.nlm.nih.gov/20132197
424. Roupret, M., et al. Laparoscopic approach for artificial urinary sphincter implantation in women with
intrinsic sphincter deficiency incontinence: a single-centre preliminary experience. Eur Urol, 2010.
57: 499.
https://pubmed.ncbi.nlm.nih.gov/19346059
425. Brennand, E.A., et al. Five years after midurethral sling surgery for stress incontinence: obesity
continues to have an impact on outcomes. Int Urogynecol J, 2017. 28: 621.
https://pubmed.ncbi.nlm.nih.gov/27686569
426. Brennand, E.A., et al. Twelve-month outcomes following midurethral sling procedures for stress
incontinence: impact of obesity. BJOG, 2015. 122: 1705.
https://pubmed.ncbi.nlm.nih.gov/25316484
427. Moore, R.D., et al. Two-year evaluation of the MiniArc in obese versus non-obese patients for
treatment of stress urinary incontinence. Int J Urol, 2013. 20: 434.
https://pubmed.ncbi.nlm.nih.gov/22989174
428. Rechberger, T., et al. The clinical effectiveness of retropubic (IVS-02) and transobturator (IVS-04)
midurethral slings: randomized trial. Eur Urol, 2009. 56: 24.
https://pubmed.ncbi.nlm.nih.gov/19285788
429. Barber, M.D., et al. Risk factors associated with failure 1 year after retropubic or transobturator
midurethral slings. Am J Obstet Gynecol, 2008. 199: 666 e1.
https://pubmed.ncbi.nlm.nih.gov/19084098
430. Richter, H.E., et al. Predictors of treatment failure 24 months after surgery for stress urinary
incontinence. J Urol, 2008. 179: 1024.
https://pubmed.ncbi.nlm.nih.gov/18206917
431. Campeau, L., et al. A multicenter, prospective, randomized clinical trial comparing tension-free
vaginal tape surgery and no treatment for the management of stress urinary incontinence in elderly
women. Neurourol Urodyn, 2007. 26: 990.
https://pubmed.ncbi.nlm.nih.gov/17638307
432. Serati, M., et al. Transobturator vaginal tape for the treatment of stress urinary incontinence in
elderly women without concomitant pelvic organ prolapse: is it effective and safe? Eur J Obstet
Gynecol Reprod Biol, 2013. 166: 107.
https://pubmed.ncbi.nlm.nih.gov/23164504
433. Groutz, A., et al. The safety and efficacy of the “inside-out” trans-obturator TVT in elderly versus
younger stress-incontinent women: a prospective study of 353 consecutive patients. Neurourol
Urodyn, 2011. 30: 380.
https://pubmed.ncbi.nlm.nih.gov/20665549
434. Chughtai, B., et al. Diagnosis, Evaluation, and Treatment of Mixed Urinary Incontinence in Women.
Rev Urol, 2015. 17: 78.
https://pubmed.ncbi.nlm.nih.gov/27222643

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 125
435. Minassian, V.A., et al. Severity of urinary incontinence and effect on quality of life in women by
incontinence type. Obstet Gynecol, 2013. 121: 1083.
https://pubmed.ncbi.nlm.nih.gov/23635747
436. Brubaker, L., et al. Mixed incontinence: comparing definitions in women having stress incontinence
surgery. Neurourol Urodyn, 2009. 28: 268.
https://pubmed.ncbi.nlm.nih.gov/19274758
437. Porena, M., et al. Mixed Incontinence: How Best to Manage It? Curr Bladder Dysfunct Rep, 2013. 8: 7.
https://pubmed.ncbi.nlm.nih.gov/23396610
438. Jung, S.Y., et al. Urethral afferent nerve activity affects the micturition reflex; implication for the
relationship between stress incontinence and detrusor instability. J Urol, 1999. 162: 204.
https://pubmed.ncbi.nlm.nih.gov/10379788
439. Mahony, D.T., et al. Integral storage and voiding reflexes. Neurophysiologic concept of continence
and micturition. Urology, 1977. 9: 95.
https://pubmed.ncbi.nlm.nih.gov/556658
440. Nygaard, I.E. Evidence-Based Treatment for Mixed Urinary Incontinence. JAMA, 2019. 322: 1049.
https://pubmed.ncbi.nlm.nih.gov/31528991
441. Myers, D.L. Female mixed urinary incontinence: a clinical review. JAMA, 2014. 311: 2007.
https://pubmed.ncbi.nlm.nih.gov/24846038
442. Nygaard, I.E., et al. Efficacy of pelvic floor muscle exercises in women with stress, urge, and mixed
urinary incontinence. Am J Obstet Gynecol, 1996. 174: 120.
https://pubmed.ncbi.nlm.nih.gov/8571994
443. Sar, D., et al. The effects of pelvic floor muscle training on stress and mixed urinary incontinence
and quality of life. J Wound Ostomy Continence Nurs, 2009. 36: 429.
https://pubmed.ncbi.nlm.nih.gov/19609165
444. Celiker Tosun, O., et al. Does pelvic floor muscle training abolish symptoms of urinary incontinence?
A randomized controlled trial. Clin Rehabil, 2015. 29: 525.
https://pubmed.ncbi.nlm.nih.gov/25142280
445. Liu, B., et al. Electroacupuncture Versus Pelvic Floor Muscle Training Plus Solifenacin for Women
With Mixed Urinary Incontinence: A Randomized Noninferiority Trial. Mayo Clin Proc, 2019. 94: 54.
https://pubmed.ncbi.nlm.nih.gov/30611454
446. Sung, V.W., et al. Effect of Behavioral and Pelvic Floor Muscle Therapy Combined With Surgery vs
Surgery Alone on Incontinence Symptoms Among Women With Mixed Urinary Incontinence: The
ESTEEM Randomized Clinical Trial. JAMA, 2019. 322: 1066.
https://pubmed.ncbi.nlm.nih.gov/31529007
447. Kaya, S., et al. Short-term effect of adding pelvic floor muscle training to bladder training for female
urinary incontinence: a randomized controlled trial. Int Urogynecol J, 2015. 26: 285.
https://pubmed.ncbi.nlm.nih.gov/25266357
448. Khullar, V., et al. Treatment of urge-predominant mixed urinary incontinence with tolterodine
extended release: a randomized, placebo-controlled trial. Urology, 2004. 64: 269.
https://pubmed.ncbi.nlm.nih.gov/15302476
449. Kreder, K.J., Jr., et al. Tolterodine is equally effective in patients with mixed incontinence and those
with urge incontinence alone. BJU Int, 2003. 92: 418.
https://pubmed.ncbi.nlm.nih.gov/12930432
450. Kelleher, C., et al. Solifenacin: as effective in mixed urinary incontinence as in urge urinary
incontinence. Int Urogynecol J Pelvic Floor Dysfunct, 2006. 17: 382.
https://pubmed.ncbi.nlm.nih.gov/16283422
451. Staskin, D.R., et al. Short- and long-term efficacy of solifenacin treatment in patients with symptoms
of mixed urinary incontinence. BJU Int, 2006. 97: 1256.
https://pubmed.ncbi.nlm.nih.gov/16686722
452. Bent, A.E., et al. Duloxetine compared with placebo for the treatment of women with mixed urinary
incontinence. Neurourol Urodyn, 2008. 27: 212.
https://pubmed.ncbi.nlm.nih.gov/17580357
453. Bump, R.C., et al. Mixed urinary incontinence symptoms: urodynamic findings, incontinence
severity, and treatment response. Obstet Gynecol, 2003. 102: 76.
https://pubmed.ncbi.nlm.nih.gov/12850610
454. Shirvan, M.K., et al. Tension-Free Vaginal Tape Plus Intradetrusor BOTOX® Injection Versus
Tension-Free Vaginal Tape Versus Intradetrusor BOTOX Injection in Equal-Weight Mixed Urinary
Incontinence: A Prospective Randomized Study. J Gynecol Surg, 2013. 29: 235.
https://www.liebertpub.com/doi/10.1089/gyn.2012.0134

126 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021
455. Kuo, H.C. Effect of detrusor function on the therapeutic outcome of a suburethral sling procedure
using a polypropylene sling for stress urinary incontinence in women. Scand J Urol Nephrol, 2007.
41: 138.
https://pubmed.ncbi.nlm.nih.gov/17454953
456. Colombo, M., et al. The Burch colposuspension for women with and without detrusor overactivity.
Br J Obstet Gynaecol, 1996. 103: 255.
https://pubmed.ncbi.nlm.nih.gov/8630311
457. Kulseng-Hanssen, S., et al. The tension free vaginal tape operation for women with mixed
incontinence: Do preoperative variables predict the outcome? Neurourol Urodyn, 2007. 26: 115.
https://pubmed.ncbi.nlm.nih.gov/16894616
458. Kulseng-Hanssen, S., et al. Follow-up of TVT operations in 1,113 women with mixed urinary
incontinence at 7 and 38 months. Int Urogynecol J Pelvic Floor Dysfunct, 2008. 19: 391.
https://pubmed.ncbi.nlm.nih.gov/17891326
459. Sung, V.W., et al. Methods for a multicenter randomized trial for mixed urinary incontinence:
rationale and patient-centeredness of the ESTEEM trial. Int Urogynecol J, 2016. 27: 1479.
https://pubmed.ncbi.nlm.nih.gov/27287818
460. Han, J.Y., et al. Effectiveness of retropubic tension-free vaginal tape and transobturator inside-out
tape procedures in women with overactive bladder and stress urinary incontinence. Int Neurourol J,
2013. 17: 145.
https://pubmed.ncbi.nlm.nih.gov/24143294
461. Natale, F., et al. Mixed urinary incontinence: A prospective study on the effect of trans-obturator
mid-urethral sling. Eur J Obstet Gynecol Reprod Biol, 2018. 221: 64.
https://pubmed.ncbi.nlm.nih.gov/29248808
462. Chapple, C.R., et al. Terminology report from the International Continence Society (ICS) Working
Group on Underactive Bladder (UAB). Neurourol Urodyn, 2018. 37: 2928.
https://pubmed.ncbi.nlm.nih.gov/30203560
463. Resnick, N.M., et al. The pathophysiology of urinary incontinence among institutionalized elderly
persons. N Engl J Med, 1989. 320: 1.
https://pubmed.ncbi.nlm.nih.gov/2909873
464. Abarbanel, J., et al. Impaired detrusor contractility in community-dwelling elderly presenting with
lower urinary tract symptoms. Urology, 2007. 69: 436.
https://pubmed.ncbi.nlm.nih.gov/17382138
465. Groutz, A., et al. Prevalence and characteristics of voiding difficulties in women: are subjective
symptoms substantiated by objective urodynamic data? Urology, 1999. 54: 268.
https://pubmed.ncbi.nlm.nih.gov/10443723
466. Jeong, S.J., et al. Prevalence and Clinical Features of Detrusor Underactivity among Elderly with
Lower Urinary Tract Symptoms: A Comparison between Men and Women. Korean J Urol, 2012. 53:
342.
https://pubmed.ncbi.nlm.nih.gov/22670194
467. Valente, S., et al. Epidemiology and demographics of the underactive bladder: a cross-sectional
survey. Int Urol Nephrol, 2014. 46 Suppl 1: S7.
https://pubmed.ncbi.nlm.nih.gov/25238889
468. Cohn, J.A., et al. Underactive bladder in women: is there any evidence? Curr Opin Urol, 2016.
26: 309.
https://pubmed.ncbi.nlm.nih.gov/26927630
469. Madersbacher, H., et al. What are the causes and consequences of bladder overdistension? ICI-RS
2011. Neurourol Urodyn, 2012. 31: 317.
https://pubmed.ncbi.nlm.nih.gov/22419355
470. Suskind, A.M., et al. A new look at detrusor underactivity: impaired contractility versus afferent
dysfunction. Curr Urol Rep, 2009. 10: 347.
https://pubmed.ncbi.nlm.nih.gov/19709481
471. Abdel Raheem, A., et al. Voiding dysfunction in women: How to manage it correctly. Arab J Urol,
2013. 11: 319.
https://pubmed.ncbi.nlm.nih.gov/26558099
472. Jiang, Y.H., et al. Urothelial Barrier Deficits, Suburothelial Inflammation and Altered Sensory Protein
Expression in Detrusor Underactivity. J Urol, 2017. 197: 197.
https://pubmed.ncbi.nlm.nih.gov/27436428
473. Osman, N.I., et al. Contemporary concepts in the aetiopathogenesis of detrusor underactivity. Nat
Rev Urol, 2014. 11: 639.
https://pubmed.ncbi.nlm.nih.gov/25330789

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 127
474. Gammie, A., et al. Signs and Symptoms of Detrusor Underactivity: An Analysis of Clinical
Presentation and Urodynamic Tests From a Large Group of Patients Undergoing Pressure Flow
Studies. Eur Urol, 2016. 69: 361.
https://pubmed.ncbi.nlm.nih.gov/26318706
475. Uren, A.D., et al. Qualitative Exploration of the Patient Experience of Underactive Bladder. Eur Urol,
2017. 72: 402.
https://pubmed.ncbi.nlm.nih.gov/28400168
476. Uren, A.D., et al. The development of the ICIQ-UAB: A patient reported outcome measure for
underactive bladder. Neurourol Urodyn, 2019. 38: 996.
https://pubmed.ncbi.nlm.nih.gov/30801826
477. Kira, S., et al. Detrusor pressures in urodynamic studies during voiding in women. Int Urogynecol J,
2017. 28: 783.
https://pubmed.ncbi.nlm.nih.gov/27999934
478. Osman, N.I., et al. Detrusor Underactivity and the Underactive Bladder: A Systematic Review of
Preclinical and Clinical Studies. Eur Urol, 2018. 74: 633.
https://pubmed.ncbi.nlm.nih.gov/30139634
479. Jeong, S.J., et al. How do we diagnose detrusor underactivity? Comparison of diagnostic criteria
based on an urodynamic measure. Investig Clin Urol, 2017. 58: 247.
https://pubmed.ncbi.nlm.nih.gov/28681034
480. Griffiths, D.J., et al. Urinary bladder function and its control in healthy females. Am J Physiol, 1986.
251: R225.
https://pubmed.ncbi.nlm.nih.gov/3740303
481. Tan, T.L., et al. Stop test or pressure-flow study? Measuring detrusor contractility in older females.
Neurourol Urodyn, 2004. 23: 184.
https://pubmed.ncbi.nlm.nih.gov/15098212
482. Tan, T.L., et al. Which stop test is best? Measuring detrusor contractility in older females. J Urol,
2003. 169: 1023.
https://pubmed.ncbi.nlm.nih.gov/12576837
483. Griffiths, D.J. Assessment of detrusor contraction strength or contractility. Neurourol Urodyn, 1991.
10: 1.
https://onlinelibrary.wiley.com/doi/abs/10.1002/nau.1930100102
484. van Koeveringe, G.A., et al. Detrusor underactivity: a plea for new approaches to a common bladder
dysfunction. Neurourol Urodyn, 2011. 30: 723.
https://pubmed.ncbi.nlm.nih.gov/21661020
485. Schäfer, W. Analysis of bladder-outlet function with the linearized passive urethral resistance
relation, linPURR, and a disease-specific approach for grading obstruction: from complex to simple.
World J Urol, 1995. 13: 47.
https://pubmed.ncbi.nlm.nih.gov/7773317
486. Ameda, K., et al. The long-term voiding function and sexual function after pelvic nerve-sparing
radical surgery for rectal cancer. Int J Urol, 2005. 12: 256.
https://pubmed.ncbi.nlm.nih.gov/15828952
487. Espino-Strebel, E.E., et al. A comparison of the feasibility and safety of nerve-sparing radical
hysterectomy with the conventional radical hysterectomy. Int J Gynecol Cancer, 2010. 20: 1274.
https://pubmed.ncbi.nlm.nih.gov/21495251
488. Salvatore S, et al. Pathophysiology of urinary incontinence, faecal incontinence and pelvic organ
prolapse, In: Incontinence, 6th Edn. 2017, Abrams, P., Cardozo, L., Wagg, A., Wein, A. (Eds). Bristol, UK.
https://www.ics.org/publications/ici_6/Incontinence_6th_Edition_2017_eBook_v2.pdf
489. El Akri, M., et al. Risk of prolapse and urinary complications in adult spina bifida patients with
neurogenic acontractile detrusor using clean intermittent catheterization versus Valsalva voiding.
Neurourol Urodyn, 2019. 38: 269.
https://pubmed.ncbi.nlm.nih.gov/30311685
490. Naess, I., et al. Can maximal voluntary pelvic floor muscle contraction reduce vaginal resting
pressure and resting EMG activity? Int Urogynecol J, 2018. 29: 1623.
https://pubmed.ncbi.nlm.nih.gov/29532122
491. Mercier, J., et al. Pelvic floor muscle training: mechanisms of action for the improvement of
genitourinary syndrome of menopause. Climacteric, 2020. 23: 468.
https://pubmed.ncbi.nlm.nih.gov/32105155
492. Ladi-Seyedian, S., et al. Management of non-neuropathic underactive bladder in children with
voiding dysfunction by animated biofeedback: a randomized clinical trial. Urology, 2015. 85: 205.
https://pubmed.ncbi.nlm.nih.gov/25444633

128 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021
493. Apostolidis A, et al., Neurological Urinary and fecal incontinence, In: Incontinence, 6th Edn. 2017,
Abrams, P., Cardozo, L., Wagg, A., Wein, A. (Eds). Bristol, UK.
https://www.ics.org/publications/ici_6/Incontinence_6th_Edition_2017_eBook_v2.pdf
494. Huber, E.R., et al. [The value of intravesical electrostimulation in the treatment of acute prolonged
bladder overdistension]. Urologe A, 2007. 46: 662.
https://pubmed.ncbi.nlm.nih.gov/17356837
495. Barendrecht, M.M., et al. Is the use of parasympathomimetics for treating an underactive urinary
bladder evidence-based? BJU Int, 2007. 99: 749.
https://pubmed.ncbi.nlm.nih.gov/17233798
496. Chang, S.J., et al. The effectiveness of tamsulosin in treating women with voiding difficulty. Int
J Urol, 2008. 15: 981.
https://pubmed.ncbi.nlm.nih.gov/18721208
497. Costantini, E., et al. Open-label, longitudinal study of tamsulosin for functional bladder outlet
obstruction in women. Urol Int, 2009. 83: 311.
https://pubmed.ncbi.nlm.nih.gov/19829032
498. Yamanishi, T., et al. Combination of a cholinergic drug and an alpha-blocker is more effective than
monotherapy for the treatment of voiding difficulty in patients with underactive detrusor. Int J Urol,
2004. 11: 88.
https://pubmed.ncbi.nlm.nih.gov/14706012
499. Buckley, B.S., et al. Drugs for treatment of urinary retention after surgery in adults. Cochrane
Database Syst Rev, 2010: CD008023.
https://pubmed.ncbi.nlm.nih.gov/20927768
500. Jonas, U., et al. Efficacy of sacral nerve stimulation for urinary retention: results 18 months after
implantation. J Urol, 2001. 165: 15.
https://pubmed.ncbi.nlm.nih.gov/11125353
501. Gross, C., et al. Sacral neuromodulation for nonobstructive urinary retention: a meta-analysis.
Female Pelvic Med Reconstr Surg, 2010. 16: 249.
https://pubmed.ncbi.nlm.nih.gov/22453352
502. Gani, J., et al. The underactive bladder: diagnosis and surgical treatment options. Transl Androl Urol,
2017. 6: S186.
https://pubmed.ncbi.nlm.nih.gov/28791238
503. Swinn, M.J., et al. Sacral neuromodulation for women with Fowler’s syndrome. Eur Urol, 2000.
38: 439.
https://pubmed.ncbi.nlm.nih.gov/11025383
504. Panicker, J.N., et al. Lower urinary tract dysfunction in the neurological patient: clinical assessment
and management. Lancet Neurol, 2015. 14: 720.
https://pubmed.ncbi.nlm.nih.gov/26067125
505. Kuo, H.C. Recovery of detrusor function after urethral botulinum A toxin injection in patients with
idiopathic low detrusor contractility and voiding dysfunction. Urology, 2007. 69: 57.
https://pubmed.ncbi.nlm.nih.gov/17270614
506. Kuo, H.C. Effect of botulinum a toxin in the treatment of voiding dysfunction due to detrusor
underactivity. Urology, 2003. 61: 550.
https://pubmed.ncbi.nlm.nih.gov/12639645
507. Lee, Y.-K., et al. Therapeutic Efficacy and Quality of Life Improvement in Women with Detrusor
Underactivity Following Transurethral Incision of the Bladder Ne. Urol Sci, 2019. 30: 266.
https://www.e-urol-sci.com/article.asp?issn=1879-5226;year=2019;volume=30;issue=6;spage=266;
epage=271;aulast=Lee;type=0
508. Thorner, D.A., et al. Outcomes of reduction cystoplasty in men with impaired detrusor contractility.
Urology, 2014. 83: 882.
https://pubmed.ncbi.nlm.nih.gov/24548706
509. Gakis, G., et al. Functional detrusor myoplasty for bladder acontractility: long-term results. J Urol,
2011. 185: 593.
https://pubmed.ncbi.nlm.nih.gov/21168866
510. Malde, S., et al. Female bladder outlet obstruction: Common symptoms masking an uncommon
cause. Low Urin Tract Symptoms, 2019. 11: 72.
https://pubmed.ncbi.nlm.nih.gov/28990728
511. Irwin, D.E., et al. Worldwide prevalence estimates of lower urinary tract symptoms, overactive
bladder, urinary incontinence and bladder outlet obstruction. BJU Int, 2011. 108: 1132.
https://pubmed.ncbi.nlm.nih.gov/21231991

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 129
512. Moossdorff-Steinhauser, H., et al. A Survey on Voiding Complaints in Women Presenting at a Pelvic
Care Center. Curr Urol, 2019. 13: 31.
https://pubmed.ncbi.nlm.nih.gov/31579228
513. Kupelian, V., et al. Prevalence of lower urinary tract symptoms and effect on quality of life in a
racially and ethnically diverse random sample: the Boston Area Community Health (BACH) Survey.
Arch Intern Med, 2006. 166: 2381.
https://pubmed.ncbi.nlm.nih.gov/17130393
514. Haylen, B.T., et al. Has the true prevalence of voiding difficulty in urogynecology patients been
underestimated? Int Urogynecol J Pelvic Floor Dysfunct, 2007. 18: 53.
https://pubmed.ncbi.nlm.nih.gov/16596458
515. Hoffman, D.S., et al. Female Bladder Outlet Obstruction. Curr Urol Rep, 2016. 17: 31.
https://pubmed.ncbi.nlm.nih.gov/26902625
516. Sussman, R.D., et al. Primary Bladder Neck Obstruction. Rev Urol, 2019. 21: 53.
https://pubmed.ncbi.nlm.nih.gov/31768132
517. Brucker, B.M., et al. Comparison of urodynamic findings in women with anatomical versus functional
bladder outlet obstruction. Female Pelvic Med Reconstr Surg, 2013. 19: 46.
https://pubmed.ncbi.nlm.nih.gov/23321659
518. Nitti, V.W. Primary bladder neck obstruction in men and women. Rev Urol, 2005. 7 Suppl 8: S12.
https://pubmed.ncbi.nlm.nih.gov/16985885
519. Panicker, J.N., et al. Do we understand voiding dysfunction in women? Current understanding and
future perspectives: ICI-RS 2017. Neurourol Urodyn, 2018. 37: S75.
https://pubmed.ncbi.nlm.nih.gov/30133794
520. Karmakar, D., et al. Current concepts in voiding dysfunction and dysfunctional voiding: A review
from a urogynaecologist’s perspective. J Midlife Health, 2014. 5: 104.
https://pubmed.ncbi.nlm.nih.gov/25316994
521. Panesar, K. Drug-Induced Urinary Incontinence. US Pharm, 2014. 39: 24.
https://www.uspharmacist.com/article/druginduced-urinary-incontinence
522. Groutz, A., et al. Bladder outlet obstruction in women: definition and characteristics. Neurourol
Urodyn, 2000. 19: 213.
https://pubmed.ncbi.nlm.nih.gov/10797578
523. Rosenblum, N., et al. Voiding dysfunction in young, nulliparous women: symptoms and urodynamic
findings. Int Urogynecol J Pelvic Floor Dysfunct, 2004. 15: 373.
https://pubmed.ncbi.nlm.nih.gov/15278258
524. Massey, J.A., et al. Obstructed voiding in the female. Br J Urol, 1988. 61: 36.
https://pubmed.ncbi.nlm.nih.gov/3342298
525. Gravina, G.L., et al. Urodynamic obstruction in women with stress urinary incontinence--do
nonintubated uroflowmetry and symptoms aid diagnosis? J Urol, 2007. 178: 959.
https://pubmed.ncbi.nlm.nih.gov/17632142
526. Klijer, R., et al. Bladder outlet obstruction in women: difficulties in the diagnosis. Urol Int, 2004.
73: 6.
https://pubmed.ncbi.nlm.nih.gov/15263784
527. Hsiao, S.M., et al. Videourodynamic Studies of Women with Voiding Dysfunction. Sci Rep, 2017.
7: 6845.
https://pubmed.ncbi.nlm.nih.gov/28754926
528. Qian, M., et al. Value of Real-Time Shear Wave Elastography Versus Acoustic Radiation Force
Impulse Imaging in the Diagnosis of Female Bladder Neck Obstruction. J Ultrasound Med, 2019.
38: 2427.
https://pubmed.ncbi.nlm.nih.gov/30680774
529. Galica, V., et al. Use of transvaginal ultrasound in females with primary bladder neck obstruction. A
preliminary study. Arch Ital Urol Androl, 2015. 87: 158.
https://pubmed.ncbi.nlm.nih.gov/26150036
530. Osman, N.I., et al. A systematic review of surgical techniques used in the treatment of female
urethral stricture. Eur Urol, 2013. 64: 965.
https://pubmed.ncbi.nlm.nih.gov/23937829
531. Webb, R.J., et al. Electromyographic abnormalities in the urethral and anal sphincters of women
with idiopathic retention of urine. Br J Urol, 1992. 70: 22.
https://pubmed.ncbi.nlm.nih.gov/1638369
532. Fowler, C.J., et al. Abnormal electromyographic activity of the urethral sphincter, voiding
dysfunction, and polycystic ovaries: a new syndrome? BMJ, 1988. 297: 1436.
https://pubmed.ncbi.nlm.nih.gov/3147005

130 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021
533. Tawadros, C., et al. External urethral sphincter electromyography in asymptomatic women and the
influence of the menstrual cycle. BJU Int, 2015. 116: 423.
https://pubmed.ncbi.nlm.nih.gov/25600712
534. Osman, N.I., et al. Contemporary surgical management of female urethral stricture disease. Curr
Opin Urol, 2015. 25: 341.
https://pubmed.ncbi.nlm.nih.gov/26049879
535. Rademakers, K., et al. Male bladder outlet obstruction: Time to re-evaluate the definition and
reconsider our diagnostic pathway? ICI-RS 2015. Neurourol Urodyn, 2017. 36: 894.
https://pubmed.ncbi.nlm.nih.gov/28444709
536. Meier, K., et al. Female bladder outlet obstruction: an update on diagnosis and management. Curr
Opin Urol, 2016. 26: 334.
https://pubmed.ncbi.nlm.nih.gov/27214578
537. Blaivas, J.G., et al. Bladder outlet obstruction nomogram for women with lower urinary tract
symptomatology. Neurourol Urodyn, 2000. 19: 553.
https://pubmed.ncbi.nlm.nih.gov/11002298
538. Lemack, G.E., et al. Pressure flow analysis may aid in identifying women with outflow obstruction.
J Urol, 2000. 163: 1823.
https://pubmed.ncbi.nlm.nih.gov/10799191
539. Solomon, E., et al. Developing and validating a new nomogram for diagnosing bladder outlet
obstruction in women. Neurourol Urodyn, 2018. 37: 368.
https://pubmed.ncbi.nlm.nih.gov/28666055
540. Lindsay, J., et al. Treatment validation of the Solomon-Greenwell nomogram for female bladder
outlet obstruction. Neurourol Urodyn, 2020. 39: 1371.
https://pubmed.ncbi.nlm.nih.gov/32249980
541. Cormier, L., et al. Diagnosis of female bladder outlet obstruction and relevance of the parameter area
under the curve of detrusor pressure during voiding: preliminary results. J Urol, 2002. 167: 2083.
https://pubmed.ncbi.nlm.nih.gov/11956445
542. Deindl, F.M., et al. Dysfunctional voiding in women: which muscles are responsible? Br J Urol, 1998.
82: 814.
https://pubmed.ncbi.nlm.nih.gov/9883217
543. Romanzi, L.J., et al. The effect of genital prolapse on voiding. J Urol, 1999. 161: 581.
https://pubmed.ncbi.nlm.nih.gov/9915453
544. Haq, S., et al. Comparison of effectiveness of clean intermittent self catheterization with no
catheterization after internal optical urethrotomy for urethral stricture. J Postgrad Med Inst, 2019.
33: 64.
https://www.researchgate.net/publication/332712706
545. Jackson, M.J., et al. Intermittent self-dilatation for urethral stricture disease in males. Cochrane
Database Syst Rev, 2014: CD010258.
https://pubmed.ncbi.nlm.nih.gov/25523166
546. Bailey, C., et al. Conservative management as an initial approach for post-operative voiding
dysfunction. Eur J Obstet Gynecol Reprod Biol, 2012. 160: 106.
https://pubmed.ncbi.nlm.nih.gov/22000341
547. Rijal, A., et al. Bladder outflow problems in females. Nepal Med Coll J, 2013. 15: 46.
https://pubmed.ncbi.nlm.nih.gov/24592794
548. Madjar, S., et al. A remote controlled intraurethral insert for artificial voiding: a new concept for
treating women with voiding dysfunction. J Urol, 1999. 161: 895.
https://pubmed.ncbi.nlm.nih.gov/10022709
549. Madjar, S., et al. Long-term follow-up of the in-flowtrade mark intraurethral insert for the treatment
of women with voiding dysfunction. Eur Urol, 2000. 38: 161.
https://pubmed.ncbi.nlm.nih.gov/10895007
550. Koh, J.S., et al. Comparison of alpha-blocker, extracorporeal magnetic stimulation alone and in
combination in the management of female bladder outlet obstruction. Int Urogynecol J, 2011.
22: 849.
https://pubmed.ncbi.nlm.nih.gov/21107813
551. Sigala, S., et al. Alpha1 adrenoceptor subtypes in human urinary bladder: sex and regional
comparison. Life Sci, 2004. 76: 417.
https://pubmed.ncbi.nlm.nih.gov/15530504
552. Kim, D.K., et al. Alpha-1 Adrenergic Receptor Blockers for the Treatment of Lower Urinary Tract
Symptoms in Women: A Systematic Review and Meta-Analysis. Int Neurourol J, 2019. 23: 56.
https://pubmed.ncbi.nlm.nih.gov/30943695

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 131
553. Boyd, K., et al. α-adrenergic blockers for the treatment of lower-urinary-tract symptoms and
dysfunction in women. Ann Pharmacother, 2014. 48: 711.
https://pubmed.ncbi.nlm.nih.gov/24615630
554. Meyer, L.E., et al. Tamsulosin for voiding dysfunction in women. Int Urol Nephrol, 2012. 44: 1649.
https://pubmed.ncbi.nlm.nih.gov/22983886
555. Pischedda, A., et al. Use of alpha1-blockers in female functional bladder neck obstruction. Urol Int,
2005. 74: 256.
https://pubmed.ncbi.nlm.nih.gov/15812214
556. Kessler, T.M., et al. The effect of terazosin on functional bladder outlet obstruction in women: a pilot
study. J Urol, 2006. 176: 1487.
https://pubmed.ncbi.nlm.nih.gov/16952666
557. Athanasopoulos, A., et al. Effect of alfuzosin on female primary bladder neck obstruction. Int
Urogynecol J Pelvic Floor Dysfunct, 2009. 20: 217.
https://pubmed.ncbi.nlm.nih.gov/18982236
558. Lee, Y.S., et al. Efficacy of an Alpha-Blocker for the Treatment of Nonneurogenic Voiding
Dysfunction in Women: An 8-Week, Randomized, Double-Blind, Placebo-Controlled Trial. Int
Neurourol J, 2018. 22: 30.
https://pubmed.ncbi.nlm.nih.gov/29609420
559. Hajebrahimi, S., et al. Effect of tamsulosin versus prazosin on clinical and urodynamic parameters in
women with voiding difficulty: a randomized clinical trial. Int J Gen Med, 2011. 4: 35.
https://pubmed.ncbi.nlm.nih.gov/21403790
560. Xu, D., et al. Dysfunctional voiding confirmed by transdermal perineal electromyography, and
its effective treatment with baclofen in women with lower urinary tract symptoms: a randomized
double-blind placebo-controlled crossover trial. BJU Int, 2007. 100: 588.
https://pubmed.ncbi.nlm.nih.gov/17511770
561. Chen, C.H., et al. Clinical and urodynamic effects of baclofen in women with functional bladder
outlet obstruction: Preliminary report. J Obstet Gynaecol Res, 2016. 42: 560.
https://pubmed.ncbi.nlm.nih.gov/27108667
562. Datta, S.N., et al. Results of double-blind placebo-controlled crossover study of sildenafil citrate
(Viagra) in women suffering from obstructed voiding or retention associated with the primary
disorder of sphincter relaxation (Fowler’s Syndrome). Eur Urol, 2007. 51: 489.
https://pubmed.ncbi.nlm.nih.gov/16884844
563. Rosario, D.J., et al. Effects of intravenous thyrotropin-releasing hormone on urethral closure
pressure in females with voiding dysfunction. Eur Urol, 1995. 28: 64.
https://pubmed.ncbi.nlm.nih.gov/8521898
564. Kao, Y.L., et al. The Therapeutic Effects and Pathophysiology of Botulinum Toxin A on Voiding
Dysfunction Due to Urethral Sphincter Dysfunction. Toxins (Basel), 2019. 11.
https://pubmed.ncbi.nlm.nih.gov/31847090
565. Jiang, Y.H., et al. OnabotulinumtoxinA Urethral Sphincter Injection as Treatment for Non-neurogenic
Voiding Dysfunction - A Randomized, Double-Blind, Placebo-Controlled Study. Sci Rep, 2016.
6: 38905.
https://pubmed.ncbi.nlm.nih.gov/27958325
566. Pradhan, A.A. Botulinum toxin: An emerging therapy in female bladder outlet obstruction. Indian
J Urol, 2009. 25: 318.
https://pubmed.ncbi.nlm.nih.gov/19881122
567. Peeters, K., et al. Long-term follow-up of sacral neuromodulation for lower urinary tract dysfunction.
BJU Int, 2014. 113: 789.
https://pubmed.ncbi.nlm.nih.gov/24238278
568. Fletcher, S.G., et al. Demographic and urodynamic factors associated with persistent OAB after
anterior compartment prolapse repair. Neurourol Urodyn, 2010. 29: 1414.
https://pubmed.ncbi.nlm.nih.gov/20623545
569. Espuña Pons, M., et al. Post-void residual and voiding dysfunction symptoms in women with pelvic
organ prolapse before and after vaginal surgery. A multicenter cohort study. Actas Urol Esp, 2020.
https://pubmed.ncbi.nlm.nih.gov/32593638
570. Togo, M., et al. Lower urinary tract function improves after laparoscopic sacrocolpopexy for elderly
patients with pelvic organ prolapse. Low Urin Tract Symptoms, 2020. 12: 260.
https://pubmed.ncbi.nlm.nih.gov/32347664
571. Fitzgerald, M.P., et al. Postoperative resolution of urinary retention in patients with advanced pelvic
organ prolapse. Am J Obstet Gynecol, 2000. 183: 1361.
https://pubmed.ncbi.nlm.nih.gov/11120497

132 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021
572. Hirata, H., et al. Does surgical repair of pelvic prolapse improve patients’ quality of life? Eur Urol,
2004. 45: 213.
https://pubmed.ncbi.nlm.nih.gov/14734009
573. Basu, M., et al. Urethral dilatation: Is there any benefit over cystoscopy and distension? A
randomized trial in women with overactive bladder symptoms. Neurourol Urodyn, 2014. 33: 283.
https://pubmed.ncbi.nlm.nih.gov/23636866
574. Heidari, F., et al. On demand urethral dilatation versus intermittent urethral dilatation: results and
complications in women with urethral stricture. Nephrourol Mon, 2014. 6: e15212.
https://pubmed.ncbi.nlm.nih.gov/24783171
575. Grivas, N., et al. The effectiveness of otis urethrotomy combined with six weeks urethral dilations
until 40 Fr in the treatment of bladder outlet obstruction in women: a prospective study. Urol J,
2014. 10: 1063.
https://pubmed.ncbi.nlm.nih.gov/24469651
576. Yee, C.H., et al. The effect of urethral calibration on female primary bladder outlet obstruction. Int
Urogynecol J, 2012. 23: 217.
https://pubmed.ncbi.nlm.nih.gov/21809157
577. Popat, S., et al. Long-term management of luminal urethral stricture in women. Int Urogynecol J,
2016. 27: 1735.
https://pubmed.ncbi.nlm.nih.gov/27026141
578. Choudhury, A. Incisional treatment of obstruction of the female bladder neck. Ann R Coll Surg Engl,
1978. 60: 404.
https://pubmed.ncbi.nlm.nih.gov/697298
579. Zhang, P., et al. Bladder neck incision for female bladder neck obstruction: long-term outcomes.
Urology, 2014. 83: 762.
https://pubmed.ncbi.nlm.nih.gov/24680443
580. Fu, Q., et al. Transurethral incision of the bladder neck using KTP in the treatment of bladder neck
obstruction in women. Urol Int, 2009. 82: 61.
https://pubmed.ncbi.nlm.nih.gov/19172099
581. Shen, W., et al. Controlled transurethral resection and incision of the bladder neck to treat female
primary bladder neck obstruction: Description of a novel surgical procedure. Int J Urol, 2016. 23: 491.
https://pubmed.ncbi.nlm.nih.gov/27037830
582. Jin, X.B., et al. Modified transurethral incision for primary bladder neck obstruction in women: a
method to improve voiding function without urinary incontinence. Urology, 2012. 79: 310.
https://pubmed.ncbi.nlm.nih.gov/22310746
583. Hoag, N., et al. Surgical management of female urethral strictures. Transl Androl Urol, 2017. 6: S76.
https://pubmed.ncbi.nlm.nih.gov/28791225
584. Nayyar, R., et al. A Novel Anterior Bladder Tube for Traumatic Bladder Neck Contracture in Females:
Initial Results. Urology, 2020. 139: 201.
https://pubmed.ncbi.nlm.nih.gov/32061615
585. Ram-Liebig, G., et al. Results of Use of Tissue-Engineered Autologous Oral Mucosa Graft for Urethral
Reconstruction: A Multicenter, Prospective, Observational Trial. EBioMedicine, 2017. 23: 185.
https://pubmed.ncbi.nlm.nih.gov/28827035
586. Hampson, L.A., et al. Dorsal buccal graft urethroplasty in female urethral stricture disease: a multi-
center experience. Transl Androl Urol, 2019. 8: S6.
https://pubmed.ncbi.nlm.nih.gov/31143666
587. Gomez, R.G., et al. Female urethral reconstruction: dorsal buccal mucosa graft onlay. World J Urol,
2019.
https://pubmed.ncbi.nlm.nih.gov/31542825
588. Tao, T.-T et al. Novel surgical technique for female distal urethral stricture disease: an evaluation of
efficacy and safety compared with urethral dilatation. Int J Clin Exp Med 2018. 11: 12002.
http://www.ijcem.com/files/ijcem0080490.pdf
589. Mouracade, P., et al. Transvaginal tape lysis for urinary obstruction after suburethral tape placement.
When to do an immediate replacement? Int Urogynecol J Pelvic Floor Dysfunct, 2008. 19: 1271.
https://pubmed.ncbi.nlm.nih.gov/18461268
590. McCrery, R., et al. Transvaginal urethrolysis for obstruction after antiincontinence surgery. Int
Urogynecol J Pelvic Floor Dysfunct, 2007. 18: 627.
https://pubmed.ncbi.nlm.nih.gov/17036167
591. Giannis, G., et al. Can urethrolysis resolve outlet obstruction related symptoms after Burch
colposuspension for stress urinary incontinence? Eur J Obstet Gynecol Reprod Biol, 2015. 195: 103.
https://pubmed.ncbi.nlm.nih.gov/26512435

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 133
592. Leng, W.W., et al. Delayed treatment of bladder outlet obstruction after sling surgery: association
with irreversible bladder dysfunction. J Urol, 2004. 172: 1379.
https://pubmed.ncbi.nlm.nih.gov/15371849
593. Agnew, G., et al. Functional outcomes for surgical revision of synthetic slings performed for voiding
dysfunction: a retrospective study. Eur J Obstet Gynecol Reprod Biol, 2012. 163: 113.
https://pubmed.ncbi.nlm.nih.gov/22579029
594. Van den Broeck, T., et al. The value of surgical release after obstructive anti-incontinence surgery:
An aid for clinical decision making. Neurourol Urodyn, 2015. 34: 736.
https://pubmed.ncbi.nlm.nih.gov/25212178
595. Hashim, H., et al. International Continence Society (ICS) report on the terminology for nocturia and
nocturnal lower urinary tract function. Neurourol Urodyn, 2019. 38: 499.
https://pubmed.ncbi.nlm.nih.gov/30644584
596. Bedretdinova, D., et al. What Is the Most Effective Treatment for Nocturia or Nocturnal Incontinence
in Adult Women? Eur Urol Focus, 2020.
https://pubmed.ncbi.nlm.nih.gov/32061540
597. Bosch, J.L., et al. The prevalence and causes of nocturia. J Urol, 2010. 184: 440.
https://pubmed.ncbi.nlm.nih.gov/20620395
598. Burgio, K.L., et al. Prevalence and correlates of nocturia in community-dwelling older adults. J Am
Geriatr Soc, 2010. 58: 861.
https://pubmed.ncbi.nlm.nih.gov/20406317
599. Fitzgerald, M.P., et al. The association of nocturia with cardiac disease, diabetes, body mass index,
age and diuretic use: results from the BACH survey. J Urol, 2007. 177: 1385.
https://pubmed.ncbi.nlm.nih.gov/17382738
600. Pesonen, J.S., et al. The Impact of Nocturia on Mortality: A Systematic Review and Meta-Analysis.
J Urol, 2020. 203: 486.
https://pubmed.ncbi.nlm.nih.gov/31364920
601. Smith, A.L., et al. Outcomes of pharmacological management of nocturia with non-antidiuretic
agents: does statistically significant equal clinically significant? BJU Int, 2011. 107: 1550.
https://pubmed.ncbi.nlm.nih.gov/21518417
602. Gordon, D.J., et al. Management Strategies for Nocturia. Curr Urol Rep, 2019. 20: 75.
https://pubmed.ncbi.nlm.nih.gov/31707521
603. Weiss, J.P., et al. Management of Nocturia and Nocturnal Polyuria. Urology, 2019. 133S: 24.
https://pubmed.ncbi.nlm.nih.gov/31586470
604. Abraham, L., et al. Development and validation of a quality-of-life measure for men with nocturia.
Urology, 2004. 63: 481.
https://pubmed.ncbi.nlm.nih.gov/15028442
605. Holm-Larsen, T., et al. The Nocturia Impact Diary: a self-reported impact measure to complement
the voiding diary. Value Health, 2014. 17: 696.
https://pubmed.ncbi.nlm.nih.gov/25236993
606. Bower, W.F., et al. TANGO - a screening tool to identify comorbidities on the causal pathway of
nocturia. BJU Int, 2017. 119: 933.
https://pubmed.ncbi.nlm.nih.gov/28075514
607. van Kerrebroeck, P., et al. The standardisation of terminology in nocturia: report from the
Standardisation Sub-committee of the International Continence Society. Neurourol Urodyn, 2002.
21: 179.
https://pubmed.ncbi.nlm.nih.gov/11857672
608. Weiss, J.P., et al. Excessive nocturnal urine production is a major contributing factor to the etiology
of nocturia. J Urol, 2011. 186: 1358.
https://pubmed.ncbi.nlm.nih.gov/21855948
609. Denys, M.A., et al. ICI-RS 2015-Is a better understanding of sleep the key in managing nocturia?
Neurourol Urodyn, 2018. 37: 2048.
https://pubmed.ncbi.nlm.nih.gov/27653805
610. Johnson, T.M., 2nd, et al. Effects of behavioral and drug therapy on nocturia in older incontinent
women. J Am Geriatr Soc, 2005. 53: 846.
https://pubmed.ncbi.nlm.nih.gov/15877562
611. Aslan, E., et al. Bladder training and Kegel exercises for women with urinary complaints living in a
rest home. Gerontology, 2008. 54: 224.
https://pubmed.ncbi.nlm.nih.gov/18483451

134 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021
612. Lo, S.K., et al. Additive Effect of Interferential Therapy Over Pelvic Floor Exercise Alone in the
Treatment of Female Urinary Stress and Urge Incontinence: A Randomized Controlled Trial. Hong
Kong Phys J, 2003. 21: 37.
https://www.sciencedirect.com/science/article/pii/S1013702509700387
613. Fitzgerald, M.P., et al. Nocturia, nocturnal incontinence prevalence, and response to anticholinergic
and behavioral therapy. Int Urogynecol J Pelvic Floor Dysfunct, 2008. 19: 1545.
https://pubmed.ncbi.nlm.nih.gov/18704249
614. But, I., et al. Functional magnetic stimulation for mixed urinary incontinence. J Urol, 2005.
173: 1644.
https://pubmed.ncbi.nlm.nih.gov/15821527
615. Wang, T., et al. The Efficacy of Continuous Positive Airway Pressure Therapy on Nocturia in Patients
With Obstructive Sleep Apnea: A Systematic Review and Meta-Analysis. Int Neurourol J, 2015.
19: 178.
https://pubmed.ncbi.nlm.nih.gov/26620900
616. Hilton, P., et al. The use of desmopressin (DDAVP) in nocturnal urinary frequency in the female. Br
J Urol, 1982. 54: 252.
https://pubmed.ncbi.nlm.nih.gov/7049302
617. Lose, G., et al. Efficacy of desmopressin (Minirin) in the treatment of nocturia: a double-blind
placebo-controlled study in women. Am J Obstet Gynecol, 2003. 189: 1106.
https://pubmed.ncbi.nlm.nih.gov/14586363
618. Yamaguchi, O., et al. Gender difference in efficacy and dose response in Japanese patients
with nocturia treated with four different doses of desmopressin orally disintegrating tablet in a
randomized, placebo-controlled trial. BJU Int, 2013. 111: 474.
https://pubmed.ncbi.nlm.nih.gov/23046147
619. Rovner, E.S., et al. Low-dose Desmopressin and Tolterodine Combination Therapy for Treating
Nocturia in Women with Overactive Bladder: A Double-blind, Randomized, Controlled Study. Low
Urin Tract Symptoms, 2018. 10: 221.
https://pubmed.ncbi.nlm.nih.gov/28560762
620. Rembratt, A., et al. Desmopressin treatment in nocturia; an analysis of risk factors for hyponatremia.
Neurourol Urodyn, 2006. 25: 105.
https://pubmed.ncbi.nlm.nih.gov/16304673
621. Yuan, Z., et al. Comparison of Tolterodine with Estazolam versus Tolterodine alone for the treatment
of women with overactive Bladder Syndrome and Nocturia: A non-randomized prospective
comparative study. Pakistan J Med Sci Online, 2011. 27: 763.
https://www.researchgate.net/publication/290027958
622. Lose, G., et al. Oestradiol-releasing vaginal ring versus oestriol vaginal pessaries in the treatment of
bothersome lower urinary tract symptoms. BJOG, 2000. 107: 1029.
https://pubmed.ncbi.nlm.nih.gov/10955437
623. Reynard, J.M., et al. A novel therapy for nocturnal polyuria: a double-blind randomized trial of
frusemide against placebo. Br J Urol, 1998. 81: 215.
https://pubmed.ncbi.nlm.nih.gov/9488061
624. Cameron, A.P. Systematic review of lower urinary tract symptoms occurring with pelvic organ
prolapse. Arab J Urol, 2019. 17: 23.
https://www.tandfonline.com/doi/full/10.1080/2090598X.2019.1589929
625. Wu, J.M., et al. Lifetime risk of stress urinary incontinence or pelvic organ prolapse surgery. Obstet
Gynecol, 2014. 123: 1201.
https://pubmed.ncbi.nlm.nih.gov/24807341
626. Vergeldt, T.F., et al. Risk factors for pelvic organ prolapse and its recurrence: a systematic review. Int
Urogynecol J, 2015. 26: 1559.
https://pubmed.ncbi.nlm.nih.gov/25966804
627. DeLancey, J.O. What’s new in the functional anatomy of pelvic organ prolapse? Curr Opin Obstet
Gynecol, 2016. 28: 420.
https://pubmed.ncbi.nlm.nih.gov/27517338
628. Chen, A., et al. Management of Postoperative Lower Urinary Tract Symptoms (LUTS) After Pelvic
Organ Prolapse (POP) Repair. Curr Urol Rep, 2018. 19: 74.
https://pubmed.ncbi.nlm.nih.gov/30043287
629. Bump, R.C., et al. The standardization of terminology of female pelvic organ prolapse and pelvic
floor dysfunction. Am J Obstet Gynecol, 1996. 175: 10.
https://pubmed.ncbi.nlm.nih.gov/8694033

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 135
630. Theofrastous, J.P., et al. The clinical evaluation of pelvic floor dysfunction. Obstet Gynecol Clin
North Am, 1998. 25: 783.
https://pubmed.ncbi.nlm.nih.gov/9921557
631. Ellerkmann, R.M., et al. Correlation of symptoms with location and severity of pelvic organ prolapse.
Am J Obstet Gynecol, 2001. 185: 1332.
https://pubmed.ncbi.nlm.nih.gov/11744905
632. Grob, A.T.M., et al. Underestimation of pelvic organ prolapse in the supine straining position, based
on magnetic resonance imaging findings. Int Urogynecol J, 2019. 30: 1939.
https://pubmed.ncbi.nlm.nih.gov/30656361
633. Espuña-Pons, M., et al. Cough stress tests to diagnose stress urinary incontinence in women with
pelvic organ prolapse with indication for surgical treatment. Neurourol Urodyn, 2020. 39: 819.
https://pubmed.ncbi.nlm.nih.gov/32040873
634. Visco, A.G., et al. The role of preoperative urodynamic testing in stress-continent women
undergoing sacrocolpopexy: the Colpopexy and Urinary Reduction Efforts (CARE) randomized
surgical trial. Int Urogynecol J Pelvic Floor Dysfunct, 2008. 19: 607.
https://pubmed.ncbi.nlm.nih.gov/18185903
635. Wei, J.T., et al. A midurethral sling to reduce incontinence after vaginal prolapse repair. N Engl
J Med, 2012. 366: 2358.
https://pubmed.ncbi.nlm.nih.gov/22716974
636. van der Ploeg, J.M., et al. The predictive value of demonstrable stress incontinence during basic
office evaluation and urodynamics in women without symptomatic urinary incontinence undergoing
vaginal prolapse surgery. Neurourol Urodyn, 2018. 37: 1011.
https://pubmed.ncbi.nlm.nih.gov/28834564
637. Jelovsek, J.E., et al. A model for predicting the risk of de novo stress urinary incontinence in women
undergoing pelvic organ prolapse surgery. Obstet Gynecol, 2014. 123: 279.
https://pubmed.ncbi.nlm.nih.gov/24402598
638. Lo, T.S., et al. Predictors for detrusor overactivity following extensive vaginal pelvic reconstructive
surgery. Neurourol Urodyn, 2018. 37: 192.
https://pubmed.ncbi.nlm.nih.gov/28370456
639. Lo, T.S., et al. Clinical outcomes of detrusor underactivity in female with advanced pelvic organ
prolapse following vaginal pelvic reconstructive surgery. Neurourol Urodyn, 2018. 37: 2242.
https://pubmed.ncbi.nlm.nih.gov/29664135
640. Frawley, H.C., et al. Physiotherapy as an adjunct to prolapse surgery: an assessor-blinded
randomized controlled trial. Neurourol Urodyn, 2010. 29: 719.
https://pubmed.ncbi.nlm.nih.gov/19816918
641. Braekken, I.H., et al. Can pelvic floor muscle training reverse pelvic organ prolapse and reduce
prolapse symptoms? An assessor-blinded, randomized, controlled trial. Am J Obstet Gynecol, 2010.
203: 170 e1.
https://pubmed.ncbi.nlm.nih.gov/20435294
642. Cheung, R.Y., et al. Vaginal Pessary in Women With Symptomatic Pelvic Organ Prolapse: A
Randomized Controlled Trial. Obstet Gynecol, 2016. 128: 73.
https://pubmed.ncbi.nlm.nih.gov/27275798
643. Due, U., et al. Lifestyle advice with or without pelvic floor muscle training for pelvic organ prolapse:
a randomized controlled trial. Int Urogynecol J, 2016. 27: 555.
https://pubmed.ncbi.nlm.nih.gov/26439114
644. Due, U., et al. The 12-month effects of structured lifestyle advice and pelvic floor muscle training for
pelvic organ prolapse. Acta Obstet Gynecol Scand, 2016. 95: 811.
https://pubmed.ncbi.nlm.nih.gov/26910261
645. Hagen, S., et al. Individualised pelvic floor muscle training in women with pelvic organ prolapse
(POPPY): a multicentre randomised controlled trial. Lancet, 2014. 383: 796.
https://pubmed.ncbi.nlm.nih.gov/24290404
646. Panman, C.M., et al. Effectiveness and cost-effectiveness of pessary treatment compared with
pelvic floor muscle training in older women with pelvic organ prolapse: 2-year follow-up of a
randomized controlled trial in primary care. Menopause, 2016. 23: 1307.
https://pubmed.ncbi.nlm.nih.gov/27504918
647. Duarte, T.B., et al. Perioperative pelvic floor muscle training did not improve outcomes in women
undergoing pelvic organ prolapse surgery: a randomised trial. J Physiother, 2020. 66: 27.
https://pubmed.ncbi.nlm.nih.gov/31843420

136 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021
648. Pauls, R.N., et al. Pelvic floor physical therapy: impact on quality of life 6 months after vaginal
reconstructive surgery. Female Pelvic Med Reconstr Surg, 2014. 20: 334.
https://pubmed.ncbi.nlm.nih.gov/25185628
649. Barber, M.D., et al. Comparison of 2 transvaginal surgical approaches and perioperative behavioral
therapy for apical vaginal prolapse: the OPTIMAL randomized trial. JAMA, 2014. 311: 1023.
https://pubmed.ncbi.nlm.nih.gov/24618964
650. Weidner, A.C., et al. Perioperative Behavioral Therapy and Pelvic Muscle Strengthening Do Not
Enhance Quality of Life After Pelvic Surgery: Secondary Report of a Randomized Controlled Trial.
Phys Ther, 2017. 97: 1075.
https://pubmed.ncbi.nlm.nih.gov/29077924
651. Hagen, S., et al. Pelvic floor muscle training for secondary prevention of pelvic organ prolapse
(PREVPROL): a multicentre randomised controlled trial. Lancet, 2017. 389: 393.
https://pubmed.ncbi.nlm.nih.gov/28010994
652. Panman, C., et al. Two-year effects and cost-effectiveness of pelvic floor muscle training in mild
pelvic organ prolapse: a randomised controlled trial in primary care. BJOG, 2017. 124: 511.
https://pubmed.ncbi.nlm.nih.gov/26996291
653. de Boer, T.A., et al. Pelvic organ prolapse and overactive bladder. Neurourol Urodyn, 2010. 29: 30.
https://pubmed.ncbi.nlm.nih.gov/20025017
654. de Boer, T.A., et al. Predictive factors for overactive bladder symptoms after pelvic organ prolapse
surgery. Int Urogynecol J, 2010. 21: 1143.
https://pubmed.ncbi.nlm.nih.gov/20419366
655. Kim, M.S., et al. The association of pelvic organ prolapse severity and improvement in overactive
bladder symptoms after surgery for pelvic organ prolapse. Obstet Gynecol Sci, 2016. 59: 214.
https://pubmed.ncbi.nlm.nih.gov/27200312
656. Costantini, E., et al. Urgency, detrusor overactivity and posterior vault prolapse in women who
underwent pelvic organ prolapse repair. Urol Int, 2013. 90: 168.
https://pubmed.ncbi.nlm.nih.gov/23327990
657. van der Ploeg, J.M., et al. Prolapse surgery with or without incontinence procedure: a systematic
review and meta-analysis. BJOG, 2018. 125: 289.
https://pubmed.ncbi.nlm.nih.gov/28941138
658. Baessler, K., et al. Surgery for women with pelvic organ prolapse with or without stress urinary
incontinence. Cochrane Database Syst Rev, 2018. 8: CD013108.
https://pubmed.ncbi.nlm.nih.gov/30121956
659. Borstad, E., et al. Surgical strategies for women with pelvic organ prolapse and urinary stress
incontinence. Int Urogynecol J, 2010. 21: 179.
https://pubmed.ncbi.nlm.nih.gov/19940978
660. van der Ploeg, J.M., et al. Transvaginal prolapse repair with or without the addition of a midurethral
sling in women with genital prolapse and stress urinary incontinence: a randomised trial. BJOG,
2015. 122: 1022.
https://pubmed.ncbi.nlm.nih.gov/25754458
661. Costantini, E., et al. Burch colposuspension does not provide any additional benefit to pelvic organ
prolapse repair in patients with urinary incontinence: a randomized surgical trial. J Urol, 2008.
180: 1007.
https://pubmed.ncbi.nlm.nih.gov/18639302
662. Costantini, E., et al. Pelvic Organ Prolapse Repair with and without Concomitant Burch
Colposuspension in Incontinent Women: A Randomised Controlled Trial with at Least 5-Year
Followup. Obstet Gynecol Int, 2012. 2012: 967923.
https://pubmed.ncbi.nlm.nih.gov/22028719
663. Trabuco, E.C., et al. Treatment success of burch and midurethral sling 2 years following combined
procedure with sacrocolpopexy. Int Urogynecol J, 2016. 27: S46. [No abstract available].
664. Meschia, M., et al. A randomized comparison of tension-free vaginal tape and endopelvic fascia
plication in women with genital prolapse and occult stress urinary incontinence. Am J Obstet
Gynecol, 2004. 190: 609.
https://pubmed.ncbi.nlm.nih.gov/15041988
665. Fuentes, A.E. A prospective randomised controlled trial comparing vaginal prolapse repair with and
without tensionfree vaginal tape transobturator tape (TVTO) in women with severe genital prolapse
and occult stress incontinence: long term follow up. Int Urogynecol J, 2011. 22: Abstract 059:S60.
[No abstract available].

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 137
666. Schierlitz, L., et al. Pelvic organ prolapse surgery with and without tension-free vaginal tape in
women with occult or asymptomatic urodynamic stress incontinence: a randomised controlled trial.
Int Urogynecol J, 2014. 25: 33.
https://pubmed.ncbi.nlm.nih.gov/23812579
667. van der Ploeg, J.M., et al. Vaginal prolapse repair with or without a midurethral sling in women with
genital prolapse and occult stress urinary incontinence: a randomized trial. Int Urogynecol J, 2016.
27: 1029.
https://pubmed.ncbi.nlm.nih.gov/26740197
668. Costantini, E., et al. Must colposuspension be associated with sacropexy to prevent postoperative
urinary incontinence? Eur Urol, 2007. 51: 788.
https://pubmed.ncbi.nlm.nih.gov/17011699
669. Brubaker, L., et al. Abdominal sacrocolpopexy with Burch colposuspension to reduce urinary stress
incontinence. N Engl J Med, 2006. 354: 1557.
https://pubmed.ncbi.nlm.nih.gov/16611949
670. De Ridder, D., et al. Fistula (Committee 18), In: Incontinence, 6th Edn. 2017, Abrams, P., Cardozo,
L., Wagg, A., Wein, A. (Eds). Bristol, UK.
https://www.ics.org/publications/ici_6/Incontinence_6th_Edition_2017_eBook_v2.pdf
671. de Bernis, L. Obstetric fistula: Guiding principles for clinical management and programme
development, a new WHO guideline. Int J Gynecol Obstet, 2007. 99: S117.
https://obgyn.onlinelibrary.wiley.com/doi/abs/10.1016/j.ijgo.2007.06.032
672. Mselle, L.T., et al. Perceived Health System Causes of Obstetric Fistula from Accounts of Affected
Women in Rural Tanzania: A Qualitative Study. Afr J Reprod Health, 2015. 19: 124.
https://pubmed.ncbi.nlm.nih.gov/26103702
673. Mwini-Nyaledzigbor, P.P., et al. Lived experiences of Ghanaian women with obstetric fistula. Health
Care Women Int, 2013. 34: 440.
https://pubmed.ncbi.nlm.nih.gov/23641897
674. Jokhio, A.H., et al. Prevalence of obstetric fistula: a population-based study in rural Pakistan. BJOG,
2014. 121: 1039.
https://pubmed.ncbi.nlm.nih.gov/24684695
675. Farid, F.N., et al. Psychosocial experiences of women with vesicovaginal fistula: a qualitative
approach. J Coll Physicians Surg Pak, 2013. 23: 828.
https://pubmed.ncbi.nlm.nih.gov/24169399
676. Imoto, A., et al. Health-related quality of life among women in rural Bangladesh after surgical repair
of obstetric fistula. Int J Gynaecol Obstet, 2015. 130: 79.
https://pubmed.ncbi.nlm.nih.gov/25935472
677. Tancer, M.L. Observations on prevention and management of vesicovaginal fistula after total
hysterectomy. Surg Gynecol Obstet, 1992. 175: 501.
https://pubmed.ncbi.nlm.nih.gov/1448730
678. Hadzi-Djokic, J., et al. Vesico-vaginal fistula: report of 220 cases. Int Urol Nephrol, 2009. 41: 299.
https://pubmed.ncbi.nlm.nih.gov/18810652
679. Forsgren, C., et al. Risk of pelvic organ fistula in patients undergoing hysterectomy. Curr Opin
Obstet Gynecol, 2010. 22: 404.
https://pubmed.ncbi.nlm.nih.gov/20739885
680. Likic, I.S., et al. Analysis of urologic complications after radical hysterectomy. Am J Obstet Gynecol,
2008. 199: 644 e1.
https://pubmed.ncbi.nlm.nih.gov/18722569
681. Hilton, P., et al. The risk of vesicovaginal and urethrovaginal fistula after hysterectomy performed
in the English National Health Service--a retrospective cohort study examining patterns of care
between 2000 and 2008. BJOG, 2012. 119: 1447.
https://pubmed.ncbi.nlm.nih.gov/22901248
682. Liedl, B., et al. Outcomes of a bladder preservation technique in female patients undergoing pelvic
exenteration surgery for advanced gynaecological tumours. Int Urogynecol J, 2014. 25: 953.
https://pubmed.ncbi.nlm.nih.gov/24633066
683. Kucera, H., et al. [Complications of postoperative radiotherapy in uterine cancer]. Geburtshilfe
Frauenheilkd, 1984. 44: 498.
https://pubmed.ncbi.nlm.nih.gov/6566638
684. Biewenga, P., et al. Can we predict vesicovaginal or rectovaginal fistula formation in patients with
stage IVA cervical cancer? Int J Gynecol Cancer, 2010. 20: 471.
https://pubmed.ncbi.nlm.nih.gov/20375815

138 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021
685. Adaji, S.E., et al. Vaginally inserted herbs causing vesico-vaginal fistula and vaginal stenosis. Int
Urogynecol J, 2013. 24: 1057.
https://pubmed.ncbi.nlm.nih.gov/22797463
686. Donaldson, J.F., et al. Obstructive uropathy and vesicovaginal fistula secondary to a retained sex toy
in the vagina. J Sex Med, 2014. 11: 2595.
https://pubmed.ncbi.nlm.nih.gov/24919434
687. Penrose, K.J., et al. Delayed vesicovaginal fistula after ring pessary usage. Int Urogynecol J, 2014.
25: 291.
https://pubmed.ncbi.nlm.nih.gov/23801483
688. Huang, L.K., et al. Evaluation of the extent of ketamine-induced uropathy: the role of CT urography.
Postgrad Med J, 2014. 90: 185.
https://pubmed.ncbi.nlm.nih.gov/24443558
689. Browning, A. Risk factors for developing residual urinary incontinence after obstetric fistula repair.
BJOG, 2006. 113: 482.
https://pubmed.ncbi.nlm.nih.gov/16489933
690. Goh, J.T. A new classification for female genital tract fistula. Aust N Z J Obstet Gynaecol, 2004. 44: 502.
https://pubmed.ncbi.nlm.nih.gov/15598284
691. Goh, J.T., et al. Predicting the risk of failure of closure of obstetric fistula and residual urinary
incontinence using a classification system. Int Urogynecol J Pelvic Floor Dysfunct, 2008. 19: 1659.
https://pubmed.ncbi.nlm.nih.gov/18690403
692. Beardmore-Gray, A., et al. Does the Goh classification predict the outcome of vesico-vaginal fistula
repair in the developed world? Int Urogynecol J, 2017. 28: 937.
https://pubmed.ncbi.nlm.nih.gov/27822888
693. Ostrzenski, A., et al. Bladder injury during laparoscopic surgery. Obstet Gynecol Surv, 1998. 53: 175.
https://pubmed.ncbi.nlm.nih.gov/9513988
694. Narayanan, P., et al. Fistulas in malignant gynecologic disease: etiology, imaging, and management.
Radiographics, 2009. 29: 1073.
https://pubmed.ncbi.nlm.nih.gov/19605657
695. Latzko, W. Postoperative vesicovaginal fistulas. Am J Surg, 1942. 58: 211.
https://www.sciencedirect.com/science/article/abs/pii/S0002961042900096
696. Rajaian, S., et al. Vesicovaginal fistula: Review and recent trends. Indian J Urol, 2019. 35: 250.
https://pubmed.ncbi.nlm.nih.gov/31619862
697. Yokoyama, M., et al. Successful management of vesicouterine fistula by luteinizing hormone-
releasing hormone analog. Int J Urol, 2006. 13: 457.
https://pubmed.ncbi.nlm.nih.gov/16734874
698. Goh, J.T., et al. Oestrogen therapy in the management of vesicovaginal fistula. Aust N Z J Obstet
Gynaecol, 2001. 41: 333.
https://pubmed.ncbi.nlm.nih.gov/11592553
699. Hemal, A.K., et al. Youssef’s syndrome: an appraisal of hormonal treatment. Urol Int, 1994. 52: 55.
https://pubmed.ncbi.nlm.nih.gov/8140684
700. Tarhan, F., et al. Minimal invasive treatment of vesicouterine fistula: a case report. Int Urol Nephrol,
2007. 39: 791.
https://pubmed.ncbi.nlm.nih.gov/17006733
701. Kumar, A., et al. Management of vesico-uterine fistulae: a report of six cases. Int J Gynaecol Obstet,
1988. 26: 453.
https://pubmed.ncbi.nlm.nih.gov/2900177
702. Rubino, S.M. Vesico-uterine fistula treated by amenorrhoea induced with contraceptive steroids.
Two case reports. Br J Obstet Gynaecol, 1980. 87: 343.
https://pubmed.ncbi.nlm.nih.gov/7426505
703. Jóźwik, M., et al. Spontaneous closure of vesicouterine fistula. Account for effective hormonal
treatment. Urol Int, 1999. 62: 183.
https://pubmed.ncbi.nlm.nih.gov/10529673
704. Safan, A., et al. Fibrin glue versus martius flap interpositioning in the repair of complicated obstetric
vesicovaginal fistula. A prospective multi-institution randomized trial. Neurourol Urodyn, 2009. 28: 438.
https://pubmed.ncbi.nlm.nih.gov/19475577
705. Garrido-Ruiz, M.C., et al. Vulvar pseudoverrucous papules and nodules secondary to a urethral--
vaginal fistula. Am J Dermatopathol, 2011. 33: 410.
https://pubmed.ncbi.nlm.nih.gov/21285858
706. Onuora, V.C., et al. Iatrogenic urogenital fistulae. Br J Urol, 1993. 71: 176.
https://pubmed.ncbi.nlm.nih.gov/8461950

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 139
707. Tsai, M.S., et al. Surgery is justified in patients with bowel obstruction due to radiation therapy.
J Gastrointest Surg, 2006. 10: 575.
https://pubmed.ncbi.nlm.nih.gov/16627224
708. Shackley, D.C., et al. The staged management of complex entero-urinary fistulae. BJU Int, 2000.
86: 624.
https://pubmed.ncbi.nlm.nih.gov/11069366
709. Eyre, R.C., et al. Management of urinary and bowel complications after ileal conduit diversion.
J Urol, 1982. 128: 1177.
https://pubmed.ncbi.nlm.nih.gov/7154168
710. Waaldijk, K., et al. The obstetric fistula and peroneal nerve injury: An analysis of 947 consecutive
patients. Int Urogynecol J, 1994. 5: 12.
https://link.springer.com/article/10.1007/BF00451704
711. Hilton, P. Vesico-vaginal fistulas in developing countries. Int J Gynaecol Obstet, 2003. 82: 285.
https://pubmed.ncbi.nlm.nih.gov/14499975
712. Niël-Weise, B.S., et al. Antibiotic policies for short-term catheter bladder drainage in adults.
Cochrane Database Syst Rev, 2005: CD005428.
https://pubmed.ncbi.nlm.nih.gov/16034973
713. Wall, L.L. Dr. George Hayward (1791-1863): a forgotten pioneer of reconstructive pelvic surgery. Int
Urogynecol J Pelvic Floor Dysfunct, 2005. 16: 330.
https://pubmed.ncbi.nlm.nih.gov/15976986
714. Hilton, P., et al. Epidemiological and surgical aspects of urogenital fistulae: a review of 25 years’
experience in southeast Nigeria. Int Urogynecol J Pelvic Floor Dysfunct, 1998. 9: 189.
https://pubmed.ncbi.nlm.nih.gov/9795822
715. Shaker, H., et al. Obstetric vesico-vaginal fistula repair: should we trim the fistula edges? A
randomized prospective study. Neurourol Urodyn, 2011. 30: 302.
https://pubmed.ncbi.nlm.nih.gov/21308748
716. Krause, S., et al. Surgery for urologic complications following radiotherapy for gynecologic cancer.
Scand J Urol Nephrol, 1987. 21: 115.
https://pubmed.ncbi.nlm.nih.gov/3616502
717. Langkilde, N.C., et al. Surgical repair of vesicovaginal fistulae--a ten-year retrospective study. Scand
J Urol Nephrol, 1999. 33: 100.
https://pubmed.ncbi.nlm.nih.gov/10360449
718. Lumen, N., et al. Review of the current management of lower urinary tract injuries by the EAU
Trauma Guidelines Panel. Eur Urol, 2015. 67: 925.
https://pubmed.ncbi.nlm.nih.gov/25576009
719. Brandes, S., et al. Diagnosis and management of ureteric injury: an evidence-based analysis. BJU
Int, 2004. 94: 277.
https://pubmed.ncbi.nlm.nih.gov/15291852
720. Morton, H.C., et al. Urethral injury associated with minimally invasive mid-urethral sling procedures
for the treatment of stress urinary incontinence: a case series and systematic literature search.
BJOG, 2009. 116: 1120.
https://pubmed.ncbi.nlm.nih.gov/19438488
721. Narang, V., et al. Ureteroscopy: savior to the gynecologist? Ureteroscopic management of post
laparoscopic-assisted vaginal hysterectomy ureterovaginal fistulas. J Minim Invasive Gynecol, 2007.
14: 345.
https://pubmed.ncbi.nlm.nih.gov/17478367
722. Poletajew, S., et al. Kidney removal: the past, presence, and perspectives: a historical review. Urol J,
2010. 7: 215.
https://pubmed.ncbi.nlm.nih.gov/21170847
723. Matz, M. [Prognosis-oriented diagnosis, treatment protocol and after care in uretero-vaginal fistula
following gynecological surgery from the urological viewpoint]. Zentralbl Gynakol, 1990. 112: 1345.
https://pubmed.ncbi.nlm.nih.gov/2278219
724. Benchekroun, A., et al. [Uretero-vaginal fistulas. 45 cases]. Ann Urol (Paris), 1998. 32: 295.
https://pubmed.ncbi.nlm.nih.gov/9827201
725. el Ouakdi, M., et al. [Uretero-vaginal fistula. Apropos of 30 cases]. J Gynecol Obstet Biol Reprod
(Paris), 1989. 18: 891.
https://pubmed.ncbi.nlm.nih.gov/2614029
726. Murphy, D.M., et al. Ureterovaginal fistula: a report of 12 cases and review of the literature. J Urol,
1982. 128: 924.
https://pubmed.ncbi.nlm.nih.gov/7176053

140 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021
727. Abou-El-Ghar, M.E., et al. Radiological diagnosis of vesicouterine fistula: role of magnetic resonance
imaging. J Magn Reson Imaging, 2012. 36: 438.
https://pubmed.ncbi.nlm.nih.gov/22535687
728. Quiroz, L.H., et al. Three-dimensional ultrasound imaging for diagnosis of urethrovaginal fistula. Int
Urogynecol J, 2010. 21: 1031.
https://pubmed.ncbi.nlm.nih.gov/20069418
729. Goodwin, W.E., et al. Vesicovaginal and ureterovaginal fistulas: a summary of 25 years of
experience. J Urol, 1980. 123: 370.
https://pubmed.ncbi.nlm.nih.gov/7359641
730. Pushkar, D.Y., et al. Management of urethrovaginal fistulas. Eur Urol, 2006. 50: 1000.
https://pubmed.ncbi.nlm.nih.gov/16945476
731. Pushkar, D. Editorial comment on: Transpubic access using pedicle tubularized labial urethroplasty
for the treatment of female urethral strictures associated with urethrovaginal fistulas secondary to
pelvic fracture. Eur Urol, 2009. 56: 200.
https://pubmed.ncbi.nlm.nih.gov/18468776
732. Xu, Y.M., et al. Transpubic access using pedicle tubularized labial urethroplasty for the treatment of
female urethral strictures associated with urethrovaginal fistulas secondary to pelvic fracture. Eur
Urol, 2009. 56: 193.
https://pubmed.ncbi.nlm.nih.gov/18468778
733. Huang, C.R., et al. The management of old urethral injury in young girls: analysis of 44 cases.
J Pediatr Surg, 2003. 38: 1329.
https://pubmed.ncbi.nlm.nih.gov/14523814
734. Candiani, P., et al. Repair of a recurrent urethrovaginal fistula with an island bulbocavernous
musculocutaneous flap. Plast Reconstr Surg, 1993. 92: 1393.
https://pubmed.ncbi.nlm.nih.gov/8248420
735. McKinney, D.E. Use of full thickness patch graft in urethrovaginal fistula. J Urol, 1979. 122: 416.
https://pubmed.ncbi.nlm.nih.gov/381691
736. Browning, A. Lack of value of the Martius fibrofatty graft in obstetric fistula repair. Int J Gynaecol
Obstet, 2006. 93: 33.
https://pubmed.ncbi.nlm.nih.gov/16530766
737. Baskin, D., et al. Martius repair in urethrovaginal defects. J Pediatr Surg, 2005. 40: 1489.
https://pubmed.ncbi.nlm.nih.gov/16150356
738. Atan, A., et al. Treatment of refractory urethrovaginal fistula using rectus abdominis muscle flap in a
six-year-old girl. Urology, 2007. 69: 384 e11.
https://pubmed.ncbi.nlm.nih.gov/17320687
739. Bruce, R.G., et al. Use of rectus abdominis muscle flap for the treatment of complex and refractory
urethrovaginal fistulas. J Urol, 2000. 163: 1212.
https://pubmed.ncbi.nlm.nih.gov/10737499
740. Koraitim, M. A new retropubic retrourethral approach for large vesico-urethrovaginal fistulas. J Urol,
1985. 134: 1122.
https://pubmed.ncbi.nlm.nih.gov/4057401
741. Rovner, E.S. Urethral diverticula: a review and an update. Neurourol Urodyn, 2007. 26: 972.
https://pubmed.ncbi.nlm.nih.gov/17654566
742. Crescenze, I.M., et al. Female Urethral Diverticulum: Current Diagnosis and Management. Curr Urol
Rep, 2015. 16: 71.
https://pubmed.ncbi.nlm.nih.gov/26267225
743. El-Nashar, S.A., et al. Incidence of female urethral diverticulum: a population-based analysis and
literature review. Int Urogynecol J, 2014. 25: 73.
https://pubmed.ncbi.nlm.nih.gov/23857063
744. Adelowo, A., et al. The role of preoperative urodynamics in urogynecologic procedures. J Minim
Invasive Gynecol, 2014. 21: 217.
https://pubmed.ncbi.nlm.nih.gov/24144925
745. Reeves, F.A., et al. Management of symptomatic urethral diverticula in women: a single-centre
experience. Eur Urol, 2014. 66: 164.
https://pubmed.ncbi.nlm.nih.gov/24636677
746. Athanasopoulos, A., et al. Urethral diverticulum: a new complication associated with tension-free
vaginal tape. Urol Int, 2008. 81: 480.
https://pubmed.ncbi.nlm.nih.gov/19077415

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 141
747. Hammad, F.T. TVT can also cause urethral diverticulum. Int Urogynecol J Pelvic Floor Dysfunct,
2007. 18: 467.
https://pubmed.ncbi.nlm.nih.gov/16821000
748. Mahdy, A., et al. Urethral diverticulum after tension-free vaginal tape procedure: case report.
Urology, 2008. 72: 461 e5.
https://pubmed.ncbi.nlm.nih.gov/18355901
749. Han, D.H., et al. Outcomes of surgery of female urethral diverticula classified using magnetic
resonance imaging. Eur Urol, 2007. 51: 1664.
https://pubmed.ncbi.nlm.nih.gov/17335961
750. Leach, G.E., et al. L N S C3: a proposed classification system for female urethral diverticula.
Neurourol Urodyn, 1993. 12: 523.
https://pubmed.ncbi.nlm.nih.gov/8312937
751. Cameron, A.P. Urethral diverticulum in the female: a meta-analysis of modern series. Minerva
Ginecol, 2016. 68: 186.
https://pubmed.ncbi.nlm.nih.gov/26545036
752. Blander, D.S., et al. Endoluminal magnetic resonance imaging in the evaluation of urethral diverticula
in women. Urology, 2001. 57: 660.
https://pubmed.ncbi.nlm.nih.gov/11306374
753. Pathi, S.D., et al. Utility of clinical parameters, cystourethroscopy, and magnetic resonance imaging
in the preoperative diagnosis of urethral diverticula. Int Urogynecol J, 2013. 24: 319.
https://pubmed.ncbi.nlm.nih.gov/22707007
754. Dwarkasing, R.S., et al. MRI evaluation of urethral diverticula and differential diagnosis in
symptomatic women. AJR Am J Roentgenol, 2011. 197: 676.
https://pubmed.ncbi.nlm.nih.gov/21862811
755. Chung, D.E., et al. Urethral diverticula in women: discrepancies between magnetic resonance
imaging and surgical findings. J Urol, 2010. 183: 2265.
https://pubmed.ncbi.nlm.nih.gov/20400161
756. Blander, D.S., et al. Images in clinical urology. Magnetic resonance imaging of a “saddle bag”
urethral diverticulum. Urology, 1999. 53: 818.
https://pubmed.ncbi.nlm.nih.gov/10197865
757. Khati, N.J., et al. MR imaging diagnosis of a urethral diverticulum. Radiographics, 1998. 18: 517.
https://pubmed.ncbi.nlm.nih.gov/9536494
758. Rovner, E.S., et al. Diagnosis and reconstruction of the dorsal or circumferential urethral
diverticulum. J Urol, 2003. 170: 82.
https://pubmed.ncbi.nlm.nih.gov/12796650
759. Chou, C.P., et al. Imaging of female urethral diverticulum: an update. Radiographics, 2008. 28: 1917.
https://pubmed.ncbi.nlm.nih.gov/19001648
760. Gugliotta, G., et al. Use of trans-labial ultrasound in the diagnosis of female urethral diverticula: A
diagnostic option to be strongly considered. J Obstet Gynaecol Res, 2015. 41: 1108.
https://pubmed.ncbi.nlm.nih.gov/25772163
761. Baradaran, N., et al. Female Urethral Diverticula in the Contemporary Era: Is the Classic Triad of the
“3Ds” Still Relevant? Urology, 2016. 94: 53.
https://pubmed.ncbi.nlm.nih.gov/27079128
762. Stav, K., et al. Urinary symptoms before and after female urethral diverticulectomy--can we predict
de novo stress urinary incontinence? J Urol, 2008. 180: 2088.
https://pubmed.ncbi.nlm.nih.gov/18804229
763. Ockrim, J.L., et al. A tertiary experience of urethral diverticulectomy: diagnosis, imaging and surgical
outcomes. BJU Int, 2009. 103: 1550.
https://pubmed.ncbi.nlm.nih.gov/19191783
764. Bhatia, N.N., et al. Urethral pressure profiles of women with urethral diverticula. Obstet Gynecol,
1981. 58: 375.
https://pubmed.ncbi.nlm.nih.gov/7196559
765. Reid, R.E., et al. Role of urodynamics in management of urethral diverticulum in females. Urology,
1986. 28: 342.
https://pubmed.ncbi.nlm.nih.gov/3094219
766. Summitt, R.L., Jr., et al. Urethral diverticula: evaluation by urethral pressure profilometry,
cystourethroscopy, and the voiding cystourethrogram. Obstet Gynecol, 1992. 80: 695.
https://pubmed.ncbi.nlm.nih.gov/1407897

142 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021
767. Bass, J.S., et al. Surgical treatment of concomitant urethral diverticulum and stress incontinence.
Urol Clin North Am, 1991. 18: 365.
https://pubmed.ncbi.nlm.nih.gov/1902010
768. Ganabathi, K., et al. Experience with the management of urethral diverticulum in 63 women. J Urol,
1994. 152: 1445.
https://pubmed.ncbi.nlm.nih.gov/7933181
769. Wagner, U., et al. [Significance of the urethral pressure profile in the diagnosis of urethral
diverticulum]. Geburtshilfe Frauenheilkd, 1986. 46: 456.
https://pubmed.ncbi.nlm.nih.gov/3093308
770. Thomas, A.A., et al. Urethral diverticula in 90 female patients: a study with emphasis on neoplastic
alterations. J Urol, 2008. 180: 2463.
https://pubmed.ncbi.nlm.nih.gov/18930487
771. Greiman, A.K., et al. Urethral diverticulum: A systematic review. Arab J Urol, 2019. 17: 49.
https://pubmed.ncbi.nlm.nih.gov/31258943
772. Moran, P.A., et al. Urethral diverticula in pregnancy. Aust N Z J Obstet Gynaecol, 1998. 38: 102.
https://pubmed.ncbi.nlm.nih.gov/9521406
773. Chang, Y.L., et al. Presentation of female urethral diverticulum is usually not typical. Urol Int, 2008.
80: 41.
https://pubmed.ncbi.nlm.nih.gov/18204232
774. Ljungqvist, L., et al. Female urethral diverticulum: 26-year followup of a large series. J Urol, 2007.
177: 219.
https://pubmed.ncbi.nlm.nih.gov/17162049
775. Enemchukwu, E., et al. Autologous Pubovaginal Sling for the Treatment of Concomitant Female
Urethral Diverticula and Stress Urinary Incontinence. Urology, 2015. 85: 1300.
https://pubmed.ncbi.nlm.nih.gov/26099875
776. Faerber, G.J. Urethral diverticulectomy and pubovaginal sling for simultaneous treatment of urethral
diverticulum and intrinsic sphincter deficiency. Tech Urol, 1998. 4: 192.
https://pubmed.ncbi.nlm.nih.gov/9892000
777. Romanzi, L.J., et al. Urethral diverticulum in women: diverse presentations resulting in diagnostic
delay and mismanagement. J Urol, 2000. 164: 428.
https://pubmed.ncbi.nlm.nih.gov/10893602
778. Swierzewski, S.J., 3rd, et al. Pubovaginal sling for treatment of female stress urinary incontinence
complicated by urethral diverticulum. J Urol, 1993. 149: 1012.
https://pubmed.ncbi.nlm.nih.gov/8483202
779. Dmochowski, R.R., et al. Update of AUA guideline on the surgical management of female stress
urinary incontinence. J Urol, 2010. 183: 1906.
https://pubmed.ncbi.nlm.nih.gov/20303102
780. Nickles, S.W., et al. Simple vs complex urethral diverticulum: presentation and outcomes. Urology,
2014. 84: 1516.
https://pubmed.ncbi.nlm.nih.gov/25432847
781. Bodner-Adler, B., et al. Surgical management of urethral diverticula in women: a systematic review.
Int Urogynecol J, 2016. 27: 993.
https://pubmed.ncbi.nlm.nih.gov/26564222
782. Ingber, M.S., et al. Surgically corrected urethral diverticula: long-term voiding dysfunction and
reoperation rates. Urology, 2011. 77: 65.
https://pubmed.ncbi.nlm.nih.gov/20800882
783. Rovner, E. Bladder and Female Urethral Diverticula. 2020.
https://abdominalkey.com/bladder-and-female-urethral-diverticula/
784. Lee, U.J., et al. Rate of de novo stress urinary incontinence after urethal diverticulum repair. Urology,
2008. 71: 849.
https://pubmed.ncbi.nlm.nih.gov/18355904
785. Laudano, M.A., et al. Pathologic Outcomes following Urethral Diverticulectomy in Women. Adv Urol,
2014. 2014: 861940.
https://pubmed.ncbi.nlm.nih.gov/24860605
786. Tsivian, M., et al. Female urethral diverticulum: a pathological insight. Int Urogynecol J Pelvic Floor
Dysfunct, 2009. 20: 957.
https://pubmed.ncbi.nlm.nih.gov/19582385
787. Thomas, R.B., et al. Adenocarcinoma in a female urethral diverticulum. Aust N Z J Surg, 1991.
61: 869.
https://pubmed.ncbi.nlm.nih.gov/1750825

MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021 143
6. CONFLICT OF INTEREST
All members of the Non-neurogenic Female LUTS Panel have provided disclosure statements of all
relationships that they have that might be perceived as a potential source of a conflict of interest. This
information is publically accessible through the European Association of Urology website: https://uroweb.org/
guideline/non-neurogenic-female-luts/.
This guidelines document was developed with the financial support of the European Association of
Urology. No external sources of funding and support have been involved. The EAU is a non-profit organisation
and funding is limited to administrative assistance and travel and meeting expenses. No honoraria or other
reimbursements have been provided.

7. CITATION INFORMATION
The format in which to cite the EAU Guidelines will vary depending on the style guide of the journal in which the
citation appears. Accordingly, the number of authors or whether, for instance, to include the publisher, location,
or an ISBN number may vary.

The compilation of the complete Guidelines should be referenced as:


EAU Guidelines. Edn. presented at the EAU Annual Congress Milan 2021. ISBN 978-94-92671-13-4

If a publisher and/or location is required, include:


EAU Guidelines Office, Arnhem, The Netherlands. http://uroweb.org/guidelines/compilations-of-all-guidelines/

References to individual guidelines should be structured in the following way:


Contributors’ names. Title of resource. Publication type. ISBN. Publisher and publisher location, year.

144 MANAGEMENT OF NON-NEUROGENIC FEMALE LOWER URINARY TRACT SYMPTOMS (LUTS) - 2021
EAU Guidelines on
Paediatric
Urology
C. Radmayr (Chair), G. Bogaert, H.S. Dogan,
J.M. Nijman (Vice-chair), Y.F.H. Rawashdeh, M.S. Silay,
R. Stein, S. Tekgül
Guidelines Associates: L.A. ‘t Hoen, J. Quaedackers, N. Bhatt

© European Association of Urology 2021


TABLE OF CONTENTS PAGE
1. INTRODUCTION 9
1.1 Aim 9
1.2 Panel composition 9
1.3 Available publications 9
1.4 Publication history 9
1.5 Summary of changes 9
1.5.1 New recommendations 10

2. METHODS 11
2.1 Introduction 11
2.2 Peer review 11

3. THE GUIDELINE 11
3.1 Phimosis 11
3.1.1 Epidemiology, aetiology and pathophysiology 12
3.1.2 Classification systems 12
3.1.3 Diagnostic evaluation 12
3.1.4 Management 12
3.1.5 Complications 13
3.1.6 Follow-up 13
3.1.7 Summary of evidence and recommendations for the management of phimosis 13
3.2 Management of undescended testes 13
3.2.1 Background 13
3.2.2 Classification 13
3.2.2.1 Palpable testes 14
3.2.2.2 Non-palpable testes 14
3.2.3 Diagnostic evaluation 15
3.2.3.1 History 15
3.2.3.2 Physical examination 15
3.2.3.3 Imaging studies 15
3.2.4 Management 15
3.2.4.1 Medical therapy 15
3.2.4.1.1 Medical therapy for testicular descent 15
3.2.4.1.2 Medical therapy for fertility potential 16
3.2.4.2 Surgical therapy 16
3.2.4.2.1 Palpable testes 16
3.2.4.2.1.1 Inguinal orchidopexy 16
3.2.4.2.1.2 Scrotal orchidopexy 17
3.2.4.2.2 Non-palpable testes 17
3.2.4.2.3 Complications of surgical therapy 17
3.2.4.2.4 Surgical therapy for undescended testes after puberty 17
3.2.5 Undescended testes and fertility 18
3.2.6 Undescended testes and malignancy 19
3.2.7 Summary of evidence and recommendations for the management of
undescended testes 19
3.3 Testicular Tumours in prepubertal boys 19
3.3.1 Introduction 19
3.3.2 Clinical presentation 20
3.3.3 Evaluation 20
3.3.4 Treatment/Management 20
3.3.5 Tumour entities in prepubertal boys 21
3.3.6 Follow-up 22
3.3.7 Congenital Adrenal Hyperplasia (CAH) 22
3.4 Hydrocele 22
3.4.1 Epidemiology, aetiology and pathophysiology 22
3.4.2 Diagnostic evaluation 23
3.4.3 Management 23
3.4.4 Summary of evidence and recommendations for the management of hydrocele 23

2 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


3.5 Acute scrotum 24
3.5.1 Epidemiology, aetiology and pathophysiology 24
3.5.2 Diagnostic evaluation 24
3.5.3 Management 25
3.5.3.1 Epididymitis 25
3.5.3.2 Testicular torsion 25
3.5.3.3 Surgical treatment 25
3.5.4 Follow-up 26
3.5.4.1 Fertility 26
3.5.4.2 Subfertility 26
3.5.4.3 Androgen levels 26
3.5.4.4 Unanswered questions 26
3.6 Hypospadias 26
3.6.1 Epidemiology, aetiology and pathophysiology 26
3.6.1.1 Epidemiology 26
3.6.2 Risk factors 27
3.6.3 Classification systems 27
3.6.4 Diagnostic evaluation 27
3.6.5 Management 27
3.6.5.1 Indication for reconstruction and therapeutic objectives 27
3.6.5.2 Pre-operative hormonal treatment 28
3.6.5.3 Age at surgery 28
3.6.5.4 Penile curvature 28
3.6.5.5 Urethral reconstruction 29
3.6.5.6 Re-do hypospadias repairs 29
3.6.5.7 Penile reconstruction following formation of the neo-urethra 30
3.6.5.8 Urine drainage and wound dressing 30
3.6.5.9 Outcome 30
3.6.6 Follow-up 31
3.6.7 Summary of evidence and recommendations for the management of
hypospadias 31
3.7 Congenital penile curvature 32
3.7.1 Epidemiology, aetiology and pathophysiology 32
3.7.2 Diagnostic evaluation 32
3.7.3 Management 32
3.7.4 Summary of evidence and recommendations for the management of congenital
penile curvature 33
3.8 Varicocele in children and adolescents 33
3.8.1 Epidemiology, aetiology and pathophysiology 33
3.8.2 Classification systems 33
3.8.3 Diagnostic evaluation 34
3.8.4 Management 34
3.8.5 Summary of evidence and recommendations for the management of varicocele 35
3.9 Urinary tract infections in children 36
3.9.1 Epidemiology, aetiology and pathophysiology 36
3.9.2 Classification systems 36
3.9.2.1 Classification according to site 36
3.9.2.2 Classification according to severity 36
3.9.2.3 Classification according to episode first/persistent/recurrent/
breakthrough 36
3.9.2.4 Classification according to symptoms 37
3.9.2.5 Classification according to complicating factors 37
3.9.3 Diagnostic evaluation 37
3.9.3.1 Medical history 37
3.9.3.2 Clinical signs and symptoms 37
3.9.3.3 Physical examination 37
3.9.3.4 Urine sampling, analysis and culture 37
3.9.3.4.1 Urine sampling 38
3.9.3.4.2 Urinalysis 38
3.9.3.4.3 Urine culture 39

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 3


3.9.3.5 Imaging 39
3.9.3.5.1 Ultrasound 39
3.9.3.5.2 Radionuclide scanning/MRI 40
3.9.3.5.3 Voiding cystourethrography / urosonography 40
3.9.4 Management 41
3.9.4.1 Administration route of antibacterial therapy 41
3.9.4.2 Duration of therapy 41
3.9.4.3 Antimicrobial agents 41
3.9.4.4 Preventative measures 42
3.9.4.4.1 Chemoprophylaxis 42
3.9.4.4.2 Dietary supplements 42
3.9.4.4.3 Preputium 42
3.9.4.4.4 Bladder and bowel dysfunction 42
3.9.4.5 Monitoring of UTI 43
3.9.5 Summary of evidence and recommendations for the management of UTI in
children 43
3.10 Day-time lower urinary tract conditions 44
3.10.1 Terminology, classification, epidemiology and pathophysiology 44
3.10.1.1 Filling-phase (storage) dysfunctions 44
3.10.1.2 Voiding-phase (emptying) dysfunctions 44
3.10.2 Diagnostic evaluation 45
3.10.3 Management 45
3.10.3.1 Specific interventions 46
3.10.4 Summary of evidence and recommendations for the management of day-time
lower urinary tract conditions 47
3.11 Monosymptomatic nocturnal enuresis - bedwetting 47
3.11.1 Epidemiology, aetiology and pathophysiology 47
3.11.2 Diagnostic evaluation 48
3.11.3 Management 48
3.11.3.1 Supportive treatment measures 48
3.11.3.2 Conservative “wait and see” approach 48
3.11.3.3 Nocturnal enuresis wetting alarm treatment 48
3.11.3.4 Medical therapy 49
3.11.4 Summary of evidence and recommendations for the management of
monosymptomatic enuresis 50
3.12 Management of neurogenic bladder 50
3.12.1 Epidemiology, aetiology and pathophysiology 50
3.12.2 Classification systems 51
3.12.3 Diagnostic evaluation 51
3.12.3.1 History and clinical evaluation 51
3.12.3.2 Laboratory and urinalysis 51
3.12.3.3 Ultrasound 51
3.12.3.4 Urodynamic studies/videourodynamic 51
3.12.3.4.1 Preparation before urodynamic studies 52
3.12.3.4.2 Uroflowmetry 52
3.12.3.5 Urodynamic studies 52
3.12.3.6 Voiding cystourethrogram 52
3.12.3.7 Renal scan 52
3.12.4 Management 52
3.12.4.1 Early management with intermittent catheterisation 53
3.12.4.2 Medical therapy 53
3.12.4.3 Management of faecal incontinence 54
3.12.4.4 Urinary tract infection 54
3.12.4.4.1 Urinary tract infection and clean intermittent catherisation 55
3.12.4.5 Sexuality 55
3.12.4.6 Bladder augmentation 55
3.12.4.7 Bladder outlet procedures 56
3.12.4.8 Catheterisable cutaneous channel. 57
3.12.4.9 Continent and incontinent cutaneous urinary diversion 57
3.12.5 Follow-up 57

4 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


3.12.6 Self-organisation of patients 57
3.12.7 Summary of evidence and recommendations for the management of neurogenic
bladder 62
3.13 Dilatation of the upper urinary tract (UPJ and UVJ obstruction) 63
3.13.1 Epidemiology, aetiology and pathophysiology 63
3.13.2 Diagnostic evaluation 63
3.13.2.1 Antenatal ultrasound 63
3.13.2.2 Postnatal ultrasound 63
3.13.2.3 Voiding cystourethrogram 63
3.13.2.4 Diuretic renography 63
3.13.3 Management 64
3.13.3.1 Prenatal management 64
3.13.3.1.1 Antibiotic prophylaxis for antenatal hydronephrosis 64
3.13.3.2 UPJ obstruction 65
3.13.3.3 Megaureter 65
3.13.3.3.1 Non-operative management 65
3.13.3.3.2 Surgical management 65
3.13.4 Conclusion 66
3.13.5 Summary of evidence and recommendations for the management of UPJ-,
UVJ-obstruction 66
3.14 Vesicoureteric reflux 66
3.14.1 Epidemiology, aetiology and pathophysiology 66
3.14.2 Diagnostic evaluation 67
3.14.2.1 Infants presenting with prenatally diagnosed hydronephrosis 68
3.14.2.2 Siblings and offspring of reflux patients 69
3.14.2.3 Recommendations for paediatric screening of VUR 69
3.14.2.4 Children with febrile urinary tract infections 69
3.14.2.5 Children with lower urinary tract symptoms and vesicoureteric reflux 69
3.14.3 Disease management 70
3.14.3.1 Non-surgical therapy 70
3.14.3.1.1 Follow-up 70
3.14.3.1.2 Continuous antibiotic prophylaxis 70
3.14.3.2 Surgical treatment 70
3.14.3.2.1 Subureteric injection of bulking materials 70
3.14.3.2.2 Open surgical techniques 71
3.14.3.2.3 Laparoscopy and robot-assisted 71
3.14.4 Summary of evidence and recommendations for the management of
vesicoureteric reflux in childhood 72
3.15 Urinary stone disease 74
3.15.1 Epidemiology, aetiology and pathophysiology 74
3.15.2 Classification systems 74
3.15.2.1 Calcium stones 74
3.15.2.2 Uric acid stones 75
3.15.2.3 Cystine stones 76
3.15.2.4 Infection stones (struvite stones) 76
3.15.3 Diagnostic evaluation 76
3.15.3.1 Imaging 76
3.15.3.2 Metabolic evaluation 77
3.15.4 Management 78
3.15.4.1 Extracorporeal shockwave lithotripsy 79
3.15.4.2 Percutaneous nephrolithotomy 80
3.15.4.3 Ureterorenoscopy 80
3.15.4.4 Open or laparoscopic stone surgery 81
3.15.5 Summary of evidence and recommendations for the management of urinary
stones 82
3.16 Obstructive pathology of renal duplication: ureterocele and ectopic ureter 83
3.16.1 Epidemiology, aetiology and pathophysiology 83
3.16.1.1 Ureterocele 83
3.16.1.2 Ectopic ureter 83
3.16.2 Classification systems 83

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 5


3.16.2.1 Ureterocele 83
3.16.2.1.1 Ectopic (extravesical) ureterocele 83
3.16.2.1.2 Orthotopic (intravesical) ureterocele 83
3.16.2.2 Ectopic ureter 83
3.16.3 Diagnostic evaluation 84
3.16.3.1 Ureterocele 84
3.16.3.2 Ectopic ureter 84
3.16.4 Management 84
3.16.4.1 Ureterocele 84
3.16.4.1.1 Early treatment 84
3.16.4.1.2 Re-evaluation 85
3.16.4.2 Ectopic ureter 85
3.16.5 Summary of evidence and recommendations for the management of
obstructive pathology of renal duplication: ureterocele and ectopic ureter 86
3.17 Disorders of sex development 86
3.17.1 Introduction 86
3.17.2 Current classification of DSD conditions 87
3.17.3 Diagnostic evaluation 88
3.17.3.1 The neonatal emergency 88
3.17.3.2 Family history and clinical examination 88
3.17.4 Gender assignment 89
3.17.5 Risk of tumour development 90
3.17.6 Recommendations for the management of disorders of sex development 90
3.18 Congenital lower urinary tract obstruction (CLUTO) 90
3.18.1 Posterior urethral valves 91
3.18.1.1 Epidemiology, aetiology and pathophysiology 91
3.18.2 Classification systems 91
3.18.2.1 Urethral valve 91
3.18.3 Diagnostic evaluation 91
3.18.4 Management 92
3.18.4.1 Antenatal treatment 92
3.18.4.2 Postnatal treatment 92
3.18.5 Follow-up 93
3.18.6 Summary 94
3.18.7 Summary of evidence and recommendations for the management of posterior
urethral valves 96
3.19 Rare Conditions in Childhood 96
3.19.1 Urachal remnants 96
3.19.1.1 Introduction 96
3.19.1.2 Epidemiology 97
3.19.1.3 Symptoms 97
3.19.1.4 Diagnosis 97
3.19.1.5 Treatment 97
3.19.1.6 Pathology of removed remnants 98
3.19.1.7 Urachal cancer 98
3.19.1.8 Conclusion 98
3.19.1.9 Recommendation for management of urachal remnants 98
3.19.2 Papillary tumours of the bladder in children and adolescents (Papillary urothelial
neoplasm of low malignant potential or transitional cell carcinoma) 99
3.19.2.1 Incidence 99
3.19.2.2 Differences and similarities of papillary tumours of the bladder in
children and adults 99
3.19.2.3 Risk factors 99
3.19.2.4 Presentation 99
3.19.2.5 Investigations and treatment 99
3.19.2.6 Histology 99
3.19.2.7 Additional treatment 99
3.19.3 Penile rare conditions 100
3.19.3.1 Cystic lesions 100
3.19.3.2 Vascular malformations 101

6 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


3.19.3.3 Neurogenic lesions 101
3.19.3.4 Soft tissue tumours of penis 101
3.19.3.5 Penile Lymphedema 102
3.20 Paediatric urological trauma 103
3.20.1 Paediatric renal trauma 103
3.20.1.1 Epidemiology, aetiology and pathophysiology 103
3.20.1.2 Classification systems 103
3.20.1.3 Diagnostic evaluation 103
3.20.1.3.1 Haematuria 103
3.20.1.3.2 Blood pressure 103
3.20.1.3.3 Choice of imaging method 104
3.20.1.4 Disease management 104
3.20.1.5 Recommendations for the diagnosis and management of paediatric
renal trauma 104
3.20.2 Paediatric ureteral trauma 104
3.20.2.1 Diagnostic evaluation 104
3.20.2.2 Management 105
3.20.2.3 Recommendations for the diagnosis and management of paediatric
ureteral trauma 105
3.20.3 Paediatric bladder injuries 105
3.20.3.1 Diagnostic evaluation 105
3.20.3.2 Management 105
3.20.3.2.1 Intraperitoneal injuries 106
3.20.3.2.2 Extraperitoneal injuries 106
3.20.3.3 Recommendations for the diagnosis and management of paediatric
bladder injuries 106
3.20.4 Paediatric urethral injuries 106
3.20.4.1 Diagnostic evaluation 106
3.20.4.2 Disease management 106
3.20.4.3 Recommendations for the diagnosis and management of paediatric
trauma 107
3.21 Peri-operative fluid management 107
3.21.1 Epidemiology, aetiology and pathophysiology 107
3.21.2 Disease management 107
3.21.2.1 Pre-operative fasting 107
3.21.2.2 Maintenance therapy and intra-operative fluid therapy 107
3.21.2.3 Peri-operative feeding and fluid management 108
3.21.3 Summary of evidence and recommendations for the management of
peri-operative fluid management 109
3.22 Post-operative pain management: general information 109
3.22.1 Epidemiology, aetiology and pathophysiology 109
3.22.2 Diagnostic evaluation 109
3.22.3 Disease management 109
3.22.3.1 Drugs and route of administration 109
3.22.3.2 Circumcision 110
3.22.3.2.1 Penile, inguinal and scrotal surgery 110
3.22.3.3 Bladder and kidney surgery 110
3.22.4 Summary of evidence and recommendations for the management of
post-operative pain 110
3.23 Basic principles of laparoscopic surgery in children 110
3.23.1 Epidemiology, aetiology and pathophysiology 110
3.23.2 Technical considerations and physiological consequences 111
3.23.2.1 Pre-operative evaluation 111
3.23.2.2 Abdominal insufflation 111
3.23.2.3 Pulmonary effects 111
3.23.2.4 Cardiovascular effects 112
3.23.2.5 Effects on renal function 112
3.23.2.6 Effects on neurological system 112
3.23.3 Summary of evidence and recommendations for laparoscopy in children 112

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 7


4. REFERENCES 113

5. CONFLICT OF INTEREST 181

6. CITATION INFORMATION 181

8 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


1. INTRODUCTION
1.1 Aim
The European Association of Urology (EAU) Paediatric Urology Guidelines Panel has prepared these Guidelines
with the aim of increasing the quality of care for children with urological conditions. This Guideline document is
limited to a number of common clinical pathologies in paediatric urological practice, as covering the entire field
of paediatric urology in a single guideline document is unattainable.
The majority of urological clinical problems in children are specialised and in many ways differ to
those in adults. This publication intends to outline a practical and preliminary approach to paediatric urological
conditions. Complex and rare conditions that require special care with experienced doctors should be referred
to designated centres where paediatric urology practice has been fully established and a multidisciplinary team
is available.
Over time, paediatric urology has developed and matured, establishing its diverse body of
knowledge and expertise and may now be ready to distinguish itself from its parent specialties. Thus,
paediatric urology has recently emerged in many European countries as a distinct subspecialty of both urology
and paediatric surgery and presents a unique challenge in the sense that it covers a large area with many
different schools of thought and a huge diversity in management.
Knowledge gained by increasing experience, new technological advances and non-invasive
diagnostic screening modalities has had a profound influence on treatment modalities in paediatric urology, a
trend that is likely to continue in the years to come.

It must be emphasised that clinical guidelines present the best evidence available to the experts but following
guideline recommendations will not necessarily result in the best outcome. Guidelines can never replace
clinical expertise when making treatment decisions for individual patients, but rather help to focus decisions
- also taking personal values and preferences/individual circumstances of children and their caregivers into
account. Guidelines are not mandates and do not purport to be a legal standard of care.

1.2 Panel composition


The EAU Paediatric Urology Guidelines Panel consists of an international group of clinicians with particular
expertise in this area. All experts involved in the production of this document have submitted potential
conflict of interest statements, which can be viewed on the EAU Website: http://uroweb.org/guideline/
paediatric-urology/.

1.3 Available publications


A quick reference document (Pocket guidelines) is available, both in print and as an app for iOS and Android
devices. These are abridged versions which may require consultation together with the full text version. A
number of translated versions, alongside several scientific publications are also available [1-7]. All documents
can be viewed through the EAU website: http://uroweb.org/guideline/paediatric-urology/.

1.4 Publication history


The Paediatric Urology Guidelines were first published in 2001 [8]. This 2021 publication includes a number of
updated chapters and sections as detailed below.

1.5 Summary of changes


The literature for the complete document has been assessed and updated, wherever relevant. Key changes in
the 2021 publication:

• Section 3.2 - Management of undescended testes: Both the literature and the text have been updated;
• Section 3.3 - Testicular Tumours in prepubertal boys: This is a new section in the Guideline;
• Section 3.5 - Acute Scrotum: Both the literature and the text have been updated;
• Section 3.6 - Hypospadias: Both the literature and the text have been updated;
• Section 3.7 - Congenital penile curvature: Both the literature and the text have been updated;
• Section 3.9 - Urinary Tract Infections: Both the literature and the text have been extensively updated;
• Section 3.10 - Day-time lower urinary tract conditions: The literature has been updated resulting in minor
amendments to the text;
• Section 3.13 - Dilatation of the upper urinary tract (UPJ and UVJ obstruction): The literature has been
updated resulting in minor amendments to the text:
• Section 3.15 - Urinary stone disease: The literature has been updated resulting in minor amendments to
the text;
• Section 3.18 - Congenital lower urinary tract obstruction (CLUTO): Both the literature and the text have
been updated;

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 9


• Section 3.19 – Rare Conditions - This is a new section in the Guideline comprising of urachal remnants,
papillary tumours of the bladder and penile rare conditions.

1.5.1 New recommendations

3.3 Recommendations for testicular tumours in prepubertal boys

Recommendations LE Strength rating


High-resolution ultrasound (7.5 – 12.5 MHz), preferably a doppler ultrasound, should 3 Strong
be performed to confirm the diagnosis.
Alpha-fetoprotein (AFP) should be determined in prepubertal boys with a testicular 2b Strong
tumour before surgery.
Surgical exploration should be done with the option for frozen section, but not as an 3 Strong
emergency operation.
Organ-preserving surgery should be performed in all benign tumours. 3 Strong
Staging (MRI abdomen /CT chest) should only be performed in patients with a 3 Strong
malignant tumour to exclude metastases.
MRI should only be performed in patients with the potential malignant Leydig or 4 Weak
Sertoli-cell-tumours to rule out lymph node enlargement.
Patients with a non-organ confined tumour should be referred to paediatric 4 Weak
oncologists post-operatively.

3.19 Recommendations for rare conditions in children


3.19.1 Urachal remnants

Recommendations Strength rating


Urachal remnants with no epithelial tissue carry little risk of malignant transformation. Strong
Asymptomatic and non-specific atretic urachal remnants can safely be managed non- Strong
operatively.
Urachal remnants (UR) incidentally identified during diagnostic imaging for non-specific Strong
symptoms should also be observed non-operatively since they tend to resolve
spontaneously.
A small urachal remnant, especially at birth, may be viewed as physiological. Strong
Urachal remnants in patients younger than 6 months are likely to resolve with non-operative Strong
management.
Follow-up is necessary only when symptomatic for 6 to 12 months. Strong
Surgical excision of urachal remnants solely as a preventive measure against later Strong
malignancy appears to have minimal support in the literature.
Only symptomatic URs should be safely removed by open or laparoscopic approach. Strong
A voiding cystourethrogram is only recommended when presenting with febrile UTIs. Strong

3.19.2 Papillary tumours of the bladder

Recommendations LE Strength rating


Ultrasound is the first investigation of choice for the diagnosis of paediatric bladder 3 Strong
tumours.
Cystoscopy should be reserved if a bladder tumour is suspected on imaging for 3 Strong
diagnosis and treatment.
After histological confirmation, inflammatory myofibroblastic bladder tumours should 4 Weak
be resected locally.
Follow-up should be every 3-6 months in the first year, and thereafter at least 4 Weak
annually with urinanalysis and an ultrasound for at least 5 years.

10 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


3.19.3 Penile conditions

Recommendations LE Strength rating


Treatment of penile cystic lesions is by total surgical excision, it is mainly indicated 4 Weak
for cosmetic or symptomatic (e.g. infection) reasons.
Propranolol is currently first line treatment for infantile haemangiomas. 2b Strong
Conservative management is the first-line treatment for penile lymphedema. 4 Weak
In symptomatic cases or in patients with functional impairment, surgical intervention 4 Weak
may become necessary for penile lymphedema.

2. METHODS
2.1 Introduction
These Guidelines were compiled based on current literature following a structured review. Databases covered by
the searches included Pubmed, Ovid, EMBASE and the Cochrane Central Register of Controlled Trials and the
Cochrane Database of Systematic Reviews. Application of a structured analysis of the literature was not possible
in many conditions due to a lack of well-designed studies. The limited availability of large randomised controlled
trials (RCTs) - influenced also by the fact that a considerable number of treatment options relate to surgical
interventions on a large spectrum of different congenital problems - means this document is largely a consensus
document. Clearly there is a need for continuous re-evaluation of the information presented in this document.

For each recommendation within the guidelines there is an accompanying online strength rating form, the
basis of which is a modified GRADE methodology [9, 10]. Each strength rating form addresses a number of key
elements namely:
1. the overall quality of the evidence which exists for the recommendation, references used in
this text are graded according to a classification system modified from the Oxford Centre for
Evidence-Based Medicine Levels of Evidence [11];
2. the magnitude of the effect (individual or combined effects);
3. the certainty of the results (precision, consistency, heterogeneity and other statistical or study
related factors);
4. the balance between desirable and undesirable outcomes;
5. the impact of patient values and preferences on the intervention;
6. the certainty of those patient values and preferences.

These key elements are the basis which panels use to define the strength rating of each recommendation. The
strength of each recommendation is represented by the words ‘strong’ or ‘weak’ [12]. The strength of each
recommendation is determined by the balance between desirable and undesirable consequences of alternative
management strategies, the quality of the evidence (including certainty of estimates), and nature and variability
of patient values and preferences. The strength rating forms will be available online.
Additional information can be found in the general Methodology section of this print, and online at
the EAU websit; http://www.uroweb.org/guideline/.

A list of Associations endorsing the EAU Guidelines can also be viewed online at the above address.

2.2 Peer review


All chapters of the Paediatric Urology Guidelines were peer-reviewed in 2015.

3. THE GUIDELINE
3.1 Phimosis
This chapter does not deal with neonatal circumcision as practised in the USA, nor mass circumcision as
practised in many African countries as part of a national program to prevent HIV. Also “religious and cultural”
circumcision is not discussed. At present in some European countries professional organisations do not

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 11


support circumcision for these reasons and it is no longer covered by insurance in these countries. Special
centres have been set up, where well-trained doctors perform circumcisions under sedation and local
anaesthetics at a lower cost. In all circumstances facilities have to comply with national regulations regarding
hygiene, special equipment, pain protocols and follow-up. Usually these clinics have an agreement with a
nearby hospital for the immediate treatment of possible complications. It is estimated that 37-39% of men
worldwide are circumcised [13].

3.1.1 Epidemiology, aetiology and pathophysiology


At the end of the first year of life, retraction of the foreskin behind the glandular sulcus is possible in
approximately 50% of boys; this increases to approximately 89% by the age of three years. The incidence of
phimosis is 8% in six to seven year olds and in just 1% in males aged sixteen to eighteen years [14].

3.1.2 Classification systems


Phimosis is either primary with no sign of scarring, or secondary (pathological) to a scarring such as balanitis
xerotica obliterans (BXO) [14]. Balanitis xerotica obliterans, also termed lichen sclerosis, has been recently
found in 35% of circumcised prepuce in children and adolescents and in 17% of boys younger than ten years
presenting with phimosis. The clinical appearance of BXO in children may be confusing and does not always
correlate with the final histopathological results. Lymphocyte-mediated chronic inflammatory disease was the
most common finding [15, 16] (LE: 2b).
Phimosis has to be distinguished from normal agglutination (adhesion) of the foreskin to the glans,
which is a more or less lasting physiological phenomenon with clearly-visible meatus and partial retraction
[17]. Separation of the prepuce from the glans is based on accumulated epithelial debris (smegma) and penile
erections. Forceful preputial retraction should be discouraged to avoid cicatrix formation [18].
Paraphimosis must be regarded as an emergency situation: retraction of a too narrow prepuce
behind the glans penis into the glanular sulcus may constrict the shaft and lead to oedema of the glans and
retracted foreskin. It interferes with perfusion distally from the constrictive ring and brings a risk of preputial
necrosis.

3.1.3 Diagnostic evaluation


The diagnosis of phimosis and paraphimosis is made by physical examination. If the prepuce is not retractable,
or only partly retractable, and shows a constrictive ring on drawing back over the glans penis, a disproportion
between the width of the foreskin and the diameter of the glans penis has to be assumed. In addition to the
constricted foreskin, there may be adhesions between the inner surface of the prepuce and the glanular
epithelium and/or a fraenulum breve. Paraphimosis is characterised by a retracted foreskin with the constrictive
ring localised at the level of the sulcus, which prevents replacement of the foreskin over the glans.

3.1.4 Management
Conservative treatment is an option for primary phimosis. The class 4 therapies were more effective over
placebo and manual stretching [19]. A corticoid ointment or cream (0.05-0.1%) can be administered twice a
day over a period of 4-8 weeks with a success rate of > 80% [20-23] (LE: 1b). A recurrence rate of up to 17%
can be expected [24]. This treatment has no side effects and the mean bloodcortisol levels are not significantly
different from an untreated group of patients [25] (LE: 1b). The hypothalamic pituitary-adrenal axis was not
influenced by local corticoid treatment [26]. Adhesion of the foreskin to the glans does not respond to steroid
treatment [20] (LE: 2).
Operative treatment of phimosis in children is dependent on the caregivers’ preferences and can be
plastic or radical circumcision after completion of the second year of life. Alternatively, the Shang Ring may be
used especially in developing countries [27]. Plastic circumcision has the objective of achieving a wide foreskin
circumference with full retractability, while the foreskin is preserved (dorsal incision, partial circumcision, trident
preputial plasty, combining 2 Z-plasties and a Y plasty) [28, 29]. However, this procedure carries the potential
for recurrence of the phimosis [30]. In the same session, adhesions are released and an associated fraenulum
breve is corrected by fraenulotomy. Meatoplasty is added if necessary. In all cases meticulous haemostasis is
mandatory and absorbable interrupted sutures are most often used.
An absolute indication for circumcision is secondary phimosis. In primary phimosis (including
those not responding to medical treatment), recurrent balanoposthitis and recurrent urinary tract infections
(UTIs) in patients with urinary tract abnormalities are indications for surgical intervention [31-34] (LE: 2b).
Male circumcision significantly reduces the bacterial colonisation of the glans penis with regard to both
non-uropathogenic and uropathogenic bacteria [35] (LE: 2b). Simple ballooning of the foreskin during
micturition is not a strict indication for circumcision.
Routine neonatal circumcision to prevent penile carcinoma is not indicated. A meta-analysis could
not find any risk in uncircumcised patients without a history of phimosis [36]. Contraindications for circumcision

12 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


are: an acute local infection and congenital anomalies of the penis, particularly hypospadias or buried penis,
as the foreskin may be required for a reconstructive procedure [37, 38]. Circumcision can be performed in
children with coagulopathy with 1-5% suffering complications (bleeding), if haemostatic agents or a diathermic
blade are used [39, 40]. Childhood circumcision has an appreciable morbidity and should not be recommended
without a medical reason and also taking into account epidemiological and social aspects [41-45] (LE: 1b).
Balanitis xerotica obliterans is associated with meatal pathology (stenosis) after circumcision in up to 20% of
boys and adjuvant local steroid treatment is advised [16, 46].
Treatment of paraphimosis consists of manual compression of the oedematous tissue with a
subsequent attempt to retract the tightened foreskin over the glans penis. Injection of hyaluronidase beneath
the narrow band or 20% mannitol may be helpful to release the foreskin [47, 48] (LE: 3-4). If this manoeuvre
fails, a dorsal incision of the constrictive ring is required. Depending on the local findings, a circumcision is
carried out immediately or can be performed in a second session.

3.1.5 Complications
Complications following circumcision vary and have been reported between 0-30% [45]. In a recent study
Hung et al. found during a 5-year follow-up period 2.9% complications in non-neonates of which 2.2% were early
(within 30 days after circumcision). Non-healing wounds, haemorrhage, wound infection, meatal stenosis, redundant
skin and non-satisfying cosmetic appearance as well as cicatrix formation and trapped penis all may occur [49].

3.1.6 Follow-up
Any surgery done on the prepuce requires an early follow-up of four to six weeks after surgery.

3.1.7 Summary of evidence and recommendations for the management of phimosis

Summary of evidence LE
Treatment for phimosis usually starts after two years of age or according to caregivers’ preference. 3
In primary phimosis, conservative treatment with a third generation corticoid ointment or cream is a 1b
first-line treatment with a success rate of more than 80%.

Recommendations LE Strength rating


Offer corticoid ointment or cream to treat primary symptomatic phimosis. 1b Strong
Circumcision will also solve the problem. 2b Strong
Treat primary phimosis in patients with recurrent urinary tract infection and/or with 2b Strong
urinary tract abnormalities.
Circumcise in case of lichen sclerosus or scarred phimosis. 2b Strong
Treat paraphimosis by manual reposition and proceed to surgery if it fails. 3 Strong
Avoid retraction of asymptomatic preputial adhesions. 2b Weak

3.2 Management of undescended testes


3.2.1 Background
Cryptorchidism or undescended testis is one of the most common congenital malformations of male neonates.
Incidence varies and depends on gestational age, affecting 1.0-4.6% of full-term and 1.1-45% of preterm
neonates. Following spontaneous descent within the first months of life, nearly 1.0% of all full-term male
infants still have undescended testes at one year of age [50]. This congenital malformation may affect both
sides in up to 30% of cases [51]. In newborn cases with non-palpable or undescended testes on both sides
and any sign of disorders of sex development (DSDs) like concomitant hypospadias, urgent endocrinological
and genetic evaluation is required [52].

3.2.2 Classification
The term cryptorchidism is most often used synonymously for undescended testes. The most useful
classification of undescended testes is distinguishing into palpable and non-palpable testes, and clinical
management is decided by the location and presence of the testes (see Figure 1). Approximately 80% of all
undescended testes are palpable [53]. Acquired undescended testes can be caused by entrapment after
herniorrhaphy or spontaneously referred to as ascending testis.
Palpable testes include true undescended testes and ectopic testes. Non-palpable testes include
intra-abdominal, inguinal, absent, and sometimes also some ectopic testes. Most importantly, the diagnosis of
palpable or non-palpable testis needs to be confirmed once the child is under general anaesthesia, as this is
the first step of any surgical procedure for undescended testes.

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 13


Figure 1: Classification of undescended testes

Undescended testis

Palpable Non-palpable

Intra-
Inguinal Inguinal Ectopic Absent
abdominal

Ectopic Agenesis

Vanishing
Retractile
testis

3.2.2.1 Palpable testes


Undescended testes
A true undescended testis is on its normal path of descent but is halted on its way down to the scrotum.
Depending on the location, the testes may be palpable or not, as in the case of testes arrested in the inguinal
canal.

Ectopic testes
If the position of a testis is outside its normal path of descent and outside the scrotum, the testis is considered
to be ectopic. The most common aberrant position is in the superficial inguinal pouch. Sometimes an ectopic
testis can be identified in a femoral, perineal, pubic, penile or even contralateral position. Usually, there is no
possibility for an ectopic testis to descend spontaneously to the correct position; therefore, it requires surgical
intervention. In addition, an ectopic testis might not be palpable due to its position.

Retractile testes
Retractile testes have completed their descent into a proper scrotal position but can be found again in a
suprascrotal position along the path of their normal descent. This is due to an overactive cremasteric reflex
[54]. Retractile testes can be easily manipulated down to the scrotum and remain there at least temporarily.

They are typically normal in size and consistency. However, they may not be normal and should be monitored
carefully since up to one-third can ascend and become undescended [55].

3.2.2.2 Non-palpable testes


Among the 20% of non-palpable testes, 50-60% are intra-abdominal, canalicular or peeping (right inside the
internal inguinal ring). The remaining 20% are absent and 30% are atrophic or rudimentary.

Intra-abdominal testes
Intra-abdominal testes can be located in different positions, with most of them being found close to the internal
inguinal ring. However, possible locations include the kidney, anterior abdominal wall, and retrovesical space. In
the case of an open internal inguinal ring, the testis may be peeping into the inguinal canal.

14 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


Absent testes
Monorchidism can be identified in up to 4% of boys with undescended testes, and anorchidism (bilateral
absence) in < 1%. Possible pathogenic mechanisms include testicular agenesis and atrophy after intrauterine
torsion with the latter one most probably due to an in utero infarction of a normal testis by gonadal vessel
torsion. The term “vanishing testis” is commonly used for this condition [56].

3.2.3 Diagnostic evaluation


History taking and physical examination are key in evaluating boys with undescended testes. Localisation
studies using different imaging modalities are usually without any additional benefit.

3.2.3.1 History
Caregivers should be asked for maternal and paternal risk factors, including hormonal exposure and genetic
or hormonal disorders. If the child has a history of previously descended testes this might be suggestive of
testicular ascent [57]. Prior inguinal surgery is indicative of secondary undescended testes due to entrapment.

3.2.3.2 Physical examination


An undescended testis is pursued by carefully advancing the examining fingers along the inguinal canal
towards the pubis region, perhaps with the help of lubricant. A possible inguinal testis can be felt to bounce
under the fingers [58]. A non-palpable testis in the supine position may become palpable once the child is in
a sitting or squatting position. If no testis can be identified along the normal path of descent, possible ectopic
locations must be considered.
In the event of unilateral non-palpable testis, the contralateral testis needs to be examined. Its
size and location can have important prognostic implications. Any compensatory hypertrophy suggests
testicular absence or atrophy [59]. Nevertheless, this does not preclude surgical exploration since the sign of
compensatory hypertrophy is not specific enough [60, 61].
In the event of bilateral undescended testes and any evidence or sign of DSDs, such as genital
ambiguity, or scrotal hyperpigmentation, further evaluation including endocrinological and genetic assessment
becomes mandatory [62].

3.2.3.3 Imaging studies


Imaging studies cannot determine with certainty that a testis is present or not [63]. Ultrasound (US) lacks the
diagnostic sensitivity to detect the testis confidently or establish the absence of an intra-abdominal testis
[64]. Consequently, the use of different imaging modalities, such as US or Magnetic resonance imaging (MRI)
[65], for undescended testes is limited and only recommended in specific and selected clinical scenarios
(e.g. identification of Müllerian structures in cases with suspicion of DSDs) [64].

3.2.4 Management
Treatment should be started at the age of six months. After that age, undescended testes rarely descend [66].
Any kind of treatment leading to a scrotally positioned testis should be finished by twelve months, or eighteen
months at the latest, because histological examination of undescended testes at that age has already revealed
a progressive loss of germ cells and Leydig cells [67]. The early timing of treatment is also driven by the final
adult results on spermatogenesis and hormone production, as well as on the risk of tumour development [68].

3.2.4.1 Medical therapy


Unfortunately, most of the studies on hormonal treatment have been of poor quality, with heterogeneous and
mixed patient populations, testis location, schedules and dosages of hormonal administration. Additionally,
long-term data are almost completely lacking.
Short-term side effects of hormonal treatment include increased scrotal erythema and pigmentation,
and induction of pubic hair and penile growth. Some boys experience pain after intramuscular injection of
human chorionic gonadotropin (hCG). All of these tend to regress after treatment cessation [69, 70].

3.2.4.1.1 Medical therapy for testicular descent


Hormonal therapy using hCG or gonadotropin-releasing hormone (GnRH) is based on the hormonal
dependence of testicular descent, but has a limited success rate of only 20% [71]. However, it must be taken
into account that almost 20% of these descended testes have the risk of re-ascending later [72]. In general,
success rates depend on testicular location. The higher the testis is located prior to therapy, the lower the
success rate, suggesting that testicular position is an important determinant of success [69]. Some authors
recommend combined hCG-GnRH treatment. Unfortunately, it is poorly documented and the treatment groups
were diverse. Some studies reported successful descent in up to 38% of non-responders to monotherapy [73].
The Panel consensus is that endocrine treatment to achieve testicular descent is not recommended (LE: 4).

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 15


Human chorionic gonadotropin
Human chorionic gonadotropin (hCG) stimulates endogenous testosterone production and is administered by
intramuscular injection. Several dose and administration schedules are reported. There is no proven difference
between 1.5 IU and weight-based doses up to 3.0 IU every other day for fourteen days [74]. Similar response
rates were achieved with 500 IU once weekly and 1.50 IU three times weekly [75]. However, there is evidence
that dosing frequency might affect testicular descent rates. Fewer lower dose injections per week for five
weeks seem to be superior to one higher dose every seven to ten days for three weeks with regard to testicular
descent [76].

Gonadotropin-releasing hormone
Gonadotropin-releasing hormone (GnRH) analogues (e.g., buserelin and gonadorelin) are available as nasal
sprays, thus avoiding painful intramuscular injections. A typical dosage regimen consists of 1.2 mg per day in
three divided doses, for four weeks. Success rates are wide ranging, from 9 to 60%, due to multiple treatment
strategies and heterogeneous patient populations [77].

3.2.4.1.2 Medical therapy for fertility potential


Hormonal treatment may improve fertility indices [77, 78] and therefore serve as an additional tool to
orchidopexy. There is no difference in treatment with GnRH before (neo-adjuvant) or after (adjuvant) surgical
orchidolysis and orchidopexy in terms of increasing fertility index, which may be a predictor for fertility later
in life [79]. It is still unknown whether this effect on testicular histology persists into adulthood but it has been
shown that men who were treated in childhood with buserelin had better semen analyses compared with men
who had childhood orchidopexy alone or placebo treatment [77].
It is reported that hCG treatment may be harmful to future spermatogenesis through increased
apoptosis of germ cells, including acute inflammatory changes in the testes and reduced testicular volume in
adulthood [80].
Identification of specific subgroups of boys with undescended testes who would benefit from such
an approach using hormones is difficult. Since these important data on specific groups as well as additional
support on the long-term effects are still lacking, the Nordic consensus does not recommend hormonal therapy
[81]. The consensus of the Panel is to recommend endocrine treatment with GnRH analogues in a dosage
described above for boys with bilateral undescended testes to preserve the fertility potential (LE: 4).

3.2.4.2 Surgical therapy


If a testis has not concluded its descent at the age of six months (corrected for gestational age), and since
spontaneous testicular descent is unlikely to occur after that age, surgery should be performed within
the subsequent year, and by age eighteen months at the latest [68]. In addition, early orchidopexy can be
followed by partial catch-up testicular growth, which is not the case in delayed surgery [79]. All these findings
recommend performing early orchidopexy between the ages of six and twelve months [66]. But despite early
and successful orchiopexy within the first year of life up to 25% of boys with nonsyndromic undescended
testes may be at risk for infertility based on hormonal and histological data, as a recently published series on
333 boys showed. This is especially true for bilateral cases, but in addition in about 5% of unilateral cases
reduced numbers of germ cells were detected in testicular biopsies as well [82].

3.2.4.2.1 Palpable testes


Surgery for palpable testes includes orchidofunicolysis and orchidopexy, either via an inguinal or scrotal
approach. The latter approach is mainly reserved for low-positioned, undescended testes, with the pros and
cons of each method being weighed against each other [83].

3.2.4.2.1.1 Inguinal orchidopexy


Inguinal orchidopexy is a widely used technique with a high success rate of up to 92% [84]. Important steps
include mobilisation of the testis and spermatic cord to the level of the internal inguinal ring, with dissection
and division of all cremasteric fibres, to prevent secondary retraction and detachment of the gubernaculum
testis. The patent processus vaginalis needs to be ligated proximally at the level of the internal ring, because
an unidentified or inadequately repaired patent processus vaginalis is an important factor leading to failure
of orchidopexy [85]. Any additional pathology has to be taken care of, such as removal of an appendix testis
(hydatid of Morgagni). At this moment the size of the testis can be measured and the connection of the
epididimis to the testis can be judged and described in the protocol. Some boys have a significant dissociation
between testis and epididymis which is prognostically bad for fertility. Finally, the mobilised testicle needs to
be placed in a sub-dartos pouch within the hemi-scrotum without any tension. If the length achieved using the
above-mentioned technique is still inadequate, the Prentiss manoeuvre, which consists of dividing the inferior
epigastric vessels and transposing the spermatic cord medially, in order to provide a straight course to the

16 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


scrotum, might be an option [86]. With regard to fixation sutures, if required, they should be made between the
tunica vaginalis and the dartos musculature [87]. Lymph drainage of a testis that has undergone surgery for
orchidopexy may have changed from high retroperitoneal drainage to iliac and inguinal drainage, which might
become important in the event of later malignancy [88].

3.2.4.2.1.2 Scrotal orchidopexy


Low-positioned, palpable undescended testis can be fixed through a scrotal incision including division of the
gubernaculum, and the processus vaginalis needs to be probed to check for patency [89]. Otherwise, fixation in
the scrotum is carried out correspondingly to the inguinal approach. In up to 20% of cases, an inguinal incision
will be compulsory to correct an associated inguinal hernia [90]. Any testicular or epididymal appendages
can be easily identified and removed. A systematic review shows that the overall success rates ranged from
88 to 100%, with rates of recurrence and post-operative testicular atrophy or hypotrophy < 1% [83]. Another
recently published systematic review and meta-analysis revealed similar outcome data regarding post-operative
complications, including wound infection, testicular atrophy, testicular reascent, and hernia for palpable low
positioned undescended testes. The only siginificant difference was the shorter operative time [91].

3.2.4.2.2 Non-palpable testes


For non-palpable testes, surgery must clearly determine whether a testis is present or not [92]. If a testis is
found, the decision has to be made to remove it or bring it down to the scrotum. An important step in surgery
is a thorough re-examination once the boy is under general anaesthesia, since a previously non-palpable
testis might be identifiable and subsequently change the surgical approach to standard inguinal orchidopexy,
as described above. Otherwise, the easiest and most accurate way to locate an intra-abdominal testis is
diagnostic laparoscopy [93]. Subsequent removal or orchidolysis and orchidopexy can be carried out using the
same approach to achieve the therapeutic aims [94]. Some tend to start with inguinal surgical exploration, with
possible laparoscopy during the procedure [95]. If an ipsilateral scrotal nubbin is suspected, and contralateral
compensatory testicular hypertrophy is present, a scrotal incision with removal of the nubbin, thus confirming
the vanishing testis, is an option avoiding the need for laparoscopy [96].
During laparoscopy for non-palpable testes, possible anatomical findings include spermatic vessels
entering the inguinal canal (40%), an intra-abdominal (40%) or peeping (10%) testis, or blind-ending spermatic
vessels confirming vanishing testis (10%) [97].
If there is a vanishing testis, the procedure is finished once blind-ending spermatic vessels
are clearly identified. If the vessels enter the inguinal canal, an atrophic testis may be found upon inguinal
exploration or a healthy testis that needs to undergo standard orchidopexy [98]. A peeping testis can be
placed down in the scrotum laparoscopically or via an inguinal incision [99]. Placement of an intra-abdominal
testis can sometimes be a surgical challenge. Usually, testes lying > 2 cm above the internal inguinal ring may
not reach the scrotum without division of the testicular vessels [100]. Under such circumstances, a Fowler-
Stephens orchidopexy may be an option [101] (see Figure 2).

Proximal cutting and transection of the testicular vessels, with conservation of the collateral arterial blood
supply, via the deferential artery and cremasteric vessels comprise the key features of the Fowler-Stephens
procedure. Recently, a modification with low spermatic vessel ligation has gained popularity, allowing blood
supply from the testicular artery to the deferential artery. An additional advantage is the position of the
peritoneal incision, leading to a longer structure, to ease later scrotal placement [102]. Due to the nature of
these approaches the testis is at risk of hypotrophy or atrophy if the collateral blood supply is insufficient [103].
The testicular survival rate in the one-stage Fowler-Stephens technique varies between 50 and 65% based
on post-operative Doppler-ultrasound findings [104]. For two-stage procedures success rates increase up
to 90% [105]. The advantages of two-stage orchidopexy, with the second part done usually six months after
the first, are to allow for development of collateral blood supply and to create greater testicular mobility [106].
In addition, preservation of the gubernaculum may also decrease the chance of testicular atrophy [107]. An
alternative might be microsurgical auto-transplantation, which has a success rate of up to 90%. However, this
approach requires skilled and experienced surgeons and is performed in a limited number of centres [108].

3.2.4.2.3 Complications of surgical therapy


Surgical complications are usually uncommon, with testicular atrophy being the most serious. A systematic
review revealed an overall atrophy rate for primary orchidopexy of 1.83%, 28.1% for one-stage Fowler-
Stephens procedure, and 8.2% for the two-stage approach [109]. Other rare complications comprise testicular
ascent and vas deferens injury besides local wound infection, dehiscence, and haematoma.

3.2.4.2.4 Surgical therapy for undescended testes after puberty


A study on 51 men diagnosed with inguinal unilateral undescended testis and a normal contralateral one, with

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 17


no history of any previous therapy, demonstrated a wide range of changes upon histological evaluation. Nearly
half of the study population still had significant germ cell activity at different maturation levels. Importantly, the
incidence of intratubular germ cell neoplasia was 2% [110].
The Panel consensus recommends orchiectomy in post-pubertal boys with an undescended testis
and a normal contralateral one in a scrotal position.

Figure 2: Treatment of unilateral non-palpable undescended testes

Unilateral non-palpable
testis

Re-exam under
anaesthesia

Still
Palpable
non-palpable

Inguinal
Diagnostic exploration Standard
laparoscopy with possible orchidopexy
laparoscopy

Blind ending Spermatic


Testis close to Testis too high
spermatic vessels enter
internal ring for orchidopexy
vessels inguinal ring

Laparoscopic or Staged Fowler- Vanishing testis


Inguinal
inguinal Stephens no further
exploration
orchidopexy procedure steps

3.2.5 Undescended testes and fertility


The association of undescended testes with compromised fertility [111] is extensively discussed in the literature
and seems to be a result of multiple factors, including germ cell loss, impaired germ cell maturation [112],
Leydig cell diminution and testicular fibrosis [113].
Although boys with one undescended testis have a lower fertility rate, they have the same paternity
rate as those with bilateral descended testes. Boys with bilateral undescended testes suffer both lower fertility
and paternity rates. Fertility rate is the number of offspring born per mating pair, individual or population,
whereas paternity reflects the actual potential of fatherhood [114]. The age at which surgical intervention for an
undescended testis occurs seems to be an important predictive factor for fertility later in life. Endocrinological
studies revealed higher inhibin-B and lower follicle-stimulating hormone (FSH) levels in men who underwent
orchidopexy at two years of age compared to individuals who had surgery later, which is indicative of a
benefit of earlier orchidopexy [115]. In addition, others demonstrated a relation between undescended testes
and increased loss of germ cells and Leydig cells, which is also suggestive of prompt orchidopexy being a
significant factor for fertility preservation [116]. Outcome studies for untreated bilateral undescended testes
revealed that 100% are oligospermic and 75% azoospermic. Among those successfully treated for bilateral
undescended testes, 75% still remain oligospermic and 42% azoospermic [113].

In summary, early surgical correction of undescended testes is highly recommended before twelve months of
age, and by eighteen months at the latest for preservation of fertility potential [67].

18 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


3.2.6 Undescended testes and malignancy
Boys who are treated for an undescended testis have an increased risk of developing testicular malignancy.
Screening and self-examination both during and after puberty is therefore recommended [117]. A Swedish
study, with a cohort of almost 17,000 men (56 developed a testicular tumour) who were treated surgically for
undescended testes and followed for 210,000 person-years, showed that management of undescended testes
before the onset of puberty decreased the risk of testicular cancer. The relative risk of testicular cancer among
those who underwent orchidopexy before thirteen years of age was 2.2 compared to the Swedish general
population; this increased to 5.4 for those treated after thirteen years of age [118].
A systematic review and meta-analysis of the literature have also concluded that pre-pubertal
orchidopexy may reduce the risk of testicular cancer and that early surgical intervention is indicated in boys
with undescended testes [119].

3.2.7 Summary of evidence and recommendations for the management of undescended testes

Summary of evidence LE
An undescended testis justifies treatment early in life to avoid loss of spermatogenic potential. 2a
A failed or delayed orchidopexy may increase the risk of testicular malignancy later in life. 2a
The earlier the treatment, the lower the risk of impaired fertility and testicular cancer. 2a
In unilateral undescended testis, fertility rate is reduced whereas paternity rate is not. 1b
In bilateral undescended testes, fertility and paternity rates are impaired. 1b
The treatment of choice for undescended testis is surgical replacement in the scrotum. 1b
The palpable testis is usually treated surgically using an inguinal approach. 2b
The non-palpable testis is most commonly approached laparoscopically. 2b
There is no consensus on the use of hormonal treatment. 2b

Recommendations LE Strength rating


Do not offer medical or surgical treatment for retractile testes instead undertake 2a Strong
close follow-up on a yearly basis until puberty.
Perform surgical orchidolysis and orchidopexy before the age of twelve months, and 2b Strong
by eighteen months at the latest.
Evaluate male neonates with bilateral non-palpable testes for possible disorders of 1b Strong
sex development.
Perform a diagnostic laparoscopy to locate an intra-abdominal testicle. 1a Strong
Hormonal therapy in unilateral undescended testes is of no benefit for future 2a Strong
paternity.
Offer endocrine treatment in case of bilateral undescended testes. 4 Weak
Inform the patient/caregivers about the increased risk of a later malignancy with an 3 Weak
undescended testis in a post-pubertal boy or older and discuss removal in case of a
contralateral normal testis in a scrotal position.

3.3 Testicular Tumours in prepubertal boys


3.3.1 Introduction
Testicular tumours account for approximately 1-2% of all paediatric solid tumours. [120]. Testicular tumours in
prepubertal boys differ in several aspects to testicular tumours in adolescent and adult men: they have a lower
incidence, they have a different histologic distribution (teratomas and yolk sac tumours are more common and
germ cell tumours are less common) and they are more often benign. A recent epidemiological study showed
that in children under the age of 15 years the incidence is highest in Asia (4.2 per million) and South America
(5 per million) and lowest in Europe (2.1 per million) and North America (2.5 per million). This is in contrast to
the incidence in adolescent and young adults where the highest incidence is in Europe (137.4 per million), and
North America (94.9 per million), while a lower incidence was observed in South and Central America (66.5 per
million) and Asia (27.1 per million) [121]. For age distribution in prepubertal boys, there is a small peak around
the age of two years [122]. Some recent studies demonstrated that up to 60-75% of the tumours are benign
[120, 123-127]. Intratubular neoplasia (TIN) is practically non-existent in children [128-131]. Testicular tumours
can generally be classified as germ cell or stromal tumours. One specific tumour type is the gonadoblastoma,
which contains germ cell and stromal cell tumour types and will occur almost exclusively in the setting of
disorders of sexual differentiation [132].

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 19


In the past 30 years, it has clearly been shown, that there is a fundamental difference between testicular
tumours in childhood and those in adulthood - not only in terms of the difference and incidence [121] but also
in terms of histology [128]. In prepubertal boys, most intratesticular tumours are benign, whereas post puberty
the tumours are most likely malignant.

3.3.2 Clinical presentation


Clinical presentation is a painless scrotal mass in more than 90% of the patients, detected by the caregiver,
physician or the patient himself. A history of a trauma, pain or hernia is rare. A hydrocele can be found in
15 – 50% [124, 133]. In boys with early onset of puberty (e.g. early penile and prepubic hair growth) as well as
high testosterone and low gonadotropin levels, a Leydig cell tumour should be excluded [134].

In patients presenting with a scrotal mass, paratesticular tumours should also be taken into account
as a differential diagnosis. However, these are even less common compared to intratesticular tumours.
The spectrum of paratesticular tumours includes benign tumours such as leiomyoma, fibroma, lipoma,
haemangioma, cystic lymphangioma and lipoblastoma as well as malignant tumours such as the paratesticular
rhabdomyosarcoma with an excellent prognosis and the rare melanotic neuroectodermal tumour of infancy
with a high recurrence rate [135-138]. As most of them are benign, intra-operative frozen section should
be available during surgery. An organ sparing surgical approach is preferred in benign tumours, whereas in
malignant tumour standard orchiectomy is carried out.

3.3.3 Evaluation
To confirm the diagnosis, a high-resolution ultrasound (US) examination (7.5 – 12.5 MHz), preferably a
doppler ultrasound, is required. The detection rate is almost 100% [139-142]. With high-resolution US,
microlithiasis - small hyperdense areas without sound shadows - is increasingly seen in prepubertal boys.
A recent meta-analysis showed that only 4 out of 296 boys (< 19 years of age diagnosed with microlithiasis)
developed a testicular tumour of whom two previously had a testicular tumour on the opposite or ipsilateral site
[143]. If microlithiasis shows up in patients with additional risk factors for testicular tumour, then the caregivers/
patients should be informed about the increased risk and encouraged to carry out regular self-examinations -
similar to patients treated for undescended testis [144]. There is no evidence, that regular sonographic follow-
up is useful [143]. The risk for infertility may be higher in patients with microlithiasis and if these patients have
any sign of infertility later, the risk of developing a tumour seems to be higher compared to patients without
microlithiasis and infertility [145]. Due to the low incidence of a contralateral tumour, even in cases of testicular
microlithiasis, there is no indication for contralateral testicular biopsy in prepubertal boys.

Age should be taken into account, when tumour markers are used. Human chorionic gonadotropin (ß-hCG) is
derived from chorion carcinoma, embryonal carcinoma or seminoma. However, these tumours are extremely
rare in prepubertal boys and therefore ß-hCG is not useful in prepubertal boys. Alpha-fetoprotein (AFP) has
a clear limitation of its sensitivity and specificity in the first months of life [133] and sometimes takes up to
12 months before the serum concentration reaches the known standard values (< 10 ng/mL) [127, 146]. It is
produced by > 90% of yolk sac tumours. Teratomas can also produce AFP, but not to that extent of yolk sac
tumours [147]. Alpha-fetoprotein should be measured before any therapeutic intervention (tumour enucleation/
orchiectomy) and ideally should be available at the time of the procedure. Alpha-fetoprotein has a serum
biological half-life of 5 days and should be measured 5 days after tumour resection/orchiectomy in those with
an elevated AFP. There is no urgent need for pre-operative staging, as this has no consequence before the
definitive histology is available.

3.3.4 Treatment/Management
If a testicular tumour is suspected, surgery with the option of intra-operative frozen section should be
performed. It is not necessary to do this as an emergency procedure. However, in order to confirm the
diagnosis and to avoid familial anxiety, the operation should be scheduled as soon as possible, preferably
within the next few days. Organ-preserving surgery should be performed, whenever possible. A recent
published review article showed that out of 227 patients with organ-sparing surgery only two cases (one in a
patient with an epidermoid cyst and one in a patient with a mature teratoma) had a recurrence [148-150].

Orchiectomy could be considered only if normal testicular parenchyma is no longer detectable in the pre-
operatively high-resolution ultrasound and/or the AFP is > 100 ng/mL in a > 12-month-old boy: highly
suspicious of a yolk sac tumour.

For surgical technique, the Panel is in favour of an inguinal approach. Furthermore, clamping of the vessels has
the advantage of a better view, when organ sparing surgery is performed. However, there is no evidence in the
literature, that tumour-spread is prevented by clamping the vessels. Whenever possible, testis sparing surgery

20 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


should be performed along with frozen sections during surgery to confirm the diagnosis (begin vs. malignant
tumour) and to confirm if a microscopically margin-negative resection is performed, in which no gross or
microscopic tumour remains in the primary tumour bed (R0 resection). In cases of an R0 resection, the tunica
is closed and the testis is replaced in the scrotum. In case of R1 resection (removal of all macroscopic disease,
but microscopic margins are positive for tumour) confirmed by frozen section in a malignant or potential
malignant tumour, an orchiectomy should be performed at the same time of surgery. If the final pathology later
demonstrates a R1 resection in a malignant tumour despite intra-operative negative margins on frozen section,
an inguinal orchiectomy can safely be performed.

In patients with a malignant tumour (yolk sac tumour, immature teratoma) staging should be performed
including an MRI of the abdomen and a CT-scan from the chest. If there is any suspicion of a non-organ
confined tumour, the patient should be referred to a paediatric oncologist. In patients with the rare diagnosis
of a Granulosa cell tumour, imaging of the abdomen to exclude enlarged lymph nodes is reasonable as this
may be a potentially malignant tumour; in those with Sertoli or a Leydig cell tumour, an MRI is recommended,
as 10% are malignant and the metastases do not respond very well to chemotherapy or radiation in the adult
literature [151, 152]. The TNM classification from 2015 for adult testicular tumours can be used in patients with
a malignant tumour [153]. In benign tumours (mature teratoma, epidermoid cysts) no further staging is required.

3.3.5 Tumour entities in prepubertal boys


Teratomas are usually benign in prepubertal children and represent the greatest proportion of intratesticular
tumours (around 40%) [120, 154]. They present at a median age of 13 months (0-18 months). Only in
adolescent and adults, they should be considered as malignant tumours. Histologically they can consist of a
combination of the three primitive embryological germ-cell layers (ectoderm, mesoderm and endoderm). Most
of these elements shows microscopically mature elements [155]; however, some immature teratomas in this
age group have also been reported [156]. To exclude any malignant potential, like focal areas of a yolk-sac
tumour, the entire specimen should be investigated. On US examination a heterogenous picture with some
calcification is seen [157] and AFP should be less than 100 ng/mL in an infant. After organ-sparing surgery only
one recurrence was reported in the literature [150].

Epidermoid cysts are of ectodermal origin and seem to be related to well-differentiated teratomas; they are
always benign [155]. Keratin-producing epithelium is responsible for the keratinised-squamous-epithelial
deposits, which appear hyperechogenic in an US [157]. Organ-sparing surgery should be performed and
if confirmed by histology, there is no need for surveillance despite the fact that one “recurrence” has been
reported 13 years after diagnosis [149].

Juvenile granulosa cell tumours occur usually in the first year of life, typically within the first 6 months [158]. They are
well circumscribed and have a typical yellow-tan appearance; 2/3 have cystic elements, 1/3 solid [158]. The stroma
can be fibrous or fibromyxoid. So far, no recurrence has been reported after organ-sparing surgery [158, 159].

Leydig cell tumours arising from the testosterone producing Leydig cells should be suspected in boys with
early onset of puberty with high testosterone and low gonadotropin levels [134]. Patients are usually between
5 and 10 years of age; the tumours are well circumscribed with yellow-brown nodules. In children there are no
reports of malignant Leydig cell tumours and after organ sparing surgery, there are no reported recurrences to
date [160, 161]. In the adult literature, there is a malignancy rate of 10% reported and primary retroperitoneal
lymphadenectomy should be discussed in cases with enlarged lymph nodes, as these metastases do not
respond very well to chemotherapy or radiation [162].

Around 1/5 of the Sertoli-cell tumours occur in children; usually within the first year of life [163]. In the
paediatric age group, the large-cell calcifying Sertoli cell tumours (LCCSCT) are the most common tumour
variant [164, 165]. They can occur in patients with complex dysplastic syndromes, such as the Carney or
Peutz-Jeghers syndrome [165-167]. Except one case report with the histological diagnosis of a malignant
LCCSCT [164], all other reported tumours are benign, therefore organ-sparing surgery should be performed.

Yolk sac tumours are the predominant prepubertal malignant germ cell tumours and may represent around
15% of the prepubertal tumours in boys [120]. They also have a number of other names: endodermal sinus
tumours, juvenile embryonal carcinoma, clear cell carcinoma, orchioblastoma, vitellineum, archenteronoma
and sometimes extraembryonal mesoblastoma [168]. They are histologically mostly solid, yellow-grey tumours.
They occur usually within the first two years of life [169]. Up to 80-85% of the tumours are organ confined
(Stage I) [170]. The tumour usually spreads haematogenously (chest). Twenty percent of those with Stage
I disease may develop visible metastasis in 20% within the next 2 years. In a German study, 14 out of 91
patients with Stage I had a recurrence after observation – all were cured by chemotherapy alone. Four out of

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 21


five with metastatic disease initially, were cured by chemotherapy after radical orchiectomy [171]. In a recent
published series from China, 21 out of 90 paediatric patients with a Stage I yolk sac tumour received primary
chemotherapy. One of the 21 had a recurrence, whereas 29 out of 69 who underwent surveillance after initial
orchiectomy had a recurrence. The overall 4-year survival rate was 97.8% [169], almost the same recurrence
rate has also been reported by American oncology groups [172, 173]. Therefore in patients with Stage I disease
(no metastatic disease in the MRI-abdomen and CT scan of the chest as well as normal age-adapted AFP-
values) close follow-up together with the paediatric oncologists including AFP every 2-3 months and MRI of
the abdomen is recommended, at least for the first 2-3 years [133]. This is especially recommended in those
with invasions of the lymphatic vessels, as this has been shown to be a prognostic factor in one of the recent
series [169]. In cases of recurrence, chemotherapy should be performed by paediatric oncologists according to
national study protocols.

3.3.6 Follow-up
Regular US examination is recommended in the follow-up period to detect any recurrence and/or other
abnormalities. As there are only a few studies with recurrence after testicular sparing surgery or orchiectomy,
no clear recommendation can be made concerning the interval and the duration of follow-up. However, doing
an US examination every 3-6 months within the first year seems reasonable, as few recurrences have been
detected at this time and the rate of atrophy is extremely low after organ-sparing surgery [148]. Only in patients
with a malignant tumour, regular follow-up examination after the first year of surgery seems reasonable (see
above). The follow-up in patients with a Leydig cell tumour should include endocrinological examinations.
Using the SEER data base, the 5-year relative survival for testicular malignancies for patients < 14 years of age
diagnosed with localised testicular cancer was 97.4%, and for those with distant disease 72.6% [174].

3.3.7 Congenital Adrenal Hyperplasia


Boys with a congenital adrenal hyperplasia (CAH) represent a special group. Up to a third of the patients have
so-called testicular adrenal rest tumours (TARTs) This proportion increase with age [175, 176]. It is most likely to
be ectopic adrenal cells, which are growing under pathological stimulation from Adrenocorticotropic Hormone
(ACTH) [177]. They have no malignant potential, but they can have a lasting impact on fertility by displacing
the normal testicular parenchyma [177, 178]. These patients should be offered US screening and advice on
fertility with the option of cryopreservation [178]. As far as is known, no malignant tumour has been described
in patients with a typical TART. As a result, the indication for surgical intervention in these patients to rule out a
malignant tumour should be offered very cautiously.

Summary of evidence LE
Testicular tumours in prepubertal boys have a lower incidence and a different histologic distribution 2a
compared to the adolescent and adult patients.
In prepupertal boys up to 60-75% of testicular tumours are benign. 3

Recommendations LE Strength rating


High-resolution ultrasound (7.5 – 12.5 MHz), preferably a doppler ultrasound, should 3 Strong
be performed to confirm the diagnosis.
Alpha-fetoprotein should be determined in prepubertal boys with a testicular tumour 2b Strong
before surgery.
Surgical exploration should be done with the option for frozen section, but not as an 3 Strong
emergency operation.
Organ-preserving surgery should be performed in all benign tumours. 3 Strong
Staging (MRI abdomen/CT chest) should only be performed in patients with a 3 Strong
malignant tumour to exclude metastases.
Magnetic resonance imaging should only be performed in patients with the potential 4 Weak
malignant Leydig or Sertoli-cell-tumours to rule out lymph node enlargement.
Patients with a non-organ confined tumour should be referred to paediatric 4 Weak
oncologists post-operatively.

3.4 Hydrocele
3.4.1 Epidemiology, aetiology and pathophysiology
Hydrocele is defined as a collection of fluid between the parietal and visceral layers of the tunica vaginalis [179].
Pathogenesis of primary hydrocele is based on patency of the processus vaginalis in contrast with secondary
hydrocele. Incomplete obliteration of the processus vaginalis peritonei results in formation of various types of

22 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


communicating hydrocele; a large open processus vaginalis allowing passage of abdominal viscera results in
clinical hernia [180]. The exact time of spontaneous closure of the processus vaginalis is not known. It persists
in approximately 80-94% of newborns and in 20% of adults [181]. If complete obliteration of the processus
vaginalis occurs with patency of mid-portion, a hydrocele of the cord occurs. Scrotal hydroceles without
associated patency of the processus vaginalis are also encountered in newborns [182]. Non-communicating
hydroceles, based on an imbalance between the secretion and re-absorption of this fluid, are found secondary
to minor trauma, testicular torsion, epididymitis, varicocele operation (due to ligation of the lymphatics) or may
appear as a recurrence after primary repair of a communicating or non-communicating hydrocele.

3.4.2 Diagnostic evaluation


The classic description of a communicating hydrocele is that of a hydrocele that fluctuates in size,
and is usually related to ambulation. It may be diagnosed by history-taking and physical investigation.
Transillumination of the scrotum provides the diagnosis in the majority of cases, keeping in mind that fluid
filled intestine and some pre-pubertal tumours may transilluminate as well [183, 184]. If the diagnosis is that
of a hydrocele, there will be no history of reducibility and no associated symptoms; the swelling is translucent,
smooth and usually not tender. If there are any doubts about the character of an intrascrotal mass, scrotal US
should be performed and has nearly 100% sensitivity in detecting intrascrotal lesions. Doppler US studies
help to distinguish hydroceles from varicocele and testicular torsion, although these conditions may also be
accompanied by a hydrocele.

3.4.3 Management
In the majority of infants, surgical treatment of hydrocele is not indicated within the first twelve months because
of the tendency for spontaneous resolution [185] (LE: 2). Little risk is taken by initial observation as progression
to hernia is rare and does not result in incarceration [185]. Early surgery is indicated if there is suspicion of a
concomitant inguinal hernia or underlying testicular pathology [186, 187] (LE: 2). Persistence of a simple scrotal
hydrocele beyond twelve months of age may be an indication for surgical correction. There is no evidence that
this type of hydrocele risks testicular damage. The natural history of hydrocele is poorly documented beyond
the age of two years and according to a systematic review there is no good evidence to support current
practice. Delaying surgery may reduce the number of procedures necessary without increasing morbidity [188].
The question of contralateral disease should be addressed by both history-taking and physical
examination at the time of initial consultation [189] (LE: 2). In late-onset hydrocele, suggestive of a non-
communicating hydrocele, there is a reasonable chance of spontaneous resolution (75%) and expectant
management of six to nine months is recommended [190]. In the paediatric age group, the operation consists
of ligation of the patent processus vaginalis or scrotal via inguinal incision and the distal stump is left open,
whereas in hydrocele of the cord the cystic mass is excised or unroofed [184, 186, 191, 192] (LE: 4). In expert
hands, the incidence of testicular damage during hydrocele or inguinal hernia repair is very low (0.3%) (LE: 3).
Laparoscopic hernia repair with percutaneous ligation of the patent processes vaginalis is a minimally invasive
alternative to open inguinal herniorrhaphy [193, 194]. Sclerosing agents should not be used because of the risk
of chemical peritonitis in communicating processus vaginalis peritonei [184, 186] (LE: 4). The scrotal approach
(Lord or Jaboulay technique) is used in the treatment of a secondary non-communicating hydrocele.

3.4.4 Summary of evidence and recommendations for the management of hydrocele

Summary of evidence LE
In the majority of infants, surgical treatment of hydrocele is not indicated within the first twelve months 2a
due to the tendency for spontaneous resolution. Little risk is taken by initial observation as progression
to hernia is rare.
In the paediatric age group, an operation would generally involve ligation of the patent processus 4
vaginalis via inguinal incision.

Recommendations LE Strength rating


In the majority of infants, observe hydrocele for twelve months prior to considering 2a Strong
surgical treatment.
Perform early surgery if there is suspicion of a concomitant inguinal hernia or 2b Strong
underlying testicular pathology.
Perform a scrotal ultrasound in case of doubt about the character of an intrascrotal 4 Strong
mass.
Do not use sclerosing agents because of the risk for chemical peritonitis. 4 Strong

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 23


3.5 Acute scrotum
3.5.1 Epidemiology, aetiology and pathophysiology
Acute scrotum is a paediatric urological emergency, most commonly caused by torsion of the testis or
appendix testis, or epididymitis/epididymo-orchitis [195-200]. Other causes of acute scrotal pain are idiopathic
scrotal oedema, mumps orchitis, varicocele, scrotal haematoma, incarcerated hernia, appendicitis or systemic
disease (e.g. Henoch-Schönlein purpura) [201-213]. Trauma can also be a cause of acute scrotum due to post-
traumatic haematomas, testicular contusion, rupture, dislocation or torsion [214-219]. Scrotal fat necrosis has
also been reported to be an uncommon cause of mild-to-moderate scrotal pain in pre-pubertal overweight
boys after exposure to cold [220].
In this chapter testicular torsion and epididymitis are discussed, while recurrent epididymitis is
discussed in the chapter dealing with infections. Torsion of the testis occurs most often in the neonatal period
and around puberty, whereas torsion of the appendix testis occurs over a wider age range [221]. Epididymitis
affects two age groups: less than one year and twelve to fifteen years [222, 223]. One study predicted the
annual incidence of epididymitis around 1.2 per 1,000 children [224].
Perinatal torsion of the testis most often occurs prenatally. Bilateral torsion comprises 11-21% of
all perinatal cases [225]. Most cases of perinatal torsion are extravaginal, in contrast to the usual intravaginal
torsion which occurs during puberty.

3.5.2 Diagnostic evaluation


Patients usually present with scrotal pain, except in neonatal torsion. The sudden onset of invalidating pain in
combination with vomiting is typical for torsion of the testis or appendix testis [226, 227].
In general, the duration of symptoms at presentation is shorter in testicular torsion (69% present
within twelve hours) and torsion of the appendix testis (62%) compared to epididymitis (31%) [197, 198, 223].
Prepubertal males are more likely to present with atypical symptoms and delayed presentation and diagnosis,
leading to delayed surgical intervention and a higher rate of orchiectomy, compared to postpubertal boys [228].

In the early phase, location of the pain can lead to diagnosis. Patients with acute epididymitis experience a
tender epididymis, whereas patients with testicular torsion are more likely to have a tender testicle, in case of
torsion of the appendix testis there may be isolated tenderness of the superior pole of the testis [223].
An abnormal (horizontal) position of the testis is more frequent in testicular torsion than
epididymitis [197]. Looking for absence of the cremasteric reflex is a simple method with 100% sensitivity
and 66% specificity for testicular torsion [222, 227] (LE: 3). Elevation of the scrotum may reduce complaints in
epididymitis, but not in testicular torsion.
Fever occurs more often in epididymitis (11-19%). The classical sign of a “blue dot” was found
only in 10-23% of patients with torsion of the appendix testis [196, 197, 222, 229]. In many cases, it is not
easy to determine the cause of acute scrotum based on history and physical examination alone [195-200, 222,
229]. A positive urine culture is only found in a few patients with epididymitis [199, 222, 229, 230]. It should be
remembered that a normal urinalysis does not exclude epididymitis. Similarly, an abnormal urinalysis does not
exclude testicular torsion.
Doppler US is useful to evaluate acute scrotum, with 63.6-100% sensitivity and 97-100% specificity,
a positive predictive value of 100% and negative predictive value of 97.5% [231-236] (LE: 3). The use of
Doppler US may reduce the number of patients with acute scrotum undergoing scrotal exploration, but it is
operator-dependent and can be difficult to perform in pre-pubertal patients [233, 237]. It may also show a
misleading arterial flow in the early phases of torsion and in partial or intermittent torsion. Of key importance,
persistent arterial flow does not exclude testicular torsion. In a multicentre study of 208 boys with torsion of the
testis, 24% had normal or increased testicular vascularisation [233]. A comparison with the other side should
always be done.

Better results were reported using high-resolution US (HRUS) for direct visualisation of the spermatic cord twist
with a sensitivity of 97.3% and specificity of 99% [233, 238] (LE: 2). A so-called positive whirlpool sign (the
presence of a spiral-like pattern), has a pooled sensitivity and specificity of 0.73 (95% CI, 0.65-0.79) and 0.99
(95% CI, 0.92-0.99), respectively, and may be viewed as a definitive sign for testicular torsion. But its role in
neonates is limited [239].
Scintigraphy and, more recently, dynamic contrast-enhanced subtraction MRI of the scrotum
also provide a comparable sensitivity and specificity to US [240-243]. These investigations may be used
when diagnosis is less likely and if torsion of the testis still cannot be excluded from history and physical
examination. This should be done without inordinate delays for emergency intervention [229].
The diagnosis of acute epididymitis in boys is mainly based on clinical judgement and adjunctive
investigation. However, it should be remembered that findings of secondary inflammatory changes in the
absence of evidence of an extra-testicular nodule by Doppler US might suggest an erroneous diagnosis of

24 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


epididymitis in children with torsion of the appendix testes [244]. Pre-pubertal boys with acute epididymitis
have an incidence of underlying urogenital anomalies of 25-27.6%. Complete urological evaluation in all
children with acute epididymitis is still debatable [199, 222, 224].

3.5.3 Management
3.5.3.1 Epididymitis
In pre-pubertal boys, the aetiology is usually unclear, with an underlying pathology in about 25%. A urine
culture is usually negative, and unlike in older boys, a sexually transmitted disease is very rare.
Antibiotic treatment, although often started, is not indicated in most cases unless urinalysis and
urine culture show a bacterial infection [224, 245]. Epididymitis is usually self-limiting and with supportive
therapy (i.e. minimal physical activity and analgesics) heals without any sequelae (LE: 3). However, bacterial
epididymitis can be complicated by abscess or necrotic testis and surgical exploration is required [246].

3.5.3.2 Testicular torsion


Manual detorsion of the testis is done without anaesthesia, and should be attempted in all patients if possible,
because it is associated with improved surgical salvage rates [247]. It should initially be done by outward
rotation of the testis - like opening a book -, unless the pain increases or if there is obvious resistance. Success
is defined as the immediate relief of all symptoms and normal findings at physical examination [248] (LE: 3;).
Doppler US may be used for guidance [249]. Bilateral orchiopexy is still required after successful detorsion.
This should not be done as an elective procedure, but rather immediately following detorsion. One study
reported residual torsion during exploration in 17 out of 53 patients, including eleven patients who had reported
pain relief after manual detorsion [248, 250].
External cooling before exploration may be effective in reducing ischaemia reperfusion injury and
preserving the viability of the torsed and the contralateral testis [251]. Medical treatments aimed at limiting such
injury remain experimental [252-255].

Torsion of the appendix testis can be managed non-operatively with the use of anti-inflammatory
analgesics (LE: 4). During the six-week follow-up, clinically and with US, no testicular atrophy was revealed.
Surgical exploration is done in equivocal cases and in patients with persistent pain [236]. Although metachronous
torsion of the appendix testis may occur in up to 8.5%, it is not necessary to explore the contralateral side, given
the benign nature ot the problem. Besides it has been demonstrated that the NNT is 24 [256].

3.5.3.3 Surgical treatment


Testicular torsion is an urgent condition which requires prompt surgical treatment. The two most important
determinants of early salvage rate of the testis are the time between onset of symptoms and detorsion, and the
degree of cord twisting [257]. Severe testicular atrophy occurred after torsion for as little as four hours when
the turn was > 360°. In cases of incomplete torsion (180-360°), with symptom duration up to twelve hours, no
atrophy was observed. However, a necrotic or severely atrophied testis was found in all cases of torsion > 360°
and symptom duration > 24 hours [258].
Early surgical intervention with detorsion (mean torsion time less than thirteen hours) was found
to preserve fertility [259]. Urgent surgical exploration is mandatory in all cases of testicular torsion within 24
hours of symptom onset. In patients with testicular torsion > 24 hours, exploration may be performed as a
semi-elective exploration procedure [257, 258] (LE: 3), unless there is a clear history of torsion-detorsion in
which urgent exploration should still be considered. In case of prolonged torsion (> 24 hours) it is still subject
to debate whether the surgically detorsed testis should be preserved. An alternative to detorsion and fixation
may be to perform orchiectomy. A study found that sperm quality was preserved after both orchiectomy and
orchidopexy in comparison to normal control men, although orchiectomy resulted in better sperm morphology
[260]. Incision of the tunica albuginea with tunica vaginalis graft to prevent or treat compartment syndrome has
also been suggested [261].
In neonates with signs of testicular torsion at birth the duration of symptoms will not be clear. The
decision to perform surgical exploration should be weighed against the general condition of the child. In this
age group the operation can safely be done under spinal anesthesia. New onset of symptoms of testicular
torsion in neonates should be considered a surgical emergency similar to older boys.

During exploration, fixation of the contralateral testis is also performed. It is good clinical practice to
also perform fixation of the contralateral testis in prenatal and neonatal torsion, although there is no literature
to support this, and to remove an atrophied testicle [262]. Recurrence after orchidopexy is rare (4.5%) and
may occur several years later. There is no consensus recommendation about the preferred type of fixation and
suture material [260, 261, 264].

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 25


3.5.4 Follow-up
Patients require follow-up mainly for fertility issues and hormonal consequences. Despite timely and adequate
detorsion and fixation of the testicle, up to half of the patients may develop testicular atrophy, even when
intraoperatively assessed as viable, and should be counselled accordingly [264, 265].

3.5.4.1 Fertility
The results vary and are conflicting. In one study, unilateral torsion of the testis seriously intervened with
subsequent spermatogenesis in about 50% of the patients and produced borderline impairment in another
20% [242]. Although, 30% of affected testicles with mumps orchitis show a degree of atrophy, long-term
outcome in terms of fertility is not conclusive [266].
A recent study showed a normal pregnancy rate after unilateral testicular torsion, with no difference
between the patients undergoing orchidopexy and those after orchidectomy [267].

3.5.4.2 Subfertility
Subfertility is found in 36-39% of patients after torsion. Semen analysis may be normal in only 5-50% in long-
term follow-up [257]. Early surgical intervention (mean torsion time less than thirteen hours) with detorsion was
found to preserve fertility, but a prolonged torsion period (mean 70 hours) followed by orchiectomy jeopardised
fertility [259].

Subfertility and infertility are consequences of direct injury to the testis after the torsion. This is caused by the
cut-off of blood supply, but also by post-ischaemia-reperfusion injury that is caused after the detorsion when
oxygen-derived free radicals are rapidly circulated within the testicular parenchyma [257].

3.5.4.3 Androgen levels


Even though the levels of FSH, luteinising hormone (LH) and testosterone are higher in patients after testicular
torsion compared to normal controls, endocrine testicular function remains in the normal range after testicular
torsion [260].

3.5.4.4 Unanswered questions


Although testicular torsion is a common problem, the mechanism of neonatal and prenatal torsion is still not
exactly known, as well as whether fixation of the contralateral testicle in these cases is really necessary. The
influence of an atrophied testicle on fertility is also unclear.

Summary of evidence LE
Diagnosis of testicular torsion is based on presentation and physical examination.
Doppler US is an effective imaging tool to evaluate acute scrotum and comparable to scintigraphy and 2a
dynamic contrast-enhanced subtraction MRI.
Neonates with acute scrotum should be treated as surgical emergencies. 3

Recommendations Strength rating


Testicular torsion is a paediatric urological emergency and requires immediate treatment. Strong
In neonates with testicular torsion perform orchidopexy of the contralateral testicle. In Weak
prenatal torsion the timing of surgery is usually dictated by clinical findings.
Base the clinical decision on physical examination. The use of Doppler ultrasound to Strong
evaluate acute scrotum is useful, but this should not delay the intervention.
Manage torsion of the appendix testis conservatively. Perform surgical exploration in Strong
equivocal cases and in patients with persistent pain.
Perform urgent surgical exploration in all cases of testicular torsion within 24 hours of Strong
symptom onset. In prenatal torsion the timing of surgery is usually dictated by clinical
findings.

3.6 Hypospadias
3.6.1 Epidemiology, aetiology and pathophysiology
3.6.1.1 Epidemiology
The total prevalence of hypospadias in Europe is 18.6 new cases per 10,000 births (5.1-36.8) according to the
recent EUROCAT registry-based study. This incidence was stable over the period of 2001 to 2010 [268, 269].
The mean worldwide prevalence of hypospadias according to an extended systematic literature review
varies: Europe 19.9 (range: 1-464), North America 34.2 (6-129.8), South America 5.2 (2.8-110), Asia 0.6-69,

26 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


Africa 5.9 (1.9-110), and Australia 17.1-34.8. There are conflicting data on the recent trends of prevalence –
different trends in Europe and an increasing trend in the USA [270, 271].

3.6.2 Risk factors


Risk factors associated with hypospadias are likely to be genetic, placental and/or environmental [268, 269]
(LE: 2b). Interactions between genetic and environmental factors may help explain non-replication in genetic
studies of hypospadias. Single nucleotide polymorphisms seemed to influence hypospadias risk only in
exposed cases [269, 272] (LE: 2b).
• An additional family member with hypospadias is found in 7% of families, but this is more predominant in
anterior and middle forms [272-275].
• Endocrine disorders can be detected in rare cases.
• Babies with a low birth weight have a higher risk of hypospadias [272-275].
• Over the last 25 years, a significant increase in the incidence of hypospadias has been found.
• Endocrines disruptors are one component of a multi-factorial model for hypospadias.
• The use of oral contraceptives prior to pregnancy has not been associated with an increased risk of
hypospadias in offspring, but their use after conception increased the risk of middle and posterior
hypospadias [273-276] (LE: 2a).

3.6.3 Classification systems


Hypospadias are usually classified based on the anatomical location of the proximally displaced urethral orifice:
• distal-anterior hypospadias (located on the glans or distal shaft of the penis and the most common type
of hypospadias);
• intermediate-middle (penile);
• proximal-posterior (penoscrotal, scrotal, perineal).

The pathology may be different after skin release and should be reclassified accordingly. Anatomical location
of the meatus may not always be enough to explain the severity and the complex nature of this pathology.
Therefore, a simple classification related to severity of the problem, which considers penile length, glans size,
shape, urethral plate quality and penile curvature is commonly used. In that classification there are two types: mild
hypospadias (glanular or penile isolated hypospadias without associated chordee, micropenis or scrotal anomaly);
severe hypospadias (penoscrotal, perineal hypospadias with associated chordee and scrotal anomalies).

3.6.4 Diagnostic evaluation


Most hypospadias patients are easily diagnosed at birth (except for the megameatus intact prepuce variant
which can only be seen after retraction of foreskin). Diagnosis includes a description of the local findings:
• position, shape and width of the orifice;
• presence of atretic urethra and division of corpus spongiosum;
• appearance of the preputial hood and scrotum;
• size of the penis;
• curvature of the penis on erection.

The diagnostic evaluation also includes an assessment of associated anomalies, which are:
• cryptorchidism (in up to 10% of cases of hypospadias);
• open processus vaginalis or inguinal hernia (in 9-15%).

Severe hypospadias with unilaterally or bilaterally impalpable testis, or with ambiguous genitalia, requires a
complete genetic and endocrine work-up immediately after birth to exclude DSD, especially congenital adrenal
hyperplasia. Urine trickling and ballooning of the urethra requires exclusion of meatal stenosis. The relationship
between the severity of the hypospadias and associated anomalies of the upper or lower urinary tract were not
confirmed [277] (LE: 3).

3.6.5 Management
3.6.5.1 Indication for reconstruction and therapeutic objectives
Differentiation between functionally necessary and aesthetically feasible operative procedures is important for
therapeutic decision making.

The indications for surgery are:


• proximally located (ectopic) meatus causing ventrally deflected or spraying urinary stream;
• meatal stenosis;
• anterior curvature of the penis;

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 27


• cleft glans;
• rotated penis with abnormal cutaneous raphe;
• preputial hood;
• penoscrotal transposition;
• split scrotum.

Physical examination should check all anatomic components of the penis and evaluate the degree and nature
of abnormality in each component. The examination should evaluate location of the meatus, the degree of
proximal spongiosal hypoplasia, presence and degree of penile curvature, width and depth of the urethral plate,
size of the glans, degree of ventral skin deficiency, availability of the foreskin and scrotal abnormalities like
penoscrotal transposition and bifid scrotum.

As all surgical procedures carry the risk of complications, thorough pre-operative counselling of the caregiver is
crucial.

To achieve an overall acceptable functional and cosmetic outcome, the penile curvature must be corrected and
a neo-urethra of an adequate size with opening on the glans formed with proper skin coverage of the penile
shaft [278] (LE: 4) (Figure 3). The use of magnifying spectacles and fine synthetic absorbable suture materials
(6.0-7.0) are required. As in any penile surgery, exceptional prudence should be adopted with the use of
cautery. Bipolar cautery is recommended. Knowledge of a variety of surgical reconstructive techniques, wound
care and post-operative treatment are essential for a satisfactory outcome.

3.6.5.2 Pre-operative hormonal treatment


There is a lack of high-quality evidence to support that pre-operative hormonal treatment with androgen
stimulation improves surgical outcomes. Yet, this treatment in the form of systemic testosterone, topical
testosterone, and derivatives like dihydrotestosterone (DHT) and hCG are commonly being used to increase
glans size pre-operatively to allow better tubularisation of the urethral plate and decrease the incidence of
glans dehiscence. This treatment is usually limited to patients with proximal hypospadias, a small appearing
penis, reduced glans circumference or reduced urethral plate [276, 279, 280]. Studies have shown that it leads
to significant enlargement of the glans and shaft of the penis (LE: 1b) [281, 282].

Moderate quality evidence from three randomised studies demonstrate significantly lower rates of
urethracutaneous fistulae and re-operation rates in patients who received pre-operative hormonal treatment [283].
Pre-operative testosterone administration is most often well tolerated. Transient side effects on
child´s behaviour, increased genital pigmentation, appearance of pubic hair, penile skin irritation and redness,
increased erections and peri-operative bleeding have been reported, but no persistent side effects related to
hormonal stimulation have been reported in the literature. There is also no evidence about possible effects on
bone maturation [280, 283, 284].

There are concerns regarding the negative impacts of testosterone on wound-healing and increased bleeding
during surgery. Cessation of therapy is recommended one or two months prior to surgery to avoid adverse
effects during or after surgery [285].

3.6.5.3 Age at surgery


The age at surgery for primary hypospadias repair is usually 6-18 (24) months [278, 286, 287] (LE: 3). Age at
surgery is not a risk factor for urethroplasty complication in pre-pubertal tubularised incised plate urethroplasty
(TIP) repair [286] (LE: 2b). Complication rate after primary TIP repair was 2.5 times higher in adults than in the
paediatric group according to a recent prospective controlled study [288] (LE: 2a).

3.6.5.4 Penile curvature


If present, penile curvature is often released by degloving the penis (skin chordee) and by excision of the
connective tissue of the genuine chordee on the ventral aspect of the penis in up to 70% [289]. The urethral
plate has well vascularised connective tissue and does not cause curvature in most cases [290, 291]. The
residual curvature is caused by corporeal disproportion and requires straightening of the penis, mostly using
dorsal midline plication or orthoplasty (modification of the Nesbit plication with or without elevation of the
neurovascular bundle). In more severe curvature (> 45°), which is often combined with a short urethral plate
requiring transection, ventral penile lengthening is recommended to prevent shortening of the penis. This
consists of a ventral transverse incision of the tunica albuginea extending from the 3 to 9 o´clock position
patched with tunica vaginalis flap or graft, or in several short ventral corporotomies without grafting (LE: 2b)
[292]. After the ventral lengthening, a shorter dorsal midline plication is usually added.

28 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


According to a retrospective study, dorsal plication remained significantly associated with recurrent
ventral curvature independently of the other factors. Ventral corporeal grafting for severe penile curvature gives
good long-term results and safety profiles for erectile function [293] (LE: 2b).

3.6.5.5 Urethral reconstruction


The mainstay of hypospadias repair is preservation of the well-vascularised urethral plate and its use for
urethral reconstruction has become standard practice in hypospadias repair [291]. Mobilisation of the corpus
spongiosum/urethral plate and the bulbar urethra decreases the need for urethral plate transection [292] (LE: 2b).
If the urethral plate is wide, it can be tubularised following the Thiersch-Duplay technique. If the
plate is too narrow to be simply tubularised, it is recommended relaxing the plate by a midline incision and its
subsequent tubularisation according to the Snodgrass-Orkiszewski TIP technique. This technique has become
the treatment of choice in distal- and mid-penile hypospadias [294-297]. If the incision of the plate is deep, it
is recommended to cover the raw surface with inner preputial (or buccal) inlay graft in primary and secondary
repairs [298]. This also enables extension of the incision beyond the end of the plate to prevent meatal stenosis
[299, 300] (LE: 2a).

For distal forms of hypospadias, a range of other techniques is available (e.g. Mathieu, urethral advancement)
[301] (LE: 2b). The TIP technique has become an option for proximal hypospadias as well [294-297, 302].
However, urethral plate elevation and urethral mobilisation should not be combined with TIP repair because it
results in focal devascularisation of the neo-urethra with symptomatic stricture development [303] (LE: 2b). The
onlay technique using a preputial island flap is a standard repair, preferred in proximal hypospadias, if a plate is
unhealthy or too narrow [289]. An onlay preputial graft is an option for single-stage repair [304] (LE: 2b).

If the continuity of the urethral plate cannot be preserved, single or two-stage repairs are used. For the former,
a modification of the tubularised flap (Duckett tube), such as a tube-onlay or an inlay-onlay flap, or onlay
flap on albuginea are used to prevent urethral stricture [305-307] (LE: 3); alternatively the Koyanagi-Hayashi
technique is used [308-311]. The two-stage procedure has become preferable over the past few years because
of lower recurrence of ventral curvature and more favourable results with variable long-term complication rates
[300, 305, 312-316].

3.6.5.6 Re-do hypospadias repairs


For re-do hypospadias repairs, no definitive guidelines can be given. All the above-mentioned procedures are
used in different ways and are often modified according to the individual findings and needs of the patient.

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 29


Figure 3: Algorithm for the management of hypospadias

Diagnosis at birth Exclude DSD

No
Paediatric urologist
reconstruction

Reconstruction
required

Distal Proximal

Chordee No chordee

Urethral Urethral
plate cut plate preserved

Thiersch Duplay,
TIP, Mathieu, Two-stage,
MAGPI, tube-onlay Onlay, TIP
advancement Koyanagi repair

DSD = disorders of sex development; TIP = tubularised incised plate urethroplasty; MAGPI = meatal
advancement and glanuloplasty incorporated.

3.6.5.7 Penile reconstruction following formation of the neo-urethra


Following formation of the neo-urethra, the procedure is completed by glansplasty and by reconstruction of
the penile skin. If there is a shortage of skin covering, the preputial double-face technique or placement of
the suture line into the scrotum according to Cecil-Michalowski is used. In countries where circumcision is
not routinely performed, preputial reconstruction can be considered. Preputial reconstruction carries a risk of
specific complications but does not seem to increase the risk of urethroplasty complications [317]. In TIP repair,
the use of a preputial dartos flap reduces the fistula rate [294, 295] (LE: 2b).

3.6.5.8 Urine drainage and wound dressing


Urine is drained transurethrally (e.g. dripping stent) or with a suprapubic tube. No drainage after distal
hypospadias repair is another option [318, 319]. Circular dressing with slight compression, as well as
prophylactic antibiotics during surgery, are established procedures [319] (LE: 4). Post-operative prophylaxis
after hypospadias repair has limited benefit and it only reduces the risk of asymptomatic bacteriuria [320-322]
(LE: 2b). There is no consensus on duration of stenting and dressing.

3.6.5.9 Outcome
Some studies have tried to determine risk factors for complications after hypospadias repair. An analysis of
prospectively collected data found glans size (width < 14 mm), proximal meatal location and re-operation as
independent risk factors for urethral complication [319, 323]. Low surgeon volume independently increases the
risk of fistula, stricture or diverticulum repair [319, 324] (LE: 3).

30 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


A meta-analysis of complication rates of TIP repair found lower complication rates and incidence of
re-operations in primary distal repairs (in 4.5%) than in primary proximal repairs (in 12.2%) and in secondary
repair (in 23.3%) [294-297, 301, 319]. One should expect a predictable outcome with complication rates below
10% in distal hypospadias (fistula, meatal stenosis, dehiscence, recurrent ventral curvature, and haematoma)
[324, 325]. A similar incidence of fistula (3.4-3.6%) can be expected after the Mathieu and TIP repairs of distal
hypospadias [302, 326-328].

The complication rates of TIP and onlay repairs of primary severe hypospadias are similar, 24% and 27%,
respectively. It is higher in free graft and in preputial island tube urethroplasty [289]. There is no strong evidence
to suggest that the use of inlay grafts in TIP repair improves the outcome [329].

The complication rates of single-stage Koyanagi and Hayashi modification repairs go up 61%, according to
a comparative study [308, 319]. Staged buccal mucosa graft requires a redo grafting in 13% of patients, after
the second stage more than one third of patients have complications, mostly with some degree of graft fibrosis
[327, 330]. A recent long-term study on two-stage flap repair showed a complication rate of 68% [319]; another
study showed a re-operation rate of 28% [300, 319].

3.6.6 Follow-up
Long-term follow-up is necessary up to adolescence to detect urethral stricture, voiding dysfunctions and
recurrent penile curvature, diverticula, glanular dehiscence [331]. Up to half of complications requiring
re-operation present after the first year post-operatively [332] (LE: 2b).
Obstructive flow curve is common after hypospadias repair and while most are not clinically
significant, long-term follow-up is required [333-336] (LE: 2a). Urine flow is significantly lower in patients after
hypospadias surgery, especially in those who had corrected chordee, but without significant association with
lower urinary symptoms [337] (LE: 2a).

Objective scoring systems have been developed in order to evaluate the results of hypospadias surgery (HOSE)
[338] (LE: 2b) and cosmetic appearance (HOPE-Hypospadias Objective Penile Evaluation) [339] (LE: 2a). The
Pediatric Penile Perception Score (PPPS) is a reliable instrument to assess penile self-perception in children
after hypospadias repair and for appraisal of the surgical result by caregivers and uninvolved urologists [340]
(LE: 2a). Cosmetic results were judged more optimistically by surgeons as compared to caregivers using
validated tools [341]. Current scoring systems have deficiencies in terms of patient reported outcomes, the
long term outcomes and sexual function [342].

Adolescents and adults, who have undergone hypospadias repair in childhood, have a slightly higher rate
of dissatisfaction with penile size, especially proximal hypospadias patients, but their sexual behaviour is
not different from that of control groups [343, 344] (LE: 2a-b). Another long-term follow-up of men born with
hypospadias revealed, in a controlled study, that these patients are less satisfied with penile cosmetic outcome
according to all parameters of the PPPS; there was a difference in penile length (9.7 vs. 11.6 cm) and more
patients had lower maximum urinary flow. More prominent results were found in proximal hypospadias vs.
controls [319, 345].

According to a systematic review of long-term patient satisfaction with cosmetic outcomes [346]:
• patient perception of penile size does not differ greatly from the norm;
• patients approaching puberty have a more negative perception and are more critical about the cosmetic
outcomes of surgery;
• patients report high levels of perception of deformity and social embarrassment.

There is a wide range of parameters that are measured to assess outcome after hypospadias surgery in the
literature. There is a need for age-specific core outcome set [347].

The majority of identified instruments focused on post-operative cosmetic satisfaction, with only one
instrument considering urinary function, and no instruments evaluating sexual function and psychosocial
sequelae [348].

3.6.7 Summary of evidence and recommendations for the management of hypospadias

Summary of evidence LE
The suggested age at surgery for primary hypospadias repair is 6 - 18 (24) months. 3

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 31


The therapeutic objectives are to correct the penile curvature, to form a neo-urethra of an adequate 4
size, to bring the new meatus to the tip of the glans, if possible, and to achieve an overall acceptable
cosmetic appearance.
Androgen stimulation therapy results in increased penile length and glans circumference. 1b
The complication rate is about 10% in distal and 25% in proximal hypospadias one-stage repairs. 3
Higher and variable rates (between 28 and 68%) can occur in two-stage repairs.
Sexual functions are usually well preserved but patients report high levels of perception of deformity 2b
and social embarrassment.

Recommendations Strength rating


At birth, differentiate isolated hypospadias from disorders of sex development which are Strong
mostly associated with cryptorchidism or micropenis.
Counsel caregivers on functional indications for surgery, aesthetically feasible operative Strong
procedures (psychological, cosmetic indications) and possible complications.
In children diagnosed with proximal hypospadias and a small appearing penis, reduced Weak
glans circumference or reduced urethral plate, pre-operative hormonal androgen stimulation
treatment is an option and the body of evidence to accentuate its harms and benefits is
inadequate.
For distal hypospadias, offer Duplay-Thiersch urethroplasty, original and modified Weak
tubularised incised plate urethroplasty; use the onlay urethroplasty or two-stage procedures
in more severe hypospadias. A treatment algorithm is presented (Figure 3). Correct
significant (> 30°) curvature of the penis.
Ensure long-term follow-up to detect urethral stricture, voiding dysfunctions and recurrent Strong
penile curvature, ejaculation disorder, and to evaluate patient´s satisfaction.
Use validated objective scoring systems to assist in evaluating the functional and cosmetic Strong
outcome.

3.7 Congenital penile curvature


3.7.1 Epidemiology, aetiology and pathophysiology
Congenital penile curvature presents penile bending of a normally formed penis due to corporal disproportion.
The incidence at birth is 0.6% and congenital penile curvature is caused by asymmetry of the cavernous
bodies and an orthotopic meatus [349] because of developmental arrest during embryogenesis [350]. On
the other hand, the incidence of clinically significant congenital penile curvature is much lower, because the
extent of the curvature and its associated sexual dysfunction varies widely [351]. Most of the cases are ventral
deviations (48%), followed by lateral (24%), dorsal (5%), and a combination of ventral and lateral (23%) [352].
Most ventral curvatures are associated with hypospadias due to chordee or ventral dysplasia of cavernous
bodies [353]. Similarly, dorsal curvature is mostly associated with exstrophy/epispadias complex. Congenital
penile curvature can decrease sexual quality of life in adults and successful repair can restore patients’
psychosocial and sexual wellbeing [354]

Curvature > 30° is considered clinically significant; curvature > 60° may interfere with satisfactory sexual
intercourse in adulthood (LE: 4). Minor penile curvature may be the result of ventral penile skin deficiency only
and should be distinguished from corporal anomalies. For penile curvature associated with hypospadias or
epispadias refer to the relevant chapters.

3.7.2 Diagnostic evaluation


Penile curvature is frequently not documented until later in childhood since the penis only appears abnormal
when erect. Patients are usually concerned with the aesthetic and/or functional aspects of their penis [355].
Besides exact history taking to exclude any possibility of acquired penile curvature (e.g. post-traumatic), a
thorough clinical examination is mandatory. In addition, photo documentation of the erect penis clearly showing
the curvature from different angles serves as a pre-requisite in pre-operative evaluation [356]. The exact degree
of curvature is generally determined at the time of surgery using an artificial erection test.

3.7.3 Management
The treatment is surgical, starting with an artificial erection to determine the degree of curvature and to check
symmetry after the repair [357]. The ultimate goal of any surgical method used to correct the curvature is to
achieve corpora of similar size. Various procedures are in use ranging from simple de-gloving and plication
procedures, to corporal rotation, use of free dermal or tunica vaginalis grafts, to complete penile disassembly
techniques [358, 359]. Reviews comparing the outcome of Nesbit/modified Nesbit procedures [360] to plication

32 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


procedures [361] were able to demonstrate that while there is a decreased risk of complications and loss of
sensation, it remains unclear whether plication techniques can lead to increased risk of recurrence [362, 363].
Altogether these methods include the risk of post-operative shortening of the penis with an average loss of
2.5 cm in stretched penile length depending on the pre-operative degree of curvature and the type of repair
used [364-366].
Recently the non-corporotomy technique has been introduced with promising results enabling
correction of any degree of ventral curvature with neither shortening of the penis nor the risk of post-operative
erectile dysfunction [367].

3.7.4 Summary of evidence and recommendations for the management of congenital penile
curvature

Summary of evidence LE
Isolated congenital penile curvature is relatively uncommon. 2a
Congenital penile curvature is often associated with hypospadias. 2a
Diagnosis is usually made late in childhood. 2a
The penis only appears abnormal when erect. 1b
Congenital penile curvature can cause aesthetic as well as functional sexual problems. 1b
Congenital penile curvature is treated with surgery. 1b
The goal of surgery is to achieve corpora of similar size. 1b

Recommendations LE Strength rating


Ensure that a thorough medical history is taken and a full clinical examination done 1a Strong
to rule out associated anomalies in boys presenting with congenital curvature.
Provide photo documentation of the erect penis from different angles as a 1b Strong
prerequisite in the pre-operative evaluation.
Perform surgery after weighing aesthetic as well as functional implications of the 2b Weak
curvature.
At the beginning as well as at the end of surgery, perform artificial erection tests. 2a Strong

3.8 Varicocele in children and adolescents


3.8.1 Epidemiology, aetiology and pathophysiology
Varicocele is defined as an abnormal dilatation of testicular veins in the pampiniformis plexus caused by
venous reflux. It is unusual in boys under ten years of age and becomes more frequent at the beginning of
puberty. It is found in 14-20% of adolescents, with a similar incidence during adulthood. It appears mostly on
the left side (78-93% of cases). Right-sided varicoceles are less common; they are usually noted only when
bilateral varicoceles are present and seldom occur as an isolated finding [368-370].
Varicocele develops during accelerated body growth and increased blood flow to the testes, by a
mechanism that is not clearly understood. Genetic factors may be present. An anatomic abnormality leading
to impaired venous drainage is expressed by the considerable prevalence of the left side condition where
the internal spermatic vein drains into the renal vein. Varicocele can induce apoptotic pathways because of
heat stress, androgen deprivation and accumulation of toxic materials. Severe damage is found in 20% of
adolescents affected, with abnormal findings in 46% of affected adolescents. Histological findings are similar
in children or adolescents and in infertile men. In 70% of patients with grade II and III varicocele, left testicular
volume loss was found.
Several authors reported on reversal of testicular growth after varicocelectomy in adolescents [371,
372]. An average proportion of catch-up growth of 76.4% (range: 52.6-93.8%) has been found according to a
meta-analysis [373] (LE: 2a). However, this may partly be attributable to testicular oedema associated with the
division of lymphatic vessels [374] (LE: 2).
In about 20% of adolescents with varicocele, fertility problems will arise [375]. The adverse influence
of varicocele increases with time. Improvement in sperm parameters has been demonstrated after adolescent
varicocelectomy [376-379] (LE: 1).

3.8.2 Classification systems


Varicocele is classified into 3 grades [380]:
• Grade I - Valsalva positive (palpable at Valsalva manoeuvre only);
• Grade II - palpable (palpable without the Valsalva manoeuvre);
• Grade III - visible (visible at distance).

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 33


3.8.3 Diagnostic evaluation
Varicocele is mostly asymptomatic, rarely causing pain. It may be noticed by the patient or caregivers, or
discovered by the paediatrician at a routine visit. The diagnosis depends upon the clinical finding of a collection
of dilated and tortuous veins in the upright posture; the veins are more pronounced when the patient performs
the Valsalva manoeuvre. The size of both testicles should be evaluated during palpation to detect a smaller
testis.
Venous reflux into the plexus pampiniformis is diagnosed using Doppler US colour flow mapping in
the supine and upright position [381]. Venous reflux detected on US only is classified as subclinical varicocele.
To discriminate testicular hypoplasia, the testicular volume is measured by US examination or by orchidometer.
In adolescents, a testis that is smaller by > 2 mL or 20% compared to the other testis is considered to be
hypoplastic [382] (LE: 2).
Extension of Wilms tumour into the renal vein and inferior vena cava can cause a secondary
varicocele. A renal US should be routinely added in pre-pubertal boys and in isolated right varicocele
(LE: 4). In order to assess testicular injury in adolescents with varicocele, supranormal FSH and LH responses
to the luteinising hormone-releasing hormone (LHRH) stimulation test are considered reliable, because
histopathological testicular changes have been found in these patients [378, 383].

3.8.4 Management
There is no evidence that treatment of varicocele at paediatric age will offer a better andrological outcome than
an operation performed later. Beneficial effect of pubertal screening and treatment for varicocele regarding
chance of paternity has been questioned according to a corresponding questionnaire in adult patients [384]
(LE: 4). The recommended indication criteria for varicocelectomy in children and adolescents are [369]:
• varicocele associated with a small testis;
• additional testicular condition affecting fertility;
• bilateral palpable varicocele;
• pathological sperm quality (in older adolescents);
• symptomatic varicocele [384].

Testicular (left + right) volume loss in comparison with normal testes is a promising indication criterion, once the
normal values are available [385]. Repair of a large varicocele, causing physical or psychological discomfort,
may also be considered. Other varicoceles should be followed-up until a reliable sperm analysis can be
performed (LE: 4).
Surgical intervention is based on ligation or occlusion of the internal spermatic veins. Ligation is
performed at different levels:
• inguinal (or subinguinal) microsurgical ligation;
• suprainguinal ligation, using open or laparoscopic techniques [386-389].

The advantage of the former is the lower invasiveness of the procedure, while the advantage of the latter is a
considerably lower number of veins to be ligated and safety of the incidental division of the internal spermatic
at the suprainguinal level.
For surgical ligation, some form of optical magnification (microscopic or laparoscopic) should be
used because the internal spermatic artery is 0.5 mm in diameter at the level of the internal ring [386, 388]. The
recurrence rate is usually < 10%.
Lymphatic-sparing varicocelectomy is preferred to prevent hydrocele formation and testicular
hypertrophy development and to achieve a better testicular function according to the LHRH stimulation test
[374, 386, 387, 390] (LE: 2). The methods of choice are subinguinal or inguinal microsurgical (microscopic)
repairs, or suprainguinal open or laparoscopic lymphatic-sparing repairs [386, 388, 391, 392]. Intrascrotal
application of isosulphan blue was recommended to visualise the lymphatic vessels [393, 394]. In suprainguinal
approach, an artery sparing varicocelectomy may not offer any advantage in regards to catch-up growth and is
associated with a higher incidence of recurrent varicocele [395, 396].

Angiographic occlusion of the internal spermatic veins also meets the requirements of lymphatic sparing
repair. It is based on retrograde or antegrade sclerotisation of the internal spermatic veins [397, 398]. However,
although this method is less invasive and may not require general anaesthesia, it is associated with radiation
burden, which is less controllable in the antegrade technique [369, 397, 398] (LE: 2).

There is low to moderate level of evidence that radiological or surgical treatment of adolescent varicocele is
associated with improved testicular size/growth and sperm concentration - based on current available RCTs.
The ultimate effects on fertility and paternity rates are not known [399].

34 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


Microsurgical varicocele repair in adolescents with varicocele significantly increases paternity rates and
decreases time to conception post-operatively. Patients with varicocele who underwent microsurgical
varicocele repair had increased sperm parameters and 3.63 times greater odds of paternity than controls who
did not undergo varicocele surgery [400].

The Panel recently conducted a systematic review and meta-analysis regarding the treatment of varicocele
in children and adolescents [401]. Of 1,550 articles identified, 98 articles including 16,130 patients were
eligible for inclusion (12 RCTs, 47 NRSs and 39 case series). The key findings are summarised in the following
paragraphs:

The meta-analysis of the twelve RCTs revealed that varicocele treatment improved testicular volume (mean
difference 1.52 ml, 95% CI 0.73-2.31) and increased total sperm concentration (mean difference 25.54, 95% CI
12.84-38.25) when compared with observation. Lymphatic sparing surgery significantly decreased hydrocele
rates (p=0.02) and the OR was 0.08 (95% CI 0.01, 0.67). Due to the lack of RCTs, it was not possible to identify
a surgical technique as being superior to the others. It remains unclear whether open surgery or laparoscopy is
more successful for varicocele treatment (OR ranged from 0,13 to 2,84).

The success rates of the treatment (disappearance of varicocele) were between 85.1% and 100% whereas
the complication rates were between 0% and 29% in the included studies. The most common complication
reported was hydrocele. Resolution of pain after treatment was more than 90% in the reported series.

In conclusion, moderate evidence exists on the benefits of varicocele treatment in children and adolescents in
terms of testicular volume and sperm concentration. Current evidence does not demonstrate superiority of any
of the surgical/interventional techniques regarding treatment success. Lymphatic sparing surgery significantly
decreases hydrocele formation. Long-term outcomes, including paternity and fertility, still remain unknown.

3.8.5 Summary of evidence and recommendations for the management of varicocele

Summary of evidence LE
Varicocele becomes more frequent at the onset of puberty and is found in 14-20% of adolescents.
Fertility problems are expected in up to 20% of adolescents with a varicocele.
Pubertal patients with a left grade II and III varicocele have the left testis smaller in up to 70% of 1b
cases; in late adolescence the contralateral right testis also becomes smaller.
After adolescent varicocelectomy, left testis catch-up growth and improvement in sperm parameters 1a
has been demonstrated.
There is no evidence that treatment of varicocele at paediatric age will offer a better andrological 1b
outcome than an operation performed later.
Division of testicular lymphatics leads to hydrocele in up to 40% and to testicular hypertrophy. 1b
Lymphatic sparing surgery significantly decrease hydrocele rates. 1a

Recommendations LE Strength rating


Examine varicocele in the standing position and classify into three grades. 4 Strong
Use scrotal ultrasound to detect venous reflux without Valsalva manoeuvre in the Strong
supine and upright position and to discriminate testicular hypoplasia.
In all pre-pubertal boys with a varicocele and in all isolated right varicoceles perform Strong
standard renal ultrasound to exclude a retroperitonal mass.
Inform caregivers and patients and offer surgery for: 2 Weak
• varicocele associated with a persistent small testis (size difference of > 2 mL or
20%);
• varicocele associated with additional testicular condition affecting fertility
(cryptorchidism, history of torsion, trauma);
• varicocele associated with pathological sperm quality (in older adolescents);
• symptomatic varicocele.
Use some form of optical magnification (microscopic or laparoscopic magnification) 2 Strong
for surgical ligation.
Use lymphatic-sparing varicocelectomy to prevent hydrocele formation and 1 Strong
testicular hypertrophy.

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 35


3.9 Urinary tract infections in children
3.9.1 Epidemiology, aetiology and pathophysiology
Urinary tract infections (UTIs) represent the most common bacterial infections in children [402-404]. There are
several classification systems used to define a UTI. In neonates, the symptoms differ in many aspects from
those in infants and children. The prevalence is higher; there is a male predominance; infections caused by
other organisms than Escherichia coli are more frequent; and there is a higher risk of urosepsis [405, 406].
In children presenting with urinary symptoms a pooled prevalence of UTI was 7.8% (CI: 6.6-8.9)
[405]. The incidence varies depending on age and sex. One meta-analysis showed that in children presenting
with fever in the first three months of life UTIs were present in 7.5% of girls, 2.4% (CI: 1.4-3.5) of circumcised
boys, and 20.1% (CI: 16.8-23.4) of uncircumcised boys [405]. The incidence for boys is highest during the first
6 months of life (5.3%) and decreases with age to around 2% for the ages 0-6 years. In girls, UTIs are less
common during the first 6 months of life (2%) and incidence increases with age to around 11% for the ages 0-6
years [407].
Associated risk factors for recurrent UTIs include bladder and bowel dysfunction (BBD),
vesicoureteral reflux (VUR) and obesity [408-410]. In older children a delay in treatment is more often seen
than in younger infants [411]. These risk factors in combination with delay in treatment have been associated
with renal scarring [412]. Recurrent febrile UTIs, especially in combination with high-grade VUR, lead to renal
scarring [413, 414]. Each new febrile UTI increases the risk of renal scarring with an incidence of renal scarring
after the first UTI, of 2.8% (CI:1.2-5.8), 25.7% (CI:12.5-43.3) after the second infection and up to 28.6%
(CI:8.4-58.1) after 3 or more febrile UTIs [414].
The leading causative organism for UTIs has been E. coli, but other bacteriae have been rising
in prevalence. In a large European study E. Coli was found in less than 50% of urine cultures. Klebsiella
pneumoniae, Enterobacter spp., Enterococcus spp., Pseudomonas spp., Proteus spp. and Candida spp. are
more frequent in nosocomial infections than in community-acquired UTIs, even though, their prevalence has
increased outside of the hospital setting [415]. Neonatal UTI is frequently complicated by bacteraemia. In a
retrospective study, 12.4% of blood cultures from neonates admitted for UTI were positive for bacteraemia
[416], however, it is less frequent in community-acquired than in nosocomial UTI [416, 417].

3.9.2 Classification systems


There are five widely used classification systems according to; site, severity, episode, symptoms and
complicating factors. For acute treatment, site and severity are most important.

3.9.2.1 Classification according to site


Lower urinary tract infection (cystitis) is an inflammatory condition of the urinary bladder mucosa with general
signs and symptoms including infection, dysuria, frequency, urgency, malodorous urine, enuresis, haematuria,
and suprapubic pain.
Upper urinary tract infection (pyelonephritis) is a diffuse pyogenic infection of the renal pelvis
and parenchyma. The onset of pyelonephritis is generally abrupt. Clinical signs and symptoms include fever
(> 38°C), chills, costovertebral angle or flank pain, and tenderness.

3.9.2.2 Classification according to severity


In a lower urinary tract infection, children may have only mild pyrexia; are able to take fluids and oral
medication; are only slightly or not dehydrated; and have a good expected level of compliance. When a low
level of compliance is expected, such children should be managed as those with severe UTI. In severe UTI,
infection is related to the presence of fever of > 39°C, the feeling of being ill, persistent vomiting, and moderate
or severe dehydration. Most severe UTIs are upper urinary tract infections.

3.9.2.3 Classification according to episode first/persistent/recurrent/breakthrough


The first UTI may be a sign of anatomical anomalies. Anatomical evaluation is recommended (see below).
Recurrent infection can be divided into unresolved and persistent infection.
In unresolved infection, initial therapy is inadequate for elimination of bacterial growth in the
urinary tract (inadequate therapy, inadequate antimicrobial urinary concentration [poor renal concentration/
gastrointestinal malabsorption], and infection involving multiple organisms with differing antimicrobial
susceptibilities).
Persistent infection is caused by re-emergence of bacteria from a site within the urinary tract
coming from a nidus for persistent infection that cannot be eradicated (e.g. infected stones, non-functioning
or poorly functioning kidneys/renal segments, ureteral stumps after nephrectomy, necrotic papillae, urachal
cyst, urethral diverticulum, peri-urethral gland, vesicointestinal, rectourethral or vesicovaginal fistulas). The
same pathogen is identified in recurrent infections, but episodes of sterile urine may occur during and shortly
following antimicrobial treatment.

36 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


A breakthough infection in patients under antibacterial prophylaxis is usually caused by resistent
bacteria, parental non-compliance and/or severe urogenital anomalies [418, 419].
In re-infection, each episode can be caused by a variety of new infecting organisms, in contrast
to bacterial persistence in which the same infecting organism is always isolated. However, the most common
general pathogenic species is E. coli, which occurs in many different serotypes. Therefore, recurrent E. coli UTI
does not equate to infection with the same organism.

3.9.2.4 Classification according to symptoms


Children may have typical or atypical symptoms regarding a UTI.
In neonates and infants the most commen symptoms are fever, vomiting, lethargy and/or irritability.
Infants and children may have non-specific signs such as poor appetite, failure to thrive, lethargy, irritability,
vomiting or diarrhoea. Toilet trained children may report cystitis symptoms along with fever/flank pain.
Asymptomatic bacteriuria indicates attenuation of uropathogenic bacteria by the host, or
colonisation of the bladder by non-virulent bacteria that are incapable of activating a symptomatic response
(no leukocyturia, no symptoms). Asymptomatic UTI includes leukocyturia but no other symptoms.
Symptomatic UTI, includes irritative voiding symptoms, suprapubic pain (cystitis), fever and malaise
(pyelonephritis). Cystitis may represent early recognition of an infection destined to become pyelonephritis, or
bacterial growth controlled by a balance of virulence and host response.

3.9.2.5 Classification according to complicating factors


In uncomplicated UTI, infection occurs in a patient with a morphologically and functionally normal upper
and lower urinary tract, normal renal function and competent immune system. This category includes mostly
isolated or recurrent bacterial cystitis and is usually associated with a narrow spectrum of infecting pathogens
that are easily eradicated by a short course of oral antimicrobial agents. Patients can be managed on an
outpatient basis, with an emphasis on documenting resolution of bacteriuria, followed by elective evaluation for
potential anatomical or functional abnormalities of the urinary tract [420].
A complicated UTI occurs in children with known mechanical or functional pathology of the urinary
tract. Mechanical obstruction is commonly due to the presence of posterior urethral valves, strictures or
stones, independent of their location. Functional obstruction often results from lower urinary tract dysfunction
(LUTD) of either neurogenic or non-neurogenic origin and dilating VUR. Patients with complicated UTI require
hospitalisation and parenteral antibiotics. Prompt anatomical evaluation of the urinary tract is critical to exclude
the presence of significant abnormalities [421]. If mechanical or functional abnormalities are present, adequate
drainage of the infected urinary tract is necessary.

3.9.3 Diagnostic evaluation


3.9.3.1 Medical history
Medical history includes the question of a primary (first) or secondary (recurring) infection; possible
malformations of the urinary tract (e.g. pre- or post-natal US screening); prior operation; family history; and,
whether there is constipation or presence of lower urinary tract symptoms (LUTS).

3.9.3.2 Clinical signs and symptoms


Neonates with severe UTI can present with non-specific symptoms (failure to thrive, jaundice, hyperexcitability)
and without fever. In neonates it is important to rule out a co-existing meningitis [422]. Urinary tract infection is
the cause of fever in 4.1-7.5% of children who present to a paediatric clinic [423, 424]. Septic shock is unusual,
even with very high fever. Signs of a UTI may be vague and unspecific in small children, but later on, when
they are more than two years old, frequent voiding, dysuria and suprapubic, abdominal or lumbar pain can be
detected.

3.9.3.3 Physical examination


Physical examination includes a general examination of the throat, lymph nodes, abdomen (constipation,
palpable and painful kidney, or palpable bladder), flank, the back (stigmata of spina bifida or sacral agenesis),
genitalia (phimosis, labial adhesion, vulvitis, epididymo-orchitis), measurement of body weight and temperature.

3.9.3.4 Urine sampling, analysis and culture


Urine sampling has to be performed before any antimicrobial agent is administered. The technique for
obtaining urine for urinalysis as well as culture affects the rate of contamination, which influences interpretation
of the results. Especially in early infancy, it can be challenging and depends on the mode of urine sampling
[425].

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 37


3.9.3.4.1 Urine sampling
Urine must be collected under defined conditions and investigated as soon as possible to confirm or exclude
UTI, especially in children with fever. In neonates, infants and non-toilet-trained children, there are four main
methods with varying contamination rates and invasiveness to obtain urine:

(1) Plastic bag attached to the cleaned genitalia: Although this technique is most often used in daily practice,
contamination rates are high with around 50-60% [426]. It is helpful when the culture results are negative. Also,
if the dipstick is negative for both leukocyte esterase and nitrite, or microscopic analysis is negative for both
pyuria and bacteriuria, UTI can be excluded without the need for confirmatory culture [427].

(2) Clean-catch urine (CCU) collection: The infant is placed in the lap of a caregiver or member of the nursing
staff, who holds a sterile foil bowl underneath the infant’s genitalia. The infant is offered oral fluids and urine
collection is awaited [428]. Suprapubic tapping alternated with paravertebral lumbar massage can stimulate
sponateous voiding [426, 429]. There seems to be a good correlation between the results of urine culture
obtained by this method and suprapubic aspiration (SPA), with a false-positive rate of 5% and false-negative
rate of 12% [428, 430]; however, the contamination rate is higher for CCU with up to 26% compared to
catheterisation 10% and SPA 1% [426, 431]. In one prospective cohort study in infanty below the age of 6
months, the success rate was 49% and the contamination rate 16% with some differences in the culture
results between those obtained by CCU and those by more inavsive methods [432].

(3) Transurethral bladder catheterisation: is the fastest and safest method to obtain a reliable urine sample for
microscopic and bacteriological evaluation to rule out or to document a UTI in non-toilet trained infants and
children.

(4) Suprapubic bladder aspiration: This is the most invasive but also the most sensitive method to obtain an
uncontaminated urine sample in this age group [433, 434]. For suprapubic puncture ultrasound imaging should
be performed to asses bladder filling.
A two-step procedure where the CCU is screened and a catheter or SPA confirmation of the positive
screens is used can lead to a reduction in invasive procedures [426, 431].
In older, toilet-trained children who can void on command, after carefully retracting the foreskin and
cleaning the glans penis in boys and spreading the labia and cleaning the peri-urethral area in girls, the use of
clean catch, especially midstream urine, could be an acceptable technique for obtaining urine. After cleaning
the urethral meatus and perineum with gauze and liquid soap twice, the risk of contamination was reduced
from 23.9% (41/171) to 7.8% (14/171) in a randomised trial [435].

3.9.3.4.2 Urinalysis
There are three methods that are commonly used for urinalysis:

(1) Dipsticks: These are appealing because they provide rapid results, do not require microscopy, and are
ready to use. Leukocyte esterase (as a surrogate marker for pyuria) and nitrite (which is converted from dietary
nitrates by most Gram-negative enteric bacteria in the urine) are the most frequent markers, and are usually
combined in a dipstick test. The conversion of dietary nitrates to nitrites by bacteria takes approximately four
hours in the bladder [430, 436]. Using only nitrate sticks to screen febrile children < 2 years of age has a too
low sensitivity and relevant UTIs can be missed. However, the specificity is high for children at any age [437,
438]. In febrile infants < 90 days old urine dipstick tests using CCU samples can be used for screening for a
UTI when nitrites and leukocyte esterase combined are used with a sensitivity of 86% and a specificity of 80%
[439].

(2) Microscopy: This is the standard method of assessing pyuria after centrifugation of the urine with a
threshold of five white blood cells (WBCs) per high-power field (25 WBC/μL) [440]. In uncentrifuged urine, > 10
WBC/μL has been demonstrated to be sensitive for UTI [441] and this could perform well in clinical situations
[442]. However, this is rarely done in an outpatient setting. No significant differences was found between
dipsticks and microscopy testing for UTI [438]. A meta-analysis showed, that only microscopy with Gram
staining has a higer sensitivity compared to dipsticks [443].

(3) Flow imaging analysis technology: This is being used increasingly to classify particles in uncentrifuged urine
specimens [444]. The numbers of WBCs, squamous epithelial cells and red cells correlate well with those found
by manual methods [430]. Flow cytometry-based bacterial and leukocyte count analysis when using a cut-off
value of 250 bacteria/uL in the presence of leukocyturia has a sensitivity of 0.97 and specificity of 0.91 for
diagnosing UTI [445].

38 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


3.9.3.4.3 Urine culture
After negative results for dipstick, microscopic or automated urinalysis, urine culture is generally not necessary,
especially if there is an alternative source of fever. If the dipstick result is positive, confirmation by urine culture
is strongly recommended.
It is unclear what represents a significant UTI. In patients with a severe UTI, ≥ 105 cfu/mL can be
expected. However, the count can vary and be related to the method of specimen collection, diuresis, and time
and temperature of storage until cultivation occurs [406]. Clean-catch urine, midstream and catheterisation
urine cultures can be considered positive as 103 - 104 cfu/mL in a monoculture, and any counts obtained after
SPA should be considered as significant. Mixed cultures are indicative of contamination. In febrile children
< 4 months of age a cut-off value of 103 cfu/mL can be used when clinical and laboratory findings match and a
correct sampling method has been used [446].
A negative culture with the presence of pyuria may be due to incomplete antibiotic treatment,
urolithiasis, or foreign bodies in the urinary tract, and infections caused by Mycobacterium tuberculosis or
Chlamydia trachomatis.

A flowchart was developed as guidance during the basic diagnostic evaluation and subsequent management
of febrile children with clinical symptoms of UTI, Figure 4.

Figure 4: D
 iagnostic evaluation and subsequent management of a febrile child with clinical symptoms of
UTI

Febrile child with clinical symptoms

Physical examinaon

Blood sample (Blood cell count + CRP +/- Procalcitonin

Proper urine sampling & Urinalysis


(Dipsck +/- Microscopy +/- Flow-imaging analysis)

Nitrate -ve and leukocyte -ve / Nitrate +ve and/or leukocyte


Microscopy -ve / Flow +ve / Microscopy +ve / Flow
imaging -ve imaging +ve

Exclude other causes Urine culture

Start AB treatment following


anbacterial stewardship

Figure 5

CRP = C-reactive protein; AB = antibiotic.

3.9.3.5 Imaging
3.9.3.5.1 Ultrasound
Renal and bladder US within 24 hours is advised in infants with febrile UTI to exclude obstruction of the upper
and lower urinary tract. Abnormal results are found in 15% of cases, and 1-2% have abnormalities that require
prompt action (e.g., additional evaluation, referral or surgery) [427]. When a renal US is performed in all children
presenting with a UTI, 7% will have an abnormal US warranting further investigations [447]. The sensitivity to
detect high-grade VUR with US was found to be 0.59 (CI: 0.45-0.72) with a specificity of 0.79 (CI: 0.65-0.87)
[448]. Renal ultrasound should be performed before and after voiding. Post-void residual (PVR) urine should
be measured in toilet-trained children to exclude voiding abnormalities as a cause of UTI. Elevated PVR urine
volume predicts recurrence of UTIs in toilet-trained children [449]. When peri-renal or psoas abcesses or renal
masses are seen on US, it is important to consider xanthogranulomatous pyelonephritis, and subsequent CT
imaging is proposed [450].

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 39


3.9.3.5.2 Radionuclide scanning/MRI
Changes in dimercaptosuccinic acid (DMSA) clearance during acute UTI indicate pyelonephritis or
parenchymal damage, correlated with the presence of dilating reflux and the risk of further pyelonephritis
episodes, breakthrough infections [451] and future renal scarring. In the acute phase of a febrile UTI (up to four
to six weeks), DMSA-scan can demonstrate pyelonephritis by perfusion defects. Renal scars can be detected
after three to six months [452]. Diffusion-weighted MRI has shown to accurately diagnose acute pyelonephritis
and reveal late renal scars and could be an alternative to DMSA; therefore, avoiding radion burden [453]. The
average effective radiation dose of a single DMSA scan was 2.84 (1-12) mSv in one study [454]. These findings
are different in neonates. After the first symptomatic, community-acquired UTI, the majority of renal units with
VUR grade III or higher had normal early DMSA scanning [455]. The sensitivity of the DMSA scan to detect
VUR is 0.75 (CI:0.67-0.81) with a specificity of 0.48 (CI: 0.38-0.57), and a negative DMSA scan resulting in a
very low probability of high-grade VUR [456].

3.9.3.5.3 Voiding cystourethrography/urosonography


The optimum method to exclude or confirm VUR is VCUG. The timing of VCUG does not influence the
presence or severity of VUR [457]. Performance of early VCUG in patients with proven sterile urine does not
cause any significant morbidity [458]. Using harmonic voiding urosonography may be an alternative to the
standard VCUG avoiding radiation [459]. Visulatisation of the urethra may be difficult with this technique.
It is important to diagnose high-grade VUR after the first UTI since this is an important risk for renal
scarring. On the other hand, physicians want to avoid unnecessary VCUG investigations at the same time,
given its invasive character and radiation burden [447, 460]. Various studies have investigated the risk factors
for high-grade VUR and a top down approach is feasible. The most important risk factors for high-grade VUR
and subsequent renal scarring are: abnormal renal US, high fever UTI and non-E. Coli infections. Different top
down strategies with selective VCUG investigations have been proposed [461-465]. Based on these studies we
recommend the following updated diagnostic strategy (see Figure 5).

Figure 5: Diagnosis strategy for first febrile UTI

First febrile UTI

Renal and bladder


ultrasound

Abnormal findings (e.g. dilata on


Normal findings of upper tract/abscess
forma on/obstruc ng stone)

E. Coli Non-E. Coli Complicated UTI


infec on infec on close monitoring + i.v. an bio cs

Children Good Cri cal clinical


>12 Infants clinical status/ no
months response response

Imaging Further Urinary


Exclusion
aer evalua on diversion
of VUR (by
recurrent of upper (nephrostomy/
imaging)
febrile UTI tract JJ/catheter)

Toilet trained children


Exclude BBD

UTI = urinary tract infection; VUR = vesicoureteral reflux; i.v. = intravenous.

40 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


3.9.4 Management
3.9.4.1 Administration route of antibacterial therapy
The choice between oral and parenteral therapy should be based on patient age; clinical suspicion of
urosepsis; illness severity; refusal of fluids, food and/or oral medication; vomiting; diarrhoea; non-compliance;
and complicated pyelonephritis (e.g. urinary obstruction). As a result of the increased incidence of urosepsis
and severe pyelonephritis in newborns and infants aged less than two months, parenteral antibiotic therapy
is recommended. Electrolyte disorders with life-threatening hyponatraemia and hyperkalaemia based on
pseudohypoaldosteronism can occur in these cases [466, 467].
The choice of agent is also based on local antimicrobial sensitivity patterns, and should later be
adjusted according to sensitivity-testing of the isolated uropathogen [430]. Not all available antibiotics are
approved by national health authorities, especially in infancy. When recent urinary cultures are available use
these sensitivity patterns in the choice for treatment. In children who require intravenuous treatment tobramycin
or gentamicin is recommended if there is normal kidney function. When abnormal kidney function is suspected,
ceftriaxon or cefotaxime are alternative treatment options. In children who can receive oral treatment without
any known resistant urinary cultures, cefixime or amoxicillin-clavulanate are the empirical treatment options
[468]. Some studies have demonstrated that once daily parenteral administration of gentamicin or ceftriaxone
in a day treatment centre is safe, effective and cost-effective in children with UTI [469-471]. Delaying treatment
in children with a febrile UTI for more than 48-72 hours increases the risk of renal scars [412, 472].

3.9.4.2 Duration of therapy


Prompt adequate treatment of UTI can prevent the spread of infection and renal scarring. In newborns
and young infants with a febrile UTI, up to 20% may have a positive blood culture [416, 421]. Children with
bacteremia did not show significant clinical differences with non-bacteremic infants, but did receive longer
parental treatment [473]. In late infancy, there are no differences between strategies regarding the incidence
of parenchymal scars, as diagnosed with DMSA scan [474]. Outcomes of short courses (one to three days)
are inferior to seven to fourteen-day courses [430]. However, a simple cystitis can be treated with 3-5 days of
antibiotics [468]. No significant difference in recurrent UTIs and rehospitalisation was found between seven day
parental treatment and longer regimens for bacteremic UTI in younger infants [475]. In young infants a short
course of parental treatment with early conversion to oral antibiotics may be considered. The use of exclusively
oral therapy with a third-generation cephalosporin (e.g. cefixime or ceftibuten) has been demonstrated to be
equivalent to the usual two to four days intravenous therapy followed by oral treatment [476-479]. Similar data
have been shown for amoxicillin-clavulanate [480]. If ambulatory therapy is chosen, adequate surveillance,
medical supervision and, if necessary, adjustment of therapy must be guaranteed. In the initial phase of
therapy, a close ambulant contact to the family is advised [481].
In complicated UTI, uropathogens other than E. coli, such as Proteus mirabilis, Klebsiella spp.,
Pseudomonas aeruginosa, enterococci and staphylococci are more often the causative pathogens [421]. A
temporary urinary diversion (transurethral catheter, suprapubic cystostomy, percutaneous nephrostomy or
ureteral stenting) might be required in case of failure of conservative treatment in obstructive uropathy. Children
with acute focal bacterial nephritis often present without pyyuria and significant bacteriuria. For the majority
of children, the pathogenesis is related to ascending infection due to pre-existing uropathy, especially VUR or
urinary obstruction. Initial management consists of broad-spectrum antibiotics with good tissue penetration.
A treatment regimen of a total of 3 weeks with initial intravenous and subsequently oral therapy tailored to the
pathogen identified in culture is recommended [482].

3.9.4.3 Antimicrobial agents


There is a great difference in the prevalence of antibiotic resistance of uropathogenic E. coli in different
countries, with increased high resistance patterns in countries outside of the Organisation for Economic
Co-operation and Development (OECD) [483]. There are upcoming reports of UTIs caused by extended
spectrum ß-lactamase-producing enterobacteriaceae (ESBL) in children, with pooled numbers of UTI caused
by ESBL producing bacteria of around 14% [484]. Within OECD countries the prevalence of resistance was
53% for ampicillin, 24% for trimethoprim, 8% for co-amoxiclav, 2% for ciproxin and 1% for nitrofurantoin [483].
Several risk factors and determinants for UTIs caused by ESBL and non-E. Coli bacteriae have been identified:
history of infection, recent hospitalisation, short-term exposure to antiobiotics, and prophylaxis [483, 485, 486].
Overall, oral nitrofurantoin seems to be a good empirical choice in the treatment of cystitis [487].
The choice of antibiotics should be guided by good antibiotic stewardship. It is important to be
aware of the local resistance patterns. These are variable between countries and moreover between hospitals.
Local antibiotic protocols and web-based recommendations can guide the choice for type of antibiotic therapy.
The individual patient’s previous urine cultures should also be taken into account in this decision. The daily
dosage of antibiotics is depended on the age, weight of the child as well as on renal and liver function.

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 41


3.9.4.4 Preventative measures
Recurrent UTIs are problematic because the symptoms are bothersome to children and recurrent febrile UTIs
will also result in renal scarring [414]. Therefore, it is important to prevent the incidence of recurrent UTIs.

3.9.4.4.1 Chemoprophylaxis
Chemoprophylaxis is commonly used to prevent UTIs in children. However, with the increasing bacterial
resistance numbers, it should be carefully considered which patients should receive antibacterial prophylaxis.
The evidence for the use of antibacterial prophylaxis has been conflicting. Its use causes a reduction of the
number of recurrent symptomatic UTIs, but long-term use of antibacterial prophylaxis has also been associated
with increased microbial resistance [418, 488]. Its use did not reduce newly acquired renal damage in children
with first or second UTI [488]. However, when used in patients with anatomic abnormalities of the urinary
system a reduction in UTIs and subsequent renal scarring was shown [418, 488]. In children with BBD and
VUR, a benefit was seen in the reduction of recurrent UTI with the use of antomicrobial prophylaxis [489, 490]
(see also Chapter 3.14 on VUR).

For the specific group of patients with incomplete bladder emptying with properly performed clean intermittent
catheterisation but still suffering from recurrent UTIs the intravesical application of gentamycin has proven to
be effective [491].

Table 1: Drugs for antibacterial prophylaxis*

Substance Prophylactic dosage Limitations in neonates and infants


(mg/kg bw/d)
Trimethoprim** 1 Not recommended under six weeks of age
Trimethoprim 1-2 Not recommended under two months of age
Sulfamethoxazole 10-15
Sulfamethoxazole 1-2 Until three months of age
Nitrofurantoin** 1-2 Not recommended under two months of age
Cefaclor 10 No age limitations
Cefixim 2 Preterms and newborns
* Reproduced with permission from the International Consultation on Urological Diseases (ICUD), International
Consultation on Urogenital Infections, 2009. Copyright© by the European Association of Urology [492].
** S
 ubstances of first choice are nitrofurantoin and trimethoprim. In exceptional cases, oral cephalosporin can
be used.

3.9.4.4.2 Dietary supplements


Cranberry, mostly as juice, has been shown to prevent UTIs in healthy children, while in children with urogenital
abnormalities, cranberries appear to be just as effective as antibiotic prophylaxis [493]. The results for
probiotics are somewhat more conflicting, with one systematic review not ruling out any effect [298] and a
RCT showing promising results in children with normal urogenital anatomy [494]. A meta-analysis could not
demonstrate a beneficial effect, only as an adjuvant to antibiotic prophylaxis [495].
Other supplements of interest were Vitamin A, which showed promising results in preventing renal
scarring in children with acute pyelonephritis [496, 497]. The use of Vitamin E could possibly improve the
symptoms of UTI [498]. More studies into these supplements are warranted.

3.9.4.4.3 Preputium
A risk reduction of recurrent UTI regarding the preputium has been shown in two studies. When a physiologic
phimosis is present in boys with a UTI the use of steroid cream significantly reduced recurrent UTIs [499]. In
boys with recurrent UTIs and hydronephrosis present, 10 boys would need to be circumcised to prevent one
UTI [34].

3.9.4.4.4 Bladder and bowel dysfunction


Bladder and bowel dysfunction is a risk factor for which each child with UTI should be screened upon
presentation [409]. Normalisation of micturition disorders or bladder overactivity is important to lower the
rate of UTI recurrence. If there are signs of BBD at infection-free intervals, further diagnosis and effective
treatment are strongly recommended [490]. Treatment of constipation leads to a decrease in UTI recurrence
and a multidisciplinary approach is recommended [409, 489, 490]. Therefore, exclusion of BBD is strongly
recommended in any toilet-trained child with febrile and/or recurrent UTI, and it should be treated (For
treatment see chapter for LUTS).

42 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


3.9.4.5 Monitoring of UTI
With successful treatment, urine usually becomes sterile after 24 hours, and leukocyturia normally disappears
within three to four days. Normalisation of body temperature can be expected within 24-48 hours after the
start of therapy in 90% of cases. In patients with prolonged fever and failing recovery, treatment-resistant
uropathogens or the presence of congenital uropathy or acute urinary obstruction should be considered.
Repeated US examination is recommended in these cases.
Procalcitonin (among other laboratory inflammatory parameters such as C-reactive protein and
leukocyte count) can be used as reliable serum marker for early prediction of renal parenchymal inflammation
[500]. A cut-off value of serum procalcitonin of 1.0 ng/mL has been shown to be predictive of acute
pyelonephritis in young children [501]. In patients with febrile UTI, serum electrolytes and blood cell counts
should be followed up.

3.9.5 Summary of evidence and recommendations for the management of UTI in children

Summary of evidence LE
Urinary tract infection represents the most common bacterial infection in children less than 2 years of 1b
age. The incidence varies depending on age and sex.
Classifications are made according to the site, episode, severity, symptoms and complicating factors. 2b
For acute treatment, site and severity are most important.
The number of colony forming units in the urine culture can vary, however, any colony count of one 2b
specimen indicates a high suspicion for UTI.
Due to increasing resistance numbers good antibiotic stewardship should guide the choice of 2a
antibiotics, taking into account local resistance patterns, old urine cultures (when available) and
clinical parameters.
Preventive measures against recurrent UTIs include: chemoprophylaxis (oral and intravesical), 2a
cranberries, probiotics and Vitamin A and E.
Urinalysis by dipstick yields rapid results, but it should be used with caution. Microscopic investigation 2a
is the standard method of assessing pyuria after centrifugation.
During acute UTI both DMSA and diffusion-weighted MRI can confirm pyelonephritis or parenchymal 2a
damage.

Recommendations LE Strength rating


Take a medical history, assess clinical signs and symptoms and perform a physical 3 Strong
examination to diagnose children suspected of having a urinary tract infection (UTI).
Exclude bladder- and bowel dysfunction in any toilet-trained child with febrile and/or 3 Strong
recurrent UTI.
Clean catch urine can be used for screening for UTI. Bladder catheterisation and 2a Strong
suprapubic bladder aspiration to collect urine can be used for urine cultures.
Do not use plastic bags for urine sampling in non-toilet-trained children since it has 2a Strong
a high risk of false-positive results.
Midstream urine is an acceptable technique for toilet-trained children. 2a Strong
The choice between oral and parenteral therapy should be based on patient age; 2a Strong
clinical suspicion of urosepsis; illness severity; refusal of fluids, food and/or oral
medication; vomiting; diarrhoea; non-compliance; complicated pyelonephritis.
Treat febrile UTIs with four to seven day courses of oral or parenteral therapy. 1b Strong
Treat complicated febrile UTI with broad-spectrum antibiotics. 1b Strong
Offer long-term antibacterial prophylaxis in case of high susceptibility to UTI and risk 1b Strong
of acquired renal damage and lower urinary tract symptoms.
In selected cases consider dietery supplements as an alternative or add-on 2a Strong
preventive measure.
In infants with febrile UTI use renal and bladder ultrasound to exclude obstruction of 2a Strong
the upper and lower urinary tract within 24 hours.
In infants, exclude vesicoureteral reflux after first epidose of febrile UTI with a 2a Strong
non-E. Coli infection. In children more than one year of age with an E. Coli infection,
exclude VUR after the second febrile UTI.

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 43


3.10 Day-time lower urinary tract conditions
3.10.1 Terminology, classification, epidemiology and pathophysiology
Urinary incontinence in children may be caused by congenital anatomical or neurologic abnormalities such as
ectopic ureter, bladder exstrophy or myelomeningocele (MMC). In many children, however, there is no such
obvious cause for the incontinence, and they are referred as having functional bladder problems. The most
recent International Children’s Continence Society (ICCS) document suggests using the term day-time lower
urinary tract (LUT) conditions to group together all functional bladder problems in children.

Normal storage and emptying of the bladder at a socially accepted place and time is mostly achieved by age
three to four. Children with LUT conditions would present with failure to achieve continence (being still wet
after the age of four), urgency, weak stream, hesitancy, frequency and accompanied UTIs. Isolated night-
time wetting without any day-time symptoms is known as ‘enuresis’ and considered as a different entity (see
chapter 3.11) [502].

As different studies have used varying definitions and criteria, it is difficult to give reliable percentages
regarding the incidence of this problem. Reported prevalence ranges widely from 1% to 20% [503-511]. Due
to increasing awareness and better access to specialised health care, the prevalence seems to be increasing
[512, 513].

Lower urinary tract conditions in children may be due to disturbances of the filling phase, the voiding phase
or a combination of both in varying severity. Mainly the conditions are divided into either overactive bladder
(OAB) or dysfunctional voiding. They can, of course, coincide and one may even be causative of the other.
Dysfunctional bowel emptying may also be part of the clinical problems and BBD is the term used to cover
concomitant bladder and bowel disturbances.

Lower urinary tract conditions are considered to be the result of incomplete or delayed maturation of the
bladder sphincter complex. The pons is considered to be responsible for detrusor sphincter co-ordination
while the cortical area is responsible for inhibition of the micturition reflex and voluntary initiation of micturition.
Therefore overactivity would be the result of delayed maturation of cortical control, while dysfunctional voiding
would be the result of non-maturation of the co-ordination. Detrusor overactivity should not be considered as
a sole bladder based problem but more a symptom of a centrally located dysfunction affecting bladder, bowel
and even mood and behaviour [514].

A link between LUT and behavioural disorders such as ADHD (attention deficit/ hyperactivity disorder) has also
been shown [515-517].

3.10.1.1 Filling-phase (storage) dysfunctions


In filling-phase dysfunctions, the detrusor can be overactive, as in OAB, or underactive, as in underactive
bladder (UAB). Overactivity of the bladder is the most common problem, seen mostly around five to seven
years of age. This may lead to disturbances characterised by urgency, frequency and at times urgency
incontinence. Some children habitually postpone micturition leading to voiding postponement. Therefore,
holding manoeuvres such as leg crossing and squatting can often be seen in this group. Recurrent UTIs are
common and high-pressure state of the bladder can be a cause of VUR. Constipation can be an additional
aetiological factor, which needs to be assessed. In children with an underactive detrusor, voiding occurs with
reduced or minimal detrusor contractions with post-void residuals. Urinary tract infections, straining to void,
constipation and incontinence is common. Incontinence often occurs when the bladder is over-distended in the
form of overflow incontinence.

3.10.1.2 Voiding-phase (emptying) dysfunctions


In voiding-phase (emptying), incomplete relaxation or tightening of the sphincteric mechanism and pelvic
floor muscles results in staccato voiding pattern (continuous urine flow with periodic reductions in flow
rate precipitated by bursts of pelvic floor activity) or an interrupted voiding pattern (unsustained detrusor
contractions resulting in infrequent and incomplete voiding, with micturition in fractions). The general term for
this condition is dysfunctional voiding and is associated with elevated bladder pressures and PVRs. Symptoms
will vary depending on the severity of inco-ordination between bladder and the sphincter. Staccato voiding
is in less severe forms and interrupted voiding and straining is in more severe forms. The co-existence of
constipation and LUTD and recurrent UTI is well described [518]. There is no evidence to conclude if bladder
problems or bowel problems are the leading cause. The prevalence of constipation in older children varies from
5 to 27%. Approximately 90% of them being functional constipation without an organic cause In children with
functinal constipation the prevelance of bladder symptoms have been shown to be as high as 64% [519, 520].

44 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


In incomplete emptying, high voiding pressures generated by bladder working against a functional obstruction
caused by non-relaxing sphincter may induce not only UTIs but also VUR. It is been shown that LUTD is
more significant for the occurrence of UTI than VUR itself [521]. In the majority of children with dysfunctional
voiding the recurrent infections disappear following successful treatment, which confirms the hypothesis that
dysfunctional voiding is the main factor responsible for the infections. Spontaneous resolution of VUR may also
be seen after successful treatment of dysfunctional voiding.

3.10.2 Diagnostic evaluation


The evaluation of LUT conditions includes medical and voiding history (bladder diaries and structured
questionnaires), a physical examination, a urinalysis, and uroflowmetry with PVR. The upper urinary tract (UUT)
needs to be evaluated in children with recurrent infections and dysfunctional voiding. Uroflowmetry can be
combined with pelvic floor electromyography to demonstrate overactivity of the pelvic floor muscles during
voiding. Urodynamic studies are usually reserved for patients with therapy resistant dysfunctional voiding and
those not responding to treatment who are being considered for invasive treatment [517, 522-525].

In addition to a comprehensive medical history a detailed voiding diary provides documentation of voiding
and defecation habits, frequency of micturition, voided volumes, night-time urine output, number and timing
of incontinence episodes, and fluid intake. A voiding diary should at least be done for two days, although
longer observation periods are preferred. A voiding diary provides information about storage function and
incontinence frequency, while a pad test can help to quantify the urine loss. In the paediatric age group, where
the history is taken from both the caregivers and child together, a structured approach is recommended using
a questionnaire. Many signs and symptoms related to voiding and wetting will be unknown to the caregivers
and should be specifically requested, using the questionnaire as a checklist. Some symptom scorings have
been developed and validated [526, 527]. Although the reliability of questionnaires are limited they are practical
in a clinical setting to check the presence of symptoms and have also been shown to be reliable to monitor
the response to treatment. History taking should also include assessment of bowel function. For evaluation of
bowel function in children, the Bristol Stool Scale is an easy-to-use tool [528, 529].

Urinalysis and urinary culture are essential to evaluate for UTI. Since transient voiding symptoms are common
in the presence of UTI, exclusion of UTI is essential before further management of symptoms. During clinical
examination, genital inspection and observation of the lumbosacral spine and the lower extremities are
necessary to exclude obvious uropathy and neuropathy.

Uroflowmetry with PVR evaluates the emptying ability, while an UUT US screens for (secondary) anatomical
changes. A flow rate which reaches its maximum quickly and levels off (‘tower shape’) may be indicative of
over-active bladder whereas interrupted or staccato voiding patterns may be seen in dysfunctional voiding.
Plateau uroflowmetry patterns are usually seen in anatomic obstruction of flow. A single uroflowmetry test may
not always be representative of the clinical situation and multiple uroflowmetry tests, which all give a similar
result, are more reliable. Uroflowmetry examination should be done when there is desire to empty the bladder
and the voided volume should at least be 50% of the age-expected capacity [(age in years) + 1] x 30 mL for the
children. While testing the child in a clinical environment, the impact of stress and mood changes on bladder
function should also be taken into account [530, 531].
In the case of treatment failure re-evaluation is warranted and (video)-urodynamic (VUD) studies and
neurological evaluation may be considered. Sometimes, there are minor, underlying, urological or neurological
problems, which can only be suspected using VUD. In these cases, structured psychological interviews to
assess social stress should be added [532] (LE: 1b).
Video-urodynamics may also be used as initial investigational tool in patients with suspicion of
reflux. In this case reflux may be observed along with bladder dynamics. In the case of anatomical problems,
such as posterior urethral valve (PUV) problems, syringocoeles, congenital obstructive posterior urethral
membrane (COPUM) or Moormann’s ring, it may be necessary to perform cystoscopy with treatment. If
neuropathic disease is suspected, MRI of the lumbosacral spine and medulla can help to exclude tethered
cord, lipoma or other rare conditions.

3.10.3 Management
The treatment of LUTD involves a multimodal approach, involving strategies such as behavioural modification,
and anticholinergic medication along with underlying and potentially complicating conditions such as
constipation and UTIs.

Behavioural modification, mostly referred to as urotherapy, is a term which covers all non-pharmacological and
non-surgical treatment modalities. It includes standardisation of fluid intake, bowel management; timed voiding

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 45


and basic relaxed voiding education. The child and family are educated about normal bladder function and
responses to urgency. Voiding regimens are instituted and UTIs and any constipation are treated. Treatment is
aimed at optimising bladder emptying and inducing full relaxation of the urinary sphincter or pelvic floor prior to
and during voiding.

Strategies to achieve these goals include:

1. Information and demystification, which includes explanation about normal LUT function and how a
particular child deviates from normal function.
2. Instructions about what to do about the problem:
• Regular voiding habits, sound voiding posture, pelvic floor awareness and training to relax pelvic
floor and avoiding holding manoeuvres.
• Lifestyle advice, regarding fluid intake, prevention of constipation, etc.
• Registration of symptoms and voiding habits using bladder diaries or frequency-volume charts.
• Support and encouragement via regular follow-up by the caregiver.

Recurrent UTIs and constipation should also be treated and prevented during the treatment period. In case of
combined BBD it is advised to treat the bowel dysfunction first [512] as LUTS may disappear after successful
management of bowel dysfunction.

Addition of other strategies, as below, may be needed:

• Pelvic floor muscle awareness practices with repeated sessions of biofeedback visualisation of uroflow
curves and/or pelvic floor activity and relaxation.
• Clean intermittent self-catheterisation for large PVR volumes of urine.
• Antimuscarinic drug therapy if detrusor overactivity is present.
• If the bladder neck is associated with increased resistance to voiding, α-blocker drugs may be introduced.

Treatment efficacy can be evaluated by improvement in bladder emptying and resolution of associated
symptoms. Controlled studies of the various interventions are needed. As with detrusor overactivity, the
natural history of untreated dysfunctional voiding is not well delineated and optimum duration of therapy is
poorly described. A high success rate has been described for urotherapy programmes, independent of the
components of the programme. However, the evidence level is low as most studies of urotherapy programmes
are retrospective and non-controlled [533]. A recent Cochrane analysis found very little evidence that can help
to make evidene-based treatment decisions [534].

3.10.3.1 Specific interventions


As well as urotherapy, there are some specific interventions, including physiotherapy (e.g. pelvic floor
exercises), biofeedback, alarm therapy and neuromodulation. Although good results with these treatment
modalities have been reported, the level of evidence remains low, since only a few RCTs were published [535-
541].

A systematic review reports that biofeedback is an effective, non-invasive method of treating dysfunctional
voiding, and approximately 80% of children benefited from this treatment. However, most reports were of low
level of evidence and studies of more solid design such as RCTs should be conducted [542].
A more recently published multicentre controlled trial of cognitive treatment, placebo, oxybutynin
and bladder and pelvic floor training did not report better results with oxybutynin and pelvic floor training
compared to standard urotherapy [532] (LE: 1b).

Two RCTs on underactive bladder without neurophatic disease have recently been published. Transcutaneous
interferential electrical stimulation and animated biofeedback with pelvic floor exercise have been
shown to be effective [543, 544]. In some cases, pharmacotherapy may be added. Some studies on
orthosympathicomimetics have been published with a low level of evidence [545].

Overactive bladder is common in the paediatric population. Although a stepwise approach starting with
behavioural therapy is advised, antimuscarinic agents remain the mainstay of medical treatment for OAB.
Oxybutynin is the most commonly used antimuscarinic in the paediatric population. The response to
antimuscarinics varies and many children experience serious side effects. Although there have been reports
about the use of tolterodine, fesoterodine, trospium, propiverine, and solifenacin in children, to date, most
of them are off-label depending on age and national regulations. A few RCTs have been published, one

46 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


on tolterodine showed safety but not efficacy [546], while another on propiverine showed both safety and
efficacy [547] (LE:1). The recent study on solifenacin showed its efficacy with side effects like constipation and
electrocardiogram changes [548].

The difference in results is probably due to study design. Despite the low level of evidence for the use of
anticholinergics and antimuscarinics, their use is recommended because of the large number of studies
reporting a positive effect on OAB symptoms. Although α-blocking agents are used occasionally, an RCT
showed no benefit [549]. Botulinum toxin injection seems promising, but can only be used off-label [550].

A meta-analysis reports that neuromodulation therapy may lead to better partial improvement of non-
neurogenic OAB; however, it may not render a definitive complete response. Office-based neuromodulation
seems more efficacious than self-administered neuromodulation [551]. These new treatment modalities can
only be recommended for standard therapy-resistant cases [552]. Despite early successful treatment, there is
evidence that there is a high recurrence rate of symptoms in the long term which necessitates long-term follow-
up [553]. In addition, many patients may present later in adulthood with different forms of LUTD [554].

3.10.4 Summary of evidence and recommendations for the management of day-time lower urinary
tract conditions

Summary of evidence LE
The term ‘bladder bowel dysfunction’ should be used rather than ‘dysfunctional elimination syndrome 4
and voiding dysfunction’.
Day-time LUTS has a high prevalence (1% to 20%). 2

Recommendations LE Strength rating


Use two day voiding diaries and/or structured questionnaires for objective evaluation 2 Strong
of symptoms, voiding drinking habits and response to treatment.
Use a stepwise approach, starting with the least invasive treatment in managing 4 Weak
day-time lower urinary tract dysfunction in children.
Initially offer urotherapy involving bladder rehabilitation and bowel management. 2 Weak
If bladder bowel dysfunction is present, treat bowel dysfunction first, before treating 2 Weak
the lower urinary tract condition.
Use pharmacotherapy (mainly antispasmodics and anticholinergics) as second line 1 Strong
therapy in overactive bladder.
Use antibiotic prophylaxis if there are recurrent infections. 2 Weak
Re-evaluate in case of treatment failure; this may consist of (video) urodynamics MRI 3 Weak
of lumbosacral spine and other diagnostic modalities, guiding to off-label treatment
which should only be offered in highly experienced centres.

3.11 Monosymptomatic nocturnal enuresis - bedwetting


3.11.1 Epidemiology, aetiology and pathophysiology
Monosymptomatic nocturnal enuresis, also known as bedwetting, is defined as an intermittent nocturnal
incontinence. It is a relatively frequent symptom in children, 5-10% at seven years of age and 1–2% in
adolescents. With a spontaneous yearly resolution rate of 15% (at any age), it is considered as a relatively
benign condition [530, 555]. Seven out of 100 seven-year-old bedwetting children will continue to wet their bed
into adulthood. Nocturnal enuresis is considered primary when a child has not yet had a prolonged period of
being dry (six months). The term “secondary nocturnal enuresis” is used when a child or adult begins wetting
again after having stayed dry.
Non-monosymptomatic nocturnal enuresis is defined as the condition of nocturnal enuresis in
association with day-time lower urinary tracts symptoms (LUTS, recurrent UTIs and/or bowel dysfunction) [555,
556].
Nocturnal enuresis has significant secondary stressful, emotional and social consequences for
the child and their caregivers. Therefore treatment is advised from the age of six to seven years onwards
considering mental status, family expectations, social issues and cultural background.
There is a clear hereditary factor in nocturnal enuresis. If none of the parents or their immediate
relatives has suffered from bedwetting, the child has a 15% chance of wetting its bed. If one of the parents,
or their immediate relatives have suffered from bedwetting, the chance of bedwetting increases to 44%, and
if both parents have a positive history the chance increases to 77%. However, from a genetic point of view,

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 47


enuresis is a complex and heterogeneous disorder. Loci have been described on chromosomes 12, 13 and 22
[556]. There is also a gender difference: two boys to one girl at any age.

High arousal threshold is the most important pathophysiological factor; the child does not wake up when the
bladder is full. In addition to the high arousal threshold, there needs to be an imbalance between night-time
urine output and night-time bladder capacity and activity [530, 555, 556]. Recently, attention has been given
to the chronobiology of micturition in which the existence of a circadian clock in kidney, brain and bladder is
postulated [557] (LE: 1).

A high incidence of comorbidity and correlation between nocturnal urine production and sleep disordered
breathing, such as obstructive sleep apnoea, has been found and investigated. Symptoms such as habitual
snoring, apnoeas, excessive sweating at night and mouth breathing in the patient history or via sleep
questionnaires can lead to the diagnosis of adenotonsillar hypertrophy.

3.11.2 Diagnostic evaluation


The diagnosis is mainly obtained by history-taking. Focused questions to differentiate monosymptomatic
vs. non-monosymptomatic, primary vs. secondary, comorbid factors such as behavioural or psychological
problems and sleep disorder breathing, should be asked. In addition, a two day complete voiding and drinking
diary, which records day-time bladder function and drinking habits will further exclude comorbid factors such
as LUTS and polydipsia.
The night-time urine production should be registered by weighing the night-time diapers in the
morning and adding the first morning voided volume [558]. The night-time urine production should be recorded
over (at least) a two week period to diagnose an eventual differentiation between a high night-time production
(more than 130% of the age expected bladder capacity) vs. a night-time OAB.
A physical examination should be performed with special attention to the external genitalia and
surrounding skin as well as to the condition of the clothes (wet underwear or encopresis).
Urine analysis is indicated if there is a sudden onset of bedwetting, a suspicion or history of UTIs,
or inexplicable polydipsia.

A uroflowmetry and US is indicated only if there is a history of previous urethral or bladder surgery, straining
while voiding, interrupted voiding, an abnormal weak or strong stream, or a prolonged voiding time. If the
comorbid factor of possible sleep disordered breathing occurs, a referral to an ear-nose-throat (ENT) specialist
should be advised.
If the comorbid factor is developmental, attention or learning difficulties, family problems, parental
distress and possible punishment of the child, a referral to a psychologist should be advised and followed-up.

3.11.3 Management
Before introducing any form of possible treatment, it is of utmost importance to explain the bedwetting
condition to the child and the caregivers in order to demystify the problem.

3.11.3.1 Supportive treatment measures


Initially, supportive measures including normal and regular eating and drinking habits should be reviewed,
stressing normal fluid intake during the day and reducing fluid intake in the hours before sleep. Keeping a chart
depicting wet and dry nights, also called as basic bladder advice, has not been shown to be successful in the
early treatment of nocturnal enuresis [559] (LE: 1a).

3.11.3.2 Conservative “wait and see” approach


If the child and its family is unable to comply with a treatment, if the treatment options are not possible for the
family situation, and if there is no social pressure, a “wait and see” approach can be chosen. However, in this
approach, it is important to emphasise the fact that the child should wear diapers at night to ensure a normal
quality of sleep.

3.11.3.3 Nocturnal enuresis wetting alarm treatment


The nocturnal alarm treatment is the use of a device that is activated by getting wet. The goal is that the child
wakes up by the alarm, which can be acoustic or tactile, either by itself or with the help of a caregiver. The
method of action is to repeat the awakening and therefore change the high arousal to a low arousal threshold,
specifically when a status of full bladder is reached. It is of utmost importance that the child is collaborating.
Initial success rates of 80% are realistic, with low relapse rates, especially when night-time diuresis does not
exceed age-expected bladder capacity [560]. Regular follow-up will improve the success.

48 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


3.11.3.4 Medical therapy
In the case of high night-time diuresis, success rates of 70% can be obtained with desmopressin (DDAVP),
either as tablets (200-400 μg), or as sublingual DDAVP oral lyophilisate (120-240 μg). A nasal spray is no longer
recommended due to the increased risk of overdose [561, 562] (LE: 1). Relapse rates can be high after DDAVP
discontinuation [555], however recently, structured withdrawal has shown lower relapse rates [563] (LE: 1).
In the event of desmopressin-resistant treatment for nocturnal enuresis or if a suspicion exists for
night-time OAB, combination with antispasmodics or anticholinergics is safe and efficient [558]. Imipramine,
which has been popular for treatment of enuresis, achieves only a moderate response rate of 50% and has a
high relapse rate. Furthermore, cardiotoxicity and death from overdose are described, its use should therefore
be discouraged as the first-line therapy [564] (LE: 1). Figure 6 presents stepwise assessment and management
options for nocturnal enuresis.

Although several forms of neuromodulation and acupuncture have been investigated for nocturnal enuresis
treatment, present data precludes their use because of its inefficiency, or at least no additional benefit.

Figure 6: A stepwise assessment and management options for nocturnal enuresis

Child ≥ 5 years nocturnal enuresis

2 weeks
Upon
night-time
indication
urine
• Urine
Detailed production
2 days microscopy
questions for Physical recording
day-time • Uroflow -
day-time exam and (= weight
voiding and metry
symptoms urinanalysis night-time
drinking diary • Ultrasound
diapers +
• ENT referral
morning first
• Psychologist
voided
referral
volume)

Supportive measures (not a treatment) (max 4 weeks)


• Normal and regular drinking habits
• Regular voiding and bowel habits
• Monitor night-time production (weight diapers)

Child + caregivers seek a treatment

Nocturnal enuresis
wetting alarm
desmopressin
treatment
+/- anticholinergics
with regular
follow-up

If no improvement (< 4 weeks) OR Lack of compliance


+
Re-evaluate

ENT = ear, nose and throat.

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 49


3.11.4 Summary of evidence and recommendations for the management of monosymptomatic
enuresis

Summary of evidence LE
Chronobiology of micturition, in which the existence of a circadian clock has been proven in kidney, 1
brain and bladder, and disturbances in this chronobiology play a major role in the pathophysiology of
enuresis.

Recommendations LE Strength rating


Do not treat children less than five years of age in whom spontaneous cure is 2 Strong
likely, but inform the family about the involuntary nature, the high incidence of
spontaneous resolution and the fact that punishment will not help to improve the
condition.
Use voiding diaries or questionnaires to exclude day-time symptoms. 2 Strong
Perform a urine test to exclude the presence of infection or potential causes such as 2 Strong
diabetes insipidus.
Offer supportive measures in conjunction with other treatment modalities, of which 1 Strong
pharmacological and alarm treatment are the two most important.
Offer desmopressin in proven night-time polyuria. 1 Strong
Offer alarm treatment in motivated and compliant families. 1 Strong

3.12 Management of neurogenic bladder


3.12.1 Epidemiology, aetiology and pathophysiology
Neurogenic detrusor-sphincter dysfunction (NDSD) can develop as a result of a lesion at any level in the
nervous system. This condition contributes to various forms of LUTD, which may lead to incontinence, UTIs,
VUR, and ultimately to renal scarring and renal failure requiring dialysis and/or transplantation. Conservative
treatment starting in the first year of life is the first choice, however, surgery may be required at a later stage to
establish adequate bladder storage, continence and drainage later on [565-567]. The main goals of treatment
concerning the urinary tract are prevention of UTI’s, urinary tract deterioration, achievement of continence at an
appropriate age and promoting as good a QoL as possible. With regard to the associated bowel dysfunction,
stool continence, with evacuation at a social acceptable moment, is another goal as well as education and
treatment of disturbance in sexual function. Due to the increased risk of development of latex allergy, latex-free
products (e.g., gloves, catheters etc.) should be used from the very beginning whenever possible [568].

Neurogenic bladder in children with myelodysplasia presents with various patterns of Detrusor-Sphincter-
Dyssynergia with a wide range of severity [569]. About 12% of neonates with myelodysplasia have no signs
of neuro-urological dysfunction at birth [570]. Newborns with myelodysplasia and initially normal urodynamic
studies are at risk for neurological deterioration secondary to spinal cord tethering, especially during the
first six years of life. Close follow-up of these children is important for the early diagnosis and timely surgical
correction of tethered spinal cord, and for the prevention of progressive urinary tract deterioration [570]. At
birth, the majority of patients have normal UUTs, but up to 60% develop upper tract deterioration due to
bladder changes, UTI and /or VUR, if not treated properly [571-574]. Even today in a contemporary series
around 50% of the patients are incontinent and 15% have an impaired renal function at the age of 29 years
[575]. A recent systematic review concerning the outcome of adult meningomyelocele patients demonstrated
that around 37% (8-85%) are continent, 25% have some degree of renal damage and 1.3% end stage renal
failure [576]. The term “continence” is used differently in the reports, and the definition of “always dry” was
used in only a quarter of the reports [577].

The most common presentation at birth is myelodysplasia. The incidence of neural tube defects in Europe
is 9.1 per 10,000 births and has not decreased in recent years, despite longstanding recommendations
concerning folic acid supplementations [578]. The term myelodysplasia includes a group of developmental
anomalies that result from defects in neural tube closure. Lesions include spina bifida aperta and occulta,
meningocele, lipomyelomeningocele, or myelomeningocele. Myelomeningocele is by far the most common
defect seen and the most detrimental.
With antenatal screening spina bifida can be diagnosed before birth with the possibility of
intrauterine closure of the defect [579, 580]. Traumatic and neoplastic spinal lesions of the cord are less
frequent in children, but can also cause severe urological problems. Other congenital malformations or
acquired diseases can cause a neurogenic bladder, such as total or partial sacral agenesis which can be part

50 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


of the caudal regression syndrome [581]. In any child presenting with anorectal malformation (ARM) and cloacal
malformations, the development of a neurogenic bladder is possible [582]. Patients with cerebral palsy may
also present with varying degrees of voiding dysfunction, usually in the form of uninhibited bladder contractions
(often due to spasticity of the pelvic floor and sphincter complex) and wetting. Finally, a “non-neurogenic
neurogenic” bladder, such as Hinman or Ochoa syndrome, has been described, in which no neurogenic
anomaly can be found, but severe bladder dysfunction as seen in neurogenic bladders is present [583, 584].

3.12.2 Classification systems


As bladder sphincter dysfunction is poorly correlated with the type and spinal level of the neurological lesion,
urodynamic and functional classifications are much more practical for defining LUT pathology and planning
treatment in children.

The bladder and sphincter are two units working in harmony to act as a single functional unit. In patients with
a neurogenic disorder, the storage and emptying phase of the bladder function can be disturbed. The bladder
and sphincter may function either overactive or underactive and present in four different combinations. This
classification system is based on urodynamic findings [585-587]:

• Overactive sphincter and overactive bladder.


• Overactive sphincter and underactive bladder.
• Underactive sphincter and overactive bladder.
• Underactive sphincter and underactive bladder.

3.12.3 Diagnostic evaluation


Today several guidelines and timetables are used [588-590]. The Panel advocate proactive management in
children with spinal dysraphism. In those with a safe bladder during the first urodynamic investigation, the next
urodynamic investigation can be delayed until one year of age.

3.12.3.1 History and clinical evaluation


History should include questions on clean intermittent catheterisation (CIC) frequency, urine leakage, bladder
capacity, UTI, medication, bowel function as well as changes of neurological status. A thorough clinical
evaluation is mandatory including the external genitalia and the back. A two day diary, recording drinking
volume and times as well as CIC intervals, bladder volume and leakage can provide additional information
about the efficacy of the treatment.

3.12.3.2 Laboratory and urinalysis


After the first week of life, the plasma creatinine level should be obtained, later in life; the cystatin level is
more accurate [591, 592]. If there is any sign of decreased renal function, physicians should be encouraged
to optimise the treatment as much as possible. The criteria for urine analysis are the same as for UTI (refer to
Chapter 3.9). However, it is much easier for caregivers or patients to obtain catheter urine in patients who are on
CIC. They can also perform a dip stick analysis to screen for UTI at home. (For relevance see Section 3.12.4.5)

3.12.3.3 Ultrasound
At birth, US of the kidneys and bladder should be performed and then repeated at least annually. If there are
any clinical changes in between, another US should be performed. Dilatation of the UUT should be reported
according to the classification system of the Society of Foetal Urology [593], including the measurement of the
caliceal dilatation and anterior posterior diameter of the renal pelvis. Residual urine and bladder wall thickness
should also be mentioned. A dilated ureter behind the bladder should be recorded. Bladder wall thickness has
been shown not to be predictive of high pressures in the bladder during voiding and storage and cannot be
used as a non-invasive tool to judge the risk for the UUT [594].

3.12.3.4 Urodynamic studies/videourodynamic


Urodynamic studies (UD) are one of the most important diagnostic tools in patients with neurogenic bladders.
In newborns with spina bifida aperta (failure of mesodermal in-growth over the developing spinal canal results
in an open lesion most commonly seen in the lumbosacral area including an incomplete closure of the vertebral
column and not covered by skin), the first UD should be performed after the phase of the spinal shock after
closure, usually between the second and third months of life [595]. Especially in newborns, performing and
interpretation of UD may be difficult, as no normal values exist. After that it should be repeated annually,
depending on the clinical situation. During and after puberty bladder capacity, maximum detrusor pressure
and detrusor leak point pressure increase significantly [596]. Therefore, during this time, a careful follow-up is
mandatory.

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 51


3.12.3.4.1 Preparation before urodynamic studies
Before any UD a urine analysis should be undertaken. The first assessment should be done under antibiotic
prophylaxis. A Cochrane analysis of nine randomised controlled trials showed, that the administration of
prophylactic antibiotics compared to placebo reduced the risk of significant bacteriuria from 12% to 4% after
UD studies. However, this was without significant difference for symptomatic UTI (20% vs. 28%), fever or
dysuria [597]. If there is a significant bacteriuria, antibacterial treatment should be discussed; especially in older
patients a single shot may be sufficient [598].

Generally UD-parameters should include:


• the bladder cystometric capacity;
• the intravesical filling pressure;
• detrusor compliance;
• the intravesical pressure at the moment of voiding or leakage;
• the presence or absence of overactive detrusor;
• the competence of the internal and external sphincter;
• the degree of synergy of the detrusor and sphincter during voiding;
• the PVR urine volume.

In infants, information on detrusor filling pressure and the pressure and bladder volume at which the child
voids or leaks can be obtained [595]. Detrusor leak point pressure is more accurate than abdominal leak point
pressure, but keeping the rectal probe in an infant in place can be challenging [595]. Addition of fluoroscopy
(video-urodynamic study) will provide information about presence of VUR, at what pressures VUR starts and
the configuration of the bladder neck during filling and leakage or voiding.

3.12.3.4.2 Uroflowmetry
Unlike in children with non-neurogenic voiding dysfunction, uroflowmetry can rarely be used since most
affected patients do not void spontaneously. In those with cerebral palsy, non-neurogenic-neurogenic bladder
or other neurological conditions allowing active voiding it may be a practical tool. It provides an objective way
of assessing the efficiency of voiding, while recording of pelvic floor activity with electromyography (EMG) can
be used to evaluate synergy between detrusor and the sphincter. The PVR urine is measured by US. The main
limitation of uroflowmetry is a compliant child to follow instructions [599-602].

3.12.3.5 Urodynamic studies


The standards of the ICCS should be applied to UDs in patients with neurogenic bladders and accordingly
reported [522, 585]. Natural fill UD in children with neurogenic bladder detects more overactivity compared with
diagnoses delivered by conventional UD [603, 604]. It may be an option in patients where the findings in the
normal UD are inconsistent with clinical symptoms and other clinical findings [604].

3.12.3.6 Voiding cystourethrogram


If video-urodynamic equipment is not available, a VCUG with UD is an alternative to confirm or exclude VUR
and visualise the LUT including the urethra.

3.12.3.7 Renal scan


DMSA (Technetium Dimercapto-Succinic Acid) Renal scan is the gold standard to evaluate renal parenchyma.
In contemporary series, renal scars can be detected in up to 46% as patients get older [605-607]. A positive
DMSA-Scan correlates well with hypertension in adulthood, whereas US has a poor correlation with renal scars
[607]. Therefore, a DMSA scan as a baseline evaluation in the first year of life is recommended.

3.12.4 Management
The medical care of children with neurogenic bladder requires an on-going multidisciplinary approach. There is
some controversy about optimal timing of the management; proactive vs. expectant management [565-567].
Even with a close expectant management e.g. in one series 11 out of 60 need augmentation within a follow-
up of 16 years and 7 out of 58 had a decrease in total renal function, which was severe in two [608]. During
the treatment it should be also taken into account with spina bifida patients, that QoL is related to urinary
incontinence independent from the type and level of spinal dysraphism and the presence or absence of a liquor
shunt [609].

Foetal open and endoscopic surgery for meningomyelocele are performed to close the defect as early as
possible to reduce the neurological, orthopaedic and urological problems [610]. In the MOMS-Trail, Brooks et
al. found no difference between those closed in utero vs. those closed after birth concerning the need for CIC

52 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


[580], but less trabeculation was found in the prenatal surgery group. Mean gestation age (28.3 vs. 35.2) seems
to have no initial impact on bladder function in the first few years of life [611]. Despite some promising reports
[611-614], caregivers need to be aware of the high risk of developing a neurogenic bladder as demonstrated by
a Brazilian group [615]. Regular and close follow-up examinations including UD are indicated in all these patients.

3.12.4.1 Early management with intermittent catheterisation


Starting intermittent catheterisation (IC) soon after birth and closure of the defect by the neurosurgeon in all
infants has shown to decrease renal complications and the need for later augmentation [616-618]. In infants
without any clear sign of outlet obstruction, this may be delayed but only in very selected cases. These infants
should be monitored very closely for UTIs and changes of the urinary tract with US and UD. The early initiation
of IC in the newborn period makes it easier for caregivers to master the procedure and for children to accept it,
as they grow older [619, 620].
A Cochrane review as well as some recent studies showed, that there is a lack of evidence to state
that the incidence of UTI is affected by use of sterile or clean technique, coated or uncoated catheters, single
(sterile) or multiple use (clean) catheters, self-catheterisation or catheterisation by others, or by any other
strategy [621-624]. Looking at the microbiological milieu of the catheter, there was a trend for reduced recovery
of potentially pathogenic bacteria with the use of hydrophilic catheters. Also, a trend for a higher patient
satisfaction with the use of hydrophilic catheters was seen [625]. Based on the current data, it is not possible to
state that one catheter type, technique or strategy is better than another.

3.12.4.2 Medical therapy


Antimuscarinic/anticholinergic medication reduces/prevents detrusor overactivity and lowers intravesical
pressure [626, 627]. Effects and side effects depend on the distribution of the M1-M5 receptors [628]. In the
bladder, the subtype M2 and M3 are present [627, 629]. Oxybutynin is the most frequently used in children
with neurogenic bladder with a success rate of up to 93% [630, 631]. Dose-dependent side-effects (such as
dry mouth, facial flushing, blurred vision heat intolerance etc.) limit its use. Intravesical administration has a
significant higher bioavailability due to the circumvention of the intestinal first pass metabolism, as well as
possible local influence on C-fiber-related activity and can be responsible for different clinical effect [632, 633].
Intravesical administration should be considered in patients with severe side-effects, as long-term results
demonstrated that it was well-tolerated and effective [634, 635]. The transdermal administration leads also to
a substantial lower ratio of N-desethyloxybutynin to oxybutynin plasma levels, however, there are treatment-
related skin reactions in 12 out of 41 patients [636]. There are some concerns about central anticholinergic
adverse effects associated with oxybutynin [637, 638]. A double blinded cross-over trial, as well as a case
control study, showed no deleterious effect on children’s attention and memory [639, 640]. Tolterodine,
solifenacin, trospium chloride and propiverine and their combinations can be also used in children [641-647].
The oral dosage for oxybutynin is up to 0.2 mg/kg/every 8 hours [627] given three times daily. The intravesical
dosage can be up to 0.7 mg/kg/daily and transdermal 1.3-3.9 mg/daily. The dosage of the other drugs is:
Tolterodine 0.5 – 4 mg/day divided in two doses, Solifenacin 1.25 up to 10 mg per day (single dose), Propiverin
0.8 mg/kg/day divided in two dosages and trospium chloride up to 3 times 15 mg starting with 3 times 5 mg.
Except for oxybutynin, all other anticholinergic drugs are off-label use, which should be explained to the caregivers.
Early prophylactic treatment with anticholinergics showed a lower rate of renal deterioration as well
as a lower rate of progression to bladder augmentation [616, 618, 648]. Beta-3 agonists like mirabegron may
also be an alternative agent and may be effective in patients with neurogenic bladders. To date, there is almost
no experience with this drug [649], therefore no recommendation can be made.
Alpha-adrenergic antagonists may facilitate emptying in children with neurogenic bladder [650].
Doxazosin with an initial dose of 0.5 to 1.0 mg or tamsulosin hydrochloride in a medium (0.0002-0.0004 mg/kg/
day) or high dose (0.0004-0,0008 mg/kg/day) has been given to children with neurogenic bladders [650-652]. It
was well tolerated but not effective at least in one study [651].

Botulinum toxin A injections: In neurogenic bladders that are refractory to anticholinergics, the off-label use
of suburothelial or intramuscular injection of onabotulinum toxin A into the detrusor muscle is a treatment
option [653, 654]. In children, continence could be achieved in 32-100% of patients, a decrease in maximum
detrusor pressure of 32% to 54%, an increase of maximum cystometric capacity from 27% to 162%, and an
improvement in bladder compliance of 28%-176% [653]. Onabotulinum toxin A seems to be more effective
in bladders with obvious detrusor muscle over-activity, whereas non-compliant bladders without obvious
contractions are unlikely to respond [655, 656]. Also, the injections into the trigone seems to be save in regard
of reflux and upper tract damage; if it has some benefit is not further investigated [657].
The most commonly used dose of onabotulinum toxin A is 10 to 12 U/kg with a maximum dose
between 200 U and 360 U [653]. However, in one study, 5 U/kg were used with comparable results [658]. To
date, no randomised dose titration study has been published in children. The optimal dose in children as well

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 53


as the time point when to inject which child is still unclear. Onabotolinum toxin A can be effective between
three to twelve (0-25) months and repeated injections are effective up to ten years in one study [654, 659, 660].
Urethral sphincter onabotulinum toxin A injection has been shown to be effective in decreasing
urethral resistance and improve voiding. The evidence is still too low to recommend its routine use in
decreasing outlet resistance, but it could be considered as an alternative in refractory cases [661, 662].

Neuromodulation
Intravesical electrical stimulation of the bladder [663-665], sacral nerve stimulation [666, 667] and
transcutaneous neuromodulation [668] are still experimental and cannot be recommended outside of clinical
trials. The same is true for the intradural somatic-to-autonomic nerve anastomosis [669, 670].

Urethral Dilatation
The aim is to lower the pop-off pressure by lowering the detrusor leak-point pressure by dilatation of the
external sphincter under general anaesthesia up to 36 Charr. Some studies showed, that especially in females,
the procedure is safe and in selected patients, effective [671-673].

Vesicostomy
Vesicostomy - preferably a Blocksom stoma [674] - is an option to reduce bladder pressure in children/
newborns, if the caregivers are incompliant with IC and/or IC through the urethra is extremely difficult or
impossible [675]. Especially in the young infant with severe upper tract dilatation or infections, a vesicostomy
should be considered. Drawbacks are the difficulty fitting and maintaining a collecting appliance in older
patients. A cystostomy button may be an alternative, with a complication rate (mostly UTI) of up to 34% within
a mean follow-up of 37 months [676].

3.12.4.3 Management of faecal incontinence


Children with neurogenic bladder usually have also a neurogenic bowel function. Faecal incontinence may have
an even greater impact on QoL, as the odour can be a reason for social isolation. The aim of each treatment is
to obtain a smooth, regular bowel emptying and to achieve continence and impendence. The regime should be
tailored to the patient’s need, which may change over time. Beside a diet with small portioned fibre food and
adequate fluid intake to keep a good fluid balance [627], follow-up options should be offered to the patients
and caregivers.
In the beginning, faecal incontinence is managed most commonly with mild laxatives, such as
mineral oil, combined with enemas to facilitate removal of bowel contents. To enable the child to defecate
once a day at a given time, rectal suppositories as well as digital stimulation by parents or caregivers can be
used. Today, transanal irrigation is one of the most important treatments for patients with neurogenic bowel
incontinence. Regular irrigations significantly reduce the risk for faecal incontinence and may have a positive
effect on the sphincter tonus as well as the rectal volume [677]. The risk of irrigation induced perforation of
the bowel is estimated as one per 50,000 [678]. During childhood, most children depend on the help of the
caregivers. Later in some patients, transanal irrigation becomes difficult or impossible due to anatomic or social
circumstances. In these patients antegrade irrigation using a MACE-stoma (Malone Antegrade Continence
Enema) is an option, which can also be placed in the left abdomen [679, 680]. In a long-term study of 105
patients, 69% had successful bowel management. They were started on normal saline, but were switched
to GoLYTELY (PEG-3350 and electrolyte solution). Additives (biscodyl, glycerin etc.) were needed in 34% of
patients. Stomal complications occurred in 63% (infection, leakage, and stenosis) of patients, 33% required
surgical revision and 6% eventually required diverting ostomies [681]. In addition, patients need to be informed,
that the antegrade irrigation is also time consuming taking at least 20-60 minutes..

3.12.4.4 Urinary tract infection


Urinary tract infections are common in children with neurogenic bladders. However, there is no consensus
in most European centres, for prevention, diagnosing and treating UTIs in children with neurogenic bladders
performing CIC [682]. Although bacteriuria is seen in more than half of children on CIC, patients who are
asymptomatic do not need treatment [683, 684]. Continuous antibiotic prophylaxis (CAP) creates more
bacterial resistance as demonstrated by a randomised study. Those that stopped the prophylaxis had
reduced bacterial resistance, however, 38 out of 88 started antibiotic prophylaxis again due to recurrent
UTIs or the caregivers request [685]. A cohort study with 20 patients confirmed these findings. Continuous
antibiotic prophylaxix was not protective against the development of symptomatic UTIs and new renal
scarring, but increased the risk of bacterial resistance [686]. A randomised study in 20 children showed that
cranberry capsules significantly reduced the UTI-rate as well as the rate of bacteriuria [687]. If VUR is present,
prophylactic antibiotics should be started when patients experience recurrent UTIs [688, 689].

54 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


3.12.4.4.1 Urinary tract infection and clean intermittent catherisation
The incidence of asymptomatic bacteriuria ranges between 42%-76% [619, 627, 690]. A cross-over study
in 40 children with neurogenic bladder demonstrated, that the reuse of CIC-catheters for up to three weeks
compared to one week increased the prevalence of bacteriuria from 34% to 74% (it was 60% at the start of the
study). During the study-period of eighteen weeks, none of the patients developed a febrile UTI [691]. There is
no medical benefit in performing CAP in children with neurogenic bladder, who perform CIC [627]. In those with
recurrent UTI, intravesical instillation of gentamycin may be an option [692, 693].

Reflux
Secondary reflux in patients with neurogenic bladder increases the risk for pyelonephritis. The treatment is
primary related to bladder function including anticholinergic therapy, CIC and may be later augmentation [694].
Those with early and post-therapy persistent reflux during videourodynamic studies at low pressure have a
higher risk of pyelonephritis [695]. Patients with a high-grade reflux before augmentation have a higher risk of
persistent symptomatic reflux after the enterocystoplasty [696]. Therefore simultaneous ureteral re-implantation
in high-grade symptomatic reflux especially in those with low-pressure high-grade reflux should be discussed
with the patient/caregivers. Endoscopic treatment has a failure rate of up to 75% after a median follow-up
of 4.5 years [697] which is in contrast to the open techniques with a higher success rate, but may have an
increased risk of inducing obstruction [698].

3.12.4.5 Sexuality
Sexuality, while not an issue in childhood, becomes progressively more important as the patient gets older.
This issue has historically been overlooked in individuals with myelodysplasia. However, patients with
myelodysplasia do have sexual encounters [699]. The prevalence of precocious puberty is higher in girls with
meningomyelocele [700]. Studies indicate that at least 15-20% of males are capable of fathering children and
70% of females can conceive and carry a pregnancy to term. It is therefore important to counsel patients about
sexual development in early adolescence.

Women seem to be more sexually active than men in some studies from the USA and the Netherlands [699,
701]; in an Italian study men were more active [701]. The level of the lesion was the main predictor to be
sexually active [702, 703]. Erectile function can be improved by sildenafil in up to 80% of the male patients
[704, 705]. Neurosurgical anastomosis between the inguinal nerve and the dorsal penile nerve in patients with
a lesion below L3 and disturbed sensation is still to be considered as an experimental treatment [701, 706].
Only 17% to one third of the patients talk to their doctors about sexuality, 25–68% were informed by their
doctors about reproductive function [699]. Therefore, early discussion about sexuality in the adolescent is
recommended and should be promoted by the paediatric urologist taking care of these patients.

3.12.4.6 Bladder augmentation


In patients where conservative treatment including onabotulinum toxin A (for indication see 3.12.4.3) fails to
keep a low-pressure reservoir with a good capacity and compliance, bladder augmentation should be offered.
For augmentation, ileal and colonic segments can be used [707]. Gastric segments are rarely used due to its
associated complications like the haematuria-dysuria syndrome as well as secondary malignancies, which
arise earlier than with other intestinal segments [708-711]. Enterocystoplasty increases bladder capacity,
reduces storage pressure and can improve UUT drainage [712]. A good socially acceptable continence rate
can be achieved with or without additional bladder outlet procedures [713]. In those, who are not able to
perform CIC through the urethra, a continent cutaneous channel should be offered. Surgical complications
and revision rate in this group of patients is high. The 30-day all over event rate in the American College
of Surgeons’ National Surgical Quality Database is approximately 30% (23-33%) with a re-operation rate
in this short time period of 13% [714, 715]. In these patients with long-life expectancy the complication
rate clearly increases with the follow-up period [714-717]. The ten-year cumulative complication incidence
from the Paediatric Health Information System showed a rate of bladder rupture in up to 6.4%, small bowel
obstruction in up to 10.3%, bladder stones in 36%, pyelonephritis in more than a third of the patients and a
re-augmentation rate of up to 13% [718]. Bladder perforation, as one of the worst complications, occurs in
6-13% [719]. The rate of VP-shunt infections after gastrointestinal and urological procedures ranges between
0-22%. In a recent study, bowel preparation seems not to have a significant influence on the infection rate
(10.5% vs. 8.3%) [720]. Not only surgical complications must be considered; also metabolic complications and
consequences after incorporating bowel segments have to be taken into account, such as imbalance of the
acid base balance, decrease in vitamin B12 levels and loss of bone density. Stool frequency can increase as
well as diarrhoea after exclusion of bowel segments [721] and last, but not least, these patients have a lifelong
increased risk to develop secondary malignancies [722, 723]. Therefore, a lifelong follow-up of these patients
is required including physical examination, US, blood gas analysis, (pH and base excess), renal function and

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 55


vitamin B12 if Ileum is used. Endoscopic evaluation starting ten years after augmentation is not cost-effective
[724, 725], but may prevent some advanced cancer. Woodhouse et al. do not recommend cystoscopy within
the first fifteen years after surgery [726]. The real value of annual cystoscopic evaluation has not been proven
by any study. Urodynamic studies after bladder augmentation are only indicated, if upper tract dilatation and/or
incontinence after the operation has not improved [727].
Adverse effects of intestinal cystoplasties can be avoided by the use of ureterocystoplasty. The
combination of a small contracted bladder, associated with a severe dilation of the ureter of a non-functioning
kidney is quite rare. The technique was first described in 1973 by Eckstein [728]; the success rate depends on
patient selection and the re-augmentation rate can reach 73% [729, 730].
Auto-augmentation with partial detrusorectomy or detrusormyotomy creating a diverticulum avoids
metabolic complications with the use of intestinal segments. The reports are conflicting, therefore, it may be
used in selected cases [731-734]. For a successful outcome, a pre-operative bladder capacity of 75-80% of
the expected volume seems necessary [735, 736]. Seromuscular cystoplasty has also not proven to be as
successful as standard augmentation with intestine [737]. Tissue engineering, even if successful in vitro and
some animal models, does not reach the results by using intestinal segments with a higher complication rate
[738, 739]. Therefore, these alternatives for bladder augmentation should be considered as experimental and
should be used only in controlled trials.

3.12.4.7 Bladder outlet procedures


No available medical treatment has been validated to increase bladder outlet resistance. Alpha-adrenergic
receptor stimulation of the bladder neck has not been effective [740-745]. Using fascial slings with autologous
fascial strip or artificial material a continence rate between 40-100% can be achieved. In most cases this is
achieved in combination with bladder augmentation [746, 747]. Catheterising through a reconstructed bladder
neck or a urethra compressed by a sling may not be easy; many surgeons prefer to combine this approach
with a catheterisable channel [565]. In contrast to the autologous slings, artificial slings in girls with CIC
through the urethra have a high complication rate [748]. In males, it may be an option [749], however as long
as long-term results are missing, this method has to be classified as experimental and should only be carried
out in studies. Artificial urinary sphincters were introduced by Scott in 1973 [750]. The continence rates in the
literature in selected patients can be up to 85% [751-754]. Post-pubertal patients, who can void voluntary
are good candidates, if they are manually dexterous. In very selected patients, CIC through the sphincter in
an augmented bladder is possible [755]. The erosion rate can be up to 29% and the revision rate up to 100%
depending on the follow-up time [756].

Patients, who underwent a bladder neck procedure only, have a chance of > 30% for an augmentation later
on; half of them developed new upper tract damage in that time [757, 758]. In patients with a good bladder
capacity and bladder compliance without an indication for bladder augmentation, there is a risk of post-
operative changes of the bladder function. Therefore, close follow-up of these patients with UD is required to
avoid upper tract damage and chronic renal failure.

Bladder neck reconstruction is used mostly in exstrophy patients with acceptable results. However, in children
with a neurogenic bladder the results are less favourable [759]. In most patients, the creation of a continent
catheterisable stoma is necessary due to difficulties in performing the CIC via the urethra. In one series, 10%
to a third still performed CIC via the urethra with a re-operation rates between 67% and 79% after a median
follow-up between seven and ten years [760]. In patients who are still incontinent after a bladder outlet
procedure, bladder neck closure with a continent catheterisable stoma is an option. The combination of a sling
procedure together with a urethral lengthening procedure may improve the continence rates [761].

Bulking agents have a low success rate (10-40%), which is in most cases only temporary [762-764]. However, it
does not adversely affect the outcome of further definite surgical procedures [762].
Bladder neck closure is often seen as the last resort to gain urinary continence in those patients
with persistent urinary incontinence through the urethra. In girls, the transection is done between bladder neck
and urethra and in boys above the prostate with preservation of the neurovascular bundle. It is an effective
method to achieve continence together with a catheterisable cutaneous channel +/- augmentation as a primary
or secondary procedure [765, 766]. A complication rate of up to a third and a vesicourethral/vesicovaginal
fistula in up to 15% should be considered [767], together with a higher risk for bladder stones, bladder
perforation and deterioration of the upper tract function, if the patient is not compliant with CIC and bladder
irrigations [767, 768].

56 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


3.12.4.8 Catheterisable cutaneous channel.
In most patients with a neurogenic bladder IC is required. If this is not possible, or very time and/or resources
consuming via the urethra, a continent cutaneous catheterisable channel should be offered as well as in those
with bladder outlet procedures. It is especially beneficial to wheelchair-bound patients who often have difficulty
with urethral catheterisation or are dependent on others to catheterise the bladder. In long-term studies the
revision rate due to stenosis or incontinence can be as high as 50-60% depending on the type of channel
[769, 770].

The stoma can be placed at the umbilicus or in the lower right abdominal wall using a VQZ plasty [771]. It
should be carefully evaluated pre-operatively: it is extremely important that the patient can reach the stoma
easily. Sometimes it has to be placed in the upper abdominal wall due to sever scoliosis mostly associated with
obesity.

3.12.4.9 Continent and incontinent cutaneous urinary diversion


Incontinent urinary diversion should be considered in patients who are not willing or able to perform a CIC and
who need urinary diversion because of upper tract deterioration or gain urinary continence due to social reasons.
In children and adolescents, the colonic conduit has shown to have less complications compared to the ileal
conduit [772-775]. Total bladder replacement is extremely rare in children and adolescents, but may be necessary
in some adults due to secondary malignancies or complications with urinary diversions. Any type of major
bladder and bladder outlet construction should be performed in centres with sufficient experience in the surgical
technique, and with experienced healthcare personnel to carry out post-operative follow-up [713, 776, 777].

Algorithms can be used for management of these patients (Figures 7 and 8).

3.12.5 Follow-up
Neurogenic bladder patients require lifelong follow-up including not only urological aspects but also
neurological and orthopaedic aspects. Regular investigation of upper and lower urinary tract is mandatory.
In patients with changes of the function of the upper and/or lower urinary tract, a complete neurological
re-investigation should be recommended including a total spine MRI to exclude a secondary tethered cord
or worsening of the hydrocephalus. In addition, if some neurological changes are observed a complete
investigation of the urinary tract should be undertaken.

In those patients with urinary tract reconstruction using bowel segments, regulatory investigations concerning
renal function, acid base balance and vitamin B12 status are mandatory to avoid metabolic complications.
There is an increased risk for secondary malignancies in patients with a neurogenic bladder either with or
without enteric bladder augmentations [723, 778-785]. Therefore, patients need to be informed of this risk and
possible signs like haematuria. Although there are poor data on follow-up schemes to discover secondary
malignancies, after a reasonable follow-up time (e.g. ten to fifteen years), an annual cystoscopy can be
considered.

3.12.6 Self-organisation of patients


As patients’ self-organisations can support the parents, caregivers and the patients in all aspects of their daily
life, patients should be encouraged to join these organisations.

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 57


Figure 7a: M
 anagement of children with myelodysplasia with a neurogenic bladder
Flowchart - First year of life

First 12 months

Birth-
discharge
6-12 weeks 6 months 9 months 1 year
from the
hospital

• Bladder-catheter • Medical history • Medical history, • Medical history • Medical history


until closure of • Clinical • Clinical • Clinical • Clinical
the back has examination examination examination examination
healed • Blood pressure • Urine analysis • Urine analysis • Blood pressure
• Then start • Urine analysis, • Check and • Check and • Urine analysis
CIC + AB after • Check + optimise bowel optimise bowel • Check and
peri-operative optimise bowel manageent management optimise bowel
antibiotic is manageent management
finished

At one week: • RBU • RBU • RBU • RBU


• RBU • Creatinine • Creatinine • Creatinine
• Creatinine

• Start CIC + teach • VUD or • CMG if first • If Reflux present


the parents CIC. • VCUG & CMG, CMG showed or febrile UTI,
• After getting if VUD is not a hostile or VUD or VCUG &
comfortable with available non-conclusive CMG if no reflux
the CIC, stop CMG or febrile UTI,
AB. CMG is ok
• Check bowel
management • Start oxybutynin
if any sign • Baseline DMSA
of bladder • If no reflux or
overactivity no UTI and low
grade reflux,
stop AB if given
due to reflux and
• Start AB if reflux monitor urine
and hostile with dip sticks at
bladder or home
non-conclusive
VUD/CMG

RBUS = Renal bladder ultrasound; UTI = urinary tract infection; VUD = videourodynamic;
VCUG = voiding cystourethrography; CMG = cystometrogram; DMSA = dimercaptosuccinic acid;
AB = antibiotics.

58 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


Figure 7b: M
 anagement of children with myelodysplasia with a neurogenic bladder
Flowchart - 18 months - 4 years of age

18 months - 4 years

18 months 2 years 2.5 years 3 years 4 years

• Medical history • Medical history • Medical history • Medical history • Medical history
• Clinical • Clinical • Clinical • Clinical • Clinical
examination examination examination examination examination
• Blood pressure • Blood pressure • Blood pressure • Blood pressure • Blood pressure
• Urine analysis • Urine analysis • Urine analysis • Urine analysis • Urine analysis
• Check and • Check and • Check and • Check and • Check and
optimise bowel optimise bowel optimise bowel optimise bowel optimise bowel
management management management management management
• Check anti- • Check anti- • Check anti- • Check anti- • Check anti-
cholinergic cholinergic cholinergic cholinergic cholinergic
medication + medication + medication + medication + medication +
adapt to weight adapt to weight adapt to weight adapt to weight adapt to weight

• RBUS • RBUS • RBUS • RBUS • RBUS


• Creatinine • Creatinine • Creatinine

• CMG only, if • If reflux present • CMG only, if • If Reflux present • If Reflux present
there is still a or febrile UTI, clinical status or febrile UTI, or febrile UTI,
hostile bladder VUD or VCUG & has changed VUD or VCUG & VUD or VCUG
or clinical status CMG if no reflux CMG if no reflux & CMG if no
has changed or febrile UTI, or febrile UTI, hostile bladder
CMG is ok CMG is ok and clinical no
change CMG at
5 yrs. is ok

• If no reflux or no • If no reflux or no
UTI + low-grade UTI + low grade
reflux, stop AB reflux, stop AB • If no reflux or no
• If AB is given • If AB is given UTI + low grade
due to reflux + due to reflux + reflux, stop AB
monitor urine monitor urine • If AB is given
with dip sticks at with dip sticks at due to reflux +
home home monitor urine
with dip sticks at
home

RBUS = Renal bladder ultrasound; UTI = urinary tract infection; VUD = videourodynamic;
VCUG = voiding cystourethrography; CMG = cystometrogram; DMSA = dimercaptosuccinic acid;
AB = antibiotics

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 59


Figure 7c: M
 anagement of children with myelodysplasia with a neurogenic bladder
Flowchart - 5 years to adulthood

5 years - adulthood

5 years 6 years - puberty yearly Adolescence yearly Adulthood yearly

• Medical history • Medical history • Medical history • Medical history


• Clinical examination • Clinical examination • Clinical examination • Clinical examination
• Blood pressure • Blood pressure • Blood pressure • Blood pressure
• Urine analysis • Urine analysis • Urine analysis • Urine analysis
• Check and optimise • Check and optimise • Check and optimise • Check and optimise
bowel management bowel management bowel management bowel management
• Check anti- • Check anti- • Discuss sexual • Discuss sexual
cholinergic cholinergic function/fertility function + treat
medication + adapt medication + adapt • Check anti- accordingly
to weight to weight cholinergic • Check anti-
medication + adapt cholinergic
to weight medication + adapt
to weight

• RBUS • RBUS • RBUS • RBUS


• Creatinine • Creatinine • Creatinine • Creatinine
• Cystatin C • Cystatin C • Cystatin C • Cystatin C

• If Reflux present or • If no hostile bladder • If no hostile bladder • If no hostile bladder


febrile UTI, VUD or or clinical changes or clinical changes in or clinical changes in
VCUG & CMG biannually CMG a compliant patient a compliant patient
• If no reflux or febrile biannually CMG biannually CMG
UTI, CMG is ok otherwise yearly otherwise yearly

• DMSA scan, if reflux • DMSA scan at age • DMSA scan at age • DMSA scan if
was/is present of 10, if reflux was/is of 15, if reflux was/is indicated
or febrile UTI has present or febrile UTI present or febrile UTI
occurred has occurred has occurred

• In patients with
bowel segments
• In patients with • In patients with incorporated into the
bowel segments bowel segments urinary tract
incorporated into the incorporated into the • Acid-base balance
urinary tract urinary tract • Vitamin B12
• Acid-base balance • Acid-base balance • If pathological -
• Vitamin B12 • Vitamin B12 substitution
• If pathological - • If pathological - • Check for secondary
substitution substitution malignancy

RBUS = Renal bladder ultrasound; UTI = urinary tract infection; VUD = videourodynamic;
VCUG = voiding cystourethrography; CMG = cystometrogram; DMSA = dimercaptosuccinic acid.

60 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


Figure 8: Algorithm for the management of children with myelodysplasia with a neurogenic bladder

Time at diagnosis

Newborn Late presentation

Early CIC

Understanding the detrussor-sphincter


relationship status: history, USG,
VUD/VCU,
nuclear medicine

Detrussor overactive, Detrussor Detrussor underactive, Detrussor underactive,


Sphincter overactive, Sphincter Sphincter underactive
under/normoactive Sphincter overactive over/normoactive

Antimuscarinic Antimuscarinic CIC if residual urine CIC if residual urine


CIC if residual urine CIC CAP if VUR present CAP if VUR present
CAP if VUR present CAP if VUR present

In cases of clinical In cases of clinical Decision given Bladder neck


failure or upper failure or upper regarding the clinical procedures
urinary tract urinary tract situation +/- augmentation
deterioration: deterioration: procedures
Botulinum toxin Botulinum toxin
injection to bladder: injection to bladder
added to treatment or sphincter: added
to treatment

Augmentation Augmentation
procedures procedures

In failed cases bladder neck closure


Or
urinary diversion

CAP = continuous antibiotic prophylaxis; CIC = clean intermittent catheterisation; US = ultrasound;


VCUG = voiding cystourethrography; VUD = videourodynamic; VUR = vesicoureteric reflux.

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 61


3.12.7 Summary of evidence and recommendations for the management of neurogenic bladder

Summary of evidence LE
Neurogenic detrusor-sphincter dysfunction (NDSD) may result in different forms of LUTD and 2a
ultimately result in incontinence, UTIs, VUR, and renal scarring.
In children, the most common cause of NDSD is myelodysplasia (a group of developmental anomalies 2
that result from defects in neural tube closure).
Bladder sphincter dysfunction correlates poorly with the type and level of the spinal cord lesion. 2a
Therefore, urodynamic and functional classifications are more practical in defining the extent of the
pathology and in guiding treatment planning.
Children with neurogenic bladder can have disturbances of bowel function as well as urinary function 2a
which require monitoring and, if needed, management.
The main goals of treatment are prevention of urinary tract deterioration and achievement of 2a
continence at an appropriate age.
Injection of botulinum toxin into the detrusor muscle in children who are refractory to anticholinergics, 2a
has been shown to have beneficial effects on clinical and urodynamic variables.

Recommendations LE Strength rating


Urodynamic studies should be performed in every patient with spina bifida as well 2 Strong
as in every child with high suspicion of a neurogenic bladder to estimate the risk for
the upper urinary tract and to evaluate the function of the detrusor and the sphincter.
In all newborns, intermittent catheterisation (IC) should be started soon after birth. 3 Strong
In those with a clear underactive sphincter and no overactivity starting IC may be
delayed. If IC is delayed, closely monitor babies for urinary tract infections, upper
tract changes (US) and the lower tract (UD).
Start early anticholinergic medication in the newborns with suspicion of an 2 Strong
overactive detrusor.
The use of suburothelial or intradetrusoral injection of onabotulinum toxin A 2 Strong
is an alternative and a less invasive option in children who are refractory to
anticholinergics in contrast to bladder augmentation.
Treatment of faecal incontinence is important to gain continence and independence. 3 Strong
Treatment should be started with mild laxatives, rectal suppositories as well as
digital stimulation. If not sufficient transanal irrigation is recommended, if not
practicable or feasible, a Malone antegrade colonic enema (MACE)/Antegrade
continence enema (ACE) stoma should be discussed.
Ileal or colonic bladder augmentation is recommended in patients with therapy 2 Strong
resistant overactivity of the detrusor, small capacity and poor compliance, which
may cause upper tract damage and incontinence. The risk of surgical and non-
surgical complications and consequences outweigh the risk of permanent damage
of the upper urinary tract +/- incontinence due to the detrusor.
In patients with a neurogenic bladder and a weak sphincter, a bladder outlet 3 Weak
procedure should be offered. It should be done in most patients together with a
bladder augmentation.
Creation of a continent cutaneous catheterisable channel should be offered to 3 Weak
patients who have difficulties in performing an IC through the urethra.
A life-long follow-up of renal and reservoir function should be available and offered 3 Weak
to every patient. Addressing sexuality and fertility starting before/during puberty
should be offered.
Urinary tract infections are common in children with neurogenic bladders, however, 3 Weak
only symptomatic UTIs should be treated.

62 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


3.13 Dilatation of the upper urinary tract (UPJ and UVJ obstruction)
3.13.1 Epidemiology, aetiology and pathophysiology
Dilatation of the upper urinary tract (UUT) remains a significant clinical challenge in deciding which patient
will benefit from treatment. Ureteropelvic junction (UPJ) obstruction is defined as impaired urine flow from the
pelvis into the proximal ureter with subsequent dilatation of the collecting system and the potential to damage
the kidney. It is the most common pathological cause of neonatal hydronephrosis [786]. It has an overall
incidence of 1:1,500 and a ratio of males to females of 2:1 in newborns.
Ureterovesical junction (UVJ) obstruction is an obstructive condition of the distal ureter as it enters
the bladder, commonly called a primary obstructive megaureter. Megaureters are the second most likely cause
of pathological neonatal hydronephrosis. They occur more often in males and are more likely to occur on the
left side [787]. It can be very difficult to define ‘obstruction’ as there is no clear division between ‘obstructed’
and ‘non-obstructed’ urinary tracts. Currently, the most popular definition is that an obstruction represents any
restriction to urinary outflow that, if left untreated, will cause progressive renal deterioration [788].

3.13.2 Diagnostic evaluation


The widespread use of US during pregnancy has resulted in a higher detection rate for antenatal
hydronephrosis [789]. The challenge in the management of dilated UUT is to decide which child should be
observed, which should be managed medically, and which requires surgical intervention. Despite the wide
range of diagnostic tests, there is no single test that can accurately distinguish obstructive from nonobstructive
cases (see Figure 9).

3.13.2.1 Antenatal ultrasound


Usually between the 16th and 18th weeks of pregnancy, the kidneys are visualised routinely, when almost all
amniotic fluid consists of urine. The most sensitive time for foetal urinary tract evaluation is the 28th week. If
dilatation is detected, US should focus on:
• laterality, severity of dilatation, and echogenicity of the kidneys;
• hydronephrosis or hydro-ureteronephrosis;
• bladder volume and bladder emptying;
• sex of the child;
• amniotic fluid volume [790].

3.13.2.2 Postnatal ultrasound


Since transitory neonatal dehydration lasts about 48 hours after birth, imaging should be performed following
this period of postnatal oliguria. However, in severe cases (bilateral dilatation, solitary kidney, oligohydramnios),
immediate postnatal sonography is recommended [791]. Ultrasound should assess the anteroposterior
diameter of the renal pelvis, calyceal dilatation, kidney size, thickness of the parenchyma, cortical echogenicity,
ureters, bladder wall and residual urine.

3.13.2.3 Voiding cystourethrogram


In newborns with identified UUT dilatation, the primary or important associated factors that must be detected
include:
• vesicoureteral reflux (found in up to 25% of affected children) [792];
• urethral valves;
• ureteroceles;
• diverticula;
• neurogenic bladder.
Conventional VCUG is the method of choice for primary diagnostic procedures [793].

3.13.2.4 Diuretic renography


Diuretic renography is the most commonly used diagnostic tool to detect the severity and functional
significance of problems with urine transport. Technetium-99m (99mTc) mercaptoacetyltriglycine (MAG3) is
the radionuclide of choice. It is important to perform the study under standardised circumstances (hydration,
transurethral catheter) after the fourth and sixth weeks of life [794]. Oral fluid intake is encouraged prior to the
examination. At fifteen minutes before the injection of the radionuclide, it is mandatory to administer normal
saline intravenous infusion at a rate of 15 mL/kg over 30 minutes, with a subsequent maintenance rate of 4 mL/
kg/h throughout the entire time of the investigation [795]. The recommended dose of furosemide is 1 mg/kg for
infants during the first year of life, while 0.5 mg/kg should be given to children aged one to sixteen years, up to
a maximum dose of 40 mg.

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 63


Figure 9: Diagnostic algorithm for dilatation of the upper urinary tract

Postnatal US

Dilatation (uni- or bilateral) No dilatation

Voiding cystourethrogram (VCUG)* Repeat US after 4 weeks

Diuretic renography

* A diagnostic work-up including VCUG must be discussed with the caregivers, as it is possible that, even if
reflux is detected, it may have absolutely no clinical impact. However, it should be borne in mind that reflux has
been detected in up to 25% of cases of prenatally detected and postnatally confirmed hydronephrosis [695].
US = ultrasound.

3.13.3 Management
3.13.3.1 Prenatal management
Counselling the caregivers of an affected child is one of the most important aspects of care. The prognosis is
hopeful for a hydronephrotic kidney, even if it is severely affected, as it may still be capable of meaningful renal
function, unlike a severely hypoplastic and dysplastic kidney.
It is important to be able to tell the caregivers exactly when they will have a definitive diagnosis for
their child and what this diagnosis will mean. In some cases, however, it will be immediately obvious that the
child is severely affected; there will be evidence of massive bilateral dilatation, bilateral hypoplastic dysplasia,
progressive bilateral dilatation with oligohydramnios, and pulmonary hypoplasia.
Intrauterine intervention is rarely indicated and should only be performed in well-experienced
centres [796].

3.13.3.1.1 Antibiotic prophylaxis for antenatal hydronephrosis


The benefits and harms of continous antibiotic prophylaxis (CAP) vs. observation in patients with antenatal
hydronephrosis are controversial. Currently, only two RCTs have been published, one of which is a pilot
trial [797] and the other publication is only available as a congress abstract [798]. Both publications present
incomplete data and outcomes.

The Panel conducted a SR assessing the literature from 1980 onwards [799]. The key findings are summarised
below.

Due to the heterogeneity of the published literature it was not possible to draw strong conclusions as to
whether CAP is superior to observation alone in children diagnosed with antibiotic prophylaxis for antenatal
hydronephrosis (ANH). In the first RCT, a prospective longitudinal study [797], female gender, uncircumcised
males, lack of CAP, high-grade hydronephrosis, hydroureteronephrosis and VUR were found to be the
independent predictors for the development of UTI. The second RCT included in the SR, was published as an
abstract only, presented limited data [798]. This trial seemed to focus mainly on patients with ANH and VUR
and did not report any beneficial effect of CAP on UTI rates, but details on the study population were limited.

Key findings of the SR are that CAP may or may not be superior to observation in children with antenatal
hydronephrosis in terms of decreasing UTI. Due to the low data quality it was also not possible to establish
whether boys or girls are at a greater risk of developing a UTI, or ascertain whether the presence or absence
of VUR impacts UTI rates. A correlation between VUR-grade and UTI could not be established either. However,
noncircumcised infants, children diagnosed with high-grade hydronephrosis and hydroureteronephrosis were
shown to be at higher risk of developing a UTI.

The SR also tried to identify the most effective antibiotic regimen and present data on adverse effects but, due
to heterogeneity, the available data could not be statistically compared. The most commonly used antibiotic in
infants with antenatal hydronephrosis is trimethoprim, but only one study reported side effects [797].

64 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


In conclusion, based on the available evidence, the benefits and harms of CAP in children with antenatal
hydronephrosis remain unproven. Uncircumcised infants and infants with hydroureteronephrosis and high-
grade hydronephrosis are more likely to develop a UTI. Continuous antibiotic prophylaxis should be reserved
for this sub-group of children who are proven to be at high risk.

3.13.3.2 UPJ obstruction


It is most important that management decisions are made on the basis of serial investigations that have used
the same technique and have been performed by the same institution under standardised circumstances.
According to a Cochrane review, non-surgical management of unilateral UPJ obstruction in infants less than
2 years old is also an option. However the high risk of bias of the included studies limits the evidence of this
systematic review [800].
Symptomatic obstruction (recurrent flank pain, UTI) requires surgical correction using a pyeloplasty,
according to the standardised open technique of Hynes and Anderson [801]. In experienced hands, laparoscopic
or retroperitoneoscopic techniques and robot-assisted techniques have the same success rates as standard
open procedures. In asymptomatic cases, conservative follow-up is the treatment of choice. A recent
interventional study suggested that, in operated infants less than six months, inserting a stent (transanastomotic
stent) decreases the complication rates compared to stentless approach [802]. However the results should be
taken cautiously since there are successful reported stentless procedures in other age groups.
Indications for surgical intervention comprise impaired split renal function (< 40%), a decrease
of split renal function of > 10% in subsequent studies, poor drainage function after the administration of
furosemide, increased anteroposterior diameter on US, and grade III and IV dilatation as defined by the Society
for Fetal Urology [593].

Well-established benefits of conventional laparoscopy over open surgery are the decreased length of hospital
stay, better cosmesis, less post-operative pain and early recovery [803, 804]. A recent meta-analysis in children
has shown that laparoscopic pyeloplasty (LP) was associated with decreased length of hospital stay and
complication rates but prolonged operative time when compared to open pyeloplasty (OP). Additionally, both LP
and OP had equal success rates [805]. Laparoscopic pyeloplasty can also be performed for re-do cases with the
same advantages of the primary cases [806]. Robotic-assisted laparoscopic pyeloplasty (RALP) has all the same
advantages as LP plus better manoeuvrability, improved vision, ease in suturing and increased ergonomics but
higher costs [807, 808]. A recent study comparing RALP and LP has shown similar postoperative outcomes with
exception of decreased operative time for RALP [809]. There does not seem to be any clear benefit of minimal
invasive procedures in a very young child but current data is insufficient to defer a cut-off age.

3.13.3.3 Megaureter
The treatment options of secondary megaureters are reviewed in Chapter 3.14.3.

3.13.3.3.1 Non-operative management


If a functional study reveals and confirms adequate ureteral drainage, conservative management is the best
option. Initially, low-dose prophylactic antibiotics within the first year of life are recommended for the prevention
of UTIs, although there are no existing prospective randomised trials evaluating the benefit of this regimen
[810]. With spontaneous remission rates of up to 85% in primary megaureter cases, surgical management is
no longer recommended, except for megaureters with recurrent UTIs, deterioration of split renal function and
significant obstruction [811].

3.13.3.3.2 Surgical management


In general, surgery is indicated for symptomatic children, if there is a drop in function in conservative follow-up
and hydroureteronephrosis is increasing [812]. Data suggest that children with a ureteric diameter of > 10-15
mm are more likely to require intervention [813].
The initial approach to the ureter can be either intravesical, extravesical or combined. Straightening
the ureter is necessary without devascularisation. Ureteral tapering should enhance urinary flow into the
bladder. The ureter must be tapered to achieve a diameter for an anti-reflux repair. Several tailoring techniques
exist, such as ureteral imbrication or excisional tapering [814]. Some institutions perform endoscopic stenting,
but there are still no long-term data and no prospective randomised trials to confirm their outcome. A
systematic review assessed the success rates of endoscopic management of primary obstructive megaureters
[815]. It was reported that endoscopic managements including; stent placement, balloon dilatation and incision
can be an alternative treatment in patients > 1 years of age. One third of those patients required further surgical
correction. Furthermore, the long-term outcome of endoscopic management is still unknown. Therefore the
EAU Paediatric Urology Guidelines Panel can not recommend endoscopic management routinely since the
type of intervention and the management outcomes are unclear.

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 65


3.13.4 Conclusion
The use of routine perinatal sonography has resulted in increased detection of hydronephrosis caused by UPJ
or UVJ obstruction. Meticulous and repeat postnatal evaluation is mandatory to try to identify obstructive cases
at risk of renal deterioration and requiring surgical reconstruction. Surgical methods are standardised and have
a good clinical outcome.

3.13.5 Summary of evidence and recommendations for the management of UPJ-, UVJ-obstruction

Summary of evidence LE
Nowadays, most hydronephrotic kidneys have already been diagnosed prenatally during a maternal 2
US investigation.
Ureteropelvic junction obstruction is the leading pathological cause of hydronephrotic kidneys (40%). 1
In children diagnosed with antenatal hydronephrosis, a systematic review could not establish any 1b
benefits or harms related to continuous antibiotic prophylaxis.
In children diagnosed with antenatal hydronephrosis, non-circumcised infants (LE: 1a), children 2
diagnosed with high-grade hydronephrosis (LE: 2) and hydroureteronephrosis (LE: 1b) were shown to
be at higher risk of developing UTI.

Recommendations LE Strength rating


Include serial ultrasound (US) and subsequent diuretic renogram and sometimes 2 Strong
voiding cystourethrography in post-natal investigations.
Offer continuous antibiotic prophylaxis to the subgroup of children with antenatal 2 Weak
hydronephrosis who are at high risk of developing urinary tract infection like
uncircumcised infants, children diagnosed with hydroureteronephrosis and high-
grade hydronephrosis, respectively.
Decide on surgical intervention based on the time course of the hydronephrosis and 2 Weak
the impairment of renal function.
Offer surgical intervention in case of an impaired split renal function due to 2 Weak
obstruction or a decrease of split renal function in subsequent studies and increased
anteroposterior diameter on the US, and grade IV dilatation as defined by the
Society for Fetal Urology.
Offer pyeloplasty when ureteropelvic junction obstruction has been confirmed 2 Weak
clinically or with serial imaging studies proving a substantially impaired or decrease
in function.
Do not offer surgery as a standard for primary megaureters since the spontaneous 2 Strong
remission rates are as high as 85%.

3.14 Vesicoureteric reflux


Lack of robust prospective RCTs limits the strength of the established guidelines for the management of VUR.
The scientific literature for reflux disease is still limited and the level of evidence is generally low. Most of the
studies are retrospective, include different patient groups, and have poor stratification of quality. Also, there
is a high risk of presenting misleading results by combining different types of studies when systematically
extracting data. Therefore, for reflux disease, it is unfortunately not possible to produce recommendations
based on high-quality studies. The Panel have assessed the current literature, but in the absence of conclusive
findings, have provided recommendations based on Panel consensus.

These Guidelines aim to provide a practical approach to the treatment of VUR based on risk analysis and
selective indications for both diagnostics and intervention. Although the Panel have tried to summarise most of
the possible scenarios in one single table, the table itself is still quite busy. The Panel strongly share the view
that making simple and practical guidelines would underestimate the complexity of VUR as a sign of a wide
range of pathologies [816].

3.14.1 Epidemiology, aetiology and pathophysiology


Vesicoureteric reflux is an anatomical and/or functional disorder with potentially serious consequences, such
as renal scarring, hypertension and renal failure. Patients with VUR present with a wide range of severity, and a
good proportion of reflux patients do not develop renal scars and probably do not need any intervention [817].
Vesicoureteric reflux is a very common urological anomaly in children, with an incidence of nearly 1%.

66 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


The main management goal is the preservation of kidney function, by minimising the risk of
pyelonephritis. By defining and analysing the risk factors for each patient (i.e. age, sex, reflux grade, LUTD,
anatomical abnormalities, and kidney status), it is possible to identify those patients with a potential risk of
UTIs and renal scarring. Controversy persists over the optimal management of VUR, particularly the choice of
diagnostic procedures, treatment (medical, endoscopic or surgical), and the timing of treatment.
Many children present without symptoms of UTI and, because invasive diagnostic procedures are
performed only when clinically indicated, the exact prevalence of VUR is unknown. However, the prevalence
of VUR in non-symptomatic children has been estimated at 0.4-1.8% [818]. Among infants prenatally identified
with hydronephrosis on US, who were screened for VUR, the prevalence was 16.2% (7-35%) [819]. Siblings
of children with VUR had a 27.4% (3-51%) risk of also having VUR, whereas the offspring of parents with VUR
had a higher incidence of 35.7% (21.2-61.4%) [819].
However, reflux detected by sibling screening is associated with lower grades [718] and significantly
earlier resolution [820]. When VUR is discovered in siblings after UTI, it is usually high-grade and associated
with a high incidence of reflux nephropathy, particularly if the sibling is male and the grade of reflux was high in
the index patient [821].
The incidence of VUR is much higher among children with UTIs (30-50%, depending on age).
Urinary tract infections are more common in girls than boys due to anatomical differences. However, among all
children with UTIs, boys are more likely to have VUR than girls (29% vs. 14%). Boys also tend to have higher
grades of VUR diagnosed at younger ages, although their VUR is more likely to resolve itself [822-825].
There is a clear co-prevalence between LUTD and VUR [826]. Lower urinary tract dysfunction refers
to the presence of LUTS, including urge, urge incontinence, weak stream, hesitancy, frequency and UTIs, which
reflect the filling and/or emptying dysfunction and may be accompanied with bowel problems [826]. Some
studies have described a prevalence of 40-60% for VUR in children with LUTD [827]. A published Swedish
reflux trial has demonstrated LUTD in 34% of patients, and subdivision into groups characteristic of children
revealed that 9% had isolated overactive bladder and 24% had voiding phase dysfunction [828].
The spontaneous resolution of VUR is dependent on age at presentation, sex, grade, laterality,
mode of clinical presentation, and anatomy [820]. Faster resolution of VUR is more likely with age less than
one year at presentation, lower grade of reflux (grade 1-3), and asymptomatic presentation with prenatal
hydronephrosis or sibling reflux. The overall resolution rate is high in congenital high-grade VUR during the
first years of life. In several Scandinavian studies, the complete resolution rate for high-grade VUR has been
reported at > 25%, which is higher than the resolution rate for VUR detected after infancy [828-830].
The presence of renal cortical abnormality, bladder dysfunction, and breakthrough febrile UTIs are
negative predictive factors for reflux resolution [831-833].
Dilating VUR increases the risk of developing acute pyelonephritis and renal scarring. Untreated
recurrent UTIs may have a negative impact on somatic growth and medical status of the child. Evidence
of renal scarring is present in 10-40% of children with symptomatic VUR, resulting from either congenital
dysplasia and/or acquired post-infectious damage, which may have a negative impact on somatic growth and
general well-being [834-836].
Scar rates vary in different patient groups. Patients with higher grades of VUR present with higher
rates of renal scars. In those with prenatal hydronephrosis, renal scarring occurs in 10% of patients [837-842],
whereas in patients with LUTD, this may increase up to 30% [605, 836, 843]. Renal scarring may adversely
affect renal growth and function, with bilateral scarring increasing the risk of insufficiency. Reflux nephropathy
(RN) may be the most common cause of childhood hypertension. Follow-up studies have shown that 10-20%
of children with RN develop hypertension or end-stage renal disease [844].

3.14.2 Diagnostic evaluation


The diagnostic work-up should aim to evaluate the overall health and development of the child, the presence
of UTIs, renal status, the presence of VUR, and LUT function. A basic diagnostic work-up comprises a detailed
medical history (including family history, and screening for LUTD), physical examination including blood
pressure measurement, urinalysis (assessing proteinuria), urine culture, and serum creatinine in patients with
bilateral renal parenchymal abnormalities.
The standard imaging tests include renal and bladder US, VCUG and nuclear renal scans. The
criterion standard in diagnosis of VUR is VCUG, especially at the initial work-up. This test provides precise
anatomical detail and allows grading of VUR [845]. In 1985, the International Reflux Study Committee
introduced a uniform system for the classification of VUR [846, 847] (Table 2). The grading system combines
two earlier classifications and is based upon the extent of retrograde filling and dilatation of the ureter, renal
pelvis and calyces on VCUG [847].
Radionuclide studies for detection of reflux have lower radiation exposure than VCUG, but the
anatomical details depicted are inferior [848]. Recent studies on alternative imaging modalities for detection
on VUR have yielded good results with voiding US and magnetic resonance VCUG [849-851]. Contrast

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 67


enhanced voiding urosonography with intravesical instillation of different ultrasound contrast agents has been
shown to be highly sensitive giving comparable results with conventional VCUG while avoiding exposure to
ionising radiation [459, 852]. However, despite the concerns about ionising radiation and its invasive nature,
conventional VCUG still remains the gold standard because it allows better determination of the grade of VUR
(in a single or duplicated kidney) and assessment of the bladder and urethral configuration.

Table 2: Grading system for VUR on VCUG, according to the International Reflux Study Committee [853]

Grade I Reflux does not reach the renal pelvis; varying degrees of ureteral dilatation
Grade II Reflux reaches the renal pelvis; no dilatation of the collecting system; normal fornices
Grade III Mild or moderate dilatation of the ureter, with or without kinking; moderate dilatation of the
collecting system; normal or minimally deformed fornices
Grade IV Moderate dilatation of the ureter with or without kinking; moderate dilatation of the collecting
system; blunt fornices, but impressions of the papillae still visible
Grade V Gross dilatation and kinking of the ureter, marked dilatation of the collecting system; papillary
impressions no longer visible; intraparenchymal reflux

Dimercaptosuccinic acid is the best nuclear agent for visualising the cortical tissue and differential function
between both kidneys. Dimercaptosuccinic acid is taken up by proximal renal tubular cells and is a good
indicator of renal parenchyma function. In areas of acute inflammation or scarring, DMSA uptake is poor and
appears as cold spots. Dimercaptosuccinic acid scans are therefore used to detect and monitor renal scarring.
A baseline DMSA scan at the time of diagnosis can be used for comparison with successive scans later during
follow-up [854]. Dimercaptosuccinic acid can also be used as a diagnostic tool during suspected episodes
of acute pyelonephritis [855]. Children with a normal DMSA scan during acute UTI have a low-risk of renal
damage [855, 856].
Video-urodynamic studies are only important in patients in whom secondary reflux is suspected,
such as those with spina bifida or boys in whom VCUG is suggestive of PUV. In the case of LUTS, diagnosis
and follow-up can be limited to non-invasive tests (e.g. voiding charts, US, or uroflowmetry) [826]. Cystoscopy
has a limited role in evaluating reflux, except for infravesical obstruction or ureteral anomalies that might
influence therapy.

3.14.2.1 Infants presenting with prenatally diagnosed hydronephrosis


Ultrasound of the kidney and bladder is the first standard evaluation tool for children with prenatally
diagnosed hydronephrosis. It is non-invasive and provides reliable information regarding kidney structure, size,
parenchymal thickness and collecting system dilatation [857, 858].
Ultrasound should be delayed until the first week after birth because of early oliguria in the neonate.
It is essential to evaluate the bladder, as well as the kidneys. The degree of dilatation in the collecting system
under US, when the bladder is both full and empty, may provide significant information about the presence of
VUR. Bladder wall thickness and configuration may be an indirect sign of LUTD and reflux. The absence of
hydronephrosis on postnatal US excludes the presence of significant obstruction; however, it does not exclude
VUR.
Monitoring with careful US avoids unnecessary invasive and irradiating examinations. The first two
US scans within the first one to two months of life are highly accurate for defining the presence or absence
of renal pathology. In infants with two normal, successive scans, VUR is rare, and if present it is likely to be
low-grade [837, 859]. The degree of hydronephrosis is not a reliable indicator for the presence of VUR, even
though cortical abnormalities are more common in high-grade hydronephrosis [819]. The presence of cortical
abnormalities on US (defined as cortical thinning and irregularity, as well as increased echogenicity) warrants
the use of VCUG for detecting VUR [819]. Dimercaptosuccinic acid provides more reliable and quantitative
measurement of the degree of cortical abnormalities, first detected with US.
The use of VCUG is recommended in patients with US findings of bilateral high-grade
hydronephrosis, duplex kidneys with hydronephrosis, ureterocele, ureteric dilatation, and abnormal bladders,
because the likelihood of VUR is much higher. In all other conditions, the use of VCUG to detect reflux is
optional [819, 839, 860-862].

When infants who are diagnosed with prenatal hydronephrosis become symptomatic with UTIs, further
evaluation with VCUG should be considered [861]. Patients with severe hydronephrosis and those whose
hydronephrosis is sustained or progressive, need further evaluation to exclude obstruction.

68 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


3.14.2.2 Siblings and offspring of reflux patients
The screening of asymptomatic siblings and offspring is controversial. Some authors think that early
identification of children with VUR may prevent episodes of UTI and therefore renal scarring, whereas others
think that screening asymptomatic individuals is likely to result in significant over-treatment of clinically
insignificant VUR. In screened populations the prevalence of VUR is 27.4% in siblings and 35.7% in offspring
[853]. The overall estimate for renal cortical abnormalities is 19.3% (11-54%), with 27.8% having renal
damage in cohorts of symptomatic and asymptomatic children combined. In asymptomatic siblings only,
the rate of renal damage is 14.4% (0-100%). Although early screening and therefore early diagnosis and
treatment appears to be more effective than late screening in preventing further renal damage [819, 821,
863, 864], screening in all siblings and offspring cannot be recommended based on the available evidence.
The lack of RCTs for screened patients to assess clinical health outcomes makes evidence-based guideline
recommendations difficult.

3.14.2.3 Recommendations for paediatric screening of VUR

Recommendations Strength rating


Inform parents of children with vesicoureteric reflux (VUR) that siblings and offspring have a Strong
high prevalence of VUR.
Use renal ultrasound (US) for screening of sibling(s). Strong
Use voiding cystourethrography if there is evidence of renal scarring on US or a history of Weak
urinary tract infection.
Do not screen older toilet-trained children since there is no added value in screening for Weak
VUR.

3.14.2.4 Children with febrile urinary tract infections


A routine recommendation of VCUG at zero to two years of age after the first proven febrile UTI is the safest
approach as the evidence for the criteria to selecting patients for reflux detection is weak. Children with febrile
infections and abnormal renal US findings may have higher risk of developing renal scars and they should all be
evaluated for reflux [462]. If reflux is diagnosed, further evaluation has traditionally consisted of a DMSA scan.
An alternative “top-down” approach is also an option, as suggested by several studies in the
literature. This approach carries out an initial DMSA scan close to the time of a febrile UTI, to determine the
presence of pyelonephritis, which is then followed by VCUG if the DMSA scan reveals kidney involvement. A
normal DMSA scan with no subsequent VCUG will fail to identify VUR in 5-27% of cases, with the missed VUR
presumably being less significant. In contrast, a normal DMSA scan with no VCUG avoids unnecessary VCUG
in > 50% of those screened [452, 865-867].

3.14.2.5 Children with lower urinary tract symptoms and vesicoureteric reflux
Detection of LUTD is essential in treating children with VUR. It is suggested that reflux with LUTD resolves
faster after LUTD correction, and that patients with LUTD are at higher risk for developing UTI and renal
scarring [825, 868]. The co-existence of both conditions should be explored in any patient who has VUR.
If there are symptoms suggestive of LUTD (e.g. urgency, wetting, constipation or holding manoeuvres), an
extensive history and examination, including voiding charts, uroflowmetry and residual urine determination, will
reliably diagnose underlying LUTD.
Among toilet-trained children, those with both LUTD and VUR are at higher risk of developing
recurrent UTIs than children with isolated VUR [490].
In LUTD, VUR is often low-grade and US findings are normal, and there is no indication for
performing VCUG in all children with LUTD, but the presence of febrile infections should be meticulously
investigated. The co-existence of LUTD and VUR means it would be better to do a test covering both
conditions, such as a VUDS. Any patient with LUTD and a history of febrile UTI should be investigated with a
VUDS, if available. Furthermore, any child who fails standard therapy for LUTD should undergo urodynamic
investigation. At this stage, combining a urodynamic study with VCUG is highly recommended.

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 69


3.14.3 Disease management
There are two main treatment approaches: conservative (non-surgical and surgical).

3.14.3.1 Non-surgical therapy


The objective of conservative therapy is prevention of febrile UTI. It is based on the understanding that:
• Vesicoureteric reflux resolves spontaneously, mostly in young patients with low-grade reflux. Resolution is
nearly 80% in VUR grades I and II and 30-50% in VUR grades III-V within four to five years of follow-up.
• Spontaneous resolution is low for bilateral high-grade reflux [869].
• Vesicoureteric reflux does not damage the kidney when patients are free of infection and have normal
LUT function.
• There is no evidence that small scars can cause hypertension, renal insufficiency or problems during
pregnancy. Indeed, these are possible only in cases of severe bilateral renal damage.
• The conservative approach includes watchful waiting, intermittent or continuous antibiotic prophylaxis,
and bladder rehabilitation in those with LUTD [605, 868, 870-872].
• Circumcision during early infancy may be considered as part of the conservative approach because it is
effective in reducing the risk of infection in normal children [873].

3.14.3.1.1 Follow-up
Regular follow-up with imaging studies (e.g. VCUG, nuclear cystography, or DMSA scan) is part of the
conservative management to monitor spontaneous resolution and kidney status. Conservative management
should be dismissed in all cases of febrile breakthrough infections, despite prophylaxis, and intervention should
be considered.

3.14.3.1.2 Continuous antibiotic prophylaxis


Vesicoureteral reflux increases the risk of UTI and renal scarring especially when in combination with LUTD.
Many prospective studies have evaluated the role of continuous antibiotic prophylaxis in the prevention of
recurrent UTI and renal scarring.
It is clear that antibiotic prophylaxis may not be needed in every reflux patient [874-876]. Trials
show the benefit of CAP is none or minimal in low-grade reflux. Continuous antibiotic prophylaxis is useful
in patients with grade III and IV reflux in preventing recurrent infections but its use in preventing further renal
damage is not proven. Toilet-trained children and children with LUTD derive better benefit from CAP [876-881].
The RIVUR trial was the largest, randomised, placebo-controlled, double blind, multi-centre study, involving
607 children aged 2-72 months with grade I-IV VUR. The RIVUR study showed that prophylaxis reduced the
risk of recurrent UTI by 50% but not renal scarring and its consequences (hypertension and renal failure), at
the cost of increased antimicrobial resistance. The benefit of prophylaxis was insignificant in patients with
grade III or IV VUR and in the absence of LUTD [882-885]. Additional review of the RIVUR data based on a risk
classification system defines a high-risk group (uncircumcised males; presence of BBD and high grade reflux)
who would benefit from a antibiotic prophylaxis significantly. Therefore selective prophylaxis for this group is
recommended [886].
It may be difficult and risky to select patients who do not need CAP. A safe approach would be
to use CAP in most cases. Decision-making may be influenced by the presence of risk factors for UTI, such
as young age, high-grade VUR, status of toilet-training/LUTS, female sex, and circumcision status. Although
the literature does not provide any reliable information about the duration of CAP in reflux patients, a practical
approach would be to use CAP until after children have been toilet-trained and ensuring that there is no
LUTD. Continuous antibiotic prophylaxis is mandatory in patients with LUTD and reflux. Active surveillance
of UTI is needed after CAP is discontinued. The follow-up scheme and the decision to perform an anti-reflux
procedure or discontinuation of CAP may also depend on personal preferences and the attitude of patients and
caregivers. It is strongly advised that the advantages and disadvantages should be discussed in detail with the
family.

3.14.3.2 Surgical treatment


Surgical treatment can be carried out by endoscopic injection of bulking agents or ureteral re-implantation.

3.14.3.2.1 Subureteric injection of bulking materials


With the availability of biodegradable substances, endoscopic subureteric injection of bulking agents has
become an alternative to long-term antibiotic prophylaxis and open surgical intervention in the treatment of
VUR in children. Using cystoscopy, a bulking material is injected beneath the intramural part of the ureter in
a submucosal location. The injected bulking agent elevates the ureteral orifice and the distal ureter, so that
coaptation is increased. This results in narrowing of the lumen, which prevents reflux of urine into the ureter,
while still allowing its antegrade flow.

70 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


Several bulking agents have been used over the past two decades, including
polytetrafluoroethylene (PTFE or Teflon™), collagen, autologous fat, polydimethylsiloxane, silicone,
chondrocytes, a solution of dextranomer/hyaluronic acid (Deflux™, Dexell®) and more recently
polyacrylatepolyalcohol copolymer hydrogel (Vantris®) [887, 888].
Although the best results have been obtained with PTFE [889], due to concerns about particle
migration, PTFE has not been approved for use in children [890]. Although they are all biocompatible, other
compounds such as collagen and chondrocytes have failed to provide a good outcome. Deflux™ was
approved by the USA FDA in 2001 for the treatment of VUR in children. Initial clinical trials have demonstrated
that this method is effective in treating reflux [891]. Studies with long-term follow-up have shown that there
is a high recurrence rate which may reach as high as 20% in two years [876]. In a meta-analysis [892] of
5,527 patients and 8,101 renal units, the reflux resolution rate (by ureter) following one treatment for grades
I and II reflux was 78.5%, 72% for grade III, 63% for grade IV, and 51% for grade V. If the first injection was
unsuccessful, the second treatment had a success rate of 68% and the third treatment 34%. The aggregate
success rate with one or more injections was 85%. The success rate was significantly lower for duplicated
(50%) vs. single (73%) systems, and neuropathic (62%) vs. normal (74%) bladders.

Obstruction at UVJ may happen in the long term follow-up after endoscopic correction of reflux. Patients
with high-grade reflux and dilated ureters are at risk of late obstruction. It is significantly more common when
polyacrylate-polyalcohol copolymer is used as bulking substance [893-895].
Clinical validation of the effectiveness of anti-reflux endoscopy is currently hampered by the lack
of methodologically appropriate studies. In the most recent prospective, randomised trials comparing three
treatment arms: i) endoscopic injection; ii) antibiotic prophylaxis; iii) surveillance without antibiotic prophylaxis
in 203 children aged one to two years with grade III/IV reflux, endoscopic treatment gave the highest resolution
rate of 71% compared to 39% and 47% for treatment arms ii and iii, respectively, after two years’ follow-up.
The recurrence rate at two years after endoscopic treatment was 20%. The occurrence of febrile UTIs and
scar formation was highest in the surveillance group at 57% and 11%, respectively. New scar formation rate
was higher with endoscopic injection (7%) compared with antibiotic prophylaxis (0%) [896]. Longer follow-up
studies are needed to validate these findings.

3.14.3.2.2 Open surgical techniques


Various intra- and extravesical techniques have been described for the surgical correction of reflux. Although
different methods have specific advantages and complications, they all share the basic principle of lengthening
the intramural part of the ureter by submucosal embedding of the ureter. All techniques have been shown to be
safe with a low rate of complications and excellent success rates (92-98%) [897].
The most popular and reliable open procedure is cross trigonal re-implantation described by
Cohen [895]. The main concern with this procedure is the difficulty of accessing the ureters endoscopically, if
needed, when the child is older. Alternatives are suprahiatal re-implantation (Politano-Leadbetter technique)
and infrahiatal re-implantation (Glenn-Anderson technique). If an extravesical procedure (Lich-Gregoir) is
planned, cystoscopy should be performed pre-operatively to assess the bladder mucosa and the position and
configuration of the ureteric orifices. In bilateral reflux, an intravesical anti-reflux procedure may be considered,
because simultaneous bilateral extravesical reflux repair carries an increased risk of temporary post-operative
urine retention [898]. Overall, all surgical procedures offer very high and similar success rates for correcting
VUR.

3.14.3.2.3 Laparoscopy and robot-assisted


There have been a considerable number of case series of transperitoneal, extravesical and
pneumovesicoscopic intravesical ureteral re-implantation, which have shown the feasibility of the techniques.
Various anti-reflux surgeries have been performed with the robot and the extravesical approach is the most
commonly used. Although initial reports give comparable outcomes to their open surgical counterparts in terms
of successful resolution of reflux, recent meta-analysis of results of Robotic-Assisted Laparoscopic Ureteral
Reimplantation (RALUR) are within a wide range of variation and on average they are poor compared to open
surgery. Operative times, costs and post-operative complications leading to secondary interventions are higher
with RALUR but post-operative pain and hospital stay is less compared to open surgery [899-902].
Also, laparoscopic- or robotic-assisted approaches are more invasive than endoscopic correction
and their advantages over open surgery are still debated. Therefore, at present, a laparoscopic approach
cannot be recommended as a routine procedure. It can be offered as an alternative to the caregivers in centres
where there is established experience [873, 903-911].

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 71


3.14.4 Summary of evidence and recommendations for the management of vesicoureteric reflux in
childhood

Summary of evidence
There is no evidence that correction of persistent low-grade reflux (grades I-III) without symptoms and normal
kidneys offers a significant benefit.
The traditional approach of initial medical treatment after diagnosis and shifting to interventional treatment
in case of breakthrough infections and new scar formation needs to be challenged, because the treatment
should be tailored to different risk groups.
Surgical correction should be considered in patients with persistent high-grade reflux (grades IV/V). There is
no consensus about the timing and type of surgical correction. The outcome of open surgical correction is
better than endoscopic correction for higher grades of reflux, whereas satisfactory results can be achieved by
endoscopic injection for lower grades.
The choice of management depends on the presence of renal scars, clinical course, grade of reflux, ipsilateral
renal function, bilaterality, bladder function, associated anomalies of the urinary tract, age, compliance, and
parental preference. Febrile UTI, high-grade reflux, bilaterality, and cortical abnormalities are considered to be
risk factors for possible renal damage. The presence of LUTD is an additional risk factor for new scars.

Recommendations Strength rating


Initially treat all patients diagnosed within the first year of life with continuous antibiotic Weak
prophylaxis, regardless of the grade of reflux or presence of renal scars.
Offer immediate, parenteral antibiotic treatment for febrile breakthrough infections. Strong
Offer definitive surgical or endoscopic correction to patients with frequent breakthrough Weak
infections.
Offer open surgical correction to patients with persistent high-grade reflux and endoscopic Strong
correction for lower grades of reflux.
Initially manage all children presenting at age one to five years conservatively. Strong
Offer surgical repair to children above the age of one presenting with high-grade reflux and Weak
abnormal renal parenchyma.
Offer close surveillance without antibiotic prophylaxis to children presenting with lower Strong
grades of reflux and without symptoms.
Ensure that a detailed investigation for the presence of lower urinary tract dysfunction Strong
(LUTD) is done in all and especially in children after toilet-training. If LUTD is found, the
initial treatment should always be for LUTD.
Offer surgical correction, if parents prefer definitive therapy to conservative management. Strong
Select the most appropriate management option based on: Weak
• the presence of renal scars;
• clinical course;
• the grade of reflux;
• ipsilateral renal function;
• bilaterality;
• bladder function;
• associated anomalies of the urinary tract;
• age and gender;
• compliance;
• parental preference.
Refer to Table 3 for risk factors and follow-up.
In high-risk patients who already have renal impairment, a more aggressive, multidisciplinary Strong
approach is needed.

72 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


Table 3: Management and follow-up according to different risk groups

Risk Groups Presentation Initial treatment Comment Follow-up


High Symptomatic male Initial treatment is Greater possibility of More aggressive
or female patients always for LUTD with earlier intervention follow-up for UTI
after toilet-training CAP; intervention and LUTD; full
with high-grade may be considered in re-evaluation after 6
reflux (grades IV-V), cases of BT infections months
abnormal kidneys and or persistent reflux
LUTD
High Symptomatic male Intervention should be Open surgery has Post-operative VCUG
or female patients considered better results than on indication only;
after toilet-training endoscopic surgery follow-up of kidney
with high-grade reflux status until after
(grade IV-V), abnormal puberty
kidneys and no LUTD
Moderate Symptomatic male or CAP is the Spontaneous Follow-up for UTI/
female patients beforeinitial treatment. resolution is higher in hydronephrosis; full
toilet-training, with Intervention may be males re-evaluation after
high-grade reflux and considered in cases 12-24 months
abnormal kidneys of BT infections or
persistent reflux
Moderate Asymptomatic CAP is the Follow-up for UTI/
patients (PNH or initial treatment. hydronephrosis; full
sibling) with high- Intervention may be re-evaluation after
grade reflux and considered in cases 12-24 months
abnormal kidneys of BT, infections or
persistent reflux
Moderate Symptomatic male or Initial treatment is In case of persistent Follow-up for UTI and
female patients after always for LUTD with LUTD, despite LUTD, kidney status;
toilet-training, with CAP. Intervention urotherapy, full re-evaluation after
high-grade reflux and may be considered in intervention should successful urotherapy
normal kidneys with cases of BT infections be considered. The
LUTD or persistent reflux choice of intervention
is controversial
Moderate Symptomatic male Choice of treatment Follow-up for UTI,
or female patients is controversial. LUTD, and kidney
after toilet-training Endoscopic treatment status until after
with low-grade reflux, may be an option. puberty
abnormal kidneys LUTD treatment
with or without LUTD should be given if
needed
Moderate All symptomatic Initial treatment is Follow-up for UTI and
patients with normal always for LUTD with LUTD
kidneys, with low- or without CAP
grade reflux, with
LUTD
Low All symptomatic No treatment or CAP If no treatment is Follow-up for UTI
patients with normal given, parents should
kidneys, with low- be informed about
grade reflux, with no risk of infection
LUTD
Low All asymptomatic No treatment or CAP If no treatment is Follow-up for UTI
patients with normal in infants given, parents should
kidneys with low- be informed about
grade reflux risk of infection
BT = breakthrough; CAP = continuous antibiotic prophylaxis; LUTD = lower urinary tract dysfunction;
PNH = prenatal diagnosed hydronephrosis; UTI = urinary tract infection; VCUG = voiding cystourethrography.

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 73


3.15 Urinary stone disease
3.15.1 Epidemiology, aetiology and pathophysiology
Paediatric stone disease is an important clinical problem in paediatric urology practice. Due to its recurrent
nature, every effort should be made to discover the underlying metabolic abnormality so that it can be treated
appropriately. Obtaining a stone-free state with close follow-up are of the utmost importance, although, it may
not be possible in some circumstances (e.g. oxalosis or nephrocalcinosis).

Bladder stones are still common in underdeveloped areas of the world and are usually ammonium acid urate
and uric acid stones, strongly implicating dietary factors [912]. Patients with augmented bladder constitute
another important group with a risk of up to 15% [913].

The incidence and characteristics of stones show a wide geographical variation in children. Although urinary
stone disease is generally considered to be a relatively rare disease, it is quite common in some parts of the
world. Paediatric stone disease is endemic in Turkey, Pakistan and in some South Asian, African and South
American countries. However, recent epidemiological studies have shown that the incidence of paediatric
stone disease is also increasing in the Western world [914-916], especially in girls, Caucasian ethnicity, African
Americans and older children [917]. More than 70% of stones in children contain calcium oxalate, while
infection stones are found more frequently in younger children [918].

3.15.2 Classification systems


Urinary stone formation is the result of a complex process involving metabolic, anatomical factors and
presence of infection.

3.15.2.1 Calcium stones


Calcium stones are usually made from calcium oxalate or calcium phosphate. Super-saturation of calcium
(hypercalciuria) and oxalate (hyperoxaluria) or decreased concentration of inhibitors, such as citrate
(hypocitraturia) or magnesium (hypomagnesemia) play a major role in the formation of calcium oxalate stones.
Higher super-saturations of calcium oxalate was shown to be associated with multiple stone disease [919].

Hypercalciuria: This is defined by a 24-hour urinary calcium excretion of more than 4 mg/kg/day (0.1 mmol/kg/
day) in a child weighing < 60 kg. In infants younger than three months, 5 mg/kg/day (0.125 mmol/kg/day) is
considered to be the upper limit for normal calcium excretion [920].
Hypercalciuria can be classified as either idiopathic or secondary. Idiopathic hypercalciuria is
diagnosed when clinical, laboratory, and radiographic investigations fail to delineate an underlying cause
leading to hypercalcaemia. Urinary calcium may increase in patients with high sodium chloride intake.
Secondary hypercalciuria occurs when a known process produces excessive urinary calcium. In secondary
hypercalcaemic hypercalciuria, a high serum calcium level may be due to increased bone resorption
(hyperparathyroidism, hyperthyroidism, immobilisation, acidosis, metastatic disease) or gastrointestinal
hyperabsorption (hypervitaminosis D) [921].

A good screening test for hypercalciuria compares the ratio of urinary calcium to creatinine. The normal
calcium-to-creatinine ratio in children is less than 0.2. If the calculated ratio is higher than 0.2, repeat-testing
is indicated. Neonates and infants have a higher calcium excretion and lower creatinine excretion than older
children [920, 921]. If the follow-up ratios are normal, then no additional testing for hypercalciuria is needed.

However, if the ratio remains elevated, a timed 24-hour urine collection should be obtained and the calcium
excretion calculated.
The 24-hour calcium excretion test is the standard criterion for the diagnosis of hypercalciuria. If
calcium excretion is higher than 4 mg/kg/day (0.1 mmol/kg/day), the diagnosis of hypercalciuria is confirmed
and further evaluation is warranted: levels of serum bicarbonate, creatinine, alkaline phosphatase, calcium,
phosphorus, magnesium, pH, and parathyroid hormone. Freshly voided urine should be measured for pH [920-
922]. In addition to calcium, the 24-hour urine analysis should also include phosphorus, sodium, magnesium,
uric acid, citrate and oxalate.
Initial management is always to increase fluid intake and urinary flow. Dietary modification is
a mandatory part of effective therapy. The child should be referred to a dietician to assess accurately the
daily intake of calcium, animal protein, and sodium. Dietary sodium restriction is recommended as well
as maintenance of calcium intake consistent with the daily needs of the child [923]. A brief trial of a low
calcium diet can be carried out to determine if exogenous calcium intake and/or calcium hyperabsorption is
contributing to high urinary calcium. Any recommendation to restrict calcium intake below the daily needs of
the child should be avoided. Moreover, low calcium intake is a risk factor for stone formation [924] (LE: 3).

74 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


Hydrochlorothiazide and other thiazide-type diuretics may be used to treat idiopathic hypercalciuria,
especially with calcium renal leak, at a starting dosage of 0.5-1 mg/kg/day [925-928] (LE: 3). In long-term use
of thiazide-type diuretics, a decrease in hypocalciuric effect may be seen after the third month and may cause
hypokalemia, hypocitraturia, hyperuricaemia and hypomagnesaemia. Therefore, control of blood and serum
values should be performed with regular intervals. Citrate therapy is also useful if citrate levels are low or if
hypercalciuria persists, despite other therapies [925, 929] (LE: 4).

Hyperoxaluria: Only 10-15% of oxalate comes from diet.


The average child excretes less than 50 mg (0.57 mmol)/1.73 m2/day [930-932], while infants excrete four times
as much. Hyperoxaluria may result from increased dietary intake, enteric hyperabsorption (as in short bowel
syndrome) or an inborn error of metabolism.

In rare primary hyperoxaluria, one of the two liver enzymes that play a role in the metabolism of oxalate may
be deficient. With increased deposition of calcium oxalate in the kidneys, renal failure may ensue in resulting
deposition of calcium oxalate in other tissues (oxalosis). The diagnosis is made upon laboratory findings of
severe hyperoxaluria and clinical symptoms. The definitive diagnosis requires liver biopsy to assay the enzyme
activity.
Other forms of hyperoxaluria, as mentioned earlier, may be due to hyperabsorption of oxalate in
inflammatory bowel syndrome, pancreatitis and short bowel syndrome. Yet, the majority of children have
‘mild’ (idiopathic) hyperoxaluria, with urine oxalate levels elevated only mildly in these cases. The treatment
of hyperoxaluria consists of the promotion of high urine flow, restriction of dietary oxalate and regular
calcium intake. Pyridoxine may be useful in reducing urine levels, especially in primary hyperoxaluria. Citrate
administration increases inhibitory urine activity [925, 933] (LE: 4).

Hypocitraturia: Citrate is a urinary stone inhibitor. Citrate acts by binding to calcium and by directly inhibiting
the growth and aggregation of calcium oxalate as well as calcium phosphate crystals. Thus low urine citrate
may be a significant cause of calcium stone disease. In adults, hypocitraturia is the excretion of citrate in urine
of less than 320 mg/day (1.5 mmol/day) for adults; this value must be adjusted for children depending on body
size [934-936].
Hypocitraturia usually occurs in the absence of any concurrent symptoms or any known metabolic
derangements. It may also occur in association with any metabolic acidosis, distal tubular acidosis or
diarrhoeal syndromes.
Environmental factors that lower urinary citrate include a high protein intake and excessive salt
intake. Many reports emphasise the significance of hypocitraturia in paediatric calcium stone disease. The
presence of hypocitraturia ranges from 30% to 60% in children with calcium stone disease [935, 937]. The
urine calcium-to-citrate ratios were higher in recurrent calcium stone forming children than solitary formers
[934, 938].

The restoration of normal citrate levels is advocated to reduce stone formation, although there are few relevant
studies in children. Hypocitraturia is treated by potassium citrate at a starting dose of 1 mEq/kg, given in two
divided doses [926] (LE: 3). The side effects of potassium citrate are very rare and most of the time they include
non-specific gastrointestinal complaints. Potassium citrate should be used with caution in hyperkalemic and
chronic renal failure conditions.

3.15.2.2 Uric acid stones


Uric acid stones are responsible for urinary calculi in 4-8% of children. Uric acid is the end product of purine
metabolism. Hyperuricosuria is the main cause of uric acid stone formation in children. A daily output of uric
acid of more than 10 mg/kg/day (0.6 mmol/kg/day) is considered to be hyperuricosuria [925].
The formation of uric acid stones is mainly dependent on the presence of acidic urinary
composition. Uric acid dissociation and solubility is strongly reduced at pH of < 5.8. As the pH becomes more
alkaline, uric acid crystals become more soluble and the risk of uric acid stone formation is reduced.
In the familial or idiopathic form of hyperuricosuria, children usually have normal serum uric
acid levels. In other children, it can be caused by uric acid overproduction secondary to inborn errors of
metabolism, myeloproliferative disorders or other causes of cell breakdown. Hyperuricosuria is also caused
by high purine and protein intake. Although hyperuricosuria is a risk factor for calcium oxalate stone formation
in adults, this does not appear to be a significant risk factor in children. Uric acid stones are non-opaque
stones. Plain X-rays are insufficient to show uric acid stones, and renal sonography and spiral CT are used for
diagnosis.

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 75


Alkalinisation of urine is the mainstay of therapy and prevention for uric acid stones. Citrate preparations are
useful as alkalinising agents. Maintaining a urine pH of 6 to 6.5 is sufficient to prevent uric acid stones [925].
In patients who failed with conservative measures with sustaining hyperuricosuria and hyperuricemia, stone
recurrences or myeloproliferative diseases, allopurinol (10 mg/kg) may be used. This medication may cause
several drug reactions (rash, diarrhoea, eosinophilia) and should be cautiously used in chronic renal failure
patients.

3.15.2.3 Cystine stones


Cystinuria is the cause of cystine stone formation and accounts for 2-6% of all urinary stones in children.
Cystinuria is an incompletely recessive autosomal disorder characterised by failure of renal tubules to reabsorb
four basic amino acids: cystine, ornithine, lysine and arginine.
Of these four amino acids, only cystine has poor solubility in urine, so that only cystine stones
may form in the case of excessive excretion in urine. Cystine solubility is pH-dependent, with cysteine
precipitation beginning at pH levels < 7.0. Other metabolic conditions, such as hypercalciuria, hypocitraturia
and hyperuricosuria, may accompany cystinuria, so leading to the formation of mixed-composition stones.
Cystine stones are faintly radiopaque and may be difficult to visualise on regular radiograph studies. They are
also hard in texture and more difficult to disintegrate by extracorporeal shockwave lithotripsy (SWL). Cystinuric
patients present with larger stones at the time of diagnosis, higher new stone formation rates, and are at higher
risk of surgery [939].
The medical treatment for cystine stones aims to reduce cystine saturation in urine and increase
its solubility. The initial treatment consists of maintaining a high urine flow and the use of alkalinising agents,
such as potassium citrate to maintain urine pH at above 7.0 (better above 7.5). If this treatment fails, the use
of α-mercaptopropionyl glycine or D-penicillamin may increase cystine solubility and reduce cystine levels in
urine and prevent stone formation. Side effects of these drugs are mostly mild and include gastrointestinal
complaints (alterations in taste and odour), fever and rash, however they can be associated with severe side-
effects, such as bone marrow depression, nephrotic syndrome and epidermolysis [940] (LE: 4).

3.15.2.4 Infection stones (struvite stones)


Infection-related stones constitute nearly 5% of urinary stones in children, though incidence increases over
10% in younger ages [941] and in non-endemic regions [918, 942]. Bacteria capable of producing urease
enzyme (Proteus, Klebsiella, Pseudomonas) are responsible for the formation of such stones.
Urease converts urea into ammonia and bicarbonate, alkalinising the urine and further converting
bicarbonate into carbonate. In the alkaline environment, triple phosphates form, eventually resulting in a
supersaturated environment of magnesium ammonium phosphate and carbonate apatite, which in turn leads to
stone formation.
In addition to bacterial elimination, stone elimination is essential for treatment, as stones will
harbour infection and antibiotic treatment will not be effective. Consideration should be given to investigating
any congenital problem that causes stasis and infection. Genitourinary tract anomalies predispose to formation
of such stones.

3.15.3 Diagnostic evaluation


Presentation tends to be age-dependent, with symptoms such as flank pain and haematuria being more
common in older children. Non-specific symptoms (e.g. irritability, vomiting) are common in very young
children. Haematuria, usually visible, occurring with or without pain, is less common in children. However,
nonvisible haematuria may be the sole indicator and is more common in children. In some cases, urinary
infection may be the only finding leading to radiological imaging in which a stone is identified [943, 944].

3.15.3.1 Imaging
Generally, US should be used as a first approach. Renal US is very effective for identifying stones in the
kidney. Many radiopaque stones can be identified with a simple abdominal flat-plate examination. The most
sensitive test for identifying stones in the urinary system (especially for ureteric stones) is non-contrast helical
CT scanning. It is safe and rapid, with 97% sensitivity and 96% specificity [945-947] (LE: 2). Despite its high
diagnostic accuracy, because of the potential radiation hazards, its use should be reserved for cases with non-
informative US and/or plain abdominal roentgenogram. Low dose protocols have also been developed with the
goal of reducing radiation dose with adequate image quality [948]. Intravenous pyelography is rarely used in
children, but may be needed to delineate the caliceal anatomy prior to percutaneous or open surgery.

76 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


3.15.3.2 Metabolic evaluation
Due to the high incidence of predisposing factors for urolithiasis in children and high stone recurrence rates,
every child with a urinary stone should be given a complete metabolic evaluation [912, 940, 949, 950]. A limited
urinary metabolic evaluation (24-h calcium, citrate, and oxalate and low urinary volume) is able to detect the
vast majority of clinically significant metabolic abnormalities [951]. However collections are most of the time
inadequate and should be repeated in this case [951, 952].

Metabolic evaluation includes:


• family and patient history of metabolic problems and dietary habits;
• analysis of stone composition (following stone analysis, metabolic evaluation can be modified according
to the specific stone type);
• electrolytes, blood/urea/nitrogen (BUN), creatinine, calcium, phosphorus, alkaline phosphatase, uric acid,
total protein, carbonate, albumin, and parathyroid hormone (if there is hypercalcaemia);
• spot urinalysis and culture, including ratio of calcium to creatinine;
• urine tests, including a 24-hour urine collection for calcium, phosphorus, magnesium, oxalate, uric acid
citrate, protein, and creatinine clearance;
• 24-hour cystine analysis if cystinuria is suspected (positive sodium nitroprusside test, cystine stone,
cystine hexagonal crystals in urine).

Figure 10 provides an algorithm of how to perform metabolic investigations in urinary stone disease in children
and how to plan medical treatment accordingly.

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 77


Figure 10: Algorithm for metabolic investigations in urinary stone disease in children

Paediatric stone patient

Elimination of stones by spontaneous


Exclude obstructive uropathy
passage or active removal (SWL, surgery)

Stone analysis

Mg Ammonium
Calcium stones
phosphate Uric acid stone Cystine
CaOX-CaPO
(struvite)

urine pH urine pH
urine culture urine and serum urine cystine
uric acid levels level

possibly acidicurine
urease producing hyperuricosuria cystinuria
bacteria hyperuricemia

Alkali High fluid intake


Total elimination replacement- potassium citrate
of stone k- citrate 3-4 mEq/kg/d
(surgery/SWL) Allopurinol mercaptopropiyonilglycine
antibiotics (10 mg/kg) 10-15 mg/kg/d
low purine diet penicillamin 30 mg/kg/day

urine - blood pH
serum PTH hypercalcaemia urine - blood Ca - uric acid levels,urine
Mg,pH
Phosphate
> 5.5
urine Ca-Oxalate-Citrate-Mg-Uric A -Phosphate

Elevated First morning urine First morning urine


pH < 5.5 pH > 5.5

Hyperparathyroidism Further investigation


Surgical treatment for RTA

hypercalciuria hyperoxaluria hyperuricosuria hypocitraturia

K-citrate
Regular calcium
diet Alkali
intake Citrate
(normal calcium replacement
Diet low in ox. replacement
low sodium (K-citrate)
K-citrate K- citrate
intake) allopurinol
pyridoxine
HCTZ (diuretic)

Ca = calcium; HCTZ = hydrochlorothiazide; Mg = magnesium; Ox = oxalate; PTH = parathyroid hormone;


SWL = extracorporeal shockwave lithotripsy; RTA = renal tubular acidosis; Uric A = uric acid.

3.15.4 Management
Adequate fluid intake and restricting the use of salt within daily allowance range are the general
recommendations besides the specific medical treatment against the detected metabolic abnormalities. With
the advance of technology, stone management has changed from open surgical approaches to endoscopic
techniques that are less invasive. Deciding on the type of treatment depends on the number, size, location,
stone composition and the anatomy of the urinary tract [950, 953, 954]. Expectant management is the initial
management in children with asymptomatic small size stones (< 4-5 mm ) with a possibility of spontaneous
clearance. There is no consensus on the size of stones for different ages eligible for clearance and the
duration of conservative follow-up. Adult literature reveals the benefits of medical expulsive therapy (MET)
using α-blockers. Although, experience in children is limited showing different results [955], a meta-analysis
of three randomised and two retrospective studies demonstrate that treatment with MET results in increased
odds of spontaneous ureteral stone passage and a low rate of adverse events [956]. Currently, most paediatric

78 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


stones can easily be managed by SWL. Endoscopic treatment can be applied for ureteric and bladder stones.
Percutaneous removal of stones is also possible for kidney stones in children. Only a small portion of children
will require open surgery but all attempts must be made to completely remove all stones since post-operative
residual fragments pass spontaneously in only 20-25% of cases [957, 958]. A congenital obstructive uropathy
should be managed together with stone removal therapy to prevent recurrence.

3.15.4.1 Extracorporeal shockwave lithotripsy


Many reports confirm that SWL can be performed in children with no suspicion of long-term morbidity of the
kidney [959-966].
The mean number of shockwaves for each treatment is approximately 1,800 and 2,000 (up to 4,000
if needed) and the mean power settings vary between 14 kV and 21 kV. The use of US and digital fluoroscopy
has significantly decreased the radiation exposure and it has been shown that children are exposed to
significantly lower doses of radiation compared to adults [953, 967, 968]. Concerns about anaesthesia no
longer present a problem due to advances in technique and medication, even in the infant age group. The type
of anaesthesia should be general or dissociative for children under ten years of age, whereas conventional
intravenous sedation or patient-controlled analgesia is an option for older children who are able to co-operate
[969] (LE: 2b).
Stone-free rates are significantly affected by various factors. Regardless of the location, as the
stone size increases, the stone-free rates decrease and retreatment rate increases. The stone-free rates for < 1
cm, 1-2 cm, > 2 cm and overall, were reported as nearly 90%, 80%, 60% and 80%, respectively. As the stone
size increases, the need for additional sessions increases [953, 967, 968, 970-974].
Localisation of the calculi has been described as a significant factor affecting the success rates
in different studies. Stones in the renal pelvis and upper ureter seem to respond better to SWL. For these
locations, the stone clearance rates are nearly 90%. However, SWL was found to be less effective for caliceal
stones; particularly the lower caliceal stones. Several studies reported stone-free rates for isolated lower
caliceal stones varying between 50% and 62% [975-977].
Shockwave lithotripsy can also be used to treat ureteral calculi. However, this is a more specific
issue and controversial. The success rates with SWL are less for distal ureteric stones. There may also be
technical problems with localisation and focusing of ureteric stones in children [976-979].
The type of machine used significantly influences success rates and complications. First-generation
machines can deliver more energy to a larger focal zone, resulting in higher fragmentation rates in a single
therapy. However, general anaesthesia is usually required due to the intolerable discomfort associated with a
first-generation machine. Later-generation machines have a smaller focal zone and deliver less energy, and
have a lower risk of pulmonary trauma, however, additional treatments may be needed. The success rate is
higher in younger children [972].
Although stenting does not affect stone clearance, overall complication rates are higher and
hospital stay is longer in the unstented patient [972, 974]. Stenting is essential in solitary kidneys undergoing
SWL treatment. Children with a large stone burden have a high risk of developing Steinstrasse and urinary
obstruction and should be followed more closely for the risk of prolonged urinary tract obstruction after SWL.
Post-SWL stent or nephrostomy tube placement may be needed in prolonged obstruction [940, 971].
The Hounsfield Unit (HU) of stone on non-contrast tomography has also been shown to be a
predictive factor for success in children and SWL was found to be more successful in stones with HU less
than 600 [958] and 1,000 [980]. Two nomogram studies revealed male gender, younger age, smaller stone
size, single stone, non-lower pole localisation and negative history for previous intervention are favourable
factors for stone clearance in paediatric SWL [981, 982]. A recent comparative study reported that these two
nomograms are independent predictors of stone-free rate following SWL in paediatric patients [983]. Although,
the invention of miniaturised endoscopic instruments seems to reduce the importance and popularity of SWL,
it has the advantage of not carrying the risk of certain complications related to endoscopic surgeries and
moreover studies comparing SWL and RIRS showed that besides having similar stone-free rates, SWL was
cheaper, had shorter hospital stay [984],with less post-operative emergency visit, pain and anaesthetic session
[985]. Complications arising from SWL in children are usually self-limiting and transient. The most common are:
• renal colic;
• transient hydronephrosis;
• dermal ecchymosis;
• UTI;
• formation of Steinstrasse;
• sepsis;
• rarely, haemoptysis.

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 79


In children with sterile pre-operative urine cultures, antibiotic prophylaxis to decrease infectious complications
is not recommended [986]. However, every effort should be made to sterilise the urine before performing SWL,
ureteroscopy (URS), or percutaneous nephrolithotomy (PCNL).

Due to the smaller size of the probes, laser energy is easier to use in smaller instruments and is more useful for
paediatric cases [987-996].

3.15.4.2 Percutaneous nephrolithotomy


Shockwave lithotripsy is the first choice for treating most renal paediatric stones. However, percutaneous
renal surgery should be used for larger and complex stones. Pre-operative evaluation, indication and surgical
technique are similar in children and adults. In most cases, percutaneous nephrolithotomy (PCNL) is used as
monotherapy, but is also used as an adjunctive procedure to other therapies.
The use of adult-sized instruments, in association with an increased number of tracts and sheath
size, seems to increase blood loss. However, the development of small-calibre instruments means that PCNL
can be used in children. In children (particularly smaller children), PCNL has some advantages, such as smaller
skin incision, single-step dilation and sheath placement, good working access for paediatric instruments,
variable length, and lower cost [986, 997, 998].
As monotherapy, PCNL is considerably effective and safe. The reported stone-free rates in the
recent literature are between 86.9% and 98.5% after a single session. These rates increase with adjunctive
measures, such as second-look PCNL, SWL and URS. Even in complete staghorn cases, a clearance rate of
89% has been achieved following a single session [990, 999-1003].
The most frequently reported complications of PCNL in children are bleeding, post-operative fever
or infection, and persistent urinary leakage. Bleeding requiring transfusion in the modern series is reported in
less than 10% [1004-1009] and is closely associated with stone burden, operative time, sheath size and the
number of tracts [1008, 1010, 1011]. In recent studies, post-operative infectious complications, such as fever
with or without documented UTI, are reported as less than 15% [1004, 1005, 1007-1009, 1012] and the origin
of fever is not always found to be the infection. With the availability of smaller size instruments, miniaturised
PCNL (‘miniperc’) through a 13F or 14F sheath [998, 1012, 1013] as well as ultramini-PCNL (UMP) through
12F sheaths [1014] have become possible, with decreased transfusion rates [1012]. The mini- and supermini-
PCNL (SMP) were shown to have higher efficacy with acceptable complication rates which were deemed to
be a safe alternative to SWL by some authors [1015, 1016]. The SMP was shown to be advantageous over
mini-PCNL in terms of complications with similar stone-free rates [1017, 1018]. This miniaturisation has been
further developed into the technique of ‘micro-perc’ using a 4.85F ‘all-seeing needle’. This technique is still
experimental and enables the stone to be fragmented by a laser in situ and left for spontaneous passage
[1019]. A study revealed that microperc provides a similar stone-free rate with similar complication rates and a
lower additional treatment rate compared with SWL in the treatment of kidney stone disease in children [1020]
(LE: 3). For stones 10-20 mm, micro-PNL was shown to have comparable results, with less bleeding, compared
to mini-PCNL [1021] and similar outcomes with less anaesthetic sessions compared to RIRS [1022] (LE: 3). As
experience has accumulated in adult cases, new approaches have started to be applied in children, including
tubeless PCNL. This technique has been used in uncomplicated surgery for stones < 2 cm, with patients
left either with an indwelling catheter or double J stent in the ureter [1006, 1023] or totally tubeless [1024].
Moreover, use of US for establishment of access is gaining popularity [1025, 1026] and supine approach [1027]
was also reported to be feasible in children.
The mean post-operative hospital stay is similar to adults. It is reported as three to four days in all
published literature and is much shorter than open surgery. The less invasive nature of this technique has made
it a promising alternative to open surgery for treating renal stones in children (LE: 2) [1004, 1006, 1007, 1009-
1011, 1013, 1014, 1019-1021, 1023, 1024, 1027, 1028].

3.15.4.3 Ureterorenoscopy
The increasing availability of smaller size endourological equipment has made it possible to manage paediatric
ureteral stones using endoscopic techniques.
The technique used in children is similar to the one used in adults. It is strongly recommended
that guide wires are used and the procedure is performed using direct vision. Routine balloon dilation of the
ureterovesical junction and ureteral stenting are controversial. In general, ureteric dilatation is being performed
less and only in selected cases. There is a tendency to use hydrodilation more because it is similarly effective
[986, 988, 994, 1029-1032] (LE: 3).
Different lithotripsy techniques, including ultrasonic, pneumatic and laser lithotripsy, have all been
shown to be safe and effective. Due to the smaller size of the probes, laser energy is easier to use in smaller
instruments and is more useful for paediatric cases [987-996].

80 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


All studies reporting the use of endoscopy for ureteric stones in children have clearly demonstrated that there
is no significant risk of ureteric strictures or reflux with this mode of therapy (LE: 1). The risk of post-operative
hydronephrosis depends on the presence of impacted stone and ureteral injury during operation [1033]. A
multi-institutional study on the use of semi-rigid ureteroscopy for ureteral calculi in children has revealed that
the procedure is effective with a 90% stone-free rate and efficacy quotient. The study also focused on the
factors affecting the complication rates. The authors found that, although operating time, age, institutional
experience, orifice dilation, stenting and stone burden were significant on univariate analysis, multivariate
analysis revealed that operating time was the only significant parameter affecting the complication rate [1034].
However, for proximal ureteral stones semi-rigid ureteroscopy is not a good first option because of higher
complication and failure rates [1035].
A recent literature review contains a growing number of case series on the use of flexible
ureterorenoscopic interventions in children. Both intrarenal and ureteric stones can be treated using this
approach [1036-1041]. In these series, the authors generally did not use active orifice dilation, but attempted
to use a ureteral sheath where possible. However, an important problem was the inability to obtain retrograde
access to the ureter in approximately half of the cases [1037, 1039]. This problem can be overcome by stenting
and leaving the stent indwelling for passive dilation of the orifice, and performing the procedure in a second
session. The success rates varied between 60 and 100%, with a negligible number of complications [1036,
1038-1040, 1042]. The need for additional procedures was related to stone size [1040]. A comparative study
showed that retrograde intra-renal surgery (RIRS) had similar stone-free rate compared to ESWL after three
months, with fewer sessions [1043], however for stones larger than 2 cm, RIRS monotherapy has lower stone-
free rates than mini-PCNL with the advantages of decreased radiation exposure, fewer complications and
shorter hospital stay [1044] (LE: 3). In contrast, for stones between 10-20 mm, RIRS has similar success and
complication rates and shorter hospital stay and low radiation exposure when compared to micro-PNL [1045]
(LE: 3). A recent systematic review revealed that compared with the other two treatments, PCNL had a longer
operative time, fluoroscopy time and hospital stay. Shockwave lithotripsy had a shorter hospital stay, higher
retreatment rate and auxiliary rate in comparison with the other two treatments. It was also shown that PCNL
presented a higher efficacy quotient than the other two treatments, and RIRS had a lower efficiency than SWL
and PCNL. In the subgroup analysis of paediatric patients with stone ≤ 20 mm, the comparative results were
similar to those described above, except for the higher complication rate of PCNL than SWL [1046].

3.15.4.4 Open or laparoscopic stone surgery


Most stones in children can be managed by SWL and endoscopic techniques. However, in some situations,
open surgery is inevitable. Good candidates for open stone surgery include very young children with large
stones and/or a congenitally obstructed system, which also require surgical correction. Open surgery is also
necessary in children with severe orthopaedic deformities that limit positioning for endoscopic procedures.
In centres with a well-established experience, a laparoscopic approach may be a good alternative
for some cases as a last resort before open surgery. Suitable candidates include patients who have a history
of previous failed endoscopic procedures, complex renal anatomy (ectopic or retrorenal colon), concomitant
UPJ obstruction or caliceal diverticula, mega-ureter, or large impacted stones. Laparoscopic stone surgery via
conventional or a robot-assisted transperitoneal or retroperitoneal approach can be attempted. However, there
is limited experience with these techniques and they are not routine therapeutic modalities [1047-1050].
Bladder stones in children can usually be managed by endoscopic techniques. Open surgery may
also be used for very large bladder stones or for bladder stones caused by an anatomical problem.
In addition to the advantages and disadvantages of each treatment modality for the specific
size and location of the stone, consideration has to be given to the availability of the instruments and the
experience with each treatment modality before the choice of technique is made. Recommendations for
interventional management are given in Table 4.

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 81


Table 4: Recommendations for interventional management in paediatric stones

Stone size and Primary treatment Secondary treatment Comment


localisation* option options
Staghorn stones PCNL Open/SWL Multiple sessions and accesses with
PCNL may be needed. Combination with
SWL may be useful.
Pelvis < 10 mm SWL RIRS/PCNL/MicroPerc
Pelvis 10-20 mm SWL PCNL/RIRS/ Multiple sessions with SWL may
be needed. PCNL has similar
recommendation grade.
Pelvis > 20 mm PCNL MicroPerc/Open Multiple sessions with SWL may be
needed.
Lower pole calyx PCNL SWL/Open Multiple sessions with SWL may be
needed.
< 10 mm SWL SWL/Open Anatomical variations are important for
complete clearance after SWL.
Lower pole calyx SWL RIRS/PCNL/MicroPerc Anatomical variations are important for
complete clearance after SWL.
> 10 mm PCNL RIRS/PCNL/MicroPerc Anatomical variations are important for
complete clearance after SWL.
Upper ureteric SWL SWL/ MicroPerc
stones
Lower ureteric URS PCNL/URS/Open Additional intervention need is high with
stones SWL.
Bladder stones Endoscopic SWL/Open Open is easier and with less operative
time with large stones.
Bladder stones Endoscopic Open is easier and with less operative
time with large stones.
* Cystine and uric acid stones excluded.
PCNL = percutaneous nephrolithotomy; SWL = shockwave lithotripsy; RIRS = retrograde intrarenal surgery;
URS = ureteroscopy.

3.15.5 Summary of evidence and recommendations for the management of urinary stones

Summary of evidence LE
The incidence of stone disease in children is increasing. 2
Contemporary surgical treatment is based on minimally invasive modalities. Open surgery is very 2a
rarely indicated.
The term ‘clinically insignificant residual fragments’ is not appropriate for children since most of them 2b
become symptomatic and require intervention.

Recommendations LE Strength rating


Use plain abdominal X-ray and ultrasound as the primary imaging techniques for the 2b Strong
diagnosis and follow-up of stones.
Use low-dose non-contrast computed tomography in cases with a doubtful 2a Strong
diagnosis, especially of ureteral stones or complex cases requiring surgery.
Perform a metabolic evaluation in any child with urinary stone disease. Any kind 2a Strong
of interventional treatment should be supported with medical treatment for the
underlying metabolic abnormality, if detected.
Limit open surgery under circumstances in which the child is very young with 2a Strong
large stones, in association with congenital problems requiring surgical correction
and/or with severe orthopaedic deformities that limit positioning for endoscopic
procedures.

82 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


3.16 Obstructive pathology of renal duplication: ureterocele and ectopic ureter
3.16.1 Epidemiology, aetiology and pathophysiology
Ureterocele and ectopic ureter are the two main anomalies associated with complete renal duplication, but
they also occur in a single system. At present, antenatal US detects both conditions in the majority of cases if
associated with obstruction, and diagnosis is confirmed after birth by further examination. Later in life, these
anomalies are revealed by clinical symptoms: UTI, pain, calculus formation, disturbances of micturition, and
urinary incontinence. There is a wide variation of symptoms in patients with ureterocele (from the asymptomatic
patient to urosepsis, urinary retention and upper tract dilatation after birth).

3.16.1.1 Ureterocele
Ureterocele is four to seven times more frequent in female than in male patients; the overall incidence in
autopsies is around one in 4,000 children. Around 80% is associated with the upper pole ureter in duplicated
systems and 20% in single systems. About 10% of ureteroceles are bilateral [1051].

3.16.1.2 Ectopic ureter


Ectopic ureter is less frequent than ureterocele (10 in 19,046 autopsies), but is also more common in female
patients (male to female ratio is 1:5). Some remain asymptomatic, therefore, the true incidence is difficult to
determine [1052]. Eighty per cent of ectopic ureters are associated with complete renal duplication; however, in
male patients about 50% of ectopic ureters are associated with a single system [1053].

3.16.2 Classification systems


3.16.2.1 Ureterocele
Ureterocele is a cystic dilatation that develops in the intravesical part of the submucosal ureter. The aetiology
remains unclear [1054-1056]. A single-system ureterocele is associated with a kidney with one ureter, and in
duplex systems, the ureterocele belongs to the upper pole.
Ureteroceles usually cause obstruction of the upper pole, but the degree of obstruction and
functional impairment is variable according to the type of ureterocele and upper pole dysplasia. In the
orthotopic form, there is often no or only mild obstruction, and frequently the function of the moiety is normal
or slightly impaired, and the corresponding ureter may be dilated. Cystic renal dysplasia is also associated
with a single system ureterocele [1057]. Vesicoureteral reflux can be observed in 50% on the ipsilateral side
and 20% on the contralateral side. Reflux into the ureterocele is uncommon [1058]. In the ectopic form, the
upper pole is altered, frequently dysplastic, and hypo-functional or non-functional [1059]. The corresponding
ureter is a mega-ureter. In the caeco-ureterocele (see definition below), the upper pole of the renal duplication
is dysplastic and non-functional. Histological evaluation demonstrated that the changes represent a process of
maldevelopment and may not result from infections or obstruction [1059].

3.16.2.1.1 Ectopic (extravesical) ureterocele


If any portion of the ureterocele extends into the bladder neck or urethra, it is called an ectopic ureterocele.
Ectopic ureterocele is the most common form of ureterocele (> 80%). It can be voluminous, dissociating the
trigone and slipping into the urethra, and may prolapse through the urethral meatus (caeco-ureterocele). The
ureterocele orifice is tight, and located in the bladder itself or below the neck. The ureter corresponding to the
lower pole moiety is raised by the ureterocele and is frequently refluxing or compressed by the ureterocele,
leading to an obstructive mega-ureter. A contralateral renal duplication is associated with 50% of cases.
Occasionally, large ureteroceles are responsible for reflux or obstruction of the contralateral upper tract.

3.16.2.1.2 Orthotopic (intravesical) ureterocele


The intravesical or orthotopic ureterocele is completely located in the bladder. Intravesical ureteroceles are
mostly combined with a single kidney system and account for about 15% of cases. It is diagnosed more in
older children or adults.

3.16.2.2 Ectopic ureter


The term ectopic ureter describes a ureter with the orifice located at the bladder neck, in the urethra or outside
the urinary tract. The ureter can drain the upper pole of a duplex or single system. There is a fundamental
difference between the sexes. In boys, the ectopic orifice is never below the external sphincter.

In girls, the ureteral orifice may be located [1060]:


• in the urethra, from the bladder neck to the meatus (35%);
• in the vaginal vestibule (34%);
• in the vagina (25%);
• in the uterus and Fallopian tube (6%).

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 83


In boys, the ureteral orifice may be located [1060]:
• in the posterior urethra (47%);
• in the prostatic utricle (10%);
• in the seminal vesicles (33%);
• in the vas deferens or ejaculatory ducts (10%).

3.16.3 Diagnostic evaluation


3.16.3.1 Ureterocele
Prenatal US easily reveals voluminous obstructive ureteroceles [1061]. In cases with a small upper pole or a
slightly obstructive ureterocele, prenatal diagnosis is difficult.
If prenatal diagnosis is impossible, the following clinical symptoms, besides incidental findings, can
reveal the congenital anomaly at birth or later:
• At birth, a prolapsed and sometimes strangulated ureterocele may be observed in front of the urethral
orifice. In a newborn boy, it might cause acute urinary retention, simulating urethral valves.
• The early symptom of pyelonephritis in either sex may lead to the diagnosis.
• Later symptoms can include dysuria, recurrent cystitis and urgency.

In cases of prenatal diagnosis, at birth US confirms the ureteral dilatation that ends at the upper pole of a renal
duplication. It also demonstrates the presence of a ureterocele in the bladder, with a dilated ureter behind the
bladder.
At this point, it is important to assess the function of the upper pole using nuclear renography of the
region of interest. This is best assessed with DMSA, however this requires a careful systematic review of the
images [1062]. Magnetic resonance urography may visualise the morphological status of the upper pole and
lower moieties and of the contralateral kidney as well as it can detect renal scars [1063, 1064]. Using functional
MR urography, differential renal function can be assessed with low intra- and interobserver variability [1065].
Based on the prevalence of high-grade reflux, VCUG is mandatory for identifying ipsilateral or contralateral
reflux and assessing the degree of intra-urethral prolapse of the ureterocele [1066]. Urethrocystoscopy may
reveal the pathology in cases where it is difficult to make the differential diagnosis between ureterocele and
ectopic mega-ureter.

3.16.3.2 Ectopic ureter


Most of the ectopic mega-ureters are diagnosed primarily by US. In some cases, clinical symptoms can lead to
diagnosis:
• In neonates: dribbling of urine, pyuria, and acute pyelonephritis.
• In young girls: permanent urinary incontinence besides normal voiding, or significant vaginal discharge as
the equivalent of incontinence; an ectopic orifice may be found in the meatal region [1067].
• In pre-adolescent boys: epididymitis is the usual clinical presentation and the seminal vesicle may be
palpable.

Ultrasound, radionuclide studies (DMSA, VCUG, MR urography, high-resolution MRI, and cystoscopy) are the
diagnostic tools to assess function, to detect reflux and rule out ipsilateral compression of the lower pole and
urethral obstruction [1068]. In some cases, the large ectopic ureter presses against the bladder and can look
like a pseudo-ureterocele [1069].
Girls who present with life-long minimal urinary incontinence, never being dry, normal bladder
function, complete emptying, and normal US are very suspicious for ectopic ureter. This needs to be excluded
or confirmed by MRI as it is the most sensitive method [1070].

3.16.4 Management
3.16.4.1 Ureterocele
Management is controversial with a choice between a non-operative approach, endoscopic decompression,
ureteral re-implantation, partial nephroureterectomy, or complete primary reconstruction [1071-1076]. The
choice of a therapeutic modality depends on the following criteria: clinical status of the patient (e.g. urosepsis);
patient age; function of the upper pole; presence of reflux or obstruction of the ipsilateral or contralateral
ureter; presence of bladder neck obstruction caused by ureterocele; intravesical or ectopic ureterocele; and
caregivers’ and the surgeon’s preferences [1076]. When the diagnosis is made by US, prophylactic antibiotic
treatment maybe indicated until a VCUG is performed.

3.16.4.1.1 Early treatment


In the presence of febrile infection or obstruction at the bladder neck, immediate endoscopic incision or
puncture of the ureterocele is recommended. In a clinically asymptomatic child with a ureterocele and a non

84 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


or hypofunctional upper pole, without significant obstruction of the lower pole and without bladder outlet
obstruction, prophylactic antibiotic treatment is given until follow-up procedures are instigated. Decompression
of the dilated system facilitates later reconstructive surgery [1077, 1078].

3.16.4.1.2 Re-evaluation
Conservative treatment may be adopted in asymptomatic patients without any bladder outlet obstruction,
severe hydroureteronephrosis of the ureterocele moiety or high-grade (over grade III) reflux [1076, 1079]. A
meta-analysis showed that, after primary ureterocele-incision, the re-operation rate is higher in those with an
ectopic ureterocele compared to those with an intravesical ureterocele [1072]. Secondary surgery is necessary
if decompression is not effective, significant reflux is present, or there is obstruction of the ipsi- or contralateral
ureters, and/or bladder neck obstruction or retained ureterocele [1080].
Surgery may vary from upper pole nephrectomy to complete unilateral LUT reconstruction [1075,
1081-1083]. In an ectopic ureterocele with severe hydroureteronephrosis and without reflux, the primary upper
tract approach without endoscopic decompression (partial upper-pole nephroureterectomy, pyelo/ureteropyelo/
ureterostomy and upper-pole ureterectomy) has an 80% chance of being the definitive treatment [1076, 1084].
Also a LUT approach in those with a poorly or non-functioning upper pole is an option [1085]. Today, despite
successful surgery, some authors think, that surgery may not be necessary at all in some patients [1086], as
less aggressive surgical treatment and non-operative management over time can achieve the same functional
results [1087].

Figure 11: A
 lgorithm for the management of duplex system ureteroceles after the first 3-6 months of life [974]

DSU

Asymptomatic Symptomatic or severe HUN or


No severe HUN or obstruction obstruction

No VUR VUR VUR No VUR

Low grade High grade or Good No/poor


multiple function function
infections

Observation Bladder surgery or Ectopic: upper Intravesical Ectopic


endoscopic to lower tract endoscopic UPPN
management anastomosis decompression

DSU = duplex system ureterocele; HUN = hydroureteronephrosis; UPPN = upper pole partial nephrectomy;
VUR = vesicoureteric reflux to the lower pole.

Obstruction is considered to be the presence of non-refluxing dilatation of non-ureterocele-bearing moieties


(especially of the lower pole) or of an obstructive drainage pattern on diuretic renography.

3.16.4.2 Ectopic ureter


In the majority of cases, the upper pole is dysplastic and poorly functioning. There are a variety of
therapeutic options, each with its advantages and disadvantages. In non-functioning moieties with recurrent
infections, heminephro-ureterectomy is a definite solution. Ureteral reconstruction (ureteral re-implantation/
ureteroureterostomy/ureteropyelostomy and upper-pole ureterectomy) are other therapeutic options especially
in cases in which the upper pole has function worth preserving. These procedures can be performed through
an open laparoscopic or robotic assisted approach [1088-1091]. So far there is no superior approach [1092].
In patients with bilateral single ectopic ureters (a very rare condition), an individual approach depending on the
sex and renal and bladder function of the patient is necessary. Usually the bladder neck is insufficient in these
patients [1093].

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 85


3.16.5 Summary of evidence and recommendations for the management of obstructive pathology of
renal duplication: ureterocele and ectopic ureter

Summary of evidence LE
Ureterocele and ectopic ureter are associated with complete renal duplication, but they also occur in a 1
single system.
In most cases, in young children (first years of life) diagnosis is done by US. 1
In older children clinical symptoms will prompt assessment. 1
Management includes a conservative approach, endoscopic decompression, partial 3
nephroureterectomy, or complete primary reconstruction. Choice of treatment will depend on:
• clinical status of the patient (e.g., urosepsis);
• patient age;
• function of the upper pole;
• presence of reflux or obstruction of the ipsilateral or contralateral ureter;
• presence of bladder neck obstruction caused by ureterocele;
• intravesical or ectopic ureterocele;

Recommendations LE Strength rating


Ureterocele Diagnosis Use ultrasound (US), radionuclide studies 3 Weak
(mercaptoacetyltriglycine (MAG3)/dimercaptosuccinic
acid (DMSA)), voiding cystourethrography (VCUG),
magnetic resonance urography, high-resolution
magnetic resonance imaging (MRI), and cystoscopy
to assess function, to detect reflux and rule out
ipsilateral compression of the lower pole and urethral
obstruction.
Treatment Select treatment based on symptoms, function and 3 Weak
reflux as well on surgical and parenteral choices:
observation, endoscopic decompression, ureteral
re-implantation, partial nephroureteretomy, complete
primary reconstruction. Offer, early endoscopic
decompression to patients with an obstructing
ureterocele.
Ectopic ureter Diagnosis Use US, DMSA scan, VCUG or MRI for a definitive 3 Weak
diagnosis.
Treatment Treatment In non-functioning moieties with 3 Weak
recurrent infections, heminephro-ureterectomy
is a definitive solution. Ureteral reconstruction
(ureteral re-implantation/ureteroureterostomy/
ureteropyelostomy and upperpole ureterectomy) are
other therapeutic option especially in cases in which
the upper pole has function worth preserving.

3.17 Disorders of sex development


3.17.1 Introduction
The formerly called ‘intersex disorders’ were the subject of a consensus document in which it was decided that
the term ‘intersex’ should be changed to ‘disorders of sex development’ (DSD) [1094].
The new classification has arisen due to advances in knowledge of the molecular genetic causes of
abnormal sexual development, controversies inherent to clinical management and ethical issues. Controversial
and negative terminology, e.g. ‘pseudohermaphroditism’ and ‘hermaphroditism’, have been renamed according
to the new pathophysiological insights. Furthermore, some conditions presenting with severe male genital
malformation, such as penile agenesis and cloacal exstrophy, which could not be categorised, have also been
included. The term ‘disorders of sex development’ is proposed to indicate congenital conditions with atypical
development of chromosomal, gonadal or anatomical sex.

In addition, in 2017, the Parliamentary Assembly of the Council of Europe decided on a resolution called:
“Promoting the human rights of and eliminating discrimination against intersex people” [1095]. The
Parliamentary Assembly concluded that the majority of intersex people are physically healthy and only a few

86 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


suffer from medical conditions that put their health at risk. Furthermore, they state that the prevailing medical
view has been that intersex children’s bodies can and should be made to conform to either a male or a female
paradigm, often through surgical and/or hormonal intervention and that this should be done as early as
possible and that the children should then be raised in the gender corresponding to the sex assigned to their
body. The Parliamentary Assembly considers that this approach involves serious breaches of physical integrity,
in many cases concerning very young children or infants who are unable to give consent and whose gender
identity is unknown.

Therefore the Parliamentary Assembly called on Council of Europe member states with regard to effectively
protecting children’s right to physical integrity and bodily autonomy and to empowering intersex people
as regards the following rights: medically unnecessary sex-“normalising” surgery, sterilisation and other
treatments practised on intersex children without their informed consent should be prohibited and in addition
that it has to be ensured that, except in cases where the life of the child is at immediate risk, any treatment
that seeks to alter the sex characteristics of the child, including their gonads, genitals or internal sex organs,
is deferred until such time as the child is able to participate in the decision, based on the right to self-
determination and on the principle of free and informed consent.

The Panel refers to the consensus documents mentioned above as well as on the Parliamentary Assembly
resolution. This chapter will focus on what is relevant for the practising paediatric urologist as the urologist is
likely to be involved in neonates with DSD conditions.

Overall, evidence-based literature on DSD is sparse. There are no RCTs and most studies are based on
retrospective clinical descriptive studies or on expert opinion. An exception is the risk of gonadal cancer, for
which the level of evidence is higher [1096].
Disorders of sex development can present as prenatal diagnosis, neonatal diagnosis and late
diagnosis. Prenatal diagnosis can be based on karyotype or US findings; neonatal diagnosis is based on
genital ambiguity and late diagnosis is made on early or delayed puberty. In this guideline, focus is on
the neonatal presentation where the paediatric urologist plays a major role. For late diagnosis we refer to
endocrinology and gynaecology guidelines on precocious and delayed puberty where paediatric urologists play
a minor role [1097, 1098].

Dealing with neonates with DSD requires a multidisciplinary approach, which should include geneticists,
neonatologists, paediatric and adult endocrinologists, gynaecologists, psychologists, ethicists and social
workers. Each team member should be specialised in DSD and a team should have treated enough patients to
ensure experience.

3.17.2 Current classification of DSD conditions


Since the International Consensus Conference on intersex and its subsequent publications on classification of
the various conditions of DSD, several updates have been published with the latest published by the Global
DSD Update Consortium in 2016 [1099]. As the field of DSD is continuously developing and knowledge and
viewpoints change over time, an effort was made to include representatives from a broad perspective including
support and advocacy groups with the goal to focus patient care upon the best possible QoL.

According to the international consensus in 2005, DSDs were defined as congenital conditions within which the
development of chromosomal, gonadal and anatomic sex is atypical. The changes that were made according
to terminology are as follows:

46XX DSD group formerly called female pseudohermaphrodite, over-virilisation of an XX female, and
masculinisation of an XX female. In this group the vast majority is due to classic congenital adrenal hyperplasia
(CAH) with various degrees of masculinisation. Among all DSD conditions together, 46XX CAH patients
comprise approximately 80%. These conditions are extremely important since they can be potentially life
threatening after birth because of salt loss phenomenon and immediate medical care is mandatory.

46XY DSD group in the past named male pseudohermaphrodite, undervirilisation of an XY male, and
undermasculinisation of an XY male. This group is often quite heterogenous and includes the partial androgen
insensitivity syndrome (PAIS) as well as the complete androgen insensitivity syndrome (CAIS) formerly called
testicular feminisation..

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 87


Sex chromosome mosaicism DSD group (45X, 45X/46XY, 47XXY) consists of multiple variants with the
mixed gonadal dysgenesis being the most important one. Many have a normal male phenotype and others
asymmetric genitalia. One scrotal half often contains a gonad which is likely to be a testis whereas the other
side is more a labia majora with usually no palpable gonad, most likely to be a streak gonad.

Ovotesticular DSD group was in the past called true hermaphrodite because of the presence of ovarian and
testicular tissue in the same individual meaning that both – female and male structures – live together. There is
great variability in phenotype with uni- or bilateral undescended gonads which can present as one ovary and
one testis or as one or two ovotestes.

Non-hormonal/non-chromosomal DSD group was introduced as well, including newborns with cloacal
exstrophy where bladder and intestines are exposed, patients with aphallia, and severe micropenis. The latter
one is a normally formed penis with a stretched length of < 2.5 standard deviation below the mean [1094,
1100].

Micropenis should be distinguished from buried and webbed penis, which are usually of normal size. The
length of the penis is measured on the dorsal aspect, while stretching the penis, from the pubic symphysis to
the tip of the glans [1094].

3.17.3 Diagnostic evaluation


3.17.3.1 The neonatal emergency
The first step is to recognise the possibility of DSD (Table 5) and to refer the newborn baby immediately to a
tertiary paediatric centre, fully equipped with neonatal, genetics, endocrinology and paediatric urology units.
Diagnosis of a 46XX DSD due to congenital adrenal hyperplasia should not be delayed and represents a
neonatal emergency situation since the possibility of salt loss phenomenon can be fatal.

Table 5: F
 indings in a newborn suggesting the possibility of DSD
(adapted from the American Academy of Pediatrics)

Apparent male
Severe hypospadias associated with bifid scrotum
Undescended testis/testes with hypospadias
Bilateral non-palpable testes in a full-term apparently male infant
Apparent female
Clitoral hypertrophy of any degree, non-palpable gonads
Vulva with single opening
Indeterminate
Ambiguous genitalia

3.17.3.2 Family history and clinical examination


A careful family history must be taken followed by a thorough clinical examination including various laboratory
tests and imaging modalities (Table 6).

Table 6: Diagnostic work-up of neonates with disorders of sex development

History (family, maternal, neonatal)


Parental consanguinity
Previous DSD or genital anomalies
Previous neonatal deaths
Primary amenorrhoea or infertility in other family members
Maternal exposure to androgens
Failure to thrive, vomiting, diarrhoea of the neonate
Physical examination
Pigmentation of genital and areolar area
Hypospadias or urogenital sinus
Size of phallus
Palpable and/or symmetrical gonads
Blood pressure

88 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


Investigations
Blood analysis: 17-hydroxyprogesterone, electrolytes, LH, FSH, TST, cortisol, ACTH
Urine: adrenal steroids
Karyotype
Ultrasound
Genitogram
hCG stimulation test to confirm presence of testicular tissue
Androgen-binding studies
Endoscopy
ACTH = adrenocorticotropic hormone; FSH = follicle-stimulating hormone; hCG = human chorionic
gonadotropin; LH = luteinising hormone; TST = testosterone.

A thorough clinical examination in a neonate presenting with ambiguous genitalia is important. As well as
an accurate description of the ambiguous genitalia, detailed information should be given on palpability and
localisation of the gonads. Information gathered by the various examinations described below should help
the team to come to a final diagnosis. Medical photography can be useful but requires sensitivity and consent
[1101].

Palpable gonad: If it is possible to feel a gonad, it is most likely to be a testis; this clinical finding therefore
virtually excludes 46XX DSD.

Phallus: The phallus should be measured. A cotton bud placed at the suprapubic base of the implant of the
stretched phallus allows for a good measurement of phallic length.

Urogenital sinus opening: The opening of the urogenital sinus must be well evaluated. A single opening has to
be identified as well as a hymenal ring. Attention needs to be paid to the fusion of the labioscrotal folds as well
as whether they show rugae or some discolouration.

Ultrasound can help to describe the palpated gonads or to detect non-palpable gonads. However, the
sensitivity and specificity are not high. Mülllerian structures like the vagina or utricular structures can be
evaluated as well [1102, 1103].

Genitography can provide some more information on the urogenital sinus, especially on the exact position of
the confluence. Moreover, it gives evidence of possible duplication of the vagina.

Invasive diagnostics under general anaesthesia can be helpful in some cases.


On cystoscopy, the urogenital sinus can be evaluated as well as the level of confluence. It allows also for
evaluation of the vagina or utriculus, the possible presence of a cervix at the top of the vagina.

Laparoscopy is necessary to obtain a final diagnosis on the presence of impalpable gonads and on the
presence of Müllerian structures. If indicated, a gonadal biopsy can be performed [1104, 1105].

These investigations will help to distinguish the various conditions of DSD and provide quick evidence of
congenital adrenal hyperplasia (CAH), which is the most frequently occurring DSD and the one that can
become life-threatening within the first days of life because of salt loss phenomenon.

3.17.4 Gender assignment


Nowadays it is obvious and clear that open and complete communications with caregivers and eventually the
affected person are mandatory. Education and psychological support regarding the impact are needed for each
individual to make sense of the condition, relate to their community and establish relationships. The lack of
outcome data and different preferences make it extremely difficult to determine whether and when to pursue
gonadal or genital surgery. Shared decision making is necessary, combining expert healthcare knowledge and
the right of a patient or surrogate to make fully informed decisions. This entails a process of education, sharing
of risks/benefits, articulating the uncertainties in DSD care and outcomes and providing time for the patient and
family to articulate back the risks and benefits of each option. The goal of all involved should be to individualise
and prioritise each patient.

However, adverse outcomes have led to recommendations to delay unnecessary surgery to an age when the
patient can give informed consent. Surgery that alters appearance is not urgent. Recently the Parliamentary
Assembly of the Council of Europe, the European Society for Paediatric Urology (ESPU) as well as the Societies
for Pediatric Urology have taken a position in the debate on surgery for DSD [1095, 1106, 1107].

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 89


In an open letter to the Council of Europe, the European Society for Paediatric Urology expressed
its attitude to the abovementioned resolution and concentrated on a worrying issue dealing with medico-
surgical care for children with DSD. It states that surgical interventions in children with DSD only being applied
in emergency conditions is discordant with the definition of health according to the World Health Organization
(WHO), stating that health is not merely the absence of disease, but is a much broader concept, including
physical, mental, and social domains. This especially applies to children, as favourable physical, social and
emotional conditions are all critical factors for their optimal growth and development, which enables them
to reach their full potential at adult age. As social and emotional interactions with the parents or caregivers,
being the most important adults in a young child’s life, form the basis for their future, treatment of children with
DSD can best be organised in a patient- and family-centred multidisciplinary setting, in an atmosphere based
on openness, commitment and trust. Physicians, who daily take care of children with a variety of congenital
conditions, the same as their parents or caregivers, are committed to the current as well as the future health
and well-being of all children entrusted to their care. In contrast to what is alleged in the recommendation,
parents and caregivers implicitly act in the best interest of their children and should be respected as their
outstanding representatives, and should not be put aside by claiming prohibition regulations regarding the
well-informed decisions they make on their behalf. Finally in that open letter the ESPU advocate keeping the
dialogue open with the professionals active in specialised centres for multidisciplinary, patient- and family-
centred care as well as with patient societies, for which the present resolution is recognised as being a solid
starting base [1108].

3.17.5 Risk of tumour development


Individuals with DSD have an increased risk of developing cancers of the germ cell lineage, malignant germ cell
tumours or germ cell cancer in comparison with to the general population [1109].
It is well-recognised that the highest risk prevalence (30-50%) is seen in conditions characterised
by disturbed gonadal development such as incomplete testis development combined with a full block of
embryonic germ cell maturation in patients with 46XY gonadal dysgenesis and in some patients with 45X/46XY
DSDs. Conversely, patients with testosterone biosynthesis disorders and androgen action disturbances show
a much lower risk (1-15%) for carcinoma in situ (CIS) development during childhood and a limited tendency
towards invasive progression of the lesions [1110]. With regard to clinical management a gonadal biopsy at
the time of a possible orchidopexy can be obtained for an initial assessment including regular self-exams and
annual ultrasound [1096].

3.17.6 Recommendations for the management of disorders of sex development

Recommendations Strength rating


Newborns with DSD conditions warrant a multidisciplinary team approach. Strong
Refer children to experienced centres where neonatology, paediatric endocrinology, Strong
paediatric urology, child psychology and transition to adult care are guaranteed.
Do not delay diagnosis and treatment of any neonate presenting with ambiguous genitalia Strong
since salt-loss in a 46XX CAH girl can be fatal.

3.18 Congenital lower urinary tract obstruction (CLUTO)


Introduction
The term congenital lower urinary tract obstruction (CLUTO) is used for a foetus, which during intrauterine US
screening shows a dilatation of the upper and lower urinary tract. During pregnancy the diagnosis is usually
based only on US examinations. There is a broad spectrum of conditions, that could cause an intrauterine
dilatation of the urinary tract. Post-partum diagnosis comprises any anatomical and functional disorder/
anomaly/malformation causing a dilatation e.g. posterior/anterior urethral valve, urethral atresia/dysplasia/
stenosis prune belly syndrome, dilating reflux. Cloacal malformation, ureterocele, a Megacystis-Microcolon-
intestinal hypoperistalsis or Megacystis-Megaureter Syndrome [1111-1114].

Megacystis
In the first trimester, foetal megacystis is defined as a bladder with a longitudinal diameter ≥ 7 mm, and in
the 2nd and 3rd trimester as an enlarged bladder failing to empty during an extended US examination lasting
at least 40 minutes. Two thirds of cases are secondary to CLUTO and the remainder are associated with genetic
syndromes, developmental or chromosomal abnormalities including anorectal malformations; 14% were normal or
having isolated urological abnormality (e.g. VUR, Duplex system) [1115]. A more recent systematic review showed
that at least 45% of cases have oligohydramnios and 15% have chromosomal abnormalities, most of them being
trisomy 13, 18 and 21. Final diagnoses were posterior urethral valve (PUV) (57%), urethral atresia/stenosis (7%),

90 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


prune-belly syndrome (4%), megacystis-microcolon-intestinal hypoperistalsis syndrome (MMIHS) (1%), cloacal
abnormality (0.7%) and undefined pathologies (36.5%). Termination of pregnancy rate was 50% [1116].
The prognosis of the foetus depends on the underlying pathology, the timing of diagnosis, presence
of an oligo-, anhydramnios and bladder volume. Fontanella et al developed a staging system of CLUTO. They
described three groups: severe (Bladder volume ≥ 5.4 cm3 and/or oligo-, an hydramnios before 20 weeks),
moderate (Bladder volume < 5.4 cm3 and/or normal amniotic fluid at 20 weeks) and mild (Normal AF at 26
weeks) [1117]. This staging system can be used to predict perinatal mortality and post-natal estimated GFR.
Another recent systematic review on prognosis of megacystis patients revealed an overall intrauterine spontaneous
resolution of 32%, with better resolution rates in early (before 18 weeks) megacystis cases (40% vs. 12%) [1118].

3.18.1 Posterior urethral valves


3.18.1.1 Epidemiology, aetiology and pathophysiology
Posterior urethral valves are one of the few life-threatening congenital anomalies of the urinary tract found during
the neonatal period. A recent systematic review showed, that the risk for chronic kidney disease (CKD) could be
up to 32% and for end-stage kidney disease (ESKD) up to 20% [1119]. Up to 17% of paediatric ESKD can be
attributed to PUV [1120]. An incidence of PUV of 1 in 7,000-8,000 live-births has been estimated [1112, 1121].

3.18.2 Classification systems


3.18.2.1 Urethral valve
Up until today, the original classification by Hugh Hampton Young is the most commonly used classification
[1122]. Hugh Hampton Young described three categories: type I, type II and type III. However, today, only type
I and type III are found to be obstructive. As type II seems to be more like a fold and not obstructive, it is no
longer referred to as a valve. Hampton Young’s descriptions of type I and III are as follows:

Type I (90-95%). ‘In the most common type there is a ridge lying on the floor of the urethra, continuous with
the verumontanum, which takes an anterior course and divides into two fork-like processes in the region of the
bulbomembranous junction. These processes are continued as thin membranous sheets, direct upward and
forward which may be attached to the urethra throughout its entire circumference. It is generally supposed that
the valves have complete fusion anteriorly, leaving only an open channel at the posterior urethral wall. Yet, the
fusion of the valves anteriorly may not be complete in all cases, and at this point a slight separation of the folds
exists’ [1122].

Type III. ‘There is a third type which has been found at different levels of the posterior urethra and which
apparently bears no such relation to the verumontanum. This obstruction was attached to the entire
circumference of the urethra, with a small opening in the centre [1112]. The transverse membrane described
has been attributed to incomplete dissolution from the urogenital portion of the cloacal membrane [1123]. The
embryology of the urethral valves is poorly understood. The membrane may be an abnormal insertion of the
mesonephric ducts into the foetal cloaca [1124].

3.18.3 Diagnostic evaluation


An obstruction above the level of the urethra affects the whole urinary tract to varying degrees.
• The prostatic urethra is distended and the ejaculatory ducts may be dilated due to urinary reflux.
• The bladder neck is hypertrophied and rigid.
• The hypertrophied bladder occasionally has multiple diverticula.
• Nearly all valve patients have dilatation of both upper urinary tracts. This may be due to the valve
itself and the high pressure in the bladder, or due to obstruction of the ureterovesical junction by the
hypertrophied bladder.
• If there is secondary reflux, the affected kidney functions poorly in most cases.

During prenatal US screening, bilateral hydroureteronephrosis and a distended bladder are suspicious signs
of a urethral valve. A thick-walled bladder seems to be of better prediction of a PUV than a dilated posterior
urethra (‘keyhole’ sign) [1125]. However, differentiation between obstructive and non-obstructive aetiologies
on prenatal US is challenging as both have a similar US appearance [1126] In the presence of increased
echogenicity of the kidney, dilatation of the urinary tract and oligohydramnion, the diagnosis of a PUV
should strongly be considered. Prenatal US is adequate in most of the cases (90%) [1127]. However in some
circumstances when technical US conditions are poor such as oligo- or anhydramnios, large maternal body
habitus, unfavourable position of the foetus or in suspicion of complex foetal anomalies such as accompanying
gastrointestinal system, foetal MRI may provide additional information [1127-1129].
Post-natally, a voiding cystourethrogram confirms the diagnosis of a PUV. This study is essential
whenever there is a question of an infravesical obstruction, as the urethral anatomy is well-outlined during

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 91


voiding. A secondary reflux is observed in at least 50% of patients with PUV [1130]. Reflux is consistently
associated with renal dysplasia in patients with PUV. It is generally accepted that reflux in the renal units acts
as a ‘pressure pop-off valve’, which would protect the other kidney, leading to a better prognosis [1131]. Other
types of pop-off mechanisms include bladder diverticula and urinary extravasation, with or without urinary
ascites [1132]. However, in the long-term, this supposed protective effect did not show a significant difference
compared to other patients with PUV [1133, 1134].

Nuclear renography with split renal function is important to assess kidney function (DMSA or MAG3).
Creatinine, blood urea nitrogen and electrolytes should be monitored closely during the first few days. Initial
management includes a multi-disciplinary team involving a paediatric nephrologist. The clinician must be
aware of a noteworthy association between PUV and undescended testes and/or inguinal hernia [1135].
Undescended testes occurred in 12-17% of PUV which is consistent with a 10-fold increase [1136].

3.18.4 Management
3.18.4.1 Antenatal treatment
Today most of PUV are discovered before birth [1111-1114]. The intrauterine obstruction leads to a decreased
urine output, which could result in an oligo- or anhydramnios. Amniotic fluid is necessary for normal development
of the lung and its absence may lead to pulmonary hypoplasia, causing a life-threatening problem.
Kidneys start to produce urine at around 10th weeks of antenatal life. Many of the megacystis cases
(7-15 mm) with normal karyotype spontaneously resolve before 20 weeks, whereas it is unlikely in those with a
bladder length > 15mm (> 12mm before age of 18 Weeks of gestation) [1126, 1137, 1138]. Antenatal imaging
of kidneys before 20 weeks is difficult and in rare instances imaging could be done earlier via transvaginal
route [1139]. The possible spontaneous resolution chance of bladder enlargement and timing of proper kidney
imaging are possibly the main obstacles on the optimum timing for prenatal intervention.

As renal dysplasia is not reversible, it is important to identify those foetuses with good renal function. A sodium
level below 100 mmol/L, a chloride value of < 90 mmol/L and an osmolarity below 200 mOsm/L found in
three foetal urine samples gained on three different days are associated with a better prognosis [1140]. Urine
samples before 23 weeks of gestation (ß2-microglobline, sodium, chloride and calcium) may be helpful to
distinguish between those who could benefit from intrauterine therapy and those in whom the outcome is most
likely to be compromised [1141]. The status of amniotic fluid, the appearance of the kidneys as well as the
foetal urine biochemistry could be helpful in counselling the caregivers.
Prenatal interventions aim to restore amniotic fluid volume and attenuate the risk of pulmonary
hypoplasia or further renal damage [1142]. Decision for prenatal intervention can be based on a staging system
that is composed of renal ultrasonographic findings, amnion amount and foetal urine biochemistry [1113].
Early intervention – before the age of 16 weeks of gestation, may be beneficial for the renal function, however
making the correct diagnosis and the detection of other severe co-morbidties is extremly difficult at this time
point [1143]. Later interventions are mostly of benefit for the lung development, but not for renal function.

The placing of a vesicoamniotic shunt has a complication rate of 21-59% with dislocation of the shunt being
the most common one [1142]. The PLUTO-trail (randomised study) failed to show any long-term benefit on
renal function by placing a visual analogue scale (VAS) [1144]. A recent meta-analysis on interventions for
CLUTO reported that VAS resulted in a higher perinatal survival rate than conservative management (57.1% vs
38.8%) with no significant differences in 6-12 month survival, 2-year survival or postnatal renal function [1145].
Foetal cystoscopy with laser ablation has a high complication rate without evidence for the effectiveness of
these interventions [1146]. To avoid the severe complication of the laser ablation, balloon dilation is tried [1147].
The number of patients included and designs of these studies are insufficient to give any recommendations.
Parental information is very important and the natural history of CLUTO including the postnatal outcomes with
or without prenatal treatment as well as the uncertainties and/or controversies about CLUTO diagnosis and
treatment should be discussed [1142].

3.18.4.2 Postnatal treatment


Bladder drainage. If a boy is born with suspected PUV, drainage of the bladder and, if possible, an immediate
VCUG is necessary. A neonate can be catheterised with a small catheter without a balloon, preferably a feeding
tube. A VCUG is performed to see if the diagnosis is correct and whether the catheter is within the bladder and
not in the posterior urethra. An alternative option is to place a suprapubic catheter, perform a VCUG and leave
the tube until the neonate is stable enough to perform an endoscopic incision or resection of the valve.

Valve ablation. When the medical situation of the neonate has stabilised and the creatinine level decreased,
the next step is to remove the intravesical obstruction. In cases where the urethra is too small to safely pass

92 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


a small foetal cystoscope, a suprapubic diversion is performed until valve ablation can be performed. Small
paediatric cystoscopes and resectoscopes are now available either to incise, ablate or to resect the valve at the
4-5, 7-8 or 12 o’clock position, or at all three positions, depending on the surgeon’s preference. It is important
to avoid extensive electrocoagulation, as the most common complication of this procedure is stricture
formation. Two studies demonstrated a lower urethral stricture rate using the cold knife compared to diathermy
[1148, 1149]. Within the three months following initial treatment, effectiveness of the treatment should be
demonstrated either by clinical improvement (US and renal function), control VCUG or a re-look cystoscopy,
depending on the clinical course [1150-1152] .

Vesicostomy. If the child is too small and/or too ill to undergo endoscopic surgery, a suprapubic diversion is
performed to drain the bladder temporarily. If initially a suprapubic tube has been inserted, this can be left in
place for six to twelve weeks. Otherwise, a cutaneous vesicostomy provides an improvement or stabilisation
of the UUT in up tp 90% of cases [1153, 1154]. Although there has been concern that a vesicostomy could
decrease bladder compliance or capacity, so far there are no valid data to support these expectations [1155,
1156]. Moreover, it was shown in PUV patients with stage 3 CKD that adding vesicostomy to valve ablation no
long-term benefit was noted from diversion in the ultimate incidence of ESKD [1157].

High diversion. If bladder drainage is insufficient to drain the UUT, high urinary diversion should be considered.
Diversion may be suitable if there are recurrent infections of the upper tract, no improvement in renal function
and/or an increase in upper tract dilatation, despite adequate bladder drainage. The choice of urinary diversion
depends on the surgeon’s preference for high-loop ureterostomy, ring ureterostomy, end ureterostomy or
pyelostomy, with each technique having advantages and disadvantages [1158-1161]. Diversion can delay
progression to end stage renal failure [1157]. Reconstructive surgery should be delayed until the UUT has
improved as much as can be expected.

Reflux is very common in PUV patients (up to 72%) and it is described bilaterally in up to 32% [1162]. During
the first months of life, antibiotic prophylaxis may be given especially in those with high-grade reflux [876] and
in those with a phimosis, circumcision can be discussed in order to reduce the risk of UTIs [1163]. However,
there are no randomised studies to support this for patients with PUV. Early administration of oxybutynin may
improve bladder function as shown in one study with eighteen patients [1164]. High-grade reflux is associated
with a poor functioning kidney and is considered a poor prognostic factor [1165, 1166]. However, early removal
of the renal unit seems to be unnecessary, as long as it causes no problems. Moreover, in the long term it
may be necessary to augment the bladder and in this case the ureter may be used [1167]. Deterioration of
renal function without a fixed obstruction and higher urine output (polyuria) may lead to an overdistension of
the bladder during the night. Drainage of the bladder during the night by a catheter may be beneficial for the
hydronephrosis as well as for renal function [1168, 1169]. Patients with high daytime PVR urine may benefit
from CIC [1170, 1171]. In those who do not want or are not able to perform a CIC via urethra, the placement of
a Mitrofanoff is a good alternative [1172].

3.18.5 Follow-up
Several prognostic factors have been described. Different serum nadir creatinine levels are given in the
literature (0.85 mg/dl-1.2 mg/d (μmol/L) [1173-1176]. Renal parenchyma quantity (total renal parenchymal area)
and quality (corticomedullary differentiation and renal echogenicity) on initial postnatal US also have prognostic
value [1177].

Life-long monitoring of these patients is mandatory, as bladder dysfunction (‘valve bladder’) is not uncommon
and the delay in day- and night-time continence is a major problem [1178, 1179]. The literature demonstrates
that urodynamic studies plays an important role in the management of patients with valve bladder especially
in those with suspicion of bladder dysfunction [1180, 1181]. Poor bladder sensation and compliance, detrusor
instability and polyuria (especially at night) and their combination are responsible for bladder dysfunction. In
those with bladder instability, anticholinergic therapy can improve bladder function. However, there is a low risk
of reversible myogenic failure (3/37 patients in one study) [1182, 1183]. In patients with poor bladder emptying,
α-blocker can be used to reduce the PVR urine, as demonstrated in one study with 42 patients using terazosin
(mean PVR was reduced from 16 to 2 mL) [1184]; in another study tamsulosin was effective [1185]. Concerning
bladder neck incision, there is no Panel consensus concerning indication and efficacy. High creatinine nadir
(> 1 mg/dL) and severe bladder dysfunction are risk factors for renal replacement therapy [1186, 1187]. Renal
transplantation in these patients can be performed safely and effectively [1188, 1189]. Deterioration of the graft
function is mainly related to LUTD [1188]. Therefore, it is essential to have and keep a good reservoir function.
An assessment and treatment algorithm is provided in Figure 12.

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 93


There are only few reports on sexual function and fertility in patients with PUV demonstrating some impairment
especially in those who are on dialysis [1190, 1191]. In a review the majority have good erectile function (74-
94%) and a fertility comparable to the normal population [1192]. However, a negative influence of the individual
patient’s fertility has to be taken into account, as these patients have a higher risk for bilateral cryptorchidism,
recurrent epididymitis and ESRD [1192].

Figure 12: An algorithm on the assessment, management and follow-up of newborns with possible PUV

Newborn with possible PUV, UUT dilation and renal insufficiency

• USG and VCUG


• Assesment of renal function and electrolyte disorders

Confirm diagnosis

No stabilisation
Bladder drainage

Nephrological care, if needed

Valve ablation when infant is stable

Improvement in No improvement No improvement


UT dilation and RF but stable and ill

Consider diversion
• Close follow-up
• Monitor urinary infection
• Monitor renal function
• Monitor night-time polyuria
and bladder over-extension

• Progressive loss of renal Short term


function
• Recurrent infections
• Poor emptying

• Check residual PUV


• CIC if not emptying
• Consider overnight drainage
• Consider α-blockers
• Anticholinergics if OAB
Long term

Consider augmentation and Mitrofanoff

CIC = clean intermittent catheterisation; OAB = overactive bladder; PUV = posterior urethral valve;
RF = renal function; UT = urinary tract; UUT = upper urinary tract; VCUG = voiding cystourethrogram.

3.18.6 Summary
Posterior urethral valves are one of the few life-threatening congenital anomalies of the urinary tract found
during the neonatal period and despite optimal treatment result in renal insufficiency in nearly one-third of
cases. Bilateral hydroureteronephrosis and a distended bladder are suspicious signs of a PUV in neonates. A
VCUG confirms a PUV diagnosis. Nuclear renography with split renal function is important to assess kidney
function and serum creatinine nadir above 80 μmol/L is correlated with a poor prognosis. Today, antenatal
therapy is becoming more and more popular. Identification of those with an obstructive uropathy and definiton
of those who would benefit from early antenatal intervention are the major challenges. Postnatal treatment

94 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


includes bladder drainage, either transurethral or suprapubic and if the child is stable enough, endoscopic
incision of the valve is performed. If a child is too small and/or too ill to undergo endoscopic surgery, a
vesicostomy is an option for bladder drainage. If bladder drainage is insufficient to drain the UUT, high urinary
diversion should be considered.
In all patients life-long monitoring is mandatory, as bladder dysfunction is quite common and may
cause progressive upper tract deterioration, if not managed properly. In the long-term between 10 and 47% of
patients may develop end-stage renal failure. Renal transplantation in these patients can be performed safely
and effectively.

• Anterior urethral valve (AUV)


Anterior urethral valve is a semilunar or iris-like band of tissue on ventral aspect of urethra. It can be isolated,
in association with or confused with urethral diverticulum. The aetiology of isolated AUV is speculated to be
secondary to congenital urethral obstruction, malunion of glanular and penile urethra, congenital cystic dilatation
of peri-urethral glands or ruptured distal lip of a syringocele [1193]. Anterior urethral valve occurs less frequently
than PUV. It can be present in the bulbous urethra, the penoscrotal junction and penile urethra. Patients may
present with poor urinary stream, penile ballooning, UTI or haematuria. Anterior urethral valves have been
classified by Firlit et al depending on the presence of diverticulum and the dilatation of urethra and upper tract
[1194]. The diagnosis is based on VCUG with possible findings of dilated or elongated posterior urethra, a
dilatation of the anterior urethra, a thickened trabeculated bladder, a hypertrophied bladder neck, VUR, and
urethral diverticula. In doubtful cases, retrograde urethrography may be helpful showing linear filling defect along
the ventral wall, or it may show a dilated urethra ending in a smooth bulge or an abrupt change in the caliber of
the dilated urethra on VCUG [1195]. Treatment is performed mainly by endoscopic valve ablation. In selected
patients, a temporary diversion may be considered until the child is big enough for endoscopy to be possible.
Open surgery is reserved in patients with very large diverticulum and defective spongiosum. Renal failure may
develop in 22% and the risk is highest in patients with pre-treatment azotaemia, VUR and UTI [1196].

• Anterior urethral diverticulum (AUD):


Common postnatal presenting features of AUD are compressible ventral penile swelling, urinary dribble post-
micturition, voiding difficulty, poor stream, and recurrent UTIs [1197-1199]. Diagnosis is made by VCUG with or
without a retrograde urethrogram. In small AUD, endoscopic cutting or deroofing of distal lip of the diverticulum
can be used as a treatment modality. Larger diverticulum requires excision of the diverticulum with a two-
layered urethroplasty; or marsupialisation with staged urethroplasty. In cases of urosepsis and obstructive
uropathy, a suprapubic catheter may be placed. Once the infant’s condition improves, temporary urinary
diversion with vesicostomy or proximal cutaneous urethrostomy can be performed before definitive surgical
management [1200, 1201]. The diverticulum is associated with a distal lip-like tissue which may be confused
with a valve. Anatomically, AUV have normal corpus spongiosum development whereas AUD have incomplete
spongious tissue formation [1200].

• Syringocele
Cowper glands are two bulbourethral glands located within the urogenital diaphragm and secrete pre-ejaculatory
mucus on both sides through the external sphincter into the urethra 1-2 cm distal to the sphincter. Syringocele is
the cystic dilatation of these glands. The aetiology can be congenital (retention cyst of the intraurethral portion of
the duct) or acquired (trauma or infection). It has been classified as simple, imperforate, perforate and ruptured
[1202]. A simpler grouping is suggested to merge simple, perforate and ruptured into “open syringocele” and
imperforate to “closed syringocele”. Closed syringoceles cause obstructive symptoms and open ones act as
a diverticula and cause post-voiding drippling and sometimes obstruction due to orientation of one membrane
into urethra [1203] . However, it is better to simply categorise into two groups as obstructing and non-obstructing
in terms of understanding pathophysiology and management [1204]. Depending on the syringocele type,
patients present with post-void dribbling, urethral discharge, UTI, perineal pain, haematuria, obstructive voiding
symptoms, dysuria and retention. Diagnosis is based on antegrade and/ or retrograde urethrogram which
shows a cystic defect distal to prostate. If the VUC/RGU are inconclusive, US and/or MRI may be used if open
reconstruction is being planned. Endoscopic deroofing of the cyst in both obstructing and non-obstructing
syringoceles is an effective method of marsupialisation [1205]. In cases where endoscopic approach is not
feasible open correction may be considered.

• Cobb’s collar
Cobb’s collar is a congenital membranous stricture of the bulbar urethra. It is different from congenital
obstructive posterior urethral membrane (COPUM) and is independent of the verumontanum and
external sphincter and may represent a persistence of part of the urogenital membrane [1206]. Voiding
cystourethrogram shows narrowing in the proximal bulbar urethra with folds extending proximally, a dilated

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 95


posterior urethra, prominent bladder neck and other findings of infravesical obstruction. Treatment is an
endoscopic incision; using cold-knife showed lower recurrence rates than electrocautery [1207].

• Urethral atresia/hypoplasia
Male urethral atresia is a congenital, complete obstruction of the urethra caused by a membrane that is
usually located at the distal end of the prostatic urethra. The urethra distal to this point is usually hypoplastic,
presumably from lack of foetal voiding [1208]. Urethral atresia is associated with bladder distention, VUR,
hydronephrosis and renal dysplasia [1209]. Most cases reported have the phenotypic characteristics of the
prune belly syndrome. Antenatal intervention may be beneficial in terms of foetal survival [1210]. Although
progressive augmentation by dilating the urethra anterior (PADUA) procedure was described as a treatment
modality, the majority of cases requires some form of supravesical diversion [1208, 1209].

• Posterior Urethral Polyps:


Although, posterior urethral polyps (PUP) does not cause antenatal hydronephrosis, it could cause obstruction
later in life. Posterior Urethral Polyps is a polypoid, pedunculated, fibroepithelial lesion arising in posterior
urethra proximal to the verumontanum. It lies on the floor of the urethra with its tip reaching into the bladder
neck and obstruction occurs because of distal displacement of polyp during urination [1211]. Patients complain
of dysuria, haematuria and obstructive symptoms such as poor urinary stream and intermittent retention
episodes. Diagnosis can be suspected by VCUG and/or US but is confirmed during cystourethroscopy.
Treatment is usually an endoscopic resection of the polyp. The course of the disease is benign and no
recurrences were reported in the literature [1212, 1213].

3.18.7 Summary of evidence and recommendations for the management of posterior urethral valves

Summary of evidence LE
Posterior urethral valves are one of the few life-threatening congenital anomalies of the urinary tract 1b
found during the neonatal period.
Antenatal therapy could be discussed based on ultrasound findings, fetal urine biochemistry amount 4
of amnion fluid and chromosomal status.
Despite optimal treatment nearly one-third of the patients end up in renal insufficiency. 2b
Bilateral hydroureteronephrosis and a distended bladder are suspicious signs on US; a VCUG 2b
confirms the diagnosis.
Serum creatinine nadir above 85 μmol/L is correlated with a poor prognosis. 2a
In the long-term up to 20% of patients develop end-stage renal failure due to primary dysplasia and/ 2a
or further deterioration because of bladder dysfunction. Renal transplantation in these patients is safe
and effective, if the bladder function is normalised.

Recommendations LE Strength rating


Diagnose posterior urethral valves (PUV) initially by ultrasound but a voiding 3 Strong
cystourethrogram (VCUG) is required to confirm the diagnosis.
Assess split renal function by dimercaptosuccinic acid scan or Strong
mercaptoacetyltriglycine (MAG3) clearance. Use serum creatinine as a prognostic
marker.
Vesico-amniotic shunt antenatally is not recommended to improve renal outcome. 1b Weak
Offer endoscopic valve ablation after bladder drainage and stabilisation of the child. 3 Strong
Offer suprapubic diversion for bladder drainage if the child is too small for valve Strong
ablation.
Offer a high urinary diversion if bladder drainage is insufficient to drain the upper Strong
urinary tract and the child remains unstable.
Monitor bladder and renal function life-long, in all patients. 3 Strong

3.19 Rare Conditions in Childhood


3.19.1 Urachal remnants
3.19.1.1 Introduction
The urachus is an embryonic structure arising as a result of the separation of the allantois from the ventral cloaca.
The allantois appears on day sixteen as a tiny, fingerlike outpouching from the caudal wall of the yolk sac, which
is contiguous with the ventral cloaca at one end and the umbilicus at the other. The ventral portion of the cloaca
develops into the bladder after cloacal division by the urogenital septum. Thus, the bladder initially extends all

96 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


the way to the umbilicus [1214]. With progressive foetal development, as the bladder descends into the pelvis,
the attachment between the umbilicus and the urachus becomes looser and the apical portion progressively
narrows to a small, epithelialised, fibromuscular strand by the fourth or fifth month of gestation. The urachus then
obliterates completely by birth, forming the median umbilical ligament [1215-1217].

The urachus varies from 3 to 10 cm in length and from 8 to 10 mm in diameter. It is a three-layered tubular
structure, the innermost layer being lined with transitional epithelium, the middle layer composed of connective
tissue, and the outermost muscular layer in continuity with the detrusor muscle [1218].

Urachal remnants (URs) originate from failure of the obliteration of the allantois, resulting in a urachal anomaly
such as (1) urachal sinus, (2) urachal cyst, (3) vesico-urachal diverticulum, and (4) patent urachus [1215, 1216,
1219]. Most often the urachal anomaly is asymptomatic, but it occasionally may become infected, may cause
urinary symptoms, or develop a urachal carcinoma in later life [1218, 1220].

3.19.1.2 Epidemiology
Reports of occurrence rates in the literature vary broadly from a very rare disease in the older literature to a fairly
common problem. Robert et al. found that URs were present in 61.7% of patients younger than 16 years [1221].
They also noted that the frequency of URs decreased with increasing age. This supports a physiological regression
of URs with age. Stopak et al. attributed this upsurge to increased awareness among community pediatricians and
improvements in US that made visualisation of urachal remnants easier [1222].
Clinical studies and paediatric autopsy studies in the past have shown a much lower incidence. Rubin
found an incidence of 1 in 7,610 cases of patent urachus and 1 in 5,000 cases of urachal cysts [1223]. Nix et al.
noted three anomalies out of 1,168,760 hospital admissions, and Blichert-Toft et al. reported five UR cases out of
40,000 patients [1224, 1225]. The incidence rate in males is a little higher than in females [1226, 1227].
The range of the various URs reported in the literature is 10% to 48% for patent urachus, 31% to
43% for urachal cyst, 18% to 43% for urachal sinus and 3% to 4% for urachal diverticulum [1228, 1229].

3.19.1.3 Symptoms
A patent urachus causes continuous or intermittent urine leakage from the umbilicus causing umbilical
granulation and erythema in infants [1228]. A urachal cyst is usually diagnosed (1) incidentally, or (2) when it
becomes infected causing abdominal pain and discharge of pus from the umbilicus or recurrent UTIs when it
drains into the bladder.
The most common symptom is umbilical granulation, discharge and erythema in infants and
abdominal pain in older children [1228].
Other symptoms of infected urachal anomalies can vary from high fever, abdominal pain, urinary
tract infections, LUTS and/or an abdominal mass [1229-1233]. A urachal diverticulum is often asymptomatic
and is usually found incidentally during investigations for other problems. An alternating sinus can empty
either into the bladder or the umbilicus and this characteristic is responsible for various presentations [1234].
Infection has been reported as the most common complication in urachal anomalies [1235]. Severe infection
may develop into peritonitis and sepsis. Cultures from umbilical discharge usually show Staphylococcus,
Streptococcus and E. Coli [1236].

• Other congenital anomalies:


Ashley found a simultaneous anomaly in 17 of 46 children, of which VUR was the most common anomaly
(6 patients) [1237]. Other investigators reported associated anomalies in cases of persistent URs including
meatal stenosis, hypospadias, umbilical and inguinal hernias, cryptorchidism, anal atresia, omphalocele,
ureteropelvic obstruction and most frequently, VUR [1227, 1238-1240].

3.19.1.4 Diagnosis
In the majority of cases with complaints of a UR, a careful history and physical examination will confirm the
suspicion of a UR. In many patients this can be confirmed by US studies [1221]. An MRI or CT scan may be
necessary in a minority of children [1232]. Because of the association with other congenital abnormalities, other
studies such as a VCUG or cystoscopy may be undertaken as well. In general, the VCUG is only undertaken
when the child also presents with UTI or when the US shows signs of upper tract abnormalities. For the
diagnosis per se it is not necessary [1241]; however, a VCUG may be useful for defining the type of urachal
anomaly and evaluating a population that may be at higher risk for VUR.

3.19.1.5 Treatment
If a UR is symptomatic, the standard approach has been surgical removal. In most cases it should be done
as an elective procedure, following appropriate treatment of active inflammation, and infection is possible.

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 97


Pre-operative IV-dosage of antibiotic like Cefazolin is generally sufficient. A Pfannenstiel, periumbilical
or infraumbilical midline incision can all be used for the open surgical approach [1231, 1242]. Even in
symptomatic infants a more conservative approach is possible as well, especially in children less than six
months old. Observation and treatment with antibiotics if necessary and radiographic monitoring are a safe
approach [1228, 1243, 1244]. Dethlefs et al. reported a 90% successful outcome [1230], while Naiditch et al.
reported that 44 of 78 symptomatic patients resolved under observation [1233]. More recently the laparoscopic
approach has been advocated, and shown to be safe [1244-1246]. Surgery is not without risk. The rate of
complications following surgical removal varies from 0 to 20%: usually wound infections [1222, 1230-1233,
1242]. Considering the probable additional risk of anaesthesia in very young children any surgical procedure
needs to be assessed carefully [1247, 1248].

3.19.1.6 Pathology of removed remnants


Removed specimens may show inflammation or a cystic structure [1230]. Patients presenting without symptoms
are as likely to have epithelial elements in the UR as those presenting with symptoms [1232].

3.19.1.7 Urachal cancer


Urachal anomalies are thought to be associated with an increased risk of bladder adenocarcinoma in adults,
and urachal adenocarcinoma has an estimated incidence of 0.18 per 100,000 individuals yearly [1249]. These
cases account for 0.1 to 0.3% of all bladder malignancies and 20 to 39% of bladder adenocarcinomas [1250].
Urachal adenocarcinoma (UrC) is very rare, especially when one considers that up to 62% of children under 16
years of age may have a UR [1221, 1251]. A study by Copp et al. found no association between the presence of
UR symptoms and the presence or absence of epithelial tissue in pathology specimens, leading them to conclude
that UR symptoms have poor predictive value for malignancy potential in these remnants [1227].

Gleason et al. found that 5,721 URs would need to be excised to prevent a single case of urachal
adenocarcinoma out of the nearly 65,000 patients reviewed [1249]. Assuming that epithelium is required in the
development of urachal adenocarcinoma, the extrapolated Number Needed to Treat (NNT) would be more than
8,000, as nearly 30% of urachal anomalies are void of an epithelial component. Less than 5% of urachal cancers
have a non-epithelial origin such as sarcoma [1252]. The presenting symptoms in adults are different from those
in children: in a study of 130 adult patients, Ashley et al. found that 49% presented with haematuria and 27% with
pain. In 51% a urachal carcinoma was diagnosed: adenocarcinoma, with 58% high grade cancer. In addition,
20% had metastases at diagnosis, the overall 5-year cancer specific survival rate in the UrC cohort was 49%
[1253]. Stasis of urine and crystallisation promotors such as mucus or desquamated epithelium in the UR are
most likely the cause for malignant degeneration as well as stone formation in the adult patient. At present no
long-term follow-up on untreated UR in children is available and there is no evidence that urachal anomalies in
children increase the likelihood of future malignancy [1228, 1254].

3.19.1.8 Conclusion
Urachal remnants appear to be more common than previously reported. During the first 6-12 months of life
spontaneous resolution is common. Excision of symptomatic urachal anomalies is an effective and safe means
of treatment, with minimal morbidity. However, most patients with simple and asymptomatic lesions do not
appear to benefit from excision, as the risk of malignancy later in life is vanishingly remote. Early intervention
(< 6 months of age) should be reserved for patients with persistent documented urine draining from the
urachus or a documented abscess. Incidental (US) UR management remains a challenge and should be
done with patient and family involvement to make the most informed decision. While surgical intervention has
minimal risk and morbidity, it is performed unnecessarily in a large proportion of asymptomatic patients due
to the unnecessary removal of non-epithelial containing urachal anomalies and the inability to predict which
anomalies will undergo malignant transformation [1255].

3.19.1.9 Recommendation for management of urachal remnants

Recommendations Strength rating


Urachal remnants with no epithelial tissue carry little risk of malignant transformation. Strong
Asymptomatic and non-specific atretic urachal remnants can safely be managed non- Strong
operatively.
Urachal remnants (URs) incidentally identified during diagnostic imaging for non-specific Strong
symptoms should also be observed non-operatively since they tend to resolve spontaneously.
A small urachal remnant, especially at birth, may be viewed as physiological. Strong
Urachal remnants in patients younger than 6 months are likely to resolve with non-operative Strong
management.

98 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


Follow-up is necessary only when symptomatic for 6 to 12 months. Strong
Surgical excision of urachal remnants solely as a preventive measure against later Strong
malignancy appears to have minimal support in the literature.
Only symptomatic URs should be safely removed by open or laparoscopic approach. Strong
A VCUG is only recommended when presenting with febrile UTIs. Strong

3.19.2 Papillary tumours of the bladder in children and adolescents (Papillary urothelial neoplasm of
low malignant potential or transitional cell carcinoma)
3.19.2.1 Incidence
Papillary tumours of the bladder in children and adolescents are extremely rare and are different from papillary
tumours in adults. A “grape-like” papillary tumour in young children will be more likely a rhabdomyosarcoma
of the bladder, which are not the focus of this guideline. A papillary tumour in older children or adolescents will
be more likely be a papillary urothelial neoplasm of low malignant potential (PUNLMP) [1256]. Children with risk
factors, such as previous bladder surgery and immunosuppressive medication can also develop a nephrogenic
adenoma of the bladder, also presenting as a papillary tumour of the bladder.

3.19.2.2 Differences and similarities of papillary tumours of the bladder in children and adults
Gender
The overall the risk of a papillary tumour in the bladder in paediatric and young adult patients is approximately
double in males compared to females [1257].

3.19.2.3 Risk factors


The majority of these patients have no identifiable risk factors.

3.19.2.4 Presentation
The most common symptom at presentation is haematuria; other less common symptoms include abdominal
pain, storage LUTS including frequency, dysuria and at times obstructive symptoms [1257].

3.19.2.5 Investigations and treatment


Ultrasound of the genitourinary tract is the first investigation of choice. It is an excellent screening tool and
can often accurately diagnose the nature and location of lesion. In children and adolescents, a bladder US
of the full bladder is more sensitive compared with adults due to reduced abdominal fat and thinner muscle
layer [1258]. In the event of a need to differentiate the renal or bladder origin of the haematuria, a red blood
cell morphology will reveal isomorphic blood cells, differentiating a bladder origin. Urine cytology can be
performed, however it has very limited value likely due to the low-grade nature of these tumours in children.
Cystoscopy should be reserved if a bladder tumour is suspected on imaging for simultaneous diagnosis and
treatment, transurethral resection of the tumour. In children, cystoscopy requires general anaesthesia [1259].

3.19.2.6 Histology
All the lesions in the children and adolescent age-group are identified as papillary and over 85% are solitary
[1258]. Papillary bladder tumours in patients younger than twenty years of age have low-grade non-invasive
disease (WHO classification) [1260]. These findings let pathologists conclude that in children and adolescents,
a papillary bladder tumour can be classified as Papillary Urothelial Neoplasm of Low Malignant Potential
(PUNLMP). PUNLMP has minimal or no cytological atypia and it differs from low grade transitional cell
carcinoma (TCC) which has cytologic atypia, hyperchromatic nuclei and scattered mitosis [1261].

3.19.2.7 Additional treatment


Mitomycin C and Bacillus Calmette-Guerin have both been used in children but there is no evidence of their
efficacy due to the rarity of TCC, and especially of high grade TCC [1257]. Hence, as per current evidence,
there is no place for instillations in children.

3.19.2.6 Prognosis, recurrence and surveillance


The prognosis of papillary tumours of the bladder in children is overall good. The recurrence rate in children
and adolescents varies from 8 to 15% [1256-1258]. Mean time to recurrence can vary from 11 to 29 months
depending on the study, with recurrences occurring up to 90 months from diagnosis; though 64% occur in the
first year [1257]. In certain cases, recurrences can be fairly aggressive [1258].
Strategies are based on the guidelines and protocols of papillary tumours of the bladder in adults.
It is advised to follow-up children and adolescents with a history of a PUNLMP initially with a short interval of
three to six months in the first year, and thereafter at least yearly with urinanalysis for haematuria and an US
of the full bladder. In the event of sudden gross haematuria, the evaluation must be performed immediately. If

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 99


the tumour was completely resected at primary surgery, standard follow-up cystoscopy is not necessary and
may be reserved for children or adolescents with a high recurrence risk or suspected recurrence on bladder US
[1258]. The exact duration of follow-up is unknown but this Panel recommends follow-up for at least five years.

Inflammatory myofibroblastic tumours of the bladder (IMTB) are rare with nearly 200 cases reported in the
literature [1262, 1263]. Around 25% occur in children with a median age at diagnosis of 7.5 years and a median
tumour size of 5.5 cm. Boys and girls are equally affected [1264]. Usually these tumours are benign, with only
very few reported malignant cases [1265]. Treatment is mostly surgical with transurethral resection, but local
resection, or partial cystectomy maybe needed in selected cases [1264, 1266]. Additionally, a conservative
approach is reported [1267]. Histological examination is required to exclude other malignant tumours such
as a rhabdomyosarcoma. In children, no recurrence has been reported so far. However due to the malignant
potential and few recurrences in adults, follow-up as for papillary bladder tumours is recommended.

Summary of evidence LE
Majority of paediatric patients have no identifiable risk factors for bladder tumours. 3
There is no evidence on intravesical therapy for bladder tumours in children and adolescents. 4
Prognosis of papillary tumours of the bladder in children is good overall. 3
Inflammatory myofibroblastic bladder tumours are usually benign. 3

Recommendations LE Strength rating


Ultrasound is the first investigation of choice for the diagnosis of paediatric bladder 3 Strong
tumours.
Cystoscopy should be reserved if a bladder tumour is suspected on imaging for 3 Strong
diagnosis and treatment.
After histological confirmation, inflammatory myofibroblastic bladder tumours should 4 Weak
be resected locally.
Follow-up should be every 3-6 months in the first year, and thereafter at least 4 Weak
annually with urinanalysis and an ultrasound for at least 5 years.

3.19.3 Penile rare conditions


Paediatric lesions of the penis are uncommon but an important part of the paediatric urological practice. The
commonest of these lesions are cystic penile lesions followed by vascular malformations and neurogenic
lesions [1268]. Soft tissue tumours of the male external genitalia are uncommon, but have been described in
the paediatric age group and can be malignant [1269].

3.19.3.1 Cystic lesions


• Epidermal inclusion cysts are the commonest genital cystic lesion and can occur anywhere on the body
in both men and women; in the penis it occurs most commonly over the penile shaft varying from 0.1 to
1 cm in diameter. Their epithelium is lined and filled with keratin. It is a painless swelling and can present
in the age group with a history of circumcision. Treatment by total surgical excision is mainly indicated for
cosmetic or symptomatic (e.g. infection) reasons and should be performed without rupturing the cyst to
avoid recurrence [1270].

• Mucoid cyst of the penis is synonymous with parameatal cyst or genitoperineal cyst of median
raphe; they are midline developmental cysts arising from ectopic urethral mucosa filled with mucoid
material. They present since birth but are usually detected during adolescence or later. They are usually
asymptomatic developing over penile ventral surface around glans and require surgical removal for either
cosmetic, functional or symptomatic reasons [1271].

• Median raphe cysts arise from incomplete closure of genital fold during embryogenesis; they are
commonly diagnosed in the first decade of life but can present later as they tend to be asymptomatic
[1272]. They are either unilocular or multilocular fluid containing cysts, with a mean size of 0.8 cm but
cysts larger than 2 cm have also been reported [1273]. Cysts are centred in dermis, with no connection to
urethra or epidermis. Histopathologically, there are 4 types: urethral (urothelium-like epithelium, account
for 55% cases), epidermoid, glandular and mixed. They can be treated conservatively and can resolve
spontaneously or persist. Cyst aspiration is associated with high risk of recurrence and surgical excision
is the treatment of choice. Though most penile cysts are asymptomatic, they may get infected resulting
in pain and tenderness. They can also present with ulceration, rupture and urinary obstruction if they are

100 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


close to the urethral meatus. This along with cosmetic issues means that most caregivers and patients
opt for surgical excision.

• Smegmal cysts or smegmal pearls can be a differential for the cysts above; they are a benign collection
of smegma in the sub-preputial space in uncircumcised boys with anticipated spontaneous resolution [1274].

• Dermoid cyst are congenital, asymptomatic, firm, solitary, subcutaneous cystic lesions occurring
commonly in the region of the corona involving the foreskin. Histopathologically they contain sweat
and sebaceous glands with elements of hair and squamous epithelium. Pilosebaceous cysts have been
described on the glans; they are benign and usually diagnosed after excision.

3.19.3.2 Vascular malformations


A broad classification of penile vascular lesions into haemangiomas and vascular malformations was proposed
by Ramos in 1999 [1275]. Haemangiomas develop rapidly at birth and involute slowly; they also include
pyogenic granulomas which are benign outgrowths of cutaneous capillary vessels formed usually from
chronic irritation [1268]. The growth cycle of infantile haemangiomas is divided into early and late proliferative
stages, followed by a slow involution phase, completing growth by nine months of age [1276]. Propranolol is
currently first line treatment for infantile haemangiomas, the exact mechanism of action is unknown but can
include inhibition of angiogenesis, vasoconstriction among others. The dose is in the range of 1.5-2.5 mg/kg,
which needs to be continued for 12 to 18 months and then tapered through active or passive weaning to
reduce risk of rebound growth [1276]. Other factors leading to rebound growth after propranolol treatment
include deep haemangiomas, which occur in about 38% patients despite propranolol therapy, requiring local
therapy such as topical timolol, pulsed dye laser or intralesional steroids. After twelve months, the median
improvement with treatment is reported as 81% (range 70-90%) based on VAS scores of serial patient
photographs.

Vascular malformations are congenital lesions of capillary, lymphatic and venous (or slow-flow) or arterial/
arteriovenous (fast-flow) origin that enlarge slowly as the patient grows. These include glomus tumours,
which are primarily congenital arteriovenous shunts that develop from thermo-regulatory glomus bodies (fast-
flow vascular malformations). Glomus tumours of the penis can arise on the glans penis, corpora of the penis
and as periurethral masses, sometimes accompanied by glomus tumours of fingers and feet [1277]. These
are usually asymptomatic at presentation or may have symptoms such as priapism, palpitation and perineal
pain. Glomus tumours are benign despite exhibiting high grade nuclear polymorphism. Vascular malformations
are usually benign and treated either with laser, sclerotherapy or surgical excision. However, glomus tumours
specifically need surgical treatment and follow-up due to the risk of recurrence from incomplete excision [1278].

3.19.3.3 Neurogenic lesions


Penile neurofibroma is an extremely rare lesion arising from perineural and Schwann cells, and occurs
usually with evidence of systemic neurofibromatosis or von Recklinghausen syndrome [1279]. They are treated
successfully with complete excision [1268]. Rare cases of malignant schwanomas on the penis presumably
secondary to malignant transformation of benign neurofibromas have been reported in boys with a strong
family history of neurofibromatosis. This type of malignant degeneration of neurofibromatosis occurs in
reportedly 5-16% children [1279]. Hence, these patients require long-term follow-up due to risk of recurrence,
new tumour formation and malignant transformation.

3.19.3.4 Soft tissue tumours of penis


Mesenchymal tumours are rare in the external genitalia and they require excision in order to differentiate
between benign and malignant neoplasms. Histopathological characterisation is essential to ensure malignant
tumours receive radical treatment with adjuvant therapy or close follow-up [1269].
Presentation is usually of a painless penile mass, that is non-tender and rubbery on examination.
Ultrasound maybe useful in characterising the lesion but is not diagnostic; it can exclude urethral invasion if it is
close to urethra [1269]. Once an excision biopsy is performed, if aggressive malignant components are found, a
further wider resection may be needed.

Fibrosarcoma is a rare non-rhabdomyosarcoma soft tissue tumour that arises from fibrous tissue. The infantile
form of fibrosarcoma is rare and those occurring on the penis are even rarer in the paediatric age-group.
Surgical intervention has a favourable prognosis in the paediatric age group with long-term survival of 90% in
sporadic cases [1280].
Myofibroma is a benign congenital lesion that occurs either as a solitary lesion or as a part of
myofibromatosis with multiple soft tissue tumours. Excision is necessary for histological diagnosis [1269].

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 101


Primary penile teratomas are extremely rare subtype of congenital germ cell tumours, and they tend to be
asymptomatic and are subdermal on US with no blood flow on Doppler [1281]. They need aggressive treatment
with surgical resection due to their unpredictable behavior and unresponsiveness to chemotherapy. Mature
teratomas are benign but immature teratoma or even mixed teratomas with immature components can turn
malignant and have the potential to metastasise and recur.

3.19.3.5 Penile Lymphedema


Lymphedema in adults is usually secondary to malignancy or infectious disease affecting lymphatic drainage.
In the paediatric age group, however, lymphedema is usually primary and generally very rare, affecting 1.2
per 100.000 persons under the age of 20 years [1282]. Of these, only a very small fraction relates to the
genital region. Regardless of underlying aetiology, inefficient lymphatic drainage leads to accumulation of
subcutaneous lymph which causes tissue swelling and inflammation. This in turn stimulates adipose deposition
and fibrosis further exacerbating enlargement. With time the edematous tissue becomes vulnerable to
infection, chronic cutaneous changes and disfigurement [1283]. Additionally, when occurring in the genital
region urological complications may ensue; such as phimosis, haematuria, bleeding, bladder outlet obstruction,
pain, dysuria, lymphorrhea and severe psychological distress due to resultant deformity [1284, 1285].

In the largest cohort of male genital oedema in the paediatric age group, 92% of cases were primary; of these
only 25% had a discernable familial or syndromic association such as Noonan syndrome, lymphedema-
distichiasis or Milroy disease [1284]. Secondary genital lymphedema in children has been reported after
inguinal surgery, and non-caseating granulomatous lymphangitis as seen with metastatic Crohn’s disease
[1284-1286]. Average age of onset was reported to be 4.5 ± 6.3 years with 61% presenting in infancy, 13%
in childhood and the remaining 26% in adolescence. Edema is usually penoscrotal in 72%, isolated scrotal in
24% and very rarely confined exclusively to the penis in 4%. Moreover, concomitant lower limb edema is the
rule in two thirds of cases [1284].

There is no general consensus on diagnostic work-up of these patients. History and physical examination
(including family history) is usually sufficient. However lymphoscintigraphy can be used as a confirmatory test,
more so for limb than genital edema where results can be difficult to interpret [1284]. Ultrasonography is non-
specific, but has been advocated by some to exclude secondary lymphedema by examining the patency of
iliac and caval vessels [1287]. Magnetic resonance imaging is useful to exclude other differential diagnoses
such as other venous or lymphatic anomalies [1284].

Conservative treatment is the accepted first-line treatment. The mainstay is compression therapy to maintain
and prevent further swelling. This can be achieved by compression stockings and undergarments. Additionally,
close observation and protection of the skin to prevent excoriations and infection is essential [1284, 1287].
Compression therapy is however, less effective on genital oedema than it is on limb edema, especially in
growing children. When conservative management fails, and especially in symptomatic cases, or in patients
with functional impairment, surgical debulking may be necessary. This can either take the form of circumcision
in cases where the foreskin is affected or excision of affected skin and subcutaneous tissues with restructuring
and contouring for optimal cosmetic outcome. Complete skin excision and grafting may also be required
[1284-1287]. Surgical management can be challenging and needs to be restricted to patients with significant
symptoms. Complications include recurrences, continuous lymphatic leakage, haematoma, infection and poor
cosmetic outcome [1282, 1287, 1288].

Summary of evidence LE
Cystic penile lesions are the commonest paediatric penile lesions followed by vascular malformations 3
and neurogenic lesions.
Neurofibroma patients require long-term follow-up due to risk of recurrence, new tumour formation 3
and malignant transformation.
Mesenchymal tumours are rare and require excision in order to differentiate between benign and 3
malignant neoplasms.

Recommendations LE Strength rating


Treatment of penile cystic lesions is by total surgical excision, it is mainly indicated 4 Weak
for cosmetic or symptomatic (e.g. infection) reasons.
Propranolol is currently the first-line treatment for infantile haemangiomas. 2b Strong
Conservative management is the first-line treatment for penile lymphedema. 4 Weak
In symptomatic cases or in patients with functional impairment, surgical intervention 4 Weak
may become necessary for penile lymphedema.

102 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


3.20 Paediatric urological trauma
Trauma is the leading cause of morbidity and mortality in children and is responsible for more childhood deaths
than the total of all other causes [1289]. In about 3% of children seen at paediatric hospital trauma centres,
there is significant involvement of the genitourinary tract [1290]. This is caused by either blunt injuries from falls,
car accidents, sports injuries, physical assault, sexual abuse, or penetrating injuries, usually due to falls onto
sharp objects or from gunshot or knife wounds.

3.20.1 Paediatric renal trauma


3.20.1.1 Epidemiology, aetiology and pathophysiology
In blunt abdominal trauma, the kidney is the most commonly affected organ, accounting for about 10% of all
blunt abdominal injuries [1289].
Children are more likely than adults to sustain renal injuries after blunt trauma because of their
anatomy. Compared to an adult kidney, a child’s kidney is larger in relation to the rest of the body and often
retains foetal lobulations, so that blunt trauma is more likely to lead to a local parenchymal disruption. The
paediatric kidney is also less well protected than the adult kidney. Children have less peri-renal fat, much weaker
abdominal muscles, and a less ossified and therefore much more elastic and compressible thoracic cage [1291].
Blunt renal trauma is usually a result of sudden deceleration of the child’s body, particularly due
to sport accidents, falls, and contact with blunt objects. Deceleration or crush injuries result in contusion,
laceration or avulsion of the less well-protected paediatric renal parenchyma.

3.20.1.2 Classification systems


Renal injuries are classified according to the kidney injury scale of the American Association for the Surgery of
Trauma (Table 7) [1292].

Table 7: R
 enal injury classified according to the kidney injury scale of the American Association for the
Surgery of Trauma [1292]

Grade Type of injury Description


I Contusion Non-visible or visible haematuria
Haematoma Normal urological studies
II Haematoma Non-expanding subcapsular haematoma
Laceration Laceration of the cortex of < 1.0 cm
III Laceration Laceration > 1.0 cm without rupture of collecting system
IV Laceration Through the cortex, medulla and collecting system
Vascular Vascular injury
V Laceration Completely shattered kidney
Vascular Avulsion of the renal hilum

3.20.1.3 Diagnostic evaluation


In a child who has sustained blunt abdominal trauma, renal involvement can often be predicted from the history,
physical examination and laboratory evaluation. Renal involvement may be associated with abdominal or flank
tenderness, lower rib fractures, fractures or vertebral pedicles, trunk contusions and abrasions, and haematuria.

3.20.1.3.1 Haematuria
Haematuria may be a reliable finding. In severe renal injuries, 65% suffer visible haematuria and 33% non-
visible, while only 2% have no haematuria at all [1293].
The radiographic evaluation of children with suspected renal trauma remains controversial. Some
centres rely on the presence of haematuria to diagnose renal trauma, with a threshold for renal involvement
of 50 RBCs/HPF. Although this may be a reliable threshold for significant non-visible haematuria in trauma,
there have been many reports of significant renal injuries that manifest with little or even no blood in the urine
[1294]. It is therefore compulsory to consider all the clinical aspects involved, including the history, physical
examination, consciousness of the child, overall clinical status and laboratory findings to decide on the
diagnostic algorithm and whether or not a child needs further imaging studies.

3.20.1.3.2 Blood pressure


It is important to consider that children, unlike adults, are able to maintain their blood pressure, even in the
presence of hypovolaemia, due to compliance of the vascular tree and mechanisms for cardiac compensation
[1295]. As blood pressure is an unreliable predictor of renal involvement in children, some centres recommend
imaging of the urinary tract in children with any degree of haematuria following significant abdominal trauma.

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 103


3.20.1.3.3 Choice of imaging method
Nowadays, CT is the best imaging method for renal involvement in children. Computed tomography scanning
is the cornerstone of modern staging of blunt renal injuries especially when it comes to grading the severity of
renal trauma.
Computed tomography scanning is quite rapid and usually performed with the injection of contrast
media. To detect extravasation, a second series of images is necessary since the initial series usually finishes
60 seconds after injection of the contrast material and may therefore fail to detect urinary extravasation. In
acute trauma, US may be used as a screening tool and for reliably following the course of renal injury. However,
US is of limited value in the initial and acute evaluation of trauma. The standard intravenous pyelogram (IVP)
is a good alternative imaging method if a CT scan is not available. It is superior to US but not as good as CT
scanning for diagnostic purposes.

3.20.1.4 Disease management


The modern management of trauma is multidisciplinary, requiring paediatricians, emergency physicians,
surgeons, urologists, and other specialties as required.
Non-surgical conservative management with bed rest, fluids and monitoring has become the
standard approach for treating blunt renal trauma. Even in high-grade renal injuries, a conservative approach
is effective and recommended for stable children. However, this approach requires close clinical observation,
serial imaging, and frequent re-assessment of the patient’s overall condition. Therefore, a good initial trauma
CT with delayed images to check for urinary extravasation is recommended since this may prevent repeat
ionising scans. In stable patients with grade 2 or higher lesions a close follow up with US 48 to 72 hours after
the initial scan is sufficient and should be considered before repeating a CT scan [1296]. A systematic review
supports application of conservative management protocols also to high-grade blunt paediatric renal trauma.
At this time, emergent operative intervention only for haemodynamic instability is recommended. Minimally
invasive interventions including angio-embolisation, stenting, and percutaneous drainage should be used when
indicated [1297]. Absolute indications for surgery include persistent bleeding into an expanding or unconfined
haematoma. Relative indications for surgery are massive urinary extravasation and extensive non-viable
renal tissue [1298]. A recently published meta-analysis concluded with the following recommendations: (1) In
paediatric patients with blunt renal trauma of all grades, non-operative management vs. operative management
in haemodynamically stable patients is strongly recommended. (2) In haemodynamically stable paediatric
patients with high-grade (AAST grade III-V) renal injuries, angio-embolisation vs. surgical intervention for
ongoing or delayed bleeding is strongly recommended; and, (3) In paediatric patients with renal trauma, routine
blood pressure checks to diagnose hypertension is recommended in the long-term follow-up [1299]. However,
long-term data on the risk of developing hypertension is lacking.

3.20.1.5 Recommendations for the diagnosis and management of paediatric renal trauma

Recommendations Strength rating


Use imaging in all children who have sustained a blunt or penetrating trauma with any level Strong
of haematuria, especially when the history reveals a deceleration trauma, direct flank trauma
or a fall from a height.
Use rapid spiral computed tomography with delayed images scanning for diagnostic and Strong
staging purposes.
Manage most injured kidneys conservatively. Strong
Offer surgical intervention in case of haemodynamic instability and a Grade V renal injury. Strong

3.20.2 Paediatric ureteral trauma


Injuries to the ureter are rare. The ureter is well protected; the upper part is protected by its close
approximation to the vertebral column and paraspinal muscles and the lower part by its route through the
bony pelvis. In addition, the ureter is a small target, and both flexible and mobile. This also means that ureteral
injuries are caused more often by penetrating trauma than blunt trauma [1300]. Since the ureter is the sole
conduit for urinary transport between the kidney and the bladder, any ureteral injury can threaten the function
of the ipsilateral kidney.

3.20.2.1 Diagnostic evaluation


Since there are no classical clinical symptoms suggestive of ureteral trauma, it is important to carry out a
careful diagnostic work-up using different imaging modalities. Unfortunately, initial imaging studies, such as
IVP and routine CT scans, are unreliable. A study of eleven disruptions of the ureteropelvic junction found that
72% had a normal or non-diagnostic IVP on initial studies [1300]. Diagnostic accuracy of CT scanning can be

104 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


improved by performing a delayed CT scan up to ten minutes after injection of the contrast material [1301]. The
most sensitive diagnostic test is a retrograde pyelogram.
Quite a few patients present several days after the injury, when the urinoma produces flank and
abdominal pain, nausea and fever.
Due to symptoms being often vague, it is important to remain suspicious of a potential undiagnosed
urinary injury following significant blunt abdominal trauma in a child.

3.20.2.2 Management
Immediate repair during abdominal exploration is rare. Minimally invasive procedures are the method of
choice, especially since many ureteral injuries are diagnosed late after the traumatic event. Percutaneous or
nephrostomy tube drainage of urinomas can be successful, as well as internal stenting of ureteral injuries [1302].
If endoscopic management is not possible, primary repair of partial lacerations should be followed by internal
stenting. The management of complete lacerations, avulsions or crush injuries depends on the amount of ureter
lost and its location. If there is an adequate healthy length of ureter, a primary ureteroureterostomy can be
performed. If primary re-anastomosis is not achievable, distal ureteral injuries can be managed using a psoas
bladder hitch, Boari flap or even nephropexy. Proximal injuries can be managed using transureteroureterostomy,
auto-transplantation or ureteral replacement with bowel or appendix [1303].

3.20.2.3 Recommendations for the diagnosis and management of paediatric ureteral trauma

Recommendations Strength rating


Diagnose suspected ureteral injuries by retrograde pyelogram. Strong
Manage ureteral injuries endoscopically, using internal stenting or drainage of an urinoma, Weak
either percutaneously or via a nephrostomy tube.

3.20.3 Paediatric bladder injuries


The paediatric bladder is less protected than the adult bladder, and is therefore more susceptible to injuries
than the adult bladder, especially when it is full, due to:
• Its higher position in the abdomen and its exposure above the bony pelvis.
• The fact that the abdominal wall provides less muscular protection.
• The fact that there is less pelvic and abdominal fat surrounding the bladder to cushion it in trauma.

Blunt trauma is the most common cause of significant bladder injury. In adults, bladder injury is often
associated with pelvic fractures. This is less common in children because the paediatric bladder sits above
the pelvic ring. In a large prospective study, only 57% of children with pelvic fractures also had a bladder injury
compared to 89% of adults [1304].

3.20.3.1 Diagnostic evaluation


The characteristic signs of bladder injury are suprapubic pain and tenderness, an inability to urinate, and visible
haematuria (95% of injuries). Patients with a pelvic fracture and visible haematuria present with a bladder
rupture in up to 45% of cases [1305].
The diagnosis of bladder rupture can be difficult in some cases. The bladder should be imaged
both when fully distended and after drainage using standard radiography or a CT scan. The best results can
be achieved by retrograde filling of the bladder using a catheter. Despite advances in CT imaging, the bladder
must still be filled to capacity to accurately diagnose a possible bladder injury [1306].

Blunt injuries to the bladder are categorised as:


• contusions with damage to the bladder mucosa or muscle, without loss of bladder wall continuity or
extravasation;
• ruptures, which are either intraperitoneal or extraperitoneal.

Intraperitoneal bladder ruptures are more common in children because of the bladder’s exposed position and
the acute increase in pressure during trauma. These cause the bladder to burst at its weakest point, i.e. the
dome. Extraperitoneal lesions occur in the lower half of the bladder and are almost always associated with
pelvic fractures. A cystogram will show extravasation into the perivesical soft tissue in a typical flame pattern
and the contrast material is confined to the pelvis.

3.20.3.2 Management
Contusions usually present with varying degrees of haematuria and are treated with catheter drainage alone.

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 105


3.20.3.2.1 Intraperitoneal injuries
The accepted management of intraperitoneal bladder ruptures is open surgical exploration and primary repair.
Post-operative drainage with a suprapubic tube is mandatory. Recent data suggest that transurethral drainage
may be as effective, with fewer complications, resulting in shorter periods of diversion [1307]. Usually, after
about seven to ten days, a repeat cystogram is performed to ensure healing is taking place properly.

3.20.3.2.2 Extraperitoneal injuries


Non-operative management with catheter drainage for seven to ten days alone is the method of choice
for extraperitoneal bladder rupture. However, if there are bone fragments within the bladder, these must
be removed and the bladder must then be repaired and drained, according to the principles for treating
intraperitoneal ruptures [1308].

3.20.3.3 Recommendations for the diagnosis and management of paediatric bladder injuries

Recommendations Strength rating


Use retrograde cystography to diagnose suspected bladder injuries. Strong
Ensure that the bladder has been filled to its full capacity and an additional film is taken Strong
after drainage.
Manage extra-peritoneal bladder ruptures conservatively with a transurethral catheter left in Strong
place for seven to ten days.
Do not delay treatment of intra-peritoneal bladder ruptures by surgical exploration and Strong
repair as well as post-operative drainage for seven to ten days.

3.20.4 Paediatric urethral injuries


Except for the penile part of the urethra, the paediatric urethra is quite well protected. In addition, its shape and
elasticity mean the urethra is seldom injured by trauma. However, a urethral injury should be suspected in any
patient with a pelvic fracture or significant trauma to the perineum until confirmed otherwise by a diagnostic
work-up.

3.20.4.1 Diagnostic evaluation


Patients with suspected urethral trauma and pelvic fractures usually present with a history of severe trauma,
often involving other organ systems.
Signs of urethral injury are blood at the meatus, visible haematuria, and pain during voiding or an
inability to void. There may also be perineal swelling and haematoma involving the scrotum. A rectal examination
to determine the position and fixation of the prostate is important in any male with a suspected urethral injury. The
prostate, as well as the bladder, may be displaced up out of the pelvis, especially in membranous urethral trauma.
Radiographic evaluation of the urethra requires a retrograde urethrogram. It is important to expose
the entire urethral length, including the bladder neck. If a catheter has already been placed by someone else
and there is suspected urethral trauma, the catheter should be left in place and should not be removed.
Instead, a small infant feeding tube can be placed into the distal urethra along the catheter to allow the
injection of contrast material for a diagnostic scan [1309].

3.20.4.2 Disease management


Since many of these patients are unstable, the urologist’s initial responsibility is to provide a method of draining
and monitoring urine output.
A transurethral catheter should only be inserted if there is a history of voiding after the traumatic
event, and if a rectal and pelvic examination, as described above, has not suggested a urethral rupture. If the
catheter does not pass easily, an immediate retrograde urethrogram should be performed.
A suprapubic tube may be placed in the emergency department percutaneously, or even in the
operating room, if the patient has to undergo immediate exploration because of other life-threatening injuries.
There are often no associated injuries with a bulbous urethral or straddle injury and management
is therefore usually straightforward. In these cases, a transurethral catheter is the best option for preventing
urethral bleeding and/or painful voiding [1310].
The initial management of posterior urethral injuries remains controversial, mainly regarding the
long-term results with primary realignment compared to simple suprapubic drainage with later reconstruction.

The main goals in the surgical repair of posterior urethral injuries are:
• Providing a stricture-free urethra.
• Avoiding the complications of urinary incontinence and impotence.

106 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


Suprapubic drainage and late urethral reconstruction was first attempted because immediate surgical repair
had a poor outcome, with significant bleeding and high rates of incontinence (21%) and impotence in up to
56% of cases [1311]. In adults, a study of the success rates of delayed repair reported re-structure rates of
11-30%, continence rates of 90-95% and impotence rates of 62-68% [1312]. However, in children, there
is significantly less experience with delayed repair. The largest paediatric series of delayed repair in 68
boys reported a success rate of 90% [1313]. Another study reported strictures and impotence in 67% of
boys, although all the boys were continent [1129]. A recently published follow-up study on 15 patients who
underwent delayed urethroplasty for blunt urethral trauma during childhood reported high long-term success
rates with a low rate of long-term urinary and sexual dysfunction in adulthood [1314].
An alternative to providing initial suprapubic drainage and delayed repair is primary realignment of
the urethra via a catheter. The catheter is usually put in place during open cystostomy by passing it from either
the bladder neck or meatus and through the injured segment. In a series of fourteen children undergoing this
procedure, this resulted in a stricture rate of 29% and incontinence in 7% of patients [1315].

3.20.4.3 Recommendations for the diagnosis and management of paediatric trauma

Recommendations Strength rating


Assess the urethra by retrograde urethrogram in case of suspected urethral trauma. Strong
Perform a rectal examination to determine the position of the prostate. Strong
Manage bulbous urethral injuries conservatively with a transurethral catheter. Strong
Manage posterior urethral disruption by either: Weak
• primary reconstruction;
• primary drainage with a suprapubic catheter alone and delayed repair;
• primary re-alignment with a transurethral catheter.

3.21 Peri-operative fluid management


3.21.1 Epidemiology, aetiology and pathophysiology
Children have a different total body fluid distribution, renal physiology and electrolyte requirements, as well as
weaker cardiovascular compensation mechanisms, compared to adults [1316]. During development, children
have a high metabolic rate and lower fat and nutrient stores which means they are more susceptible to
metabolic disturbances caused by surgical stress [1317]. The metabolic response to anaesthesia and surgery in
infants and children is related to the severity of the operation [1318].

3.21.2 Disease management


3.21.2.1 Pre-operative fasting
Pre-operative fasting has been advocated for elective surgery to avoid the complications associated with
pulmonary aspiration during induction of anaesthesia. New regimens include a 30-60 minute limitation for
clear liquids [1319, 1320] without increased risk of pulmonary aspiration [1321]. Several studies have shown
that fasting times in clinical practice often exceed the guidelines with average fasting times of 6-10 hours
[1320-1322]. Compared to adults, children have a higher metabolic rate and low glycogen stores and impaired
gluconeogenesis, which makes hypoglycaemia an important issue to consider, especially in children < 36
months old [1320]. Therefore, it is important to prevent too long fasting times. Clear-liquid carbohydrate drinks
have been proposed to reduce these fasting times [1323].

Table 8 provides the current six, four and one hour guidelines for pre-operative fasting for elective surgery
[1320, 1322].

Table 8: Pre-operative fasting times for elective surgery

Ingested material Minimum fasting period (hours)


Clear liquids 1
Breast milk 4
Light meal 6

3.21.2.2 Maintenance therapy and intra-operative fluid therapy


Generally, the anaesthetist is responsible for intra-operative management and the surgeon is responsible
for post-operative instructions. The goal of intra-operative fluid management is to sustain homeostasis by
providing the appropriate amount of parenteral fluid; this maintains adequate intravascular volume, cardiac

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 107


output and oxygen delivery to tissues at a time when normal physiological functions have been altered by
surgical stress and anaesthetic agents.

In recent years new strategies for maintenance and replacement fluid management have been developed
and this has changed intra-operative fluid management significantly. The main goal of intra-operative fluid
management is to maintain a normal extracellular fluid volume (EFV). During the intra-operative period
fluid deficits may be induced by blood loss or pre-operative fasting. These fluid deficits can be replaced
by balanced isotonic electrolyte solutions to restore a normal EFV. It is recommended that maintenance
intravenuous (IV) fluids should consist of balanced isotonic solutions with appropriate potassium chloride
and dextrose in order to decrease the risk of hyponatraemia development [1324]. No increased risk for
hypernatraemia, fluid overload with aedema and hypertension, and hyperchloremic acidosis was found, which
was always feared for isotonic solutions [1324].
When children are clinically unstable due to third-space losses, these losses should be replaced
with crystalloids (normal saline or Ringer’s lactate). Third-space losses may vary from 1 mL/kg/h for a minor
surgical procedure to 15-20 mL/kg/h for major abdominal procedures, or even up to 50 mL/kg/h for surgery
of necrotising enterocolitis in premature infants. When this fluid management is insufficient replacement
management with colloids (albumin, gelatine and hydroxyethyl starch (HES) should be adopted, using a
restrictive approach [1325].

Clinical guidelines have been proposed by Sümpelmann et al. [1325] regarding intra-operative fluid
management (Table 9).

Table 9: Intra-operative fluid management

Solution for infusion Initial/repeated dose


Background infusion Balanced isotonic solution + 1-2% glucose 10 mL/kg/h
Fluid therapy Balanced isotonic solution X 10-20 mL/kg
Volume therapy Albumin, Gelatine, hydroxyethyl starch X 5-10 mL/kg
Transfusion Red blood cells, fresh frozen plasma, platelets X 10 mL/kg

3.21.2.3 Post-operative feeding and fluid management


It is not obligatory to check serum chemistry after uncomplicated surgery in children with normal pre-operative
renal and hepatic function. However, if oral intake has been postponed for > 24 hours (e.g. as in intestinal
surgery), there is an increased risk of electrolyte abnormalities, requiring further assessment and subsequent
management, particularly with potassium. Post-operative findings, such as decreased bowel movements and
ileus, may be signs of hypokalaemia.
Children who undergo interventions to relieve any kind of obstructive diseases deserve particular
attention, especially due to the risk of polyuria as a result of post-obstructive diuresis [1326]. In children who
develop polyuria, it is important to monitor fluid intake and urine output, as well as renal function and serum
electrolytes. If necessary, clinicians should not hesitate in consulting with a paediatric nephrologist.

In children who have undergone non-abdominal surgery, studies have suggested that gastric motility returns to
normal one hour after emergence from anaesthesia [1327]. Early post-operative intake of fluid in children who
have undergone minor or non-abdominal urological surgery is associated with reduced post-operative vomiting
and lower opioid use [1328] and is therefore encouraged.

In abdominal surgery the enhanced recovery after surgery (ERAS) protocol has been implemented in the
paediatric population following its success in adults [1323, 1329]. The ERAS protocol is a multimodal approach
to prevent the post-operative effects of the surgical stress response. This protocol includes pre- and intra-
operative element such as minimal pre-operative fasting and careful intra-operative fluid management, and also
focuses on post-operative care. The post-operative ERAS protocol suggests starting clear fluid intake on the
evening of surgery and a normal diet the day after surgery and thereby early discontinuation of IV fluids. Further
focus is on early mobilisation, preventing epidurals and omitting or early removal of external tubes [1323, 1329].
The implementation of an ERAS protocol has resulted in shorter length of hospital stays, faster bowel recovery
and opioid-free post-operative need [1323, 1329, 1330]. When implementing ERAS in children with neurological
abnormalities special attention should be given to bowel management with pre-operative treatment of
constipation and early post-operative continuation of routine bowel management.

108 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


3.21.3 Summary of evidence and recommendations for the management of peri-operative fluid
management

Summary of evidence LE
Children are not simply smaller physiological versions of adults. They have their own unique metabolic 2
features, which must be considered during surgery.
During the intra-operative period balanced isotonic electrolyte solutions can be used to maintain a 1
normal extracellular fluid volume.
Following abdominal surgery ERAS protocols can be used to reduce recovery times and complications. 1

Recommendations Strength rating


Ensure shorter pre-operative fasting periods for elective surgeries (up to one hour for clear Strong
liquids).
Use ERAS protocols for abdominal surgery in children with normal bowel movement. Strong
Use isotonic solutions in hospitalised children because they are at high risk of developing Strong
hyponatraemia.
Assess the baseline and daily levels of serum electrolytes, glucose, urea and/or creatinine Strong
in every child who receives intravenous fluids, especially in intestinal surgery (e.g. ileal
augmentation), regardless of the type of solution chosen since there is an increased risk of
electrolyte abnormalities in children undergoing such surgery.
Start early oral fluid intake in all patients scheduled for minor surgical procedures. Strong

3.22 Post-operative pain management: general information


3.22.1 Epidemiology, aetiology and pathophysiology
The provision of adequate pain control requires proper pain evaluation, accurate choice of drug and route of
administration, and consideration of age, physical condition and type of surgery and anaesthesia [1331].
Traditional medical beliefs that neonates are incapable of experiencing pain have now been
abandoned following recent and better understanding of how the pain system matures in humans, better pain
assessment methods and a knowledge of the clinical consequences of pain in neonates [1332, 1333]. Many
studies have indicated that deficient or insufficient analgesia may be the cause of future behavioural and
somatic sequelae [1334, 1335]. Our current understanding of pain management in children depends fully on the
belief that all children, irrespective of age, require adequate pain treatment.

3.22.2 Diagnostic evaluation


Assessment of pain is the first step in pain management. Several pain assessment tools have been validated
according to the child’s age, cultural background, mental status, communication skills and physiological
reactions [1336]. Depending on the child’s age, the 0-10 Numeric Rating Scale, Faces Revised Pain Scale
or Colour Analog Scale, for example, can be used [1337]. One of the most important topics in paediatric
pain management is informing and involving the child and caregivers during this process. Patient-family-
controlledanalgesia is the preferred pain management in the hospital and at home if provided with the correct
information [1337, 1338].

3.22.3 Disease management


3.22.3.1 Drugs and route of administration
Pre-emptive analgesia is an important concept that aims to induce the suppression of pain before neural
hypersensitisation occurs [1339]. Regional anaesthesia are given intra-operatively which can include a regional
nerve block, caudal blocks or local wound infiltration and has proven to reduce the need for post-operative
analgesia [1340]. The WHO’s ‘pain ladder’ is a useful tool for the pain management strategy [1341]. A three
level strategy seems practical for clinical use. Post-operative management should be based on sufficient intra-
operative pre-emptive analgesia with regional or caudal blockade followed by balanced analgesia. Paracetamol
and non-steroidal anti-inflammatory drugs (NSAIDs) are the drugs of choice at the first level. As they become
insufficient to prevent pain, weak and strong opioids are added to oral drugs to achieve balanced analgesia.
Every institute must build their own strategy for post-operative analgesia. A proposed strategy for post-
operative analgesia may be as follows:
1. Intra-operative regional or caudal block.
2. Paracetamol + NSAID.
3. Paracetamol + NSAID + weak opioid (e.g. tramadol or codeine).
4. Paracetamol + NSAID + strong opioid (e.g. morphine, fentanyl, oxycodone or pethidine).

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 109


The use of opioids in children has long held a standard role in the post-operative management of pain.
Increased recognition of the adverse effects of opioids and prolonged opioid dependency demand a balanced
intra-operative administration of opioids [1337, 1342]. Intra-operative adequate dosage of paracetamol and
NSAIDs results in a decrease in opioid requirement in children [1343, 1344]. Furthermore, opioid awareness
among physicians could reduce opioid use. When prescribing lower opioid dosage, this did not increase pain
scores in urological outpatient surgeries [1345]. Caution is necessary to take account of renal function when
using NSAIDs. Paediatric dependent dosages for most common used pain medication can be found in this
publication [1346].

3.22.3.2 Circumcision
Circumcision requires anaesthesia and proper pain management [1347]. Potential analgesic interventions
during circumcision include the use of a dorsal penile nerve block (DPNB) or ring block, topical anaesthetics
(e.g. lidocaine-prilocaine cream, or 4% liposomal lidocaine cream), and sucrose preferably in combination
[1340, 1346]. Caudal blockade methods have similar efficacy compared to DPNB. However, caregivers should
be informed about the more frequent incidence of post-operative motor weakness and micturition problems
[1348]. Ultrasound guidance can be used [1346].

3.22.3.2.1 Penile, inguinal and scrotal surgery


Caudal blocks and peripheral nerve blocks (DPNB and pudendal) are commonly used methods for analgesia
following surgery for hypospadias. Several agents with different doses, concentrations and administration
techniques have been used and shown to be adequate. Overall post-operative pain scores were lower with
pudendal nerve blocks. No increase in post-operative complications was seen with these types of blocks
[1340, 1349, 1350]. Severe bladder spasms caused by the presence of the bladder catheter may sometimes
cause more problems than pain and is managed with antimuscarinic medications. For inguinoscrotal surgery,
various regional anaesthesia methods have been investigated, such as transversus abdominis plane block,
ilioinguinal/iliohypogastric nerve blocks and caudal blocks. All have been shown to have adequate post-
operative analgesic properties. Additional local anaesthetics such as clonidine or dexmedetomidine may
improve results [1340].

3.22.3.3 Bladder and kidney surgery


Continuous local infusion reduces the need for post-operative opioids [1351-1353], as well as systemic
(intravenous) application of analgesics [1354], has been shown to be effective. Ketorolac is an effective agent
that is underused. It decreases the frequency and severity of bladder spasms and the length of post-operative
hospital stay and costs [1355, 1356]. Open kidney surgery is particularly painful because all three muscle layers
are cut during conventional loin incision. A dorsal lumbotomy incision may be a good alternative because of
the shorter post-operative hospital stay and earlier return to oral intake and unrestricted daily activity [1357].
Caudal and paravertebral blocks continuous epidural analgesia, as well as rectus sheath and transversus
abdominis plane blocks have decreased post-operative morphine requirement after abdominal and renal
surgery [1358-1360]. For laparoscopic approaches, intra-peritoneal spraying of local anaesthetic before incision
of the perirenal fascia may be beneficial [1361].

3.22.4 Summary of evidence and recommendations for the management of post-operative pain

Summary of evidence LE
Adequate paracetamol and NSAIDs use reduces opioid need post-operatively. 1
Pain may cause behavioural and somatic sequelae. 3
Every institute must develop their own well-structured strategy for post-operative analgesia. 4

Recommendations Strength rating


Prevent/treat pain in children of all ages. Strong
Evaluate pain using age-compatible assessment tools. Strong
Inform patients and caregivers accurately. Strong
Use pre-emptive and balanced analgesia in order to decrease the side effects of opioids. Strong

3.23 Basic principles of laparoscopic surgery in children


3.23.1 Epidemiology, aetiology and pathophysiology
The use of laparoscopy and robot-assisted laparoscopic surgery is rapidly increasing and has gained
widespread acceptance for many urological surgeries in children. Diagnostic laparoscopy for undescended

110 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


testis, nephrectomy, heminephrectomy, varicocelectomy, pyeloplasty and ureteral reimplantation are some
of the indications which are commonly being performed. This expanding scope related to technological
advancements allows surgeons to perform more complex procedures in a minimally invasive fashion even
in infants and younger children. Generally, well established benefits of minimally invasive surgery are
decreased pain, shorter convalescence and better cosmetics compared to traditional open surgery [807].
Additional advantages of robotic surgery over conventional laparoscopy include ergonomics, 3D vision,
better manoeuvrability, decreased tremor and easy learning curve. Limitations to be considered are increased
operative time, smaller working space at young age, cost and experience of the surgeon and anaesthesiologist.
While the success and complication rates are comparable for nephrectomy and pyeloplasty (see chapter
3.13.3.2) advantages of laparoscopy and robotic surgery for ureteral reimplantation have not been proven and
this can only be recommended for experienced centres (see chapter 3.14.3.2.3).

As worldwide experience increases, there is an accumulating awareness about the physiological consequences
related to intra- and retroperitoneal carbon dioxide (CO2) insufflation in children. In contrast to traditional open
surgery pneumoperitoneum may have physiological responses which require close monitoring during surgery
and should be taken seriously.

3.23.2 Technical considerations and physiological consequences


3.23.2.1 Pre-operative evaluation
Laparoscopy in children requires specific anaesthetic precautions. Physiological effects of CO2
pneumoperitoneum, positioning of the patient and in potentially increased operative time need to be
considered by the anaesthesiology team. Therefore, a detailed medical examination and risk assessment is
mandatory pre-operatively. Especially cardiac and pulmonary system should be assessed since increased
intra-abdominal pressure may lead to decreased ventricular preload [1362].

3.23.2.2 Abdominal insufflation


Abdominal insufflation is the main principle of laparoscopic surgery to create working space for the surgeon.
Carbon dioxide is the most commonly used insufflant in laparoscopic centres throughout the world. Other
alternatives reported are nitrous oxide, helium, argon and air. However, CO2 is considered to be the best
available gas as it is colourless, cheap, has high solubility in the vascular system [1363] and is excreted by
the pulmonary system making it the safest option. Smaller children and infants absorb more CO2 than older
children [1364], suggesting the need for more attention both during and early after laparoscopic surgery for
these children.

Most complications of laparoscopy are attributable to gaining access to the abdominal cavity. One study
reporting complications of > 5,400 paediatric laparoscopic surgeries showed that there was an overall
complication rate of 5.3% of which 4.2% were related to problematic insufflation (subcutaneous emphysema,
gas embolism, injury to the organs and vascular structures, mis-insufflation etc.) [1365]. There are two main and
well-established techniques for initial access to the abdomen or retroperitoneum: open technique (Hasson) and
Veress needle. Studies comparing these two different access techniques in paediatric laparoscopic urological
procedures showed similar complication rates [1366]. The vast majority of the complications were minor and
related to lack of surgical experience. Particularly in infants and smaller children, the open access technique is
recommended by the Panel to reduce the chance of complications.

Elasticity of the abdominal wall is age-related and is higher in infants and small children compared to older
children [1367].

Pneumoperitoneal pressure (PnP in mmHg) is one of the critical points that needs to be carefully considered by
laparoscopic surgeons. A recent RCT compared two different pneumoperitoneal pressure groups (6-8 mmHg
vs. 9-10 mmHg) in infants less than 10 kg [1368]. It demonstrated that higher pressures were associated with
more pronounced respiratory and haemodynamic changes as well as increased post-operative pain scores and
prolonged time to resume feeding.

3.23.2.3 Pulmonary effects


After intra-abdominal insufflation the diaphragm is pushed upwards due to increased abdominal pressure. This
leads to decreased total pulmonary compliance. Combined with CO2 absorption this may lead to hypercarbia
and acidosis, particularly in case of prolonged operative time or low pulmonary reserve such as in infants.
Trendelenburg position may also aggravate the situation in operations in the pelvic region, such as anti-reflux
or bladder neck surgeries. Several studies revealed increased end tidal CO2 (ET CO2) related to CO2 absorption
[1364, 1369, 1370]. One study showed a 33% increase in ET CO2 in the majority of neonatal laparoscopic and

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 111


thoracoscopic procedures [1187]. Shorter operative time and lower intra-abdominal pressures decrease the
risk of increased ET CO2. Hypoxemia is rarely seen, even in neonates and can easily be adjusted by increasing
minute ventilation. These findings highlight the importance of close monitoring of the children.

3.23.2.4 Cardiovascular effects


Intra-abdominal pressure, CO2 absorption and positioning may also affect the cardiovascular system. It has
been shown in adults that after initiation of pneumoperitoneum, cardiac output and stroke volume decrease
while mean arterial pressure, central venous pressure and systemic vascular resistance increase [1371]. Similar
outcomes have been reported during paediatric laparoscopy with some nuances. Cardiac output was 30%
decreased while blood pressure remained stable during laparoscopic orchidopexy with PnP of 10 mmHg
in children between aged 6-30 months [1372]. When PnP was lowered from 12 mmHg to 6 mmHg, cardiac
index and other vascular parameters normalised [1373]. Using high intra-abdominal pressures in infants with
congenital cardiac abnormalities may result in re-opening of cardiac shunts such as the foramen ovale and
ductus arteriosus [1374]. Although cardiovascular effects of using high PnP are clinically measurable, they may
not have a significant clinical impact on healthy children. However, it is clear that using lower pressures is safer
especially in smaller children.

3.23.2.5 Effects on renal function


Although clinical studies in children are lacking, pneumoperitoneum may also have adverse effects on renal
blood flow [1375]. High intra-abdominal pressures and reverse Trendelenburg position may cause decreased
glomerular filtration rate and decreased urine output. One study has shown that 88% of infants and 14% of
children more than one year old develop anuria within 45 minutes after initiation of PnP with 8 mmHg [1376].
However, urine output recovers with temporary polyuria after the operation. Although the clinical relevance of
decreased urine output seems insignificant, it is important to monitor the fluid and electrolyte balance of the
children during and after laparoscopic surgery.

3.23.2.6 Effects on neurological system


Another effect of pneumoperitoneum is increased intracranial pressure (ICP) which normalises after desufflation
of the abdomen [1377]. Trendelenburg position, high PnP and hypoventilation are additional risk factors for
increased ICP. Laparoscopy is therefore contraindicated in patients with intracranial space occupying lesions
[1378]. Children with ventriculo-peritoneal shunts require precautions with regards to shunt drainage, however
laparoscopy is not contraindicated [1379].

3.23.3 Summary of evidence and recommendations for laparoscopy in children

Summary of evidence LE
Laparoscopy and robotic-assisted laparoscopic surgery can safely be performed in children. 1
The general benefits of laparoscopy are decreased pain, shorter convalescence and better cosmetics 1
compared to traditional open surgery.
Limitations to be considered are increased operative time, smaller working space with young age, 1
cost, surgeon and anaesthesiologist experience.
Pneumoperitoneum may have physiological effects which require close monitoring during surgery and 2
should be taken seriously.

Recommendations Strength rating


Use lower intra-abdominal pressure (6-8 mmHg) during laparoscopic surgery in infants and Strong
smaller children.
Use open access for laparoscopy in infants and smaller children. Strong
Monitor for laparoscopy-related cardiac, pulmonary and diuretic responses. Strong

112 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


4. REFERENCES
1. Bogaert, G., et al. Practical recommendations of the EAU-ESPU guidelines committee for
monosymptomatic enuresis-Bedwetting. Neurourol Urodyn, 2019.
https://pubmed.ncbi.nlm.nih.gov/31793066
2. Dogan, H.S., et al. Are EAU/ESPU pediatric urology guideline recommendations on neurogenic
bladder well received by the patients? Results of a survey on awareness in spina bifida patients and
caregivers. Neurourol Urodyn, 2019. 38: 1625.
https://pubmed.ncbi.nlm.nih.gov/31102557
3. Radmayr, C., et al. Management of undescended testes: European Association of Urology/European
Society for Paediatric Urology Guidelines. J Pediatr Urol, 2016.
https://pubmed.ncbi.nlm.nih.gov/27687532
4. Stein, R., et al. EAU/ESPU guidelines on the management of neurogenic bladder in children and
adolescent part I diagnostics and conservative treatment. Neurourol Urodyn, 2019.
https://pubmed.ncbi.nlm.nih.gov/31724222
5. Stein, R., et al. EAU/ESPU guidelines on the management of neurogenic bladder in children and
adolescent part II operative management. Neurourol Urodyn, 2019.
https://pubmed.ncbi.nlm.nih.gov/31794087
6. Stein, R., et al. Urinary tract infections in children: EAU/ESPU guidelines. Eur Urol, 2015. 67: 546.
https://pubmed.ncbi.nlm.nih.gov/25477258
7. Tekgul, S., et al. EAU guidelines on vesicoureteral reflux in children. Eur Urol, 2012. 62: 534.
https://pubmed.ncbi.nlm.nih.gov/22698573
8. Riedmiller, H., et al. EAU Guidelines on Paediatric Urology. Eur Urol, 2001. Nov; 40 (5): 589.
https://pubmed.ncbi.nlm.nih.gov/11752871
9. Guyatt, G.H., et al. What is “quality of evidence” and why is it important to clinicians? BMJ, 2008.
336: 995.
https://pubmed.ncbi.nlm.nih.gov/18456631
10. Guyatt, G.H., et al. GRADE: an emerging consensus on rating quality of evidence and strength of
recommendations. BMJ, 2008. 336: 924.
https://pubmed.ncbi.nlm.nih.gov/18436948
11. Phillips, B., et al. Oxford Centre for Evidence-Based Medicine Levels of Evidence. Updated by
Jeremy Howick March 2009. . 1998. 2014.
https://www.cebm.net/2009/06/oxford-centre-evidence-based-medicine-levels-evidence-
march-2009/
12. Guyatt, G.H., et al. Going from evidence to recommendations. BMJ, 2008. 336: 1049.
https://pubmed.ncbi.nlm.nih.gov/18467413
13. Morris, B.J., et al. Estimation of country-specific and global prevalence of male circumcision. Popul
Health Metr, 2016. 14: 4.
https://pubmed.ncbi.nlm.nih.gov/26933388
14. Gairdner, D. The fate of the foreskin, a study of circumcision. Br Med J, 1949. 2: 1433.
https://pubmed.ncbi.nlm.nih.gov/15408299
15. Kuehhas, F.E., et al. Incidence of balanitis xerotica obliterans in boys younger than 10 years
presenting with phimosis. Urol Int, 2013. 90: 439.
https://pubmed.ncbi.nlm.nih.gov/23296396
16. Celis, S., et al. Balanitis xerotica obliterans in children and adolescents: a literature review and
clinical series. J Pediatr Urol, 2014. 10: 34.
https://pubmed.ncbi.nlm.nih.gov/24295833
17. Oster, J. Further fate of the foreskin. Incidence of preputial adhesions, phimosis, and smegma
among Danish schoolboys. Arch Dis Child, 1968. 43: 200.
https://pubmed.ncbi.nlm.nih.gov/5689532
18. Palmer, L.S., et al., Management of abnormalities of the external genitalia in boys. In: Campbell-
Walsh Urology. 11th ed. Vol. 4. 2016, Philadelphia.
19. Liu, J., et al. Is steroids therapy effective in treating phimosis? A meta-analysis. Int Urol Nephrol,
2016. 48: 335.
https://pubmed.ncbi.nlm.nih.gov/26725071
20. ter Meulen, P.H., et al. A conservative treatment of phimosis in boys. Eur Urol, 2001. 40: 196.
https://pubmed.ncbi.nlm.nih.gov/11528198
21. Elmore, J.M., et al. Topical steroid therapy as an alternative to circumcision for phimosis in boys
younger than 3 years. J Urol, 2002. 168: 1746.
https://pubmed.ncbi.nlm.nih.gov/12352350

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 113


22. Zavras, N., et al. Conservative treatment of phimosis with fluticasone proprionate 0.05%: a clinical
study in 1185 boys. J Pediatr Urol, 2009. 5: 181.
https://pubmed.ncbi.nlm.nih.gov/19097823
23. Moreno, G., et al. Topical corticosteroids for treating phimosis in boys. Cochrane Database Syst
Rev, 2014: Cd008973.
https://pubmed.ncbi.nlm.nih.gov/25180668
24. Reddy, S., et al. Local steroid therapy as the first-line treatment for boys with symptomatic phimosis
- a long-term prospective study. Acta Paediatr, 2012. 101: e130.
https://pubmed.ncbi.nlm.nih.gov/22103624
25. Golubovic, Z., et al. The conservative treatment of phimosis in boys. Br J Urol, 1996. 78: 786.
https://pubmed.ncbi.nlm.nih.gov/8976781
26. Pileggi, F.O., et al. Is suppression of hypothalamic-pituitary-adrenal axis significant during clinical
treatment of phimosis? J Urol, 2010. 183: 2327.
https://pubmed.ncbi.nlm.nih.gov/20400146
27. Wu, X., et al. A report of 918 cases of circumcision with the Shang Ring: comparison between
children and adults. Urology, 2013. 81: 1058.
https://pubmed.ncbi.nlm.nih.gov/23465168
28. Pedersini, P., et al. “Trident” preputial plasty for phimosis in childhood. J Pediatr Urol, 2017. 13: 278.e1.
https://pubmed.ncbi.nlm.nih.gov/28359779
29. Benson, M., et al. Prepuce sparing: Use of Z-plasty for treatment of phimosis and scarred foreskin.
J Pediatr Urol, 2018. 14: 545.e1.
https://pubmed.ncbi.nlm.nih.gov/29909192
30. Miernik, A., et al. Complete removal of the foreskin--why? Urol Int, 2011. 86: 383.
https://pubmed.ncbi.nlm.nih.gov/21474914
31. Wiswell, T.E. The prepuce, urinary tract infections, and the consequences. Pediatrics, 2000. 105: 860.
https://pubmed.ncbi.nlm.nih.gov/10742334
32. Hiraoka, M., et al. Meatus tightly covered by the prepuce is associated with urinary infection. Pediatr
Int, 2002. 44: 658.
https://pubmed.ncbi.nlm.nih.gov/12421265
33. To, T., et al. Cohort study on circumcision of newborn boys and subsequent risk of urinary-tract
infection. Lancet, 1998. 352: 1813.
https://pubmed.ncbi.nlm.nih.gov/9851381
34. Ellison, J.S., et al. Neonatal Circumcision and Urinary Tract Infections in Infants With
Hydronephrosis. Pediatrics, 2018. 142.
https://pubmed.ncbi.nlm.nih.gov/29880703
35. Ladenhauf, H.N., et al. Reduced bacterial colonisation of the glans penis after male circumcision in
children--a prospective study. J Pediatr Urol, 2013. 9: 1137.
https://pubmed.ncbi.nlm.nih.gov/23685114
36. Larke, N.L., et al. Male circumcision and penile cancer: a systematic review and meta-analysis.
Cancer Causes Control, 2011. 22: 1097.
https://pubmed.ncbi.nlm.nih.gov/21695385
37. Thompson, H.C., et al. Report of the ad hoc task force on circumcision. Pediatrics, 1975. 56: 610.
https://pubmed.ncbi.nlm.nih.gov/1174384
38. American Academy of Pediatrics: Report of the Task Force on Circumcision. Pediatrics, 1989. 84: 388.
https://pubmed.ncbi.nlm.nih.gov/2664697
39. Elalfy, M.S., et al. Risk of bleeding and inhibitor development after circumcision of previously
untreated or minimally treated severe hemophilia A children. Pediatr Hematol Oncol, 2012. 29: 485.
https://pubmed.ncbi.nlm.nih.gov/22866674
40. Karaman, M.I., et al. Circumcision in bleeding disorders: improvement of our cost effective method
with diathermic knife. Urol J, 2014. 11: 1406.
https://pubmed.ncbi.nlm.nih.gov/24807751
41. Christakis, D.A., et al. A trade-off analysis of routine newborn circumcision. Pediatrics, 2000. 105: 246.
https://pubmed.ncbi.nlm.nih.gov/10617731
42. Griffiths, D.M., et al. A prospective survey of the indications and morbidity of circumcision in
children. Eur Urol, 1985. 11: 184.
https://pubmed.ncbi.nlm.nih.gov/4029234
43. Morris, B.J., et al. A ‘snip’ in time: what is the best age to circumcise? BMC Pediatr, 2012. 12: 20.
https://pubmed.ncbi.nlm.nih.gov/22373281
44. Ross, J.H., Circumcision: Pro and con., in Pediatric urology for the general urologist. , J.S. Elder,
Editor. 1996, Igaku-Shoin: New York.

114 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


45. Weiss, H.A., et al. Complications of circumcision in male neonates, infants and children: a
systematic review. BMC Urol, 2010. 10: 2.
https://pubmed.ncbi.nlm.nih.gov/20158883
46. Homer, L., et al. Meatal stenosis in boys following circumcision for lichen sclerosus (balanitis
xerotica obliterans). J Urol, 2014. 192: 1784.
https://pubmed.ncbi.nlm.nih.gov/24992332
47. Anand, A., et al. Mannitol for paraphimosis reduction. Urol Int, 2013. 90: 106.
https://pubmed.ncbi.nlm.nih.gov/23257575
48. DeVries, C.R., et al. Reduction of paraphimosis with hyaluronidase. Urology, 1996. 48: 464.
https://pubmed.ncbi.nlm.nih.gov/8804504
49. Hung, Y.C., et al. A Longitudinal Population Analysis of Cumulative Risks of Circumcision. J Surg
Res, 2019. 233: 111.
https://pubmed.ncbi.nlm.nih.gov/30502236
50. Sijstermans, K., et al. The frequency of undescended testis from birth to adulthood: a review. Int
J Androl, 2008. 31: 1.
https://pubmed.ncbi.nlm.nih.gov/17488243
51. Berkowitz, G.S., et al. Prevalence and natural history of cryptorchidism. Pediatrics, 1993. 92: 44.
https://pubmed.ncbi.nlm.nih.gov/8100060
52. Kaefer, M., et al. The incidence of intersexuality in children with cryptorchidism and hypospadias:
stratification based on gonadal palpability and meatal position. J Urol, 1999. 162: 1003.
https://pubmed.ncbi.nlm.nih.gov/10458421
53. Kollin, C., et al. Cryptorchidism: a clinical perspective. Pediatr Endocrinol Rev, 2014. 11 Suppl 2: 240.
https://pubmed.ncbi.nlm.nih.gov/24683948
54. Caesar, R.E., et al. The incidence of the cremasteric reflex in normal boys. J Urol, 1994. 152: 779.
https://pubmed.ncbi.nlm.nih.gov/7912745
55. Barthold, J.S., et al. The epidemiology of congenital cryptorchidism, testicular ascent and
orchiopexy. J Urol, 2003. 170: 2396.
https://pubmed.ncbi.nlm.nih.gov/14634436
56. Turek, P.J., et al. The absent cryptorchid testis: surgical findings and their implications for diagnosis
and etiology. J Urol, 1994. 151: 718.
https://pubmed.ncbi.nlm.nih.gov/7905931
57. Rabinowitz, R., et al. Late presentation of cryptorchidism: the etiology of testicular re-ascent. J Urol,
1997. 157: 1892.
https://pubmed.ncbi.nlm.nih.gov/9112557
58. Cendron, M., et al. Anatomical, morphological and volumetric analysis: a review of 759 cases of
testicular maldescent. J Urol, 1993. 149: 570.
https://pubmed.ncbi.nlm.nih.gov/8094761
59. Braga, L.H., et al. Is there an optimal contralateral testicular cut-off size that predicts monorchism in
boys with nonpalpable testicles? J Pediatr Urol, 2014. 10: 693.
https://pubmed.ncbi.nlm.nih.gov/25008806
60. Hurwitz, R.S., et al. How well does contralateral testis hypertrophy predict the absence of the
nonpalpable testis? J Urol, 2001. 165: 588.
https://pubmed.ncbi.nlm.nih.gov/11176443
61. Hodhod, A., et al. Testicular hypertrophy as a predictor for contralateral monorchism: Retrospective
review of prospectively recorded data. J Pediatr Urol, 2016. 12: 34.e1.
https://pubmed.ncbi.nlm.nih.gov/26279100
62. Elert, A., et al. Population-based investigation of familial undescended testis and its association with
other urogenital anomalies. J Pediatr Urol, 2005. 1: 403.
https://pubmed.ncbi.nlm.nih.gov/18947580
63. Hrebinko, R.L., et al. The limited role of imaging techniques in managing children with undescended
testes. J Urol, 1993. 150: 458.
https://pubmed.ncbi.nlm.nih.gov/8100860
64. Tasian, G.E., et al. Diagnostic performance of ultrasound in nonpalpable cryptorchidism: a
systematic review and meta-analysis. Pediatrics, 2011. 127: 119.
https://pubmed.ncbi.nlm.nih.gov/21149435
65. Elder, J.S. Ultrasonography is unnecessary in evaluating boys with a nonpalpable testis. Pediatrics,
2002. 110: 748.
https://pubmed.ncbi.nlm.nih.gov/12359789

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 115


66. Wenzler, D.L., et al. What is the rate of spontaneous testicular descent in infants with
cryptorchidism? J Urol, 2004. 171: 849.
https://pubmed.ncbi.nlm.nih.gov/14713841
67. Park, K.H., et al. Histological evidences suggest recommending orchiopexy within the first year of
life for children with unilateral inguinal cryptorchid testis. Int J Urol, 2007. 14: 616.
https://pubmed.ncbi.nlm.nih.gov/17645605
68. Engeler, D.S., et al. Early orchiopexy: prepubertal intratubular germ cell neoplasia and fertility
outcome. Urology, 2000. 56: 144.
https://pubmed.ncbi.nlm.nih.gov/10869645
69. Forest, M.G., et al. Undescended testis: comparison of two protocols of treatment with human
chorionic gonadotropin. Effect on testicular descent and hormonal response. Horm Res, 1988. 30: 198.
https://pubmed.ncbi.nlm.nih.gov/2907898
70. Rajfer, J., et al. Hormonal therapy of cryptorchidism. A randomized, double-blind study comparing
human chorionic gonadotropin and gonadotropin-releasing hormone. N Engl J Med, 1986. 314: 466.
https://pubmed.ncbi.nlm.nih.gov/2868413
71. Pyorala, S., et al. A review and meta-analysis of hormonal treatment of cryptorchidism. J Clin
Endocrinol Metab, 1995. 80: 2795.
https://pubmed.ncbi.nlm.nih.gov/7673426
72. Rajfer, J., et al. The incidence of intersexuality in patients with hypospadias and cryptorchidism.
J Urol, 1976. 116: 769.
https://pubmed.ncbi.nlm.nih.gov/12377
73. Lala, R., et al. Combined therapy with LHRH and HCG in cryptorchid infants. Eur J Pediatr, 1993.
152 Suppl 2: S31.
https://pubmed.ncbi.nlm.nih.gov/8101810
74. Forest, M.G., et al. Effects of human chorionic gonadotropin, androgens, adrenocorticotropin
hormone, dexamethasone and hyperprolactinemia on plasma sex steroid-binding protein. Ann N Y
Acad Sci, 1988. 538: 214.
https://pubmed.ncbi.nlm.nih.gov/2847619
75. Aycan, Z., et al. Evaluation of low-dose hCG treatment for cryptorchidism. Turk J Pediatr, 2006.
48: 228.
https://pubmed.ncbi.nlm.nih.gov/17172066
76. Hesse, V., et al. Three injections of human chorionic gonadotropin are as effective as ten injections
in the treatment of cryptorchidism. Horm Res, 1988. 30: 193.
https://pubmed.ncbi.nlm.nih.gov/2907897
77. Hagberg, S., et al. Treatment of undescended testes with intranasal application of synthetic LH-RH.
Eur J Pediatr, 1982. 139: 285.
https://pubmed.ncbi.nlm.nih.gov/6133757
78. Hadziselimovic, F., et al. Treatment with a luteinizing hormone-releasing hormone analogue after
successful orchiopexy markedly improves the chance of fertility later in life. J Urol, 1997. 158: 1193.
https://pubmed.ncbi.nlm.nih.gov/9258170
79. Kollin, C., et al. Surgical treatment of unilaterally undescended testes: testicular growth after
randomization to orchiopexy at age 9 months or 3 years. J Urol, 2007. 178: 1589.
https://pubmed.ncbi.nlm.nih.gov/17707045
80. Cortes, D., et al. Hormonal treatment may harm the germ cells in 1 to 3-year-old boys with
cryptorchidism. J Urol, 2000. 163: 1290.
https://pubmed.ncbi.nlm.nih.gov/10737531
81. Ritzen, E.M. Undescended testes: a consensus on management. Eur J Endocrinol, 2008. 159 Suppl
1: S87.
https://pubmed.ncbi.nlm.nih.gov/18728121
82. Hildorf, S., et al. Fertility Potential is Compromised in 20% to 25% of Boys with Nonsyndromic
Cryptorchidism despite Orchiopexy within the First Year of Life. J Urol, 2020. 203: 832.
https://pubmed.ncbi.nlm.nih.gov/31642739
83. Novaes, H.F., et al. Single scrotal incision orchiopexy - a systematic review. Int Braz J Urol, 2013.
39: 305.
https://pubmed.ncbi.nlm.nih.gov/23849581
84. Docimo, S.G. The results of surgical therapy for cryptorchidism: a literature review and analysis.
J Urol, 1995. 154: 1148.
https://pubmed.ncbi.nlm.nih.gov/7637073
85. Ziylan, O., et al. Failed orchiopexy. Urol Int, 2004. 73: 313.
https://pubmed.ncbi.nlm.nih.gov/15604574

116 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


86. Prentiss, R.J., et al. Undescended testis: surgical anatomy of spermatic vessels, spermatic surgical
triangles and lateral spermatic ligament. J Urol, 1960. 83: 686.
https://pubmed.ncbi.nlm.nih.gov/14434738
87. Kozminski, D.J., et al. Orchiopexy without Transparenchymal Fixation Suturing: A 29-Year
Experience. J Urol, 2015. 194: 1743.
https://pubmed.ncbi.nlm.nih.gov/26141850
88. Martin, J.M., et al. Is radiotherapy a good adjuvant strategy for men with a history of cryptorchism
and stage I seminoma? Int J Radiat Oncol Biol Phys, 2010. 76: 65.
https://pubmed.ncbi.nlm.nih.gov/19362785
89. Na, S.W., et al. Single scrotal incision orchiopexy for children with palpable low-lying undescended
testis: early outcome of a prospective randomized controlled study. Korean J Urol, 2011. 52: 637.
https://pubmed.ncbi.nlm.nih.gov/22025961
90. Parsons, J.K., et al. The low scrotal approach to the ectopic or ascended testicle: prevalence of a
patent processus vaginalis. J Urol, 2003. 169: 1832.
https://pubmed.ncbi.nlm.nih.gov/12686856
91. Feng, S., et al. Single scrotal incision orchiopexy versus the inguinal approach in children with
palpable undescended testis: a systematic review and meta-analysis. Pediatr Surg Int, 2016. 32: 989.
https://pubmed.ncbi.nlm.nih.gov/27510940
92. Wayne, C., et al. What is the ideal surgical approach for intra-abdominal testes? A systematic
review. Pediatr Surg Int, 2015. 31: 327.
https://pubmed.ncbi.nlm.nih.gov/25663531
93. Cortesi, N., et al. Diagnosis of bilateral abdominal cryptorchidism by laparoscopy. Endoscopy,
1976. 8: 33.
https://pubmed.ncbi.nlm.nih.gov/16743
94. Jordan, G.H., et al. Laparoscopic single stage and staged orchiopexy. J Urol, 1994. 152: 1249.
https://pubmed.ncbi.nlm.nih.gov/7915336
95. Chandrasekharam, V.V. Laparoscopy vs inguinal exploration for nonpalpable undescended testis.
Indian J Pediatr, 2005. 72: 1021.
https://pubmed.ncbi.nlm.nih.gov/16388149
96. Snodgrass, W.T., et al. Scrotal exploration for unilateral nonpalpable testis. J Urol, 2007. 178: 1718.
https://pubmed.ncbi.nlm.nih.gov/17707015
97. Cisek, L.J., et al. Current findings in diagnostic laparoscopic evaluation of the nonpalpable testis.
J Urol, 1998. 160: 1145.
https://pubmed.ncbi.nlm.nih.gov/9719296
98. Patil, K.K., et al. Laparoscopy for impalpable testes. BJU Int, 2005. 95: 704.
https://pubmed.ncbi.nlm.nih.gov/15784081
99. Elderwy, A.A., et al. Laparoscopic versus open orchiopexy in the management of peeping testis: a
multi-institutional prospective randomized study. J Pediatr Urol, 2014. 10: 605.
https://pubmed.ncbi.nlm.nih.gov/25042877
100. Kirsch, A.J., et al. Surgical management of the nonpalpable testis: the Children’s Hospital of
Philadelphia experience. J Urol, 1998. 159: 1340.
https://pubmed.ncbi.nlm.nih.gov/9507881
101. Fowler, R., et al. The role of testicular vascular anatomy in the salvage of high undescended testes.
Aust N Z J Surg, 1959. 29: 92.
https://pubmed.ncbi.nlm.nih.gov/13849840
102. Koff, S.A., et al. Treatment of high undescended testes by low spermatic vessel ligation: an
alternative to the Fowler-Stephens technique. J Urol, 1996. 156: 799.
https://pubmed.ncbi.nlm.nih.gov/8683787
103. Esposito, C., et al. Exploration of inguinal canal is mandatory in cases of non palpable testis if
laparoscopy shows elements entering a closed inguinal ring. Eur J Pediatr Surg, 2010. 20: 138.
https://pubmed.ncbi.nlm.nih.gov/19746341
104. Wu, C.Q., et al. Revisiting the success rate of one-stage Fowler-Stephens orchiopexy with
postoperative Doppler ultrasound and long-term follow-up: a 15-year single-surgeon experience.
J Pediatr Urol, 2020. 16: 48.
https://pubmed.ncbi.nlm.nih.gov/31784377
105. Radmayr, C., et al. Long-term outcome of laparoscopically managed nonpalpable testes. J Urol,
2003. 170: 2409.
https://pubmed.ncbi.nlm.nih.gov/14634439
106. Baker, L.A., et al. A multi-institutional analysis of laparoscopic orchidopexy. BJU Int, 2001. 87: 484.
https://pubmed.ncbi.nlm.nih.gov/11298039

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 117


107. Dave, S., et al. Open versus laparoscopic staged Fowler-Stephens orchiopexy: impact of long loop
vas. J Urol, 2009. 182: 2435.
https://pubmed.ncbi.nlm.nih.gov/19765743
108. Wacksman, J., et al. Laparoscopically assisted testicular autotransplantation for management of the
intraabdominal undescended testis. J Urol, 1996. 156: 772.
https://pubmed.ncbi.nlm.nih.gov/8683780
109. Penson, D., et al. Effectiveness of hormonal and surgical therapies for cryptorchidism: a systematic
review. Pediatrics, 2013. 131: e1897.
https://pubmed.ncbi.nlm.nih.gov/23690511
110. Koni, A., et al. Histopathological evaluation of orchiectomy specimens in 51 late postpubertal men
with unilateral cryptorchidism. J Urol, 2014. 192: 1183.
https://pubmed.ncbi.nlm.nih.gov/24840535
111. Trussell, J.C., et al. The relationship of cryptorchidism to fertility. Curr Urol Rep, 2004. 5: 142.
https://pubmed.ncbi.nlm.nih.gov/15028208
112. Hadziselimovic, F., et al. The importance of both an early orchidopexy and germ cell maturation for
fertility. Lancet, 2001. 358: 1156.
https://pubmed.ncbi.nlm.nih.gov/11597673
113. Lee, P.A. Fertility after cryptorchidism: epidemiology and other outcome studies. Urology, 2005.
66: 427.
https://pubmed.ncbi.nlm.nih.gov/16098371
114. Chua, M.E., et al. Hormonal therapy using gonadotropin releasing hormone for improvement of
fertility index among children with cryptorchidism: a meta-analysis and systematic review. J Pediatr
Surg, 2014. 49: 1659.
https://pubmed.ncbi.nlm.nih.gov/25475814
115. Coughlin, M.T., et al. Age at unilateral orchiopexy: effect on hormone levels and sperm count in
adulthood. J Urol, 1999. 162: 986.
https://pubmed.ncbi.nlm.nih.gov/10458417
116. Tasian, G.E., et al. Age at orchiopexy and testis palpability predict germ and Leydig cell loss: clinical
predictors of adverse histological features of cryptorchidism. J Urol, 2009. 182: 704.
https://pubmed.ncbi.nlm.nih.gov/19539332
117. Dieckmann, K.P., et al. Clinical epidemiology of testicular germ cell tumors. World J Urol, 2004. 22: 2.
https://pubmed.ncbi.nlm.nih.gov/15034740
118. Pettersson, A., et al. Age at surgery for undescended testis and risk of testicular cancer. N Engl
J Med, 2007. 356: 1835.
https://pubmed.ncbi.nlm.nih.gov/17476009
119. Walsh, T.J., et al. Prepubertal orchiopexy for cryptorchidism may be associated with lower risk of
testicular cancer. J Urol, 2007. 178: 1440.
https://pubmed.ncbi.nlm.nih.gov/17706709
120. Pohl, H.G., et al. Prepubertal testis tumors: actual prevalence rate of histological types. J Urol, 2004.
172: 2370.
https://pubmed.ncbi.nlm.nih.gov/15538270
121. Kusler, K.A., et al. International testicular cancer incidence rates in children, adolescents and young
adults. Cancer Epidemiol, 2018. 56: 106.
https://pubmed.ncbi.nlm.nih.gov/30130682
122. Schneider, D.T., et al. Epidemiologic analysis of 1,442 children and adolescents registered in the
German germ cell tumor protocols. Pediatr Blood Cancer, 2004. 42: 169.
https://pubmed.ncbi.nlm.nih.gov/14752882
123. Taskinen, S., et al. Testicular tumors in children and adolescents. J Pediatric Urol, 2008. 4: 134.
https://pubmed.ncbi.nlm.nih.gov/18631909
124. Metcalfe, P.D., et al. Pediatric testicular tumors: contemporary incidence and efficacy of testicular
preserving surgery. J Urol, 2003. 170: 2412.
https://pubmed.ncbi.nlm.nih.gov/14634440
125. Shukla, A.R., et al. Experience with testis sparing surgery for testicular teratoma. J Urol, 2004.
171: 161.
https://pubmed.ncbi.nlm.nih.gov/14665867
126. Nerli, R.B., et al. Prepubertal testicular tumors: Our 10 years experience. Indian J Cancer, 2010. 47: 292.
https://pubmed.ncbi.nlm.nih.gov/20587905
127. Wu, D., et al. Prepubertal testicular tumors in China: a 10-year experience with 67 cases. Pediatr
Surg Int, 2018. 34: 1339.
https://pubmed.ncbi.nlm.nih.gov/30324570

118 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


128. Hawkins, E., et al. The prepubertal testis (prenatal and postnatal): its relationship to intratubular
germ cell neoplasia: a combined Pediatric Oncology Group and Children’s Cancer Study Group.
Hum Pathol, 1997. 28: 404.
https://pubmed.ncbi.nlm.nih.gov/9104938
129. Manivel, J.C., et al. Intratubular germ cell neoplasia in testicular teratomas and epidermoid cysts.
Correlation with prognosis and possible biologic significance. Cancer, 1989. 64: 715.
https://pubmed.ncbi.nlm.nih.gov/2663131
130. Renedo, D.E., et al. Intratubular germ cell neoplasia (ITGCN) with p53 and PCNA expression and
adjacent mature teratoma in an infant testis. An immunohistochemical and morphologic study with a
review of the literature. Am J Surg Pathol, 1994. 18: 947.
https://pubmed.ncbi.nlm.nih.gov/7741838
131. Rushton, H.G., et al. Testicular sparing surgery for prepubertal teratoma of the testis: a clinical and
pathological study. J Urol, 1990. 144: 726.
https://pubmed.ncbi.nlm.nih.gov/2388338
132. Roth, L.M., et al. Gonadoblastoma: origin and outcome. Hum Pathol, 2019.
https://pubmed.ncbi.nlm.nih.gov/31805291
133. Ahmed, H.U., et al. Testicular and paratesticular tumours in the prepubertal population. Lancet
Oncol, 2010. 11: 476.
https://pubmed.ncbi.nlm.nih.gov/20434716
134. Henderson, C.G., et al. Enucleation for prepubertal leydig cell tumor. J Urol, 2006. 176: 703.
https://pubmed.ncbi.nlm.nih.gov/16813923
135. Soles, B.S., et al. Melanotic Neuroectodermal Tumor of Infancy. Arch Pathol Lab Med, 2018. 142: 1358.
https://pubmed.ncbi.nlm.nih.gov/30407852
136. Yada, K., et al. Intrascrotal lipoblastoma: report of a case and the review of literature. Surg Case
Rep, 2016. 2: 34.
https://pubmed.ncbi.nlm.nih.gov/27059472
137. Marulaiah, M., et al. Testicular and Paratesticular Pathology in Children: a 12-Year Histopathological
Review. World J Surg, 2010: 1.
https://pubmed.ncbi.nlm.nih.gov/20151127
138. Walterhouse, D.O., et al. Demographic and Treatment Variables Influencing Outcome for Localized
Paratesticular Rhabdomyosarcoma: Results From a Pooled Analysis of North American and
European Cooperative Groups. J Clin Oncol, 2018: JCO2018789388.
https://pubmed.ncbi.nlm.nih.gov/30351998
139. Akbar, S.A., et al. Multimodality imaging of paratesticular neoplasms and their rare mimics.
Radiographics, 2003. 23: 1461.
https://pubmed.ncbi.nlm.nih.gov/14615558
140. Esen, B., et al. Should we rely on Doppler ultrasound for evaluation of testicular solid lesions? World
J Urol, 2018. 36: 1263.
https://pubmed.ncbi.nlm.nih.gov/29572727
141. Lock, G. [Contrast-enhanced ultrasonography of testicular tumours]. Urologe A, 2019. 58: 1410.
https://pubmed.ncbi.nlm.nih.gov/31712858
142. Tallen, G., et al. High reliability of scrotal ultrasonography in the management of childhood primary
testicular neoplasms. Klin Padiatr, 2011. 223: 131.
https://pubmed.ncbi.nlm.nih.gov/21462100
143. Yu, C.J., et al. Incidence characteristics of testicular microlithiasis and its association with risk of
primary testicular tumors in children: a systematic review and meta-analysis. World J Pediatr, 2019.
https://pubmed.ncbi.nlm.nih.gov/31853884
144. Ludwikowski, B., et al. (2016) S2kHodenhochstand – Maldescensus testis.
https://www.awmf.org/uploads/tx_szleitlinien/006-022l_S2k_Hodenhochstand_Maldescensus-
testis_2018-08-verlaengert..pdf
145. Barbonetti, A., et al. Testicular Cancer in Infertile Men With and Without Testicular Microlithiasis: A
Systematic Review and Meta-Analysis of Case-Control Studies. Front Endocrinol (Lausanne), 2019.
10: 164.
https://pubmed.ncbi.nlm.nih.gov/30949131
146. Schneider, D.T., et al. Diagnostic value of alpha 1-fetoprotein and beta-human chorionic
gonadotropin in infancy and childhood. Pediatr Hematol Oncol, 2001. 18: 11.
https://pubmed.ncbi.nlm.nih.gov/11205836
147. Ross, J.H., et al. Clinical behavior and a contemporary management algorithm for prepubertal testis
tumors: a summary of the Prepubertal Testis Tumor Registry. J Urol, 2002. 168: 1675.
https://pubmed.ncbi.nlm.nih.gov/12352332

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 119


148. Radford, A., et al. Testicular-sparing surgery in the pediatric population: multicenter review of
practice with review of the literature. Curr Opin Urol, 2019. 29: 481.
https://pubmed.ncbi.nlm.nih.gov/31205272
149. Friend, J., et al. Benign scrotal masses in children - some new lessons learned. J Pediatric Surg,
2016. 51: 1737.
https://pubmed.ncbi.nlm.nih.gov/27558482
150. Hisamatsu, E., et al. Prepubertal testicular tumors: A 20-year experience with 40 cases. Int J Urol,
2010. 17: 956.
https://pubmed.ncbi.nlm.nih.gov/21046693
151. Fankhauser, C.D., et al. Risk Factors and Treatment Outcomes of 1,375 Patients with Testicular
Leydig Cell Tumors: Analysis of Published Case Series Data. J Urol, 2020. 203: 949.
https://pubmed.ncbi.nlm.nih.gov/31845841
152. Grogg, J., et al. Sertoli Cell Tumors of the Testes: Systematic Literature Review and Meta-Analysis
of Outcomes in 435 Patients. Oncologist, 2020.
https://pubmed.ncbi.nlm.nih.gov/32043680
153. Albers, P., et al. Guidelines on Testicular Cancer: 2015 Update. Eur Urol, 2015. 68: 1054.
https://pubmed.ncbi.nlm.nih.gov/26297604
154. Little, T., et al. Paediatric testicular tumours in a New Zealand centre. New Zealand Med J, 2017.
130: 68.
https://pubmed.ncbi.nlm.nih.gov/29240742
155. Williamson, S.R., et al. The World Health Organization 2016 classification of testicular germ
cell tumours: a review and update from the International Society of Urological Pathology Testis
Consultation Panel. Histopathology, 2017. 70: 335.
https://pubmed.ncbi.nlm.nih.gov/27747907
156. Hasegawa, T., et al. A case of immature teratoma originating in intra-abdominal undescended testis
in a 3-month-old infant. Pediatr Surg Int, 2006. 22: 570.
https://pubmed.ncbi.nlm.nih.gov/16736229
157. Chang, M.Y., et al. Prepubertal Testicular Teratomas and Epidermoid Cysts: Comparison of Clinical
and Sonographic Features. J Ultrasound Med, 2015. 34: 1745.
https://pubmed.ncbi.nlm.nih.gov/26324756
158. Kao, C.S., et al. Juvenile granulosa cell tumors of the testis: a clinicopathologic study of 70 cases
with emphasis on its wide morphologic spectrum. Am J Surg Pathol, 2015. 39: 1159.
https://pubmed.ncbi.nlm.nih.gov/26076062
159. Shukla, A.R., et al. Juvenile granulosa cell tumor of the testis:: contemporary clinical management
and pathological diagnosis. J Urol, 2004. 171: 1900.
https://pubmed.ncbi.nlm.nih.gov/15076304
160. Luckie, T.M., et al. A Multicenter Retrospective Review of Pediatric Leydig Cell Tumor of the Testis.
J Pediatr Hematol/Oncol, 2019. 41: 74.
https://pubmed.ncbi.nlm.nih.gov/29554024
161. Emre, S., et al. Testis sparing surgery for Leydig cell pathologies in children. J Pediatr Urol, 2017. 13: 51.
https://pubmed.ncbi.nlm.nih.gov/27773621
162. Geminiani, J.J., et al. Testicular Leydig Cell Tumor with Metachronous Lesions: Outcomes after
Metastasis Resection and Cryoablation. Case Rep Urol, 2015. 2015: 748495.
https://pubmed.ncbi.nlm.nih.gov/26525589
163. Talon, I., et al. Sertoli cell tumor of the testis in children: reevaluation of a rarely encountered tumor.
J Pediatr Hematol Oncol, 2005. 27: 491.
https://pubmed.ncbi.nlm.nih.gov/16189443
164. Li, G., et al. Prepubertal Malignant Large Cell Calcifying Sertoli Cell Tumor of the Testis. Urology,
2018. 117: 145.
https://pubmed.ncbi.nlm.nih.gov/29626571
165. Borer, J.G., et al. The spectrum of Sertoli cell tumors in children. Urol Clin North Am, 2000. 27: 529.
https://pubmed.ncbi.nlm.nih.gov/10985152
166. Wilson, D.M., et al. Testicular tumors with Peutz-Jeghers syndrome. Cancer, 1986. 57: 2238.
https://pubmed.ncbi.nlm.nih.gov/3697923
167. Alleemudder, A., et al. A case of Carney complex presenting as acute testicular pain. Urol Ann,
2016. 8: 360.
https://pubmed.ncbi.nlm.nih.gov/27453662
168. Coppes, M.J., et al. Primary testicular and paratesticular tumors of childhood. Med Pediatr Oncol,
1994. 22: 329.
https://pubmed.ncbi.nlm.nih.gov/8127257

120 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


169. Ye, Y.L., et al. Relapse in children with clinical stage I testicular yolk sac tumors after initial
orchiectomy. Pediatr Surg Int, 2018. 35: 383.
https://pubmed.ncbi.nlm.nih.gov/30539226
170. Grady, R.W. Current management of prepubertal yolk sac tumors of the testis. Urol Clin North Am,
2000. 27: 503.
https://pubmed.ncbi.nlm.nih.gov/10985149
171. Haas, R.J., et al. Testicular germ cell tumors, an update. Results of the German cooperative studies
1982-1997. Klin Padiatr, 1999. 211: 300.
https://pubmed.ncbi.nlm.nih.gov/10472566
172. Rogers, P.C., et al. Treatment of children and adolescents with stage II testicular and stages I and II
ovarian malignant germ cell tumors: A Pediatric Intergroup Study--Pediatric Oncology Group 9048
and Children’s Cancer Group 8891. J Clin Oncol, 2004. 22: 3563.
https://pubmed.ncbi.nlm.nih.gov/15337806
173. Schlatter, M., et al. Excellent outcome in patients with stage I germ cell tumors of the testes: a study
of the Children’s Cancer Group/Pediatric Oncology Group. J Pediatr Surg, 2003. 38: 319.
https://pubmed.ncbi.nlm.nih.gov/12632342
174. Alanee, S.R., et al. Pelvic Lymph Node Dissection in Patients Treated for Testis Cancer: The
Memorial Sloan Kettering Cancer Center Experience. Urology, 2016. 95: 128.
https://pubmed.ncbi.nlm.nih.gov/27235751
175. Claahsen-Van der Grinten, H.L., et al. Increased prevalence of testicular adrenal rest tumours during
adolescence in congenital adrenal hyperplasia. Hormone Res Paediatr, 2014. 82: 238.
https://pubmed.ncbi.nlm.nih.gov/25195868
176. Merke, D.P., et al. Management of adolescents with congenital adrenal hyperplasia. Lancet Diabetes
Endocrinol, 2013. 1: 341.
https://pubmed.ncbi.nlm.nih.gov/24622419
177. Claahsen-van der Grinten, H.L., et al. Testicular adrenal rest tumours in congenital adrenal
hyperplasia. Int J Pediatr Endocrinol, 2009. 2009: 624823.
https://pubmed.ncbi.nlm.nih.gov/19956703
178. Chaudhari, M., et al. Testicular adrenal rest tumor screening and fertility counseling among males
with congenital adrenal hyperplasia. J Pediatr Surg, 2018. 14: 155.
https://pubmed.ncbi.nlm.nih.gov/29330018
179. Kapur, P., et al. Pediatric hernias and hydroceles. Pediatr Clin North Am, 1998. 45: 773.
https://pubmed.ncbi.nlm.nih.gov/9728185
180. Barthold, J.S., Abnormalities of the testis and scrotum and their surgical management, in Campbell-
Walsh Urology, A.J. Wein & et al., Editors. 2012, Elsevier Saunders: Philadelphia.
181. Schneck, F.X., et al., Abnormalities of the testes and scrotum and their surgical management in
Campbell’s Urology, P.C. Walsh, A.B. Retik, E.D. Vaughan & A.J. Wein, Editors. 2002, WB Saunders:
Philadelphia.
182. Rubenstein, R.A., et al. Benign intrascrotal lesions. J Urol, 2004. 171: 1765.
https://pubmed.ncbi.nlm.nih.gov/15076274
183. Lin, H.C., et al. Testicular teratoma presenting as a transilluminating scrotal mass. Urology, 2006.
67: 1290.e3.
https://pubmed.ncbi.nlm.nih.gov/16750249
184. Skoog, S.J. Benign and malignant pediatric scrotal masses. Pediatr Clin North Am, 1997. 44: 1229.
https://pubmed.ncbi.nlm.nih.gov/9326960
185. Koski, M.E., et al. Infant communicating hydroceles--do they need immediate repair or might some
clinically resolve? J Pediatr Surg, 2010. 45: 590.
https://pubmed.ncbi.nlm.nih.gov/20223325
186. Stringer, M.D., et al., Patent processus vaginalis. , in Pediatric urology, J.P. Gearhart, R.C. Rink &
P.D. Mouriquand, Editors. 2001, WB Saunders: Philadelphia.
187. Stylianos, S., et al. Incarceration of inguinal hernia in infants prior to elective repair. J Pediatr Surg,
1993. 28: 582.
https://pubmed.ncbi.nlm.nih.gov/8483072
188. Hall, N.J., et al. Surgery for hydrocele in children-an avoidable excess? J Pediatr Surg, 2011. 46: 2401.
https://pubmed.ncbi.nlm.nih.gov/22152892
189. Saad, S., et al. Ten-year review of groin laparoscopy in 1001 pediatric patients with clinical unilateral
inguinal hernia: an improved technique with transhernia multiple-channel scope. J Pediatr Surg,
2011. 46: 1011.
https://pubmed.ncbi.nlm.nih.gov/21616272

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 121


190. Christensen, T., et al. New onset of hydroceles in boys over 1 year of age. Int J Urol, 2006. 13: 1425.
https://pubmed.ncbi.nlm.nih.gov/17083397
191. Alp, B.F., et al. Comparison of the inguinal and scrotal approaches for the treatment of
communicating hydrocele in children. Kaohsiung J Med Sci, 2014. 30: 200.
https://pubmed.ncbi.nlm.nih.gov/24656161
192. Oh, J.H., et al. Hydrocelectomy via scrotal incision is a valuable alternative to the traditional inguinal
approach for hydrocele treatment in boys. Investig Clin Urol, 2018. 59: 416.
https://pubmed.ncbi.nlm.nih.gov/30402575
193. Grimsby, G.M., et al. Non-absorbable sutures are associated with lower recurrence rates in
laparoscopic percutaneous inguinal hernia ligation. J Pediatr Urol, 2015. 11: 275.e1.
https://pubmed.ncbi.nlm.nih.gov/26233553
194. Saka, R., et al. Safety and efficacy of laparoscopic percutaneous extraperitoneal closure for inguinal
hernias and hydroceles in children: a comparison with traditional open repair. J Laparoendosc Adv
Surg Tech A, 2014. 24: 55.
https://pubmed.ncbi.nlm.nih.gov/24180356
195. Cavusoglu, Y.H., et al. Acute scrotum -- etiology and management. Indian J Pediatr, 2005. 72: 201.
https://pubmed.ncbi.nlm.nih.gov/15812112
196. Klin, B., et al. Epididymitis in childhood: a clinical retrospective study over 5 years. Isr Med Assoc
J, 2001. 3: 833.
https://pubmed.ncbi.nlm.nih.gov/11729579
197. Makela, E., et al. A 19-year review of paediatric patients with acute scrotum. Scand J Surg, 2007.
96: 62.
https://pubmed.ncbi.nlm.nih.gov/17461315
198. McAndrew, H.F., et al. The incidence and investigation of acute scrotal problems in children. Pediatr
Surg Int, 2002. 18: 435.
https://pubmed.ncbi.nlm.nih.gov/12415374
199. Sakellaris, G.S., et al. Acute epididymitis in Greek children: a 3-year retrospective study. Eur
J Pediatr, 2008. 167: 765.
https://pubmed.ncbi.nlm.nih.gov/17786475
200. Varga, J., et al. Acute scrotal pain in children--ten years’ experience. Urol Int, 2007. 78: 73.
https://pubmed.ncbi.nlm.nih.gov/17192737
201. Bingol-Kologlu, M., et al. An exceptional complication following appendectomy: acute inguinal and
scrotal suppuration. Int Urol Nephrol, 2006. 38: 663.
https://pubmed.ncbi.nlm.nih.gov/17160451
202. Dayanir, Y.O., et al. Epididymoorchitis mimicking testicular torsion in Henoch-Schonlein purpura. Eur
Radiol, 2001. 11: 2267.
https://pubmed.ncbi.nlm.nih.gov/11702171
203. Diamond, D.A., et al. Neonatal scrotal haematoma: mimicker of neonatal testicular torsion. BJU Int,
2003. 91: 675.
https://pubmed.ncbi.nlm.nih.gov/12699483
204. Ha, T.S., et al. Scrotal involvement in childhood Henoch-Schonlein purpura. Acta Paediatr, 2007.
96: 552.
https://pubmed.ncbi.nlm.nih.gov/17306010
205. Hara, Y., et al. Acute scrotum caused by Henoch-Schonlein purpura. Int J Urol, 2004. 11: 578.
https://pubmed.ncbi.nlm.nih.gov/15242376
206. Klin, B., et al. Acute idiopathic scrotal edema in children--revisited. J Pediatr Surg, 2002. 37: 1200.
https://pubmed.ncbi.nlm.nih.gov/12149702
207. Krause, W. Is acute idiopathic scrotal edema in children a special feature of neutrophilic eccrine
hidradenitis? Dermatology, 2004. 208: 86; author reply 86.
https://pubmed.ncbi.nlm.nih.gov/14730248
208. Matsumoto, A., et al. Torsion of the hernia sac within a hydrocele of the scrotum in a child. Int
J Urol, 2004. 11: 789.
https://pubmed.ncbi.nlm.nih.gov/15379947
209. Myers, J.B., et al. Torsion of an indirect hernia sac causing acute scrotum. J Pediatr Surg, 2004.
39: 122.
https://pubmed.ncbi.nlm.nih.gov/14694389
210. Ng, K.H., et al. An unusual presentation of acute scrotum after appendicitis. Singapore Med J, 2002.
43: 365.
https://pubmed.ncbi.nlm.nih.gov/12437045

122 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


211. Singh, S., et al. Acute scrotum in children: a rare presentation of acute, non-perforated appendicitis.
Pediatr Surg Int, 2003. 19: 298.
https://pubmed.ncbi.nlm.nih.gov/12682749
212. van Langen, A.M., et al. Acute idiopathic scrotal oedema: four cases and a short review. Eur
J Pediatr, 2001. 160: 455.
https://pubmed.ncbi.nlm.nih.gov/11475590
213. Vlazakis, S., et al. Right acute hemiscrotum caused by insertion of an inflamed appendix. BJU Int,
2002. 89: 967.
https://pubmed.ncbi.nlm.nih.gov/12010250
214. D’Andrea, A., et al. US in the assessment of acute scrotum. Crit Ultrasound J, 2013. 5: S8.
https://pubmed.ncbi.nlm.nih.gov/23902859
215. Davis, J.E., et al. Scrotal emergencies. Emerg Med Clin North Am, 2011. 29: 469.
https://pubmed.ncbi.nlm.nih.gov/21782069
216. Jimoh, B.M., et al. Idiopathic scrotal hematoma in neonate: a case report and review of the
literature. Case Rep Urol, 2014. 2014: 212914.
https://pubmed.ncbi.nlm.nih.gov/24982811
217. Matzek, B.A., et al. Traumatic testicular dislocation after minor trauma in a pediatric patient. J Emerg
Med, 2013. 45: 537.
https://pubmed.ncbi.nlm.nih.gov/23899815
218. Wright, S., et al. Emergency ultrasound of acute scrotal pain. Eur J Emerg Med, 2015. 22: 2.
https://pubmed.ncbi.nlm.nih.gov/24910960
219. Yusuf, G.T., et al. A review of ultrasound imaging in scrotal emergencies. J Ultrasound, 2013. 16: 171.
https://pubmed.ncbi.nlm.nih.gov/24432171
220. Remer, E.M., et al. ACR Appropriateness Criteria (R) acute onset of scrotal pain--without trauma,
without antecedent mass. Ultrasound Q, 2012. 28: 47.
https://pubmed.ncbi.nlm.nih.gov/22357246
221. Tanaka, K., et al. Acute scrotum and testicular torsion in children: a retrospective study in a single
institution. J Pediatr Urol, 2020. 16: 55.
https://pubmed.ncbi.nlm.nih.gov/31874735
222. Kadish, H.A., et al. A retrospective review of pediatric patients with epididymitis, testicular torsion,
and torsion of testicular appendages. Pediatrics, 1998. 102: 73.
https://pubmed.ncbi.nlm.nih.gov/9651416
223. Sauvat, F., et al. [Age for testicular torsion?]. Arch Pediatr, 2002. 9: 1226.
https://pubmed.ncbi.nlm.nih.gov/12536102
224. Somekh, E., et al. Acute epididymitis in boys: evidence of a post-infectious etiology. J Urol, 2004.
171: 391.
https://pubmed.ncbi.nlm.nih.gov/14665940
225. Yerkes, E.B., et al. Management of perinatal torsion: today, tomorrow or never? J Urol, 2005. 174: 1579.
https://pubmed.ncbi.nlm.nih.gov/16148656
226. Boettcher, M., et al. Clinical and sonographic features predict testicular torsion in children: a
prospective study. BJU Int, 2013. 112: 1201.
https://pubmed.ncbi.nlm.nih.gov/23826981
227. Nelson, C.P., et al. The cremasteric reflex: a useful but imperfect sign in testicular torsion. J Pediatr
Surg, 2003. 38: 1248.
https://pubmed.ncbi.nlm.nih.gov/12891505
228. Goetz, J., et al. A comparison of clinical outcomes of acute testicular torsion between prepubertal
and postpubertal males. J Pediatr Urol, 2019. 15: 610.
https://pubmed.ncbi.nlm.nih.gov/31690483
229. Mushtaq, I., et al. Retrospective review of paediatric patients with acute scrotum. ANZ J Surg, 2003.
73: 55.
https://pubmed.ncbi.nlm.nih.gov/12534742
230. Murphy, F.L., et al. Early scrotal exploration in all cases is the investigation and intervention of
choice in the acute paediatric scrotum. Pediatr Surg Int, 2006. 22: 413.
https://pubmed.ncbi.nlm.nih.gov/16602024
231. Baker, L.A., et al. An analysis of clinical outcomes using color doppler testicular ultrasound for
testicular torsion. Pediatrics, 2000. 105: 604.
https://pubmed.ncbi.nlm.nih.gov/10699116
232. Gunther, P., et al. Acute testicular torsion in children: the role of sonography in the diagnostic
workup. Eur Radiol, 2006. 16: 2527.
https://pubmed.ncbi.nlm.nih.gov/16724203

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 123


233. Kalfa, N., et al. Multicenter assessment of ultrasound of the spermatic cord in children with acute
scrotum. J Urol, 2007. 177: 297.
https://pubmed.ncbi.nlm.nih.gov/17162068
234. Karmazyn, B., et al. Clinical and sonographic criteria of acute scrotum in children: a retrospective
study of 172 boys. Pediatr Radiol, 2005. 35: 302.
https://pubmed.ncbi.nlm.nih.gov/15503003
235. Lam, W.W., et al. Colour Doppler ultrasonography replacing surgical exploration for acute scrotum:
myth or reality? Pediatr Radiol, 2005. 35: 597.
https://pubmed.ncbi.nlm.nih.gov/15761770
236. Schalamon, J., et al. Management of acute scrotum in children--the impact of Doppler ultrasound.
J Pediatr Surg, 2006. 41: 1377.
https://pubmed.ncbi.nlm.nih.gov/16863840
237. Pepe, P., et al. Does color Doppler sonography improve the clinical assessment of patients with
acute scrotum? Eur J Radiol, 2006. 60: 120.
https://pubmed.ncbi.nlm.nih.gov/16730939
238. Kalfa, N., et al. Ultrasonography of the spermatic cord in children with testicular torsion: impact on
the surgical strategy. J Urol, 2004. 172: 1692.
https://pubmed.ncbi.nlm.nih.gov/15371792
239. McDowall, J., et al. The ultrasonographic “whirlpool sign” in testicular torsion: valuable tool or waste
of valuable time? A systematic review and meta-analysis. Emerg Radiol, 2018. 25: 281.
https://pubmed.ncbi.nlm.nih.gov/29335899
240. Nussbaum Blask, A.R., et al. Color Doppler sonography and scintigraphy of the testis: a prospective,
comparative analysis in children with acute scrotal pain. Pediatr Emerg Care, 2002. 18: 67.
https://pubmed.ncbi.nlm.nih.gov/11973493
241. Paltiel, H.J., et al. Acute scrotal symptoms in boys with an indeterminate clinical presentation:
comparison of color Doppler sonography and scintigraphy. Radiology, 1998. 207: 223.
https://pubmed.ncbi.nlm.nih.gov/9530319
242. Terai, A., et al. Dynamic contrast-enhanced subtraction magnetic resonance imaging in diagnostics
of testicular torsion. Urology, 2006. 67: 1278.
https://pubmed.ncbi.nlm.nih.gov/16765192
243. Yuan, Z., et al. Clinical study of scrotum scintigraphy in 49 patients with acute scrotal pain: a
comparison with ultrasonography. Ann Nucl Med, 2001. 15: 225.
https://pubmed.ncbi.nlm.nih.gov/11545192
244. Karmazyn, B., et al. Duplex sonographic findings in children with torsion of the testicular
appendages: overlap with epididymitis and epididymoorchitis. J Pediatr Surg, 2006. 41: 500.
https://pubmed.ncbi.nlm.nih.gov/16516624
245. Lau, P., et al. Acute epididymitis in boys: are antibiotics indicated? Br J Urol, 1997. 79: 797.
https://pubmed.ncbi.nlm.nih.gov/9158522
246. Abul, F., et al. The acute scrotum: a review of 40 cases. Med Princ Pract, 2005. 14: 177.
https://pubmed.ncbi.nlm.nih.gov/15863992
247. Dias Filho, A.C., et al. Improving Organ Salvage in Testicular Torsion: Comparative Study of Patients
Undergoing vs Not Undergoing Preoperative Manual Detorsion. J Urol, 2017. 197: 811.
https://pubmed.ncbi.nlm.nih.gov/27697579
248. Cornel, E.B., et al. Manual derotation of the twisted spermatic cord. BJU Int, 1999. 83: 672.
https://pubmed.ncbi.nlm.nih.gov/10233577
249. Garel, L., et al. Preoperative manual detorsion of the spermatic cord with Doppler ultrasound
monitoring in patients with intravaginal acute testicular torsion. Pediatr Radiol, 2000. 30: 41.
https://pubmed.ncbi.nlm.nih.gov/10663509
250. Sessions, A.E., et al. Testicular torsion: direction, degree, duration and disinformation. J Urol, 2003.
169: 663.
https://pubmed.ncbi.nlm.nih.gov/12544339
251. Haj, M., et al. Effect of external scrotal cooling on the viability of the testis with torsion in rats. Eur
Surg Res, 2007. 39: 160.
https://pubmed.ncbi.nlm.nih.gov/17341878
252. Akcora, B., et al. The protective effect of darbepoetin alfa on experimental testicular torsion and
detorsion injury. Int J Urol, 2007. 14: 846.
https://pubmed.ncbi.nlm.nih.gov/17760753
253. Aksoy, H., et al. Dehydroepiandrosterone treatment attenuates reperfusion injury after testicular
torsion and detorsion in rats. J Pediatr Surg, 2007. 42: 1740.
https://pubmed.ncbi.nlm.nih.gov/17923206

124 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


254. Unal, D., et al. Protective effects of trimetazidine on testicular ischemia-reperfusion injury in rats.
Urol Int, 2007. 78: 356.
https://pubmed.ncbi.nlm.nih.gov/17495496
255. Yazihan, N., et al. Protective role of erythropoietin during testicular torsion of the rats. World J Urol,
2007. 25: 531.
https://pubmed.ncbi.nlm.nih.gov/17690891
256. Koh, Y.H., et al. Testicular Appendage Torsion-To Explore the Other Side or Not? Urology, 2020.
141: 130.
https://pubmed.ncbi.nlm.nih.gov/32283168
257. Visser, A.J., et al. Testicular function after torsion of the spermatic cord. BJU Int, 2003. 92: 200.
https://pubmed.ncbi.nlm.nih.gov/12887467
258. Tryfonas, G., et al. Late postoperative results in males treated for testicular torsion during childhood.
J Pediatr Surg, 1994. 29: 553.
https://pubmed.ncbi.nlm.nih.gov/8014814
259. Anderson, M.J., et al. Semen quality and endocrine parameters after acute testicular torsion. J Urol,
1992. 147: 1545.
https://pubmed.ncbi.nlm.nih.gov/1593686
260. Arap, M.A., et al. Late hormonal levels, semen parameters, and presence of antisperm antibodies in
patients treated for testicular torsion. J Androl, 2007. 28: 528.
https://pubmed.ncbi.nlm.nih.gov/17287456
261. Figueroa, V., et al. Comparative analysis of detorsion alone versus detorsion and tunica albuginea
decompression (fasciotomy) with tunica vaginalis flap coverage in the surgical management of
prolonged testicular ischemia. J Urol, 2012. 188: 1417.
https://pubmed.ncbi.nlm.nih.gov/22906680
262. Monteilh, C., et al. Controversies in the management of neonatal testicular torsion: A meta-analysis.
J Pediatr Surg, 2019. 54: 815.
https://pubmed.ncbi.nlm.nih.gov/30098810
263. Mor, Y., et al. Testicular fixation following torsion of the spermatic cord--does it guarantee prevention
of recurrent torsion events? J Urol, 2006. 175: 171.
https://pubmed.ncbi.nlm.nih.gov/16406900
264. Lian, B.S., et al. Factors Predicting Testicular Atrophy after Testicular Salvage following Torsion. Eur
J Pediatr Surg, 2016. 26: 17.
https://pubmed.ncbi.nlm.nih.gov/26509312
265. MacDonald, C., et al. A systematic review and meta-analysis revealing realistic outcomes following
paediatric torsion of testes. J Pediatr Urol, 2018. 14: 503.
https://pubmed.ncbi.nlm.nih.gov/30404723
266. Philip, J., et al. Mumps orchitis in the non-immune postpubertal male: a resurgent threat to male
fertility? BJU Int, 2006. 97: 138.
https://pubmed.ncbi.nlm.nih.gov/16336344
267. Gielchinsky, I., et al. Pregnancy Rates after Testicular Torsion. J Urol, 2016. 196: 852.
https://pubmed.ncbi.nlm.nih.gov/27117442
268. Bergman, J.E., et al. Epidemiology of hypospadias in Europe: a registry-based study. World J Urol,
2015. 33: 2159.
https://pubmed.ncbi.nlm.nih.gov/25712311
269. Morera, A.M., et al. A study of risk factors for hypospadias in the Rhone-Alpes region (France).
J Pediatr Urol, 2006. 2: 169.
https://pubmed.ncbi.nlm.nih.gov/18947603
270. Springer, A., et al. Worldwide prevalence of hypospadias. J Pediatr Urol, 2016. 12: 152 e1.
https://pubmed.ncbi.nlm.nih.gov/26810252
271. van der Zanden, L.F., et al. Exploration of gene-environment interactions, maternal effects and
parent of origin effects in the etiology of hypospadias. J Urol, 2012. 188: 2354.
https://pubmed.ncbi.nlm.nih.gov/23088992
272. Fredell, L., et al. Heredity of hypospadias and the significance of low birth weight. J Urol, 2002.
167: 1423.
https://pubmed.ncbi.nlm.nih.gov/11832761
273. Lund, L., et al. Prevalence of hypospadias in Danish boys: a longitudinal study, 1977-2005. Eur Urol,
2009. 55: 1022.
https://pubmed.ncbi.nlm.nih.gov/19155122
274. Mouriquand, O.D., et al., Hypospadias., in Pediatric Urology, J. Gearhart, R. Rink & P.D.E.
Mouriquand, Editors. 2001, WB Saunders: Philadelphia.

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 125


275. van Rooij, I.A., et al. Risk factors for different phenotypes of hypospadias: results from a Dutch
case-control study. BJU Int, 2013. 112: 121.
https://pubmed.ncbi.nlm.nih.gov/23305310
276. Netto, J.M., et al. Hormone therapy in hypospadias surgery: a systematic review. J Pediatr Urol,
2013. 9: 971.
https://pubmed.ncbi.nlm.nih.gov/23602841
277. Chariatte, V., et al. Uroradiological screening for upper and lower urinary tract anomalies in patients
with hypospadias: a systematic literature review. Evid Based Med, 2013. 18: 11.
https://pubmed.ncbi.nlm.nih.gov/22815315
278. Belman, A.B., Hypospadias and chordee, in Clinical Pediatric Urology A.B. Belman, L.R. King & S.A.
Kramer, Editors. 2002, Martin Dunitz: London.
279. Malik, R.D., et al. Survey of pediatric urologists on the preoperative use of testosterone in the
surgical correction of hypospadias. J Pediatr Urol, 2014.
https://pubmed.ncbi.nlm.nih.gov/24726783
280. Wright, I., et al. Effect of preoperative hormonal stimulation on postoperative complication rates
after proximal hypospadias repair: a systematic review. J Urol, 2013. 190: 652.
https://pubmed.ncbi.nlm.nih.gov/23597451
281. Rynja, S.P., et al. Testosterone prior to hypospadias repair: Postoperative complication rates and
long-term cosmetic results, penile length and body height. J Pediatr Urol, 2018. 14: 31.e1.
https://pubmed.ncbi.nlm.nih.gov/29174377
282. Paiva, K.C., et al. Biometry of the hypospadic penis after hormone therapy (testosterone and
estrogen): A randomized, double-blind controlled trial. J Pediatr Urol, 2016. 12: 200.e1.
https://pubmed.ncbi.nlm.nih.gov/27321554
283. Chua, M.E., et al. Preoperative hormonal stimulation effect on hypospadias repair complications:
Meta-analysis of observational versus randomized controlled studies. J Pediatr Urol, 2017. 13: 470.
https://pubmed.ncbi.nlm.nih.gov/28939350
284. Kaya, C., et al. The role of pre-operative androgen stimulation in hypospadias surgery. Transl Androl
Urol, 2014. 3: 340.
https://pubmed.ncbi.nlm.nih.gov/26816790
285. Menon, P., et al. Outcome of urethroplasty after parenteral testosterone in children with distal
hypospadias. J Pediatr Urol, 2017. 13: 292.e1.
https://pubmed.ncbi.nlm.nih.gov/28111208
286. Bush, N.C., et al. Age does not impact risk for urethroplasty complications after tubularized incised
plate repair of hypospadias in prepubertal boys. J Pediatr Urol, 2013. 9: 252.
https://pubmed.ncbi.nlm.nih.gov/22542204
287. Perlmutter, A.E., et al. Impact of patient age on distal hypospadias repair: a surgical perspective.
Urology, 2006. 68: 648.
https://pubmed.ncbi.nlm.nih.gov/16979730
288. Bhat, A., et al. Comparison of variables affecting the surgical outcomes of tubularized incised plate
urethroplasty in adult and pediatric hypospadias. J Pediatr Urol, 2016. 12: 108 e1.
https://pubmed.ncbi.nlm.nih.gov/26778183
289. Castagnetti, M., et al. Surgical management of primary severe hypospadias in children: systematic
20-year review. J Urol, 2010. 184: 1469.
https://pubmed.ncbi.nlm.nih.gov/20727541
290. Baskin, L.S., et al. Changing concepts of hypospadias curvature lead to more onlay island flap
procedures. J Urol, 1994. 151: 191.
https://pubmed.ncbi.nlm.nih.gov/8254812
291. Hollowell, J.G., et al. Preservation of the urethral plate in hypospadias repair: extended applications
and further experience with the onlay island flap urethroplasty. J Urol, 1990. 143: 98.
https://pubmed.ncbi.nlm.nih.gov/2294275
292. Snodgrass, W., et al. Straightening ventral curvature while preserving the urethral plate in proximal
hypospadias repair. J Urol, 2009. 182: 1720.
https://pubmed.ncbi.nlm.nih.gov/19692004
293. Braga, L.H., et al. Ventral penile lengthening versus dorsal plication for severe ventral curvature in
children with proximal hypospadias. J Urol, 2008. 180: 1743.
https://pubmed.ncbi.nlm.nih.gov/18721961
294. el-Kassaby, A.W., et al. Modified tubularized incised plate urethroplasty for hypospadias repair: a
long-term results of 764 patients. Urology, 2008. 71: 611.
https://pubmed.ncbi.nlm.nih.gov/18295308

126 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


295. El-Sherbiny, M.T., et al. Comprehensive analysis of tubularized incised-plate urethroplasty in primary
and re-operative hypospadias. BJU Int, 2004. 93: 1057.
https://pubmed.ncbi.nlm.nih.gov/15142164
296. Orkiszewski, M., et al. Morphology and urodynamics after longitudinal urethral plate incision in
proximal hypospadias repairs: long-term results. Eur J Pediatr Surg, 2004. 14: 35.
https://pubmed.ncbi.nlm.nih.gov/15024677
297. Snodgrass, W.T., et al. Tubularized incised plate hypospadias repair for distal hypospadias. J Pediatr
Urol, 2010. 6: 408.
https://pubmed.ncbi.nlm.nih.gov/19837000
298. Schwentner, C., et al. Interim outcome of the single stage dorsal inlay skin graft for complex
hypospadias reoperations. J Urol, 2006. 175: 1872.
https://pubmed.ncbi.nlm.nih.gov/16600785
299. Ahmed, M., et al. Is combined inner preputial inlay graft with tubularized incised plate in
hypospadias repair worth doing? J Pediatr Urol, 2015. 11: 229 e1.
https://pubmed.ncbi.nlm.nih.gov/26119452
300. Pippi Salle, J.L., et al. Proximal hypospadias: A persistent challenge. Single institution outcome
analysis of three surgical techniques over a 10-year period. J Pediatr Urol, 2016. 12: 28 e1.
https://pubmed.ncbi.nlm.nih.gov/26279102
301. Meyer-Junghanel, L., et al. Experience with repair of 120 hypospadias using Mathieu’s procedure.
Eur J Pediatr Surg, 1995. 5: 355.
https://pubmed.ncbi.nlm.nih.gov/8773227
302. Pfistermuller, K.L., et al. Meta-analysis of complication rates of the tubularized incised plate (TIP)
repair. J Pediatr Urol, 2015. 11: 54.
https://pubmed.ncbi.nlm.nih.gov/25819601
303. Snodgrass, W.T., et al. Urethral strictures following urethral plate and proximal urethral elevation
during proximal TIP hypospadias repair. J Pediatr Urol, 2013. 9: 990.
https://pubmed.ncbi.nlm.nih.gov/23707201
304. Cambareri, G.M., et al. Hypospadias repair with onlay preputial graft: a 25-year experience with
long-term follow-up. BJU Int, 2016. 118: 451.
https://pubmed.ncbi.nlm.nih.gov/26780179
305. Castagnetti, M., et al. Primary severe hypospadias: comparison of reoperation rates and parental
perception of urinary symptoms and cosmetic outcomes among 4 repairs. J Urol, 2013. 189: 1508.
https://pubmed.ncbi.nlm.nih.gov/23154207
306. Kocvara, R., et al. Inlay-onlay flap urethroplasty for hypospadias and urethral stricture repair. J Urol,
1997. 158: 2142.
https://pubmed.ncbi.nlm.nih.gov/9366331
307. Perovic, S., et al. Onlay island flap urethroplasty for severe hypospadias: a variant of the technique.
J Urol, 1994. 151: 711.
https://pubmed.ncbi.nlm.nih.gov/8308994
308. Catti, M., et al. Original Koyanagi urethroplasty versus modified Hayashi technique: outcome in
57 patients. J Pediatr Urol, 2009. 5: 300.
https://pubmed.ncbi.nlm.nih.gov/19457720
309. DeFoor, W., et al. Results of single staged hypospadias surgery to repair penoscrotal hypospadias
with bifid scrotum or penoscrotal transposition. J Urol, 2003. 170: 1585.
https://pubmed.ncbi.nlm.nih.gov/14501667
310. Hayashi, Y., et al. Neo-modified Koyanagi technique for the single-stage repair of proximal
hypospadias. J Pediatr Urol, 2007. 3: 239.
https://pubmed.ncbi.nlm.nih.gov/18947743
311. Koyanagi, T., et al. One-stage repair of hypospadias: is there no simple method universally
applicable to all types of hypospadias? J Urol, 1994. 152: 1232.
https://pubmed.ncbi.nlm.nih.gov/8072111
312. Ahmed, S., et al. Buccal mucosal graft for secondary hypospadias repair and urethral replacement.
Br J Urol, 1997. 80: 328.
https://pubmed.ncbi.nlm.nih.gov/9284210
313. Bracka, A. Hypospadias repair: the two-stage alternative. Br J Urol, 1995. 76 Suppl 3: 31.
https://pubmed.ncbi.nlm.nih.gov/8535768
314. Lam, P.N., et al. 2-stage repair in infancy for severe hypospadias with chordee: long-term results
after puberty. J Urol, 2005. 174: 1567.
https://pubmed.ncbi.nlm.nih.gov/16148653

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 127


315. Mokhless, I.A., et al. The multistage use of buccal mucosa grafts for complex hypospadias:
histological changes. J Urol, 2007. 177: 1496.
https://pubmed.ncbi.nlm.nih.gov/17382762
316. Stanasel, I., et al. Complications following Staged Hypospadias Repair Using Transposed Preputial
Skin Flaps. J Urol, 2015. 194: 512.
https://pubmed.ncbi.nlm.nih.gov/25701546
317. Castagnetti, M., et al. Does Preputial Reconstruction Increase Complication Rate of Hypospadias
Repair? 20-Year Systematic Review and Meta-Analysis. Front Pediatr, 2016. 4: 41.
https://pubmed.ncbi.nlm.nih.gov/27200322
318. Chalmers, D.J., et al. Distal hypospadias repair in infants without a postoperative stent. Pediatr Surg
Int, 2015. 31: 287.
https://pubmed.ncbi.nlm.nih.gov/25475503
319. Hsieh, M.H., et al. Surgical antibiotic practices among pediatric urologists in the United States.
J Pediatr Urol, 2011. 7: 192.
https://pubmed.ncbi.nlm.nih.gov/20537590
320. Kanaroglou, N., et al. Is there a role for prophylactic antibiotics after stented hypospadias repair?
J Urol, 2013. 190: 1535.
https://pubmed.ncbi.nlm.nih.gov/23416639
321. Meir, D.B., et al. Is prophylactic antimicrobial treatment necessary after hypospadias repair? J Urol,
2004. 171: 2621.
https://pubmed.ncbi.nlm.nih.gov/15118434
322. Chua, M.E., et al. The use of postoperative prophylactic antibiotics in stented distal hypospadias
repair: a systematic review and meta-analysis. J Pediatr Urol, 2019. 15: 138.
https://pubmed.ncbi.nlm.nih.gov/30527683
323. Bush, N.C., et al. Glans size is an independent risk factor for urethroplasty complications after
hypospadias repair. J Pediatr Urol, 2015. 11: 355 e1.
https://pubmed.ncbi.nlm.nih.gov/26320396
324. Lee, O.T., et al. Predictors of secondary surgery after hypospadias repair: a population based
analysis of 5,000 patients. J Urol, 2013. 190: 251.
https://pubmed.ncbi.nlm.nih.gov/23376710
325. Braga, L.H., et al. Tubularized incised plate urethroplasty for distal hypospadias: A literature review.
Indian J Urol, 2008. 24: 219.
https://pubmed.ncbi.nlm.nih.gov/19468401
326. Wang, F., et al. Systematic review and meta-analysis of studies comparing the perimeatal-based
flap and tubularized incised-plate techniques for primary hypospadias repair. Pediatr Surg Int, 2013.
29: 811.
https://pubmed.ncbi.nlm.nih.gov/23793987
327. Wilkinson, D.J., et al. Outcomes in distal hypospadias: a systematic review of the Mathieu and
tubularized incised plate repairs. J Pediatr Urol, 2012. 8: 307.
https://pubmed.ncbi.nlm.nih.gov/21159560
328. Winberg, H., et al. Postoperative outcomes in distal hypospadias: a meta-analysis of the Mathieu
and tubularized incised plate repair methods for development of urethrocutaneous fistula and
urethral stricture. Pediatr Surg Int, 2019. 35: 1301.
https://pubmed.ncbi.nlm.nih.gov/31372729
329. Alshafei, A., et al. Comparing the outcomes of tubularized incised plate urethroplasty and dorsal
inlay graft urethroplasty in children with hypospadias: a systematic review and meta-analysis.
J Pediatr Urol, 2020. 16: 154.
https://pubmed.ncbi.nlm.nih.gov/32061491
330. Leslie, B., et al. Critical outcome analysis of staged buccal mucosa graft urethroplasty for prior
failed hypospadias repair in children. J Urol, 2011. 185: 1077.
https://pubmed.ncbi.nlm.nih.gov/21256520
331. Howe, A.S., et al. Management of 220 adolescents and adults with complications of hypospadias
repair during childhood. Asian J Urol, 2017. 4: 14.
https://pubmed.ncbi.nlm.nih.gov/29264201
332. Spinoit, A.F., et al. Hypospadias repair at a tertiary care center: long-term followup is mandatory to
determine the real complication rate. J Urol, 2013. 189: 2276.
https://pubmed.ncbi.nlm.nih.gov/23306089
333. Andersson, M., et al. Hypospadias repair with tubularized incised plate: Does the obstructive flow
pattern resolve spontaneously? J Pediatr Urol, 2011. 7: 441.
https://pubmed.ncbi.nlm.nih.gov/20630805

128 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


334. Andersson, M., et al. Normalized Urinary Flow at Puberty after Tubularized Incised Plate
Urethroplasty for Hypospadias in Childhood. J Urol, 2015. 194: 1407.
https://pubmed.ncbi.nlm.nih.gov/26087380
335. Gonzalez, R., et al. Importance of urinary flow studies after hypospadias repair: a systematic review.
Int J Urol, 2011. 18: 757.
https://pubmed.ncbi.nlm.nih.gov/21883491
336. Hueber, P.A., et al. Long-term functional outcomes of distal hypospadias repair: a single center
retrospective comparative study of TIPs, Mathieu and MAGPI. J Pediatr Urol, 2015. 11: 68 e1.
https://pubmed.ncbi.nlm.nih.gov/25824882
337. Perera, M., et al. Long-term urethral function measured by uroflowmetry after hypospadias surgery:
comparison with an age matched control. J Urol, 2012. 188: 1457.
https://pubmed.ncbi.nlm.nih.gov/22906660
338. Holland, A.J., et al. HOSE: an objective scoring system for evaluating the results of hypospadias
surgery. BJU Int, 2001. 88: 255.
https://pubmed.ncbi.nlm.nih.gov/11488741
339. van der Toorn, F., et al. Introducing the HOPE (Hypospadias Objective Penile Evaluation)-score: a
validation study of an objective scoring system for evaluating cosmetic appearance in hypospadias
patients. J Pediatr Urol, 2013. 9: 1006.
https://pubmed.ncbi.nlm.nih.gov/23491983
340. Weber, D.M., et al. The Penile Perception Score: an instrument enabling evaluation by surgeons and
patient self-assessment after hypospadias repair. J Urol, 2013. 189: 189.
https://pubmed.ncbi.nlm.nih.gov/23174225
341. Haid, B., et al. Penile appearance after hypospadias correction from a parent’s point of view:
Comparison of the hypospadias objective penile evaluation score and parents penile perception
score. J Pediatr Urol, 2016. 12: 33.e1.
https://pubmed.ncbi.nlm.nih.gov/26725130
342. Krull, S., et al. Outcome after Hypospadias Repair: Evaluation Using the Hypospadias Objective
Penile Evaluation Score. Eur J Pediatr Surg, 2018. 28: 268.
https://pubmed.ncbi.nlm.nih.gov/28505692
343. Moriya, K., et al. Long-term cosmetic and sexual outcome of hypospadias surgery: norm related
study in adolescence. J Urol, 2006. 176: 1889.
https://pubmed.ncbi.nlm.nih.gov/16945681
344. Rynja, S.P., et al. Functional, cosmetic and psychosexual results in adult men who underwent
hypospadias correction in childhood. J Pediatr Urol, 2011. 7: 504.
https://pubmed.ncbi.nlm.nih.gov/21429804
345. Ortqvist, L., et al. Long-term followup of men born with hypospadias: urological and cosmetic
results. J Urol, 2015. 193: 975.
https://pubmed.ncbi.nlm.nih.gov/25268894
346. Adams, J., et al. Reconstructive surgery for hypospadias: A systematic review of long-term patient
satisfaction with cosmetic outcomes. Indian J Urol, 2016. 32: 93.
https://pubmed.ncbi.nlm.nih.gov/27127350
347. Leunbach, T.L., et al. A Systematic Review of Core Outcomes for Hypospadias Surgery. Sex Dev,
2019. 13: 165.
https://pubmed.ncbi.nlm.nih.gov/31865321
348. Sullivan, K.J., et al. Assessing quality of life of patients with hypospadias: A systematic review of
validated patient-reported outcome instruments. J Pediatr Urol, 2017. 13: 19.
https://pubmed.ncbi.nlm.nih.gov/28089292
349. Nyirady, P., et al. Management of congenital penile curvature. J Urol, 2008. 179: 1495.
https://pubmed.ncbi.nlm.nih.gov/18295273
350. Baskin, L.S., et al. Neuroanatomical ontogeny of the human fetal penis. Br J Urol, 1997. 79: 628.
https://pubmed.ncbi.nlm.nih.gov/9126098
351. Ebbehoj, J., et al. Congenital penile angulation. Br J Urol, 1987. 60: 264.
https://pubmed.ncbi.nlm.nih.gov/3676675
352. Kelami, A. Congenital penile deviation and its treatment with the Nesbit-Kelami technique. Br J Urol,
1987. 60: 261.
https://pubmed.ncbi.nlm.nih.gov/3676674
353. Yachia, D., et al. The incidence of congenital penile curvature. J Urol, 1993. 150: 1478.
https://pubmed.ncbi.nlm.nih.gov/8411431
354. Mayer, M., et al. Patient satisfaction with correction of congenital penile curvature. Actas Urol Esp, 2017.
https://pubmed.ncbi.nlm.nih.gov/29292041

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 129


355. Hsieh, J.T., et al. Correction of congenital penile curvature using modified tunical plication with
absorbable sutures: the long-term outcome and patient satisfaction. Eur Urol, 2007. 52: 261.
https://pubmed.ncbi.nlm.nih.gov/17234333
356. Sasso, F., et al. Penile curvature: an update for management from 20 years experience in a high
volume centre. Urologia, 2016. 83: 130.
https://pubmed.ncbi.nlm.nih.gov/27103093
357. Gittes, R.F., et al. Injection technique to induce penile erection. Urology, 1974. 4: 473.
https://pubmed.ncbi.nlm.nih.gov/4418594
358. Schultheiss, D., et al. Congenital and acquired penile deviation treated with the essed plication
method. Eur Urol, 2000. 38: 167.
https://pubmed.ncbi.nlm.nih.gov/10895008
359. Yachia, D. Modified corporoplasty for the treatment of penile curvature. J Urol, 1990. 143: 80.
https://pubmed.ncbi.nlm.nih.gov/2294269
360. Rehman, J., et al. Results of surgical treatment for abnormal penile curvature: Peyronie’s disease
and congenital deviation by modified Nesbit plication (tunical shaving and plication). J Urol, 1997.
157: 1288.
https://pubmed.ncbi.nlm.nih.gov/9120923
361. Poulsen, J., et al. Treatment of penile curvature--a retrospective study of 175 patients operated with
plication of the tunica albuginea or with the Nesbit procedure. Br J Urol, 1995. 75: 370.
https://pubmed.ncbi.nlm.nih.gov/7735803
362. Leonardo, C., et al. Plication corporoplasty versus Nesbit operation for the correction of congenital
penile curvature. A long-term follow-up. Int Urol Nephrol, 2012. 44: 55.
https://pubmed.ncbi.nlm.nih.gov/21559790
363. Çayan, S., et al. Comparison of Patient’s Satisfaction and Long-term Results of 2 Penile Plication
Techniques: Lessons Learned From 387 Patients With Penile Curvature. Urology, 2019. 129: 106.
https://pubmed.ncbi.nlm.nih.gov/30954611
364. Cavallini, G., et al. Pilot study to determine improvements in subjective penile morphology and
personal relationships following a Nesbit plication procedure for men with congenital penile
curvature. Asian J Androl, 2008. 10: 512.
https://pubmed.ncbi.nlm.nih.gov/18097530
365. Vatne, V., et al. Functional results after operations of penile deviations: an institutional experience.
Scand J Urol Nephrol Suppl, 1996. 179: 151.
https://pubmed.ncbi.nlm.nih.gov/8908683
366. Ziegelmann, M.J., et al. Clinical characteristics and surgical outcomes in men undergoing tunica
albuginea plication for congenital penile curvature who present with worsening penile deformity.
World J Urol, 2020. 38: 305.
https://pubmed.ncbi.nlm.nih.gov/31079186
367. Shaeer, O., et al. Shaeer’s Corporal Rotation III: Shortening-Free Correction of Congenital Penile
Curvature-The Noncorporotomy Technique. Eur Urol, 2016. 69: 129.
https://pubmed.ncbi.nlm.nih.gov/26298209
368. Akbay, E., et al. The prevalence of varicocele and varicocele-related testicular atrophy in Turkish
children and adolescents. BJU Int, 2000. 86: 490.
https://pubmed.ncbi.nlm.nih.gov/10971279
369. Kogan, S.J., The pediatric varicocele. , in Pediatric urology, J.P. Gearhart, R.C. Rink & P.D.E.
Mouriquand, Editors. 2001, WB Saunders: Philadelphia.
370. Oster, J. Varicocele in children and adolescents. An investigation of the incidence among Danish
school children. Scand J Urol Nephrol, 1971. 5: 27.
https://pubmed.ncbi.nlm.nih.gov/5093090
371. Kass, E.J., et al. Reversal of testicular growth failure by varicocele ligation. J Urol, 1987. 137: 475.
https://pubmed.ncbi.nlm.nih.gov/3820376
372. Paduch, D.A., et al. Repair versus observation in adolescent varicocele: a prospective study. J Urol,
1997. 158: 1128.
https://pubmed.ncbi.nlm.nih.gov/9258155
373. Li, F., et al. Effect of varicocelectomy on testicular volume in children and adolescents: a meta-
analysis. Urology, 2012. 79: 1340.
https://pubmed.ncbi.nlm.nih.gov/22516359
374. Kocvara, R., et al. Division of lymphatic vessels at varicocelectomy leads to testicular oedema and
decline in testicular function according to the LH-RH analogue stimulation test. Eur Urol, 2003.
43: 430.
https://pubmed.ncbi.nlm.nih.gov/12667726

130 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


375. The influence of varicocele on parameters of fertility in a large group of men presenting to infertility
clinics. World Health Organization. Fertil Steril, 1992. 57: 1289.
https://pubmed.ncbi.nlm.nih.gov/1601152
376. Laven, J.S., et al. Effects of varicocele treatment in adolescents: a randomized study. Fertil Steril,
1992. 58: 756.
https://pubmed.ncbi.nlm.nih.gov/1426322
377. Nork, J.J., et al. Youth varicocele and varicocele treatment: a meta-analysis of semen outcomes.
Fertil Steril, 2014. 102: 381.
https://pubmed.ncbi.nlm.nih.gov/24907913
378. Okuyama, A., et al. Surgical repair of varicocele at puberty: preventive treatment for fertility
improvement. J Urol, 1988. 139: 562.
https://pubmed.ncbi.nlm.nih.gov/3343743
379. Pinto, K.J., et al. Varicocele related testicular atrophy and its predictive effect upon fertility. J Urol,
1994. 152: 788.
https://pubmed.ncbi.nlm.nih.gov/8022015
380. Dubin, L., et al. Varicocele size and results of varicocelectomy in selected subfertile men with
varicocele. Fertil Steril, 1970. 21: 606.
https://pubmed.ncbi.nlm.nih.gov/5433164
381. Tasci, A.I., et al. Color doppler ultrasonography and spectral analysis of venous flow in diagnosis of
varicocele. Eur Urol, 2001. 39: 316.
https://pubmed.ncbi.nlm.nih.gov/11275726
382. Diamond, D.A., et al. Relationship of varicocele grade and testicular hypotrophy to semen
parameters in adolescents. J Urol, 2007. 178: 1584.
https://pubmed.ncbi.nlm.nih.gov/17707046
383. Aragona, F., et al. Correlation of testicular volume, histology and LHRH test in adolescents with
idiopathic varicocele. Eur Urol, 1994. 26: 61.
https://pubmed.ncbi.nlm.nih.gov/7925532
384. Bogaert, G., et al. Pubertal screening and treatment for varicocele do not improve chance of
paternity as adult. J Urol, 2013. 189: 2298.
https://pubmed.ncbi.nlm.nih.gov/23261480
385. Chen, J.J., et al. Is the comparison of a left varicocele testis to its contralateral normal testis
sufficient in determining its well-being? Urology, 2011. 78: 1167.
https://pubmed.ncbi.nlm.nih.gov/21782220
386. Goldstein, M., et al. Microsurgical inguinal varicocelectomy with delivery of the testis: an artery and
lymphatic sparing technique. J Urol, 1992. 148: 1808.
https://pubmed.ncbi.nlm.nih.gov/1433614
387. Hopps, C.V., et al. Intraoperative varicocele anatomy: a microscopic study of the inguinal versus
subinguinal approach. J Urol, 2003. 170: 2366.
https://pubmed.ncbi.nlm.nih.gov/14634418
388. Kocvara, R., et al. Lymphatic sparing laparoscopic varicocelectomy: a microsurgical repair. J Urol,
2005. 173: 1751.
https://pubmed.ncbi.nlm.nih.gov/15821575
389. Riccabona, M., et al. Optimizing the operative treatment of boys with varicocele: sequential
comparison of 4 techniques. J Urol, 2003. 169: 666.
https://pubmed.ncbi.nlm.nih.gov/12544340
390. Marmar, J., et al. New scientific information related to varicoceles. J Urol, 2003. 170: 2371.
https://pubmed.ncbi.nlm.nih.gov/14634419
391. Minevich, E., et al. Inguinal microsurgical varicocelectomy in the adolescent: technique and
preliminary results. J Urol, 1998. 159: 1022.
https://pubmed.ncbi.nlm.nih.gov/9474223
392. Mirilas, P., et al. Microsurgical subinguinal varicocelectomy in children, adolescents, and adults:
surgical anatomy and anatomically justified technique. J Androl, 2012. 33: 338.
https://pubmed.ncbi.nlm.nih.gov/21835913
393. Esposito, C., et al. Technical standardization of laparoscopic lymphatic sparing varicocelectomy in
children using isosulfan blue. J Pediatr Surg, 2014. 49: 660.
https://pubmed.ncbi.nlm.nih.gov/24726132
394. Oswald, J., et al. The use of isosulphan blue to identify lymphatic vessels in high retroperitoneal
ligation of adolescent varicocele--avoiding postoperative hydrocele. BJU Int, 2001. 87: 502.
https://pubmed.ncbi.nlm.nih.gov/11298043

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 131


395. Fast, A.M., et al. Adolescent varicocelectomy: does artery sparing influence recurrence rate and/or
catch-up growth? Andrology, 2014. 2: 159.
https://pubmed.ncbi.nlm.nih.gov/24339439
396. Kim, K.S., et al. Impact of internal spermatic artery preservation during laparoscopic varicocelectomy
on recurrence and the catch-up growth rate in adolescents. J Pediatr Urol, 2014. 10: 435.
https://pubmed.ncbi.nlm.nih.gov/24314819
397. Fayad, F., et al. Percutaneous retrograde endovascular occlusion for pediatric varicocele. J Pediatr
Surg, 2011. 46: 525.
https://pubmed.ncbi.nlm.nih.gov/21376204
398. Thon, W.F., et al. Percutaneous sclerotherapy of idiopathic varicocele in childhood: a preliminary
report. J Urol, 1989. 141: 913.
https://pubmed.ncbi.nlm.nih.gov/2926889
399. Locke, J.A., et al. Treatment of varicocele in children and adolescents: A systematic review and
meta-analysis of randomized controlled trials. J Pediatr Urol, 2017. 13: 437.
https://pubmed.ncbi.nlm.nih.gov/28851509
400. Cayan, S., et al. Paternity Rates and Time to Conception in Adolescents with Varicocele Undergoing
Microsurgical Varicocele Repair vs Observation Only: A Single Institution Experience with 408
Patients. J Urol, 2017. 198: 195.
https://pubmed.ncbi.nlm.nih.gov/28153511
401. Silay, M.S., et al. Treatment of Varicocele in Children and Adolescents: A Systematic Review and
Meta-analysis from the European Association of Urology/European Society for Paediatric Urology
Guidelines Panel. Eur Urol, 2018.
https://pubmed.ncbi.nlm.nih.gov/30316583
402. Hoberman, A., et al. Prevalence of urinary tract infection in febrile infants. J Pediatr, 1993. 123: 17.
https://pubmed.ncbi.nlm.nih.gov/8320616
403. Marild, S., et al. Incidence rate of first-time symptomatic urinary tract infection in children under
6 years of age. Acta Paediatr, 1998. 87: 549.
https://pubmed.ncbi.nlm.nih.gov/9641738
404. O’Brien, K., et al. Prevalence of urinary tract infection (UTI) in sequential acutely unwell children
presenting in primary care: exploratory study. Scand J Prim Health Care, 2011. 29: 19.
https://pubmed.ncbi.nlm.nih.gov/21323495
405. Shaikh, N., et al. Prevalence of urinary tract infection in childhood: a meta-analysis. Pediatr Infect
Dis J, 2008. 27: 302.
https://pubmed.ncbi.nlm.nih.gov/18316994
406. Zorc, J.J., et al. Clinical and demographic factors associated with urinary tract infection in young
febrile infants. Pediatrics, 2005. 116: 644.
https://pubmed.ncbi.nlm.nih.gov/16140703
407. Ladomenou, F., et al. Incidence and morbidity of urinary tract infection in a prospective cohort of
children. Acta Paediatr, 2015. 104: e324.
https://pubmed.ncbi.nlm.nih.gov/25736706
408. Shaikh, N., et al. Predictors of Antimicrobial Resistance among Pathogens Causing Urinary Tract
Infection in Children. J Pediatr, 2016. 171: 116.
https://pubmed.ncbi.nlm.nih.gov/26794472
409. Zaffanello, M., et al. Management of constipation in preventing urinary tract infections in children:
A concise review. Eur Res J, 2019. 5: 236.
https://www.researchgate.net/publication/327723739
410. Grier, W.R., et al. Obesity as a Risk Factor for Urinary Tract Infection in Children. Clin Pediatr, 2016.
55: 952.
https://pubmed.ncbi.nlm.nih.gov/26810625
411. Hum S.W. et al. Risk Factors for Delayed Antimicrobial Treatment in Febrile Children with Urinary
Tract Infections. J Pediatr, 2018.
https://pubmed.ncbi.nlm.nih.gov/30340935
412. Karavanaki, K.A., et al. Delayed treatment of the first febrile urinary tract infection in early childhood
increased the risk of renal scarring. Acta Paediatr, 2017. 106: 149.
https://pubmed.ncbi.nlm.nih.gov/27748543
413. Swerkersson, S., et al. Urinary tract infection in small children: the evolution of renal damage over
time. Pediatr Nephrol, 2017. 32: 1907.
https://pubmed.ncbi.nlm.nih.gov/28681079

132 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


414. Shaikh N, et al. of Renal Scarring with Number of Febrile Urinary Tract Infections in Children. JAMA
Pediatr, 2019.
https://pubmed.ncbi.nlm.nih.gov/31381021
415. Alberici, I., et al. Pathogens causing urinary tract infections in infants: a European overview by the
ESCAPE study group. Eur J Pediatr, 2015. 174: 783.
https://pubmed.ncbi.nlm.nih.gov/25428232
416. Magin, E.C., et al. Efficacy of short-term intravenous antibiotic in neonates with urinary tract
infection. Pediatr Emerg Care, 2007. 23: 83.
https://pubmed.ncbi.nlm.nih.gov/17351406
417. Sastre, J.B., et al. Urinary tract infection in the newborn: clinical and radio imaging studies. Pediatr
Nephrol, 2007. 22: 1735.
https://pubmed.ncbi.nlm.nih.gov/17665222
418. Williams, G., et al. Long-term antibiotics for preventing recurrent urinary tract infection in children.
Cochrane Database Syst Rev, 2019. 2019: CD001534.
https://pubmed.ncbi.nlm.nih.gov/30932167
419. Yiee, J.H., et al. Prospective blinded laboratory assessment of prophylactic antibiotic compliance in
a pediatric outpatient setting. J Urol, 2012. 187: 2176.
https://pubmed.ncbi.nlm.nih.gov/22503029
420. Burns, M.W., et al. Pediatric urinary tract infection. Diagnosis, classification, and significance.
Pediatr Clin North Am, 1987. 34: 1111.
https://pubmed.ncbi.nlm.nih.gov/3658502
421. Beetz, R., et al. [Urinary tract infections in infants and children -- a consensus on diagnostic, therapy
and prophylaxis]. Urologe A, 2007. 46: 112.
https://pubmed.ncbi.nlm.nih.gov/17225140
422. Tebruegge, M., et al. The age-related risk of co-existing meningitis in children with urinary tract
infection. PLoS One, 2011. 6.
https://pubmed.ncbi.nlm.nih.gov/22096488
423. Craig, J.C., et al. The accuracy of clinical symptoms and signs for the diagnosis of serious bacterial
infection in young febrile children: prospective cohort study of 15 781 febrile illnesses. BMJ, 2010.
340: c1594.
https://pubmed.ncbi.nlm.nih.gov/20406860
424. Lin, D.S., et al. Urinary tract infection in febrile infants younger than eight weeks of Age. Pediatrics,
2000. 105: E20.
https://pubmed.ncbi.nlm.nih.gov/10654980
425. Tullus, K. Difficulties in diagnosing urinary tract infections in small children. Pediatr Nephrol, 2011.
26: 1923.
https://pubmed.ncbi.nlm.nih.gov/21773821
426. Kauffman, J.D., et al. Risk Factors and Associated Morbidity of Urinary Tract Infections in Pediatric
Surgical Patients: a NSQIP Pediatric Analysis. J Pediatric Surg, 2019.
https://pubmed.ncbi.nlm.nih.gov/31126686
427. Whiting, P., et al. Rapid tests and urine sampling techniques for the diagnosis of urinary tract
infection (UTI) in children under five years: a systematic review. BMC Pediatr, 2005. 5: 4.
https://pubmed.ncbi.nlm.nih.gov/15811182
428. Ramage, I.J., et al. Accuracy of clean-catch urine collection in infancy. J Pediatr, 1999. 135: 765.
https://pubmed.ncbi.nlm.nih.gov/10586183
429. Altuntas, N., et al. Midstream clean-catch urine culture obtained by stimulation technique versus
catheter-specimen urine culture for urinary tract infections in newborns: a paired comparison of
urine collection methods. Med Princ Pract, 2019.
https://pubmed.ncbi.nlm.nih.gov/31665720
430. Roberts, K.B., et al. Urinary tract infection: clinical practice guideline for the diagnosis and
management of the initial UTI in febrile infants and children 2 to 24 months. Pediatrics, 2011.
128: 595.
https://pubmed.ncbi.nlm.nih.gov/21873693
431. Tosif, S., et al. Contamination rates of different urine collection methods for the diagnosis of urinary
tract infections in young children: an observational cohort study. J Paediatr Child Health, 2012.
48: 659.
https://pubmed.ncbi.nlm.nih.gov/22537082
432. Labrosse, M., et al. Evaluation of a New Strategy for Clean-Catch Urine in Infants. Pediatrics, 2016.
138.
https://pubmed.ncbi.nlm.nih.gov/27542848

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 133


433. Buys, H., et al. Suprapubic aspiration under ultrasound guidance in children with fever of
undiagnosed cause. BMJ, 1994. 308: 690.
https://pubmed.ncbi.nlm.nih.gov/8142792
434. Kiernan, S.C., et al. Ultrasound guidance of suprapubic bladder aspiration in neonates. J Pediatr,
1993. 123: 789.
https://pubmed.ncbi.nlm.nih.gov/8229492
435. Vaillancourt, S., et al. To clean or not to clean: effect on contamination rates in midstream urine
collections in toilet-trained children. Pediatrics, 2007. 119: e1288.
https://pubmed.ncbi.nlm.nih.gov/17502345
436. Powell, H.R., et al. Urinary nitrite in symptomatic and asymptomatic urinary infection. Arch Dis Child,
1987. 62: 138.
https://pubmed.ncbi.nlm.nih.gov/3548604
437. Coulthard, M.G. Using urine nitrite sticks to test for urinary tract infection in children aged < 2 years:
a meta-analysis. Pediatr Nephrol, 2019.
https://pubmed.ncbi.nlm.nih.gov/30895368
438. Mori, R., et al. Diagnostic performance of urine dipstick testing in children with suspected UTI: a
systematic review of relationship with age and comparison with microscopy. Acta Paediatr, 2010.
99: 581.
https://pubmed.ncbi.nlm.nih.gov/20055779
439. Herreros, M.L., et al. Performing a urine dipstick test with a clean-catch urine sample is an accurate
screening method for urinary tract infections in young infants. Acta Paediatr, 2018. 107: 145.
https://pubmed.ncbi.nlm.nih.gov/28940750
440. Hildebrand, W.L., et al. Suprapubic bladder aspiration in infants. Am Fam Physician, 1981. 23: 115.
https://pubmed.ncbi.nlm.nih.gov/7234629
441. Hoberman, A., et al. Is urine culture necessary to rule out urinary tract infection in young febrile
children? Pediatr Infect Dis J, 1996. 15: 304.
https://pubmed.ncbi.nlm.nih.gov/8866798
442. Herr, S.M., et al. Enhanced urinalysis improves identification of febrile infants ages 60 days and
younger at low risk for serious bacterial illness. Pediatrics, 2001. 108: 866.
https://pubmed.ncbi.nlm.nih.gov/11581437
443. Williams, G.J., et al. Absolute and relative accuracy of rapid urine tests for urinary tract infection in
children: a meta-analysis. Lancet Infect Dis, 2010. 10: 240.
https://pubmed.ncbi.nlm.nih.gov/20334847
444. Mayo, S., et al. Clinical laboratory automated urinalysis: comparison among automated microscopy,
flow cytometry, two test strips analyzers, and manual microscopic examination of the urine
sediments. J Clin Lab Anal, 2008. 22: 262.
https://pubmed.ncbi.nlm.nih.gov/18623125
445. Broeren, M., et al. Urine flow cytometry is an adequate screening tool for urinary tract infections in
children. Eur J Pediatr, 2019. 178: 363.
https://pubmed.ncbi.nlm.nih.gov/30569406
446. Akagawa, Y., et al. Optimal bacterial colony counts for the diagnosis of upper urinary tract infections
in infants. Clin Exp Nephrol, 2019.
https://pubmed.ncbi.nlm.nih.gov/31712943
447. Broadis, E., et al. ‘Targeted top down’ approach for the investigation of UTI: A 10-year follow-up
study in a cohort of 1000 children. J Pediatr Surg, 2016. 12: 39.
https://pubmed.ncbi.nlm.nih.gov/29370630
448. Shaikh, N.A., et al. Is ultrasound detect renal infections? Med Forum Monthly, 2016. 27: 16.
https://www.researchgate.net/publication/316504667
449. Chang, S.J., et al. Elevated postvoid residual urine volume predicting recurrence of urinary tract
infections in toilet-trained children. Pediatr Nephrol, 2015. 30: 1131.
https://pubmed.ncbi.nlm.nih.gov/25673516
450. Stoica, I., et al. Xanthogranulomatous pyelonephritis in a paediatric cohort (1963-2016): Outcomes
from a large single-center series. J Pediatr Surg, 2018. 14: 169.
https://pubmed.ncbi.nlm.nih.gov/29233628
451. Shiraishi, K., et al. Risk factors for breakthrough infection in children with primary vesicoureteral
reflux. J Urol, 2010. 183: 1527.
https://pubmed.ncbi.nlm.nih.gov/20172558
452. Quirino, I.G., et al. Combined use of late phase dimercapto-succinic acid renal scintigraphy and
ultrasound as first line screening after urinary tract infection in children. J Urol, 2011. 185: 258.
https://pubmed.ncbi.nlm.nih.gov/21074813

134 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


453. Bosakova, A., et al. Diffusion-weighted magnetic resonance imaging is more sensitive than
dimercaptosuccinic acid scintigraphy in detecting parenchymal lesions in children with acute
pyelonephritis: A prospective study. J Pediatr Surg, 2018. 14: 269.
https://pubmed.ncbi.nlm.nih.gov/29588142
454. Michaud, J.E., et al. Cost and radiation exposure in the workup of febrile pediatric urinary tract
infections. J Surg Res, 2016. 203: 313.
https://pubmed.ncbi.nlm.nih.gov/27363638
455. Siomou, E., et al. Implications of 99mTc-DMSA scintigraphy performed during urinary tract infection
in neonates. Pediatrics, 2009. 124: 881.
https://pubmed.ncbi.nlm.nih.gov/19661052
456. Shaikh, N., et al. Dimercaptosuccinic acid scan or ultrasound in screening for vesicoureteral reflux
among children with urinary tract infections. Cochrane Database Syst Rev, 2016. 2016: CD010657.
https://pubmed.ncbi.nlm.nih.gov/27378557
457. Mazzi, S., et al. Timing of voiding cystourethrography after febrile urinary tract infection in children:
A systematic review. Arch Dis Child, 2019. 105: 264.
https://pubmed.ncbi.nlm.nih.gov/31466991
458. Spencer, J.D., et al. The accuracy and health risks of a voiding cystourethrogram after a febrile
urinary tract infection. J Pediatr Urol, 2012. 8: 72.
https://pubmed.ncbi.nlm.nih.gov/21126919
459. Ntoulia, A., et al. Contrast-enhanced voiding urosonography (ceVUS) with the intravesical
administration of the ultrasound contrast agent Optison for vesicoureteral reflux detection in
children: a prospective clinical trial. Pediatr Radiol, 2018. 48: 216.
https://pubmed.ncbi.nlm.nih.gov/29181582
460. Lee, L.C., et al. The role of voiding cystourethrography in the investigation of children with urinary
tract infections. Can Urol Assoc J, 2016. 10: 210.
https://pubmed.ncbi.nlm.nih.gov/27713802
461. Pauchard, J.-Y., et al. Avoidance of voiding cystourethrography in infants younger than 3 months with
Escherichia coli urinary tract infection and normal renal ultrasound. Arch Dis Child, 2017. 102: 804.
https://pubmed.ncbi.nlm.nih.gov/28408468
462. Shaikh, N., et al. Identification of children and adolescents at risk for renal scarring after a first
urinary tract infection: a meta-analysis with individual patient data. JAMA Pediatr, 2014. 168: 893.
https://pubmed.ncbi.nlm.nih.gov/25089634
463. Rianthavorn, P., et al. Probabilities of Dilating Vesicoureteral Reflux in Children with First Time
Simple Febrile Urinary Tract Infection, and Normal Renal and Bladder Ultrasound. J Urol, 2016.
196: 1541.
https://pubmed.ncbi.nlm.nih.gov/27181502
464. Bahat, H., et al. Predictors of grade 3-5 vesicoureteral reflux in infants <= 2 months of age with
pyelonephritis. Pediatr Nephrol, 2019. 34: 907.
https://pubmed.ncbi.nlm.nih.gov/30588547
465. Mola, G., et al. Selective imaging modalities after first pyelonephritis failed to identify significant
urological anomalies, despite normal antenatal ultrasounds. Acta Paediatr, 2017. 106: 1176.
https://pubmed.ncbi.nlm.nih.gov/28437563
466. Nandagopal, R., et al. Transient Pseudohypoaldosteronism due to Urinary Tract Infection in Infancy:
A Report of 4 Cases. Int J Pediatr Endocrinol, 2009. 2009: 195728.
https://pubmed.ncbi.nlm.nih.gov/19946403
467. Tutunculer, F., et al. Transient Pseudohypoaldosteronism in an infant with urinary tract anomaly.
Pediatr Int, 2004. 46: 618.
https://pubmed.ncbi.nlm.nih.gov/15491397
468. Robinson, J.L., et al. Management of urinary tract infections in children in an era of increasing
antimicrobial resistance. Exp Rev Anti-Infect Ther, 2016. 14: 809.
https://pubmed.ncbi.nlm.nih.gov/27348347
469. Contopoulos-Ioannidis, D.G., et al. Extended-interval aminoglycoside administration for children: a
meta-analysis. Pediatrics, 2004. 114: e111.
https://pubmed.ncbi.nlm.nih.gov/15231982
470. Dore-Bergeron, M.J., et al. Urinary tract infections in 1- to 3-month-old infants: ambulatory
treatment with intravenous antibiotics. Pediatrics, 2009. 124: 16.
https://pubmed.ncbi.nlm.nih.gov/19564278
471. Gauthier, M., et al. Treatment of urinary tract infections among febrile young children with daily
intravenous antibiotic therapy at a day treatment center. Pediatrics, 2004. 114: e469.
https://pubmed.ncbi.nlm.nih.gov/15466073

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 135


472. Shaikh, N., et al. Early Antibiotic Treatment for Pediatric Febrile Urinary Tract Infection and Renal
Scarring. JAMA Pediatr, 2016. 170: 848.
https://pubmed.ncbi.nlm.nih.gov/27455161
473. Roman, H.K., et al. Diagnosis and management of bacteremic urinary tract infection in infants. Hosp
Pediatr, 2015. 5: 1.
https://pubmed.ncbi.nlm.nih.gov/25554753
474. Bouissou, F., et al. Prospective, randomized trial comparing short and long intravenous antibiotic
treatment of acute pyelonephritis in children: dimercaptosuccinic acid scintigraphic evaluation at 9
months. Pediatrics, 2008. 121: e553.
https://pubmed.ncbi.nlm.nih.gov/18267977
475. Desai, S., et al. Parenteral antibiotic therapy duration in young infants with bacteremic urinary tract
infections. Pediatrics, 2019. 144: e20183844.
https://pubmed.ncbi.nlm.nih.gov/31431480
476. Hodson, E.M., et al. Antibiotics for acute pyelonephritis in children. Cochrane Database Syst Rev,
2007: Cd003772.
https://pubmed.ncbi.nlm.nih.gov/17943796
477. Craig, J.C., et al. Antibiotic prophylaxis and recurrent urinary tract infection in children. N Engl
J Med, 2009. 361: 1748.
https://pubmed.ncbi.nlm.nih.gov/19864673
478. Neuhaus, T.J., et al. Randomised trial of oral versus sequential intravenous/oral cephalosporins in
children with pyelonephritis. Eur J Pediatr, 2008. 167: 1037.
https://pubmed.ncbi.nlm.nih.gov/18074149
479. Hoberman, A., et al. Oral versus initial intravenous therapy for urinary tract infections in young febrile
children. Pediatrics, 1999. 104: 79.
https://pubmed.ncbi.nlm.nih.gov/10390264
480. Salomonsson, P., et al. Best oral empirical treatment for pyelonephritis in children: Do we need to
differentiate between age and gender? Infect Dis (Lond), 2016. 48: 721.
https://pubmed.ncbi.nlm.nih.gov/27300266
481. Mak, R.H., et al. Are oral antibiotics alone efficacious for the treatment of a first episode of acute
pyelonephritis in children? Nat Clin Pract Nephrol, 2008. 4: 10.
https://pubmed.ncbi.nlm.nih.gov/17971799
482. Janett, S., et al. Pyuria and microbiology in acute bacterial focal nephritis: A systematic review.
Minerva Med, 2019. 110: 232.
https://pubmed.ncbi.nlm.nih.gov/30809996
483. Bryce, A., et al. Global prevalence of antibiotic resistance in paediatric urinary tract infections
caused by Escherichia coli and association with routine use of antibiotics in primary care:
Systematic review and meta-analysis. BMJ (Online), 2016. 352: i939.
https://pubmed.ncbi.nlm.nih.gov/26980184
484. Flokas, M.E., et al. Prevalence of ESBL-producing Enterobacteriaceae in paediatric urinary tract
infections: A systematic review and meta-analysis. J Infect, 2016. 73: 547.
https://pubmed.ncbi.nlm.nih.gov/27475789
485. Fostira, E., et al. Short-term antibiotic exposure affected the type and resistance of uropathogens
similar to long-term antibiotic prophylaxis in children hospitalised for urinary tract infections. Acta
Paediatr, 2019.
https://pubmed.ncbi.nlm.nih.gov/31746494
486. Uyar Aksu, N., et al. Childhood urinary tract infection caused by extended-spectrum beta-
lactamase-producing bacteria: Risk factors and empiric therapy. Pediatr Int, 2017. 59: 176.
https://pubmed.ncbi.nlm.nih.gov/27501161
487. Kara, A., et al. The use of nitrofurantoin for children with acute cystitis caused by extended-
spectrum B-lactamase-producing Escherichia coli. J Pediatr Surg, 2019. 15: 378.
https://pubmed.ncbi.nlm.nih.gov/31014984
488. Alsubaie, S.S., et al. Current status of long-term antibiotic prophylaxis for urinary tract infections in
children: An antibiotic stewardship challenge. Kidney Res Clin Pract, 2019. 38: 441.
https://pubmed.ncbi.nlm.nih.gov/31739385
489. Keren, R., et al. Risk factors for recurrent urinary tract infection and renal scarring. Pediatrics, 2015.
136: e13.
https://pubmed.ncbi.nlm.nih.gov/26055855
490. Shaikh, N., et al. Recurrent Urinary Tract Infections in Children With Bladder and Bowel Dysfunction.
Pediatrics, 2016. 137.
https://pubmed.ncbi.nlm.nih.gov/26647376

136 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


491. Dray, E.V., et al. Recurrent urinary tract infections in patients with incomplete bladder emptying: is
there a role for intravesical therapy? Transl Androl Urol, 2017. 6: S163.
https://pubmed.ncbi.nlm.nih.gov/28791235
492. Naber, K.G., et al., EAU/International Consultation on Urological Infections 2010, European
Association of Urology: The Netherlands.
493. Durham, S.H., et al. Cranberry Products for the Prophylaxis of Urinary Tract Infections in Pediatric
Patients. Ann Pharmacother, 2015. 49: 1349.
https://pubmed.ncbi.nlm.nih.gov/26400007
494. Sadeghi-Bojd, S., et al. Efficacy of Probiotic Prophylaxis After The First Febrile Urinary Tract
Infection in Children With Normal Urinary Tracts. J Pediatr Infect Dis Soc, 2019.
https://pubmed.ncbi.nlm.nih.gov/31100124
495. Hosseini, M., et al. The efficacy of probiotics in prevention of urinary tract infection in children: A
systematic review and meta-analysis. J Pediatr Surg, 2017. 13: 581.
https://pubmed.ncbi.nlm.nih.gov/29102297
496. Kahbazi, M., et al. Vitamin A supplementation is effective for improving the clinical symptoms of
urinary tract infections and reducing renal scarring in girls with acute pyelonephritis: a randomized,
double-blind placebo-controlled, clinical trial study. Complement Ther Med, 2019. 42: 429.
https://pubmed.ncbi.nlm.nih.gov/30670279
497. Zhang, G.Q., et al. The effect of vitamin A on renal damage following acute pyelonephritis in
children: a meta-analysis of randomized controlled trials. Pediatr Nephrol, 2016. 31: 373.
https://pubmed.ncbi.nlm.nih.gov/25980468
498. Yousefichaijan, P., et al. Vitamin E as adjuvant treatment for urinary tract infection in girls with acute
pyelonephritis. Iranian J Kidney Dis, 2015. 9: 97.
https://pubmed.ncbi.nlm.nih.gov/25851287
499. Chen, C.J., et al. The use of steroid cream for physiologic phimosis in male infants with a history of
UTI and normal renal ultrasound is associated with decreased risk of recurrent UTI. J Pediatr Surg,
2019. 15: 472.
https://pubmed.ncbi.nlm.nih.gov/31345734
500. Kotoula, A., et al. Comparative efficacies of procalcitonin and conventional inflammatory markers for
prediction of renal parenchymal inflammation in pediatric first urinary tract infection. Urology, 2009.
73: 782.
https://pubmed.ncbi.nlm.nih.gov/19152962
501. Zhang, H., et al. Diagnostic value of serum procalcitonin for acute pyelonephritis in infants and
children with urinary tract infections: an updated meta-analysis. World J Urol, 2016. 34: 431.
https://pubmed.ncbi.nlm.nih.gov/26142087
502. Austin, P.F., et al. The standardization of terminology of lower urinary tract function in children and
adolescents: Update report from the standardization committee of the International Children’s
Continence Society. Neurourol Urodyn, 2016. 35: 471.
https://pubmed.ncbi.nlm.nih.gov/25772695
503. Bakker, E., et al. Voiding habits and wetting in a population of 4,332 Belgian schoolchildren aged
between 10 and 14 years. Scand J Urol Nephrol, 2002. 36: 354.
https://pubmed.ncbi.nlm.nih.gov/12487740
504. Hellstrom, A.L., et al. Micturition habits and incontinence in 7-year-old Swedish school entrants. Eur
J Pediatr, 1990. 149: 434.
https://pubmed.ncbi.nlm.nih.gov/2332015
505. Soderstrom, U., et al. Urinary and faecal incontinence: a population-based study. Acta Paediatr,
2004. 93: 386.
https://pubmed.ncbi.nlm.nih.gov/15124844
506. Sureshkumar, P., et al. A population based study of 2,856 school-age children with urinary
incontinence. J Urol, 2009. 181: 808.
https://pubmed.ncbi.nlm.nih.gov/19110268
507. Bloom, D.A., et al. Toilet habits and continence in children: an opportunity sampling in search of
normal parameters. J Urol, 1993. 149: 1087.
https://pubmed.ncbi.nlm.nih.gov/8483218
508. Bower, W.F., et al. The epidemiology of childhood enuresis in Australia. Br J Urol, 1996. 78: 602.
https://pubmed.ncbi.nlm.nih.gov/8944518
509. Mattsson, S. Urinary incontinence and nocturia in healthy schoolchildren. Acta Paediatr, 1994.
83: 950.
https://pubmed.ncbi.nlm.nih.gov/7819693

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 137


510. Sureshkumar, P., et al. Daytime urinary incontinence in primary school children: a population-based
survey. J Pediatr, 2000. 137: 814.
https://pubmed.ncbi.nlm.nih.gov/11113838
511. Vaz, G.T., et al. Prevalence of lower urinary tract symptoms in school-age children. Pediatr Nephrol,
2012. 27: 597.
https://pubmed.ncbi.nlm.nih.gov/21969094
512. Borch, L., et al. Bladder and bowel dysfunction and the resolution of urinary incontinence with
successful management of bowel symptoms in children. Acta Paediatr, 2013. 102: e215.
https://pubmed.ncbi.nlm.nih.gov/23368903
513. Veiga, M.L., et al. Constipation in children with isolated overactive bladders. J Pediatr Urol, 2013.
9: 945.
https://pubmed.ncbi.nlm.nih.gov/23462384
514. Franco, I. Overactive bladder in children. Part 1: Pathophysiology. J Urol, 2007. 178: 761.
https://pubmed.ncbi.nlm.nih.gov/17631323
515. Niemczyk, J., et al. Incontinence in children with treated attention-deficit/hyperactivity disorder.
J Pediatr Urol, 2015. 11: 141.e1.
https://pubmed.ncbi.nlm.nih.gov/25863677
516. Chang, S.J., et al. Treatment of daytime urinary incontinence: A standardization document from the
International Children’s Continence Society. Neurourol Urodyn, 2015.
https://pubmed.ncbi.nlm.nih.gov/26473630
517. Hoebeke, P., et al. Diagnostic evaluation of children with daytime incontinence. J Urol, 2010. 183: 699.
https://pubmed.ncbi.nlm.nih.gov/20022025
518. Hjalmas, K., et al. Lower urinary tract dysfunction and urodynamics in children. Eur Urol, 2000. 38: 655.
https://pubmed.ncbi.nlm.nih.gov/11096254
519. Ambartsumyan, L., et al. Simultaneous urodynamic and anorectal manometry studies in children:
insights into the relationship between the lower gastrointestinal and lower urinary tracts.
Neurogastroenterol Motil, 2016. 28: 924.
https://pubmed.ncbi.nlm.nih.gov/27214097
520. van Summeren, J., et al. Bladder Symptoms in Children With Functional Constipation: A Systematic
Review. J Pediatr Gastroenterol Nutr, 2018. 67: 552.
https://pubmed.ncbi.nlm.nih.gov/30212423
521. Chen, J.J., et al. Infant vesicoureteral reflux: a comparison between patients presenting with a
prenatal diagnosis and those presenting with a urinary tract infection. Urology, 2003. 61: 442.
https://pubmed.ncbi.nlm.nih.gov/12597964
522. Bauer, S.B., et al. International Children’s Continence Society standardization report on urodynamic
studies of the lower urinary tract in children. Neurourol Urodyn, 2015. 34: 640.
https://pubmed.ncbi.nlm.nih.gov/25998310
523. Parekh, D.J., et al. The use of radiography, urodynamic studies and cystoscopy in the evaluation of
voiding dysfunction. J Urol, 2001. 165: 215.
https://pubmed.ncbi.nlm.nih.gov/11125409
524. Pfister, C., et al. The usefulness of a minimal urodynamic evaluation and pelvic floor biofeedback in
children with chronic voiding dysfunction. BJU Int, 1999. 84: 1054.
https://pubmed.ncbi.nlm.nih.gov/10571635
525. Schewe, J., et al. Voiding dysfunction in children: role of urodynamic studies. Urol Int, 2002. 69: 297.
https://pubmed.ncbi.nlm.nih.gov/12444287
526. Akbal, C., et al. Dysfunctional voiding and incontinence scoring system: quantitative evaluation of
incontinence symptoms in pediatric population. J Urol, 2005. 173: 969.
https://pubmed.ncbi.nlm.nih.gov/15711352
527. Farhat, W., et al. The dysfunctional voiding scoring system: quantitative standardization of
dysfunctional voiding symptoms in children. J Urol, 2000. 164: 1011.
https://pubmed.ncbi.nlm.nih.gov/10958730
528. Burgers, R.E., et al. Management of functional constipation in children with lower urinary tract
symptoms: report from the Standardization Committee of the International Children’s Continence
Society. J Urol, 2013. 190: 29.
https://pubmed.ncbi.nlm.nih.gov/23313210
529. Chang, S.J., et al. Constipation is associated with incomplete bladder emptying in healthy children.
Neurourol Urodyn, 2012. 31: 105.
https://pubmed.ncbi.nlm.nih.gov/22038844

138 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


530. Neveus, T., et al. The standardization of terminology of lower urinary tract function in children and
adolescents: report from the Standardisation Committee of the International Children’s Continence
Society. J Urol, 2006. 176: 314.
https://pubmed.ncbi.nlm.nih.gov/16753432
531. Yang, S.S., et al. Home uroflowmetry for the evaluation of boys with urinary incontinence. J Urol,
2003. 169: 1505.
https://pubmed.ncbi.nlm.nih.gov/12629404
532. van Gool, J.D., et al. Multi-center randomized controlled trial of cognitive treatment, placebo,
oxybutynin, bladder training, and pelvic floor training in children with functional urinary incontinence.
Neurourol Urodyn, 2014. 33: 482.
https://pubmed.ncbi.nlm.nih.gov/23775924
533. Campos, R.M., et al. Comparative, prospective, and randomized study between urotherapy and the
pharmacological treatment of children with urinary incontinence. Einstein (Sao Paulo), 2013. 11: 203.
https://pubmed.ncbi.nlm.nih.gov/23843062
534. Buckley, B.S., et al. Conservative interventions for treating functional daytime urinary incontinence in
children. Cochrane Database Syst Rev, 2019. 9: Cd012367.
https://pubmed.ncbi.nlm.nih.gov/31532563
535. Hoebeke, P., et al. Assessment of lower urinary tract dysfunction in children with non-neuropathic
bladder sphincter dysfunction. Eur Urol, 1999. 35: 57.
https://pubmed.ncbi.nlm.nih.gov/9933796
536. Barroso, U., Jr., et al. Electrical stimulation for lower urinary tract dysfunction in children: a
systematic review of the literature. Neurourol Urodyn, 2011. 30: 1429.
https://pubmed.ncbi.nlm.nih.gov/21717502
537. Bower, W.F., et al. A review of non-invasive electro neuromodulation as an intervention for non-
neurogenic bladder dysfunction in children. Neurourol Urodyn, 2004. 23: 63.
https://pubmed.ncbi.nlm.nih.gov/14694460
538. De Paepe, H., et al. Pelvic-floor therapy in girls with recurrent urinary tract infections and
dysfunctional voiding. Br J Urol, 1998. 81 Suppl 3: 109.
https://pubmed.ncbi.nlm.nih.gov/9634033
539. Hellstrom, A.L. Urotherapy in children with dysfunctional bladder. Scand J Urol Nephrol Suppl, 1992.
141: 106.
https://pubmed.ncbi.nlm.nih.gov/1609245
540. Lordelo, P., et al. Prospective study of transcutaneous parasacral electrical stimulation for overactive
bladder in children: long-term results. J Urol, 2009. 182: 2900.
https://pubmed.ncbi.nlm.nih.gov/19846164
541. Vijverberg, M.A., et al. Bladder rehabilitation, the effect of a cognitive training programme on urge
incontinence. Eur Urol, 1997. 31: 68.
https://pubmed.ncbi.nlm.nih.gov/9032538
542. Desantis, D.J., et al. Effectiveness of biofeedback for dysfunctional elimination syndrome in
pediatrics: a systematic review. J Pediatr Urol, 2011. 7: 342.
https://pubmed.ncbi.nlm.nih.gov/21527216
543. Kajbafzadeh, A.M., et al. Transcutaneous interferential electrical stimulation for the management of
non-neuropathic underactive bladder in children: a randomised clinical trial. BJU Int, 2016. 117: 793.
https://pubmed.ncbi.nlm.nih.gov/26086897
544. Ladi-Seyedian, S., et al. Management of non-neuropathic underactive bladder in children with
voiding dysfunction by animated biofeedback: a randomized clinical trial. Urology, 2015. 85: 205.
https://pubmed.ncbi.nlm.nih.gov/25444633
545. Featherstone, N., et al. Ephedrine hydrochloride: novel use in the management of resistant non-
neurogenic daytime urinary incontinence in children. J Pediatr Urol, 2013. 9: 915.
https://pubmed.ncbi.nlm.nih.gov/23332206
546. Nijman, R.J., et al. Tolterodine treatment for children with symptoms of urinary urge incontinence
suggestive of detrusor overactivity: results from 2 randomized, placebo controlled trials. J Urol,
2005. 173: 1334.
https://pubmed.ncbi.nlm.nih.gov/15758796
547. Marschall-Kehrel, D., et al. Treatment with propiverine in children suffering from nonneurogenic
overactive bladder and urinary incontinence: results of a randomized placebo-controlled phase 3
clinical trial. Eur Urol, 2009. 55: 729.
https://pubmed.ncbi.nlm.nih.gov/18502028

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 139


548. Newgreen, D., et al. Long-Term Safety and Efficacy of Solifenacin in Children and Adolescents with
Overactive Bladder. J Urol, 2017. 198: 928.
https://pubmed.ncbi.nlm.nih.gov/28506854
549. Kramer, S.A., et al. Double-blind placebo controlled study of alpha-adrenergic receptor antagonists
(doxazosin) for treatment of voiding dysfunction in the pediatric population. J Urol, 2005. 173: 2121.
https://pubmed.ncbi.nlm.nih.gov/15879863
550. Hoebeke, P., et al. The effect of botulinum-A toxin in incontinent children with therapy resistant
overactive detrusor. J Urol, 2006. 176: 328.
https://pubmed.ncbi.nlm.nih.gov/16753434
551. Fernandez, N., et al. Neurostimulation Therapy for Non-neurogenic Overactive Bladder in Children:
A Meta-analysis. Urology, 2017. 110: 201.
https://pubmed.ncbi.nlm.nih.gov/28823638
552. Groen, L.A., et al. Sacral neuromodulation with an implantable pulse generator in children with lower
urinary tract symptoms: 15-year experience. J Urol, 2012. 188: 1313.
https://pubmed.ncbi.nlm.nih.gov/22902022
553. Beksac, A.T., et al. Postvoidal residual urine is the most significant non-invasive diagnostic test
to predict the treatment outcome in children with non-neurogenic lower urinary tract dysfunction.
J Pediatr Urol, 2016. 12: 215.e1.
https://pubmed.ncbi.nlm.nih.gov/27233211
554. Bower, W.F., et al. The transition of young adults with lifelong urological needs from pediatric to
adult services: An international children’s continence society position statement. Neurourol Urodyn,
2017. 36: 811.
https://pubmed.ncbi.nlm.nih.gov/27177245
555. Lackgren, G., et al. Nocturnal enuresis: a suggestion for a European treatment strategy. Acta
Paediatr, 1999. 88: 679.
https://pubmed.ncbi.nlm.nih.gov/10419258
556. Neveus, T., et al. Enuresis--background and treatment. Scand J Urol Nephrol Suppl, 2000: 1.
https://pubmed.ncbi.nlm.nih.gov/11196246
557. Negoro, H., et al. Chronobiology of micturition: putative role of the circadian clock. J Urol, 2013.
190: 843.
https://pubmed.ncbi.nlm.nih.gov/23429068
558. Hjalmas, K., et al. Nocturnal enuresis: an international evidence based management strategy. J Urol,
2004. 171: 2545.
https://pubmed.ncbi.nlm.nih.gov/15118418
559. Caldwell, P.H., et al. Simple behavioural interventions for nocturnal enuresis in children. Cochrane
Database Syst Rev, 2013. 7: Cd003637.
https://pubmed.ncbi.nlm.nih.gov/23881652
560. Glazener, C.M., et al. Alarm interventions for nocturnal enuresis in children. Cochrane Database Syst
Rev, 2005: Cd002911.
https://pubmed.ncbi.nlm.nih.gov/15846643
561. Dehoorne, J.L., et al. Desmopressin toxicity due to prolonged half-life in 18 patients with nocturnal
enuresis. J Urol, 2006. 176: 754.
https://pubmed.ncbi.nlm.nih.gov/16813936
562. Glazener, C.M., et al. Desmopressin for nocturnal enuresis in children. Cochrane Database Syst Rev,
2002: Cd002112.
https://pubmed.ncbi.nlm.nih.gov/12137645
563. Gokce, M.I., et al. Does structured withdrawal of desmopressin improve relapse rates in patients
with monosymptomatic enuresis? J Urol, 2014. 192: 530.
https://pubmed.ncbi.nlm.nih.gov/24518770
564. Glazener, C.M., et al. Tricyclic and related drugs for nocturnal enuresis in children. Cochrane
Database Syst Rev, 2003: Cd002117.
https://pubmed.ncbi.nlm.nih.gov/12917922
565. Snow-Lisy, D.C., et al. Update on Urological Management of Spina Bifida from Prenatal Diagnosis to
Adulthood. J Urol, 2015. 194: 288.
https://pubmed.ncbi.nlm.nih.gov/25839383
566. Lee, B., et al. British Association of Paediatric Urologists consensus statement on the management
of the neuropathic bladder. J Pediatr Surg, 2016. 12: 76.
https://pubmed.ncbi.nlm.nih.gov/26946946

140 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


567. Kessler, T.M., et al. Early proactive management improves upper urinary tract function and reduces
the need for surgery in patients with myelomeningocele. Neurourol Urodyn, 2006. 25: 758.
https://pubmed.ncbi.nlm.nih.gov/16986135
568. Rendeli, C., et al. Latex sensitisation and allergy in children with myelomeningocele. Childs Nerv
Syst, 2005.
https://pubmed.ncbi.nlm.nih.gov/15703967
569. Bauer, S.B. Neurogenic bladder: Etiology and assessment. Pediatr Nephrol, 2008. 23: 541.
https://pubmed.ncbi.nlm.nih.gov/50062986
570. Tarcan, T., et al. Long-term followup of newborns with myelodysplasia and normal urodynamic
findings: Is followup necessary? J Urol, 2001. 165: 564.
https://pubmed.ncbi.nlm.nih.gov/11176436
571. McGuire, E.J., et al. Upper urinary tract deterioration in patients with myelodysplasia and detrusor
hypertonia: a followup study. J Urol, 1983. 129: 823.
https://pubmed.ncbi.nlm.nih.gov/6842712
572. Hopps, C.V., et al. Preservation of renal function in children with myelomeningocele managed with
basic newborn evaluation and close followup. J Urol, 2003. 169: 305.
https://pubmed.ncbi.nlm.nih.gov/12478177
573. Bauer, S. Clean intermittent catheterization of infants with myelodysplasia - the argument for
early asssessment and treatment of infants with spina bifida. Dialog Pediatr Urol, 2000. 23: 2. [No
abstract available].
574. Sillen, U., et al. Development of the urodynamic pattern in infants with myelomeningocele. Br J Urol,
1996. 78: 596.
https://pubmed.ncbi.nlm.nih.gov/8944517
575. Thorup, J., et al. Urological outcome after myelomeningocele: 20 years of follow-up. BJU Int, 2011.
107: 994.
https://pubmed.ncbi.nlm.nih.gov/20860652
576. Veenboer, P.W., et al. Upper and Lower Urinary Tract Outcomes in Adult Myelomeningocele Patients:
A Systematic Review. PLoS ONE, 2012. 7: e48399.
https://pubmed.ncbi.nlm.nih.gov/23119003
577. Lloyd, J.C., et al. Reviewing definitions of urinary continence in the contemporary spina bifida
literature: A call for clarity. J Pediatr Surg, 2013. 9: 567.
https://pubmed.ncbi.nlm.nih.gov/52491875
578. Khoshnood, B., et al. Long term trends in prevalence of neural tube defects in Europe: population
based study. BMJ, 2015. 351: h5949.
https://pubmed.ncbi.nlm.nih.gov/26601850
579. Adzick, N.S., et al. A randomized trial of prenatal versus postnatal repair of myelomeningocele.
N Engl J Med, 2011. 364: 993.
https://pubmed.ncbi.nlm.nih.gov/21306277
580. Brock, J.W., 3rd, et al. Bladder Function After Fetal Surgery for Myelomeningocele. Pediatrics, 2015.
136: e906.
https://pubmed.ncbi.nlm.nih.gov/26416930
581. Torre, M., et al. Long-term urologic outcome in patients with caudal regression syndrome, compared
with meningomyelocele and spinal cord lipoma. J Pediatric Surg, 2008. 43: 530.
https://pubmed.ncbi.nlm.nih.gov/18358295
582. Maerzheuser, S., et al. German network for congenital uro-rectal malformations: first evaluation and
interpretation of postoperative urological complications in anorectal malformations. Pediatr Surg Int,
2011. 27: 1085.
https://pubmed.ncbi.nlm.nih.gov/21792651
583. Hinman, F., et al. Vesical and Ureteral Damage from Voiding Dysfunction in Boys Without Neurologic
or Obstructive Disease. J Urol, 2017. 197: S127.
https://pubmed.ncbi.nlm.nih.gov/4695119
584. Ochoa, B. Can a congenital dysfunctional bladder be diagnosed from a smile? The Ochoa
syndrome updated. Pediatr Nephrol, 2004. 19: 6.
https://pubmed.ncbi.nlm.nih.gov/14648341
585. Drzewiecki, B.A., et al. Urodynamic testing in children: Indications, technique, interpretation and
significance. J Urol, 2011. 186: 1190.
https://pubmed.ncbi.nlm.nih.gov/51573934
586. Bauer, S.B., et al. Predictive value of urodynamic evaluation in newborns with myelodysplasia.
Jama, 1984. 252: 650.
https://pubmed.ncbi.nlm.nih.gov/6737668

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 141


587. Madersbacher, H. The various types of neurogenic bladder dysfunction: an update of current
therapeutic concepts. Paraplegia, 1990. 28: 217.
https://pubmed.ncbi.nlm.nih.gov/2235029
588. Wide, P., et al. Renal preservation in children with neurogenic bladder-sphincter dysfunction
followed in a national program. J Pediatr Surg, 2012. 8: 187.
https://pubmed.ncbi.nlm.nih.gov/51318737
589. Stein, R., et al. (2013) S2k Leitlinie AWMF Register 043/047: Diagnostik und Therapie der
neurogenen Blasenfunktionsstörungen bei Patienten mit Meningomyelocele.
https://link.springer.com/article/10.1007/s00120-013-3403-2
590. Routh, J.C., et al. Design and Methodological Considerations of the Centers for Disease Control and
Prevention Urologic and Renal Protocol for the Newborn and Young Child with Spina Bifida. J Urol,
2016. 196: 1728.
https://pubmed.ncbi.nlm.nih.gov/27475969
591. Fox, J.A., et al. Cystatin C as a marker of early renal insufficiency in children with congenital
neuropathic bladder. J Urol, 2014. 191: 1602.
https://pubmed.ncbi.nlm.nih.gov/53070137
592. Dangle, P.P., et al. Cystatin C-calculated Glomerular Filtration Rate-A Marker of Early Renal
Dysfunction in Patients With Neuropathic Bladder. Urology, 2017. 100: 213.
https://pubmed.ncbi.nlm.nih.gov/27542858
593. Fernbach, S.K., et al. Ultrasound grading of hydronephrosis: introduction to the system used by the
Society for Fetal Urology. Pediatr Radiol, 1993. 23: 478.
https://pubmed.ncbi.nlm.nih.gov/8255658
594. Kim, W.J., et al. Can Bladder Wall Thickness Predict Videourodynamic Findings in Children with
Spina Bifida? J Urol., 2015. 194: 180.
https://pubmed.ncbi.nlm.nih.gov/25776909
595. Bauer, S.B., et al. International Children’s Continence Society’s recommendations for initial
diagnostic evaluation and follow-up in congenital neuropathic bladder and bowel dysfunction in
children. Neurourol Urodyn, 2012. 31: 610.
https://pubmed.ncbi.nlm.nih.gov/51980168
596. Almodhen, F., et al. Postpubertal Urodynamic and Upper Urinary Tract Changes in Children With
Conservatively Treated Myelomeningocele. J Urol, 2007. 178: 1479.
https://pubmed.ncbi.nlm.nih.gov/47368429
597. Foon, R., et al. Prophylactic antibiotics to reduce the risk of urinary tract infections after urodynamic
studies. Cochrane Database Syst Rev, 2012.
https://pubmed.ncbi.nlm.nih.gov/23076941
598. Shekarriz, B., et al. Lack of morbidity from urodynamic studies in children with asymptomatic
bacteriuria. Urology, 1999. 54: 359.
https://pubmed.ncbi.nlm.nih.gov/10443739
599. Aoki, H., et al. [Evaluation of neurogenic bladder in patients with spinal cord injury using a CMG.
EMG study and CMG.UFM.EMG study]. Hinyokika Kiyo, 1985. 31: 937.
https://pubmed.ncbi.nlm.nih.gov/4061211
600. Bradley, C.S., et al. Urodynamic evaluation of the bladder and pelvic floor. Gastroenterol Clin North
Am, 2008. 37: 539.
https://pubmed.ncbi.nlm.nih.gov/18793995
601. Casado, J.S., et al. [Urodynamic assessment of the voiding phase in childhood]. Arch Esp Urol,
2002. 55: 177.
https://pubmed.ncbi.nlm.nih.gov/12014050
602. Wen, J.G., et al. Cystometry techniques in female infants and children. Int Urogynecol J Pelvic Floor
Dysfunct, 2000. 11: 103.
https://pubmed.ncbi.nlm.nih.gov/10805268
603. Zermann, D.H., et al. Diagnostic value of natural fill cystometry in neurogenic bladder in children. Eur
Urol, 1997. 32: 223.
https://pubmed.ncbi.nlm.nih.gov/9286658
604. Jorgensen, B., et al. Natural Fill Urodynamics and Conventional Cystometrogram in Infants With
Neurogenic Bladder. J Urol, 2009. 181: 1862.
https://pubmed.ncbi.nlm.nih.gov/50430551
605. Leonardo, C.R., et al. Risk factors for renal scarring in children and adolescents with lower urinary
tract dysfunction. Pediatr Nephrol, 2007. 22: 1891.
https://pubmed.ncbi.nlm.nih.gov/47506229

142 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


606. Shiroyanagi, Y., et al. The Significance of <sup>99m</sup>Technetium Dimercapto-Succinic Acid
Renal Scan in Children With Spina Bifida During Long-Term Followup. J Urol, 2009. 181: 2262.
https://pubmed.ncbi.nlm.nih.gov/50456257
607. Veenboer, P.W., et al. Diagnostic accuracy of Tc-99m DMSA scintigraphy and renal ultrasonography
for detecting renal scarring and relative function in patients with spinal dysraphism. Neurourol
Urodyn, 2015. 34: 513.
https://pubmed.ncbi.nlm.nih.gov/24706504
608. Jorgensen, B., et al. Long-term follow-up in spinal dysraphism: Outcome of renal function and
urinary and faecal continence. Scand J Urol and Nephrol, 2010. 44: 95.
https://pubmed.ncbi.nlm.nih.gov/20187759
609. Olesen, J.D., et al. The association between urinary continence and quality of life in paediatric
patients with spina bifida and tethered cord. Paediatr Child Health (Canada), 2013. 18: e32.
https://pubmed.ncbi.nlm.nih.gov/24421717
610. Araujo, E.J., et al. Outcomes of infants followed-up at least 12 months after fetal open and
endoscopic surgery for meningomyelocele: a systematic review and meta-analysis. J Evid Based
Med, 2016.
https://pubmed.ncbi.nlm.nih.gov/27305320
611. Leal da Cruz, M., et al. A 4-year prospective urological assessment of in utero Myelomeningocele
repair: Does gestational age at birth play a role at later neurogenic bladder pattern? J Urol, 2016.
14: 14.
https://pubmed.ncbi.nlm.nih.gov/27988193
612. Carr, M.C. Urological results after fetal myelomeningocele repair in pre-MOMS trial patients at the
children’s hospital of Philadelphia. Fetal Diagn Ther, 2015. 37: 211.
https://pubmed.ncbi.nlm.nih.gov/53238520
613. Danzer, E., et al. Long-term neurofunctional outcome, executive functioning, and behavioral
adaptive skills following fetal myelomeningocele surgery. Am J Obstet Gynecol, 2016. 214: 269.e1.
https://pubmed.ncbi.nlm.nih.gov/26440692
614. Horst, M., et al. Prenatal myelomeningocele repair: Do bladders better? Neurourol Urodyn, 2016.
https://pubmed.ncbi.nlm.nih.gov/27862250
615. Macedo, A., et al. Urological evaluation of patients that had undergone in utero myelomeningocele
closure: A prospective assessment at first presentation and early follow-up. Do their bladder benefit
from it? Neurourol Urodyn, 2015. 34: 461.
https://pubmed.ncbi.nlm.nih.gov/53114791
616. Kaefer, M., et al. Improved bladder function after prophylactic treatment of the high risk neurogenic
bladder in newborns with myelomentingocele. J Urol, 1999. 162: 1068.
https://pubmed.ncbi.nlm.nih.gov/10458433
617. Park, J.M. Early reduction of mechanical load of the bladder improves compliance: experimental
and clinical observations. Dial Pediatr Urol 2000. 23: 6. [No abstract available].
618. Dik, P., et al. Early start to therapy preserves kidney function in spina bifida patients. Eur Urol, 2006.
49: 908.
https://pubmed.ncbi.nlm.nih.gov/16458416
619. Joseph, D.B., et al. Clean, intermittent catheterization of infants with neurogenic bladder. Pediatrics,
1989. 84: 78.
https://pubmed.ncbi.nlm.nih.gov/2740179
620. Lindehall, B., et al. Long-term intermittent catheterization: the experience of teenagers and young
adults with myelomeningocele. J Urol, 1994. 152: 187.
https://pubmed.ncbi.nlm.nih.gov/8201663
621. Kiddoo, D., et al. Randomized Crossover Trial of Single Use Hydrophilic Coated vs Multiple Use
Polyvinylchloride Catheters for Intermittent Catheterization to Determine Incidence of Urinary
Infection. J Urol., 2015.
https://pubmed.ncbi.nlm.nih.gov/25584995
622. Prieto, J., et al. Intermittent catheterisation for long-term bladder management [Systematic Review].
Cochrane Database Syst Rev, 2014. 9: 9. CD006008
https://pubmed.ncbi.nlm.nih.gov/25208303
623. Moore, K.N., et al. Long-term bladder management by intermittent catheterisation in adults and
children. Cochrane Database Syst Rev, 2007: CD006008.
https://pubmed.ncbi.nlm.nih.gov/17943874
624. Lindehall, B., et al. Complications of clean intermittent catheterization in young females with
myelomeningocele: 10 to 19 years of followup. J Urol, 2007. 178: 1053.
https://pubmed.ncbi.nlm.nih.gov/17632181

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 143


625. Lucas, E.J., et al. Comparison of the microbiological milieu of patients randomized to either
hydrophilic or conventional PVC catheters for clean intermittent catheterization. J Pediatr Surg,
2016. 12: 172.e1.
https://pubmed.ncbi.nlm.nih.gov/26951923
626. Andersson, K.E., et al. Pharmacological treatment of overactive bladder: report from the
International Consultation on Incontinence. Curr Opin Urol, 2009. 19: 380.
https://pubmed.ncbi.nlm.nih.gov/19448545
627. Rawashdeh, Y.F., et al. International children’s continence society’s recommendations for
therapeutic intervention in congenital neuropathic bladder and bowel dysfunction in children.
Neurourol Urodyn, 2012. 31: 615.
https://pubmed.ncbi.nlm.nih.gov/22532368
628. Abrams, P., et al. Muscarinic receptors: their distribution and function in body systems, and the
implications for treating overactive bladder. Br J Pharmacol, 2006. 148: 565.
https://pubmed.ncbi.nlm.nih.gov/16751797
629. Hegde, S.S., et al. Muscarinic receptor subtypes modulating smooth muscle contractility in the
urinary bladder. Life Sci, 1999. 64: 419.
https://pubmed.ncbi.nlm.nih.gov/10069505
630. Goessl, C., et al. Urodynamic effects of oral oxybutynin chloride in children with myelomeningocele
and detrusor hyperreflexia. Urology, 1998. 51: 94.
https://pubmed.ncbi.nlm.nih.gov/9457296
631. Lee, J.H., et al. Efficacy, tolerability, and safety of oxybutynin chloride in pediatric neurogenic
bladder with spinal dysraphism: A retrospective, multicenter, observational study. Korean J Urol,
2014. 55: 828.
https://pubmed.ncbi.nlm.nih.gov/25512818
632. Krause, P., et al. Pharmacokinetics of intravesical versus oral oxybutynin in healthy adults: results of
an open label, randomized, prospective clinical study. J Urol, 2013. 190: 1791.
https://pubmed.ncbi.nlm.nih.gov/23669567
633. Van Meel, T.D., et al. The effect of intravesical oxybutynin on the ice water test and on electrical
perception thresholds in patients with neurogenic detrusor overactivity. Neurourol Urodyn, 2010.
29: 391.
https://pubmed.ncbi.nlm.nih.gov/19787712
634. Humblet, M., et al. Long-term outcome of intravesical oxybutynin in children with detrusor-sphincter
dyssynergia: With special reference to age-dependent parameters. Neurourol Urodyn, 2015.
34: 336.
https://pubmed.ncbi.nlm.nih.gov/52968069
635. Guerra, L.A., et al. Intravesical Oxybutynin for Children With Poorly Compliant Neurogenic Bladder:
A Systematic Review. J Urol, 2008. 180: 1091.
https://pubmed.ncbi.nlm.nih.gov/50210499
636. Cartwright, P.C., et al. Efficacy and Safety of Transdermal and Oral Oxybutynin in Children With
Neurogenic Detrusor Overactivity. J Urol, 2009. 182: 1548.
https://pubmed.ncbi.nlm.nih.gov/50612864
637. Gish, P., et al. Spectrum of Central Anticholinergic Adverse Effects Associated with Oxybutynin:
Comparison of Pediatric and Adult Cases. J Pediatr, 2009. 155: 432.
https://pubmed.ncbi.nlm.nih.gov/19732583
638. Todorova, A., et al. Effects of tolterodine, trospium chloride, and oxybutynin on the central nervous
system. J Clin Pharmacol, 2001. 41: 636.
https://pubmed.ncbi.nlm.nih.gov/11402632
639. Giramonti, K.M., et al. The effects of anticholinergic drugs on attention span and short-term memory
skills in children. Neurourol Urodyn, 2008. 27: 315.
https://pubmed.ncbi.nlm.nih.gov/17828786
640. Veenboer, P.W., et al. Behavioral effects of long-term antimuscarinic use in patients with spinal
dysraphism: A case control study. J Urol, 2013. 190: 2228.
https://pubmed.ncbi.nlm.nih.gov/52821065
641. Reddy, P.P., et al. Long-term efficacy and safety of tolterodine in children with neurogenic detrusor
overactivity. J Pediatr Surg, 2008. 4: 428.
https://pubmed.ncbi.nlm.nih.gov/50195955
642. Mahanta, K., et al. Comparative efficacy and safety of extended-release and instant-release
tolterodine in children with neural tube defects having cystometric abnormalities. J Pediatr Urol,
2008. 4: 118.
https://pubmed.ncbi.nlm.nih.gov/18631906

144 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


643. Bolduc, S., et al. Double anticholinergic therapy for refractory overactive bladder. J Urol, 2009.
182: 2033.
https://pubmed.ncbi.nlm.nih.gov/19695628
644. Bolduc, S., et al. Prospective open label study of solifenacin for overactive bladder in children.
J Urol, 2010. 184: 1668.
https://pubmed.ncbi.nlm.nih.gov/51040341
645. Christoph, F., et al. Long-term efficacy of tolterodine and patient compliance in pediatric patients
with neurogenic detrusor overactivity. Urol Int, 2007. 79: 55.
https://pubmed.ncbi.nlm.nih.gov/47051820
646. Nadeau, G., et al. Double anticholinergic therapy for refractory neurogenic and nonneurogenic
detrusor overactivity in children: Long-term results of a prospective open-label study. Can Urol
Assoc J, 2014. 8: 175.
https://pubmed.ncbi.nlm.nih.gov/25024786
647. Schulte-Baukloh, H., et al. Urodynamic effects of propiverine in children and adolescents with
neurogenic bladder: Results of a prospective long-term study. J Pediatr Surg, 2012. 8: 386.
https://pubmed.ncbi.nlm.nih.gov/51611912
648. Wu, H.Y., et al. Neurogenic bladder dysfunction due to myelomeningocele: neonatal versus
childhood treatment. J Urol, 1997. 157: 2295.
https://pubmed.ncbi.nlm.nih.gov/9146656
649. Wollner, J., et al. Initial experience with the treatment of neurogenic detrusor overactivity with a new
beta-3 agonist (mirabegron) in patients with spinal cord injury. Spinal Cord, 2016. 54: 78.
https://pubmed.ncbi.nlm.nih.gov/26503222
650. Austin, P.F., et al. Alpha-adrenergic blockade in children with neuropathic and nonneuropathic
voiding dysfunction. J Urol, 1999. 162: 1064.
https://pubmed.ncbi.nlm.nih.gov/10458432
651. Homsy, Y., et al. Phase IIb/III dose ranging study of tamsulosin as treatment for children with
neuropathic bladder. J Urol, 2011. 186: 2033.
https://pubmed.ncbi.nlm.nih.gov/51634187
652. Tsuda, Y., et al. Population pharmacokinetics of tamsulosin hydrochloride in paediatric patients with
neuropathic and non-neuropathic bladder. Brit J Clin Pharmacol, 2010. 70: 88.
https://pubmed.ncbi.nlm.nih.gov/20642551
653. Hascoet, J., et al. Outcomes of intra-detrusor injections of botulinum toxin in patients with spina
bifida: A systematic review. [Review]. Neurourol Urodyn, 2016. 17: 17.
https://pubmed.ncbi.nlm.nih.gov/27187872
654. Sager, C., et al. Pharmacotherapy in pediatric neurogenic bladder intravesical botulinum toxin type
a. Isrn Urology Print, 2012. 2012: 763159.
https://pubmed.ncbi.nlm.nih.gov/22720170
655. Tiryaki, S., et al. Botulinum injection is useless on fibrotic neuropathic bladders. J Pediatr Surg,
2015. 11: 27.e1.
https://pubmed.ncbi.nlm.nih.gov/25448589
656. Horst, M., et al. Repeated Botulinum-A toxin injection in the treatment of neuropathic bladder
dysfunction and poor bladder compliance in children with myelomeningocele. Neurourol Urodyn,
2011. 30: 1546.
https://pubmed.ncbi.nlm.nih.gov/51478950
657. Mascarenhas, F., et al. Trigonal injection of botulinum toxin-A does not cause vesicoureteral reflux in
neurogenic patients. Neurourol Urodyn, 2008. 27: 311.
https://pubmed.ncbi.nlm.nih.gov/17914742
658. Altaweel, W., et al. Repeated intradetrusor botulinum toxin type A in children with neurogenic
bladder due to myelomeningocele. J Urol, 2006. 175: 1102.
https://pubmed.ncbi.nlm.nih.gov/16469632
659. Leitner, L., et al. More Than 15 Years of Experience with Intradetrusor OnabotulinumtoxinA
Injections for Treating Refractory Neurogenic Detrusor Overactivity: Lessons to Be Learned. Eur
Urol, 2016. 70: 522.
https://pubmed.ncbi.nlm.nih.gov/27106070
660. Greer, T., et al. Ten years of experience with intravesical and intrasphincteric onabotulinumtoxinA in
children. J Pediatr Surg, 2016. 12: 94.e1.
https://pubmed.ncbi.nlm.nih.gov/26472538
661. Franco, I., et al. The Use of Botulinum Toxin A Injection for the Management of External Sphincter
Dyssynergia in Neurologically Normal Children. J Urol, 2007. 178: 1775.
https://pubmed.ncbi.nlm.nih.gov/47368508

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 145


662. Mokhless, I., et al. Botulinum A toxin urethral sphincter injection in children with nonneurogenic
neurogenic bladder. J Urol, 2006. 176: 1767.
https://pubmed.ncbi.nlm.nih.gov/16945643
663. Hagerty, J.A., et al. Intravesical Electrotherapy for Neurogenic Bladder Dysfunction: A 22-Year
Experience. J Urol, 2007. 178: 1680.
https://pubmed.ncbi.nlm.nih.gov/47368511
664. Boone, T.B., et al. Transurethral intravesical electrotherapy for neurogenic bladder dysfunction in
children with myelodysplasia: a prospective, randomized clinical trial. J Urol, 1992. 148: 550.
https://pubmed.ncbi.nlm.nih.gov/1640520
665. Cheng, E.Y., et al. Bladder stimulation therapy improves bladder compliance: results from a multi-
institutional trial. J Urol, 1996. 156: 761.
https://pubmed.ncbi.nlm.nih.gov/8683778
666. Guys, J.M., et al. Sacral neuromodulation for neurogenic bladder dysfunction in children. J Urol,
2004. 172: 1673.
https://pubmed.ncbi.nlm.nih.gov/15371787
667. Lansen-Koch, S.M.P., et al. Sacral nerve modulation for defaecation and micturition disorders in
patients with spina bifida. Colorectal Disease, 2012. 14: 508.
https://pubmed.ncbi.nlm.nih.gov/21689346
668. Capitanucci, M.L., et al. Long-term efficacy of percutaneous tibial nerve stimulation for different
types of lower urinary tract dysfunction in children. J Urol, 2009. 182: 2056.
https://pubmed.ncbi.nlm.nih.gov/19695611
669. Xiao, C.G., et al. An artificial somatic-central nervous system-autonomic reflex pathway for
controllable micturition after spinal cord injury: preliminary results in 15 patients. J Urol, 2003.
170: 1237.
https://pubmed.ncbi.nlm.nih.gov/14501733
670. Tuite, G.F., et al. Urological Outcome of the Xiao Procedure in Children with Myelomeningocele and
Lipomyelomeningocele Undergoing Spinal Cord Detethering. J Urol, 2016. 196: 1735.
https://pubmed.ncbi.nlm.nih.gov/27288694
671. Bloom, D.A., et al. Urethral dilation improves bladder compliance in children with myelomeningocele
and high leak point pressures. J Urol, 1990. 144: 430.
https://pubmed.ncbi.nlm.nih.gov/2374216
672. Park, J.M., et al. External urethral sphincter dilation for the management of high risk
myelomeningocele: 15-year experience. J Urol, 2001. 165: 2383.
https://pubmed.ncbi.nlm.nih.gov/11371982
673. Wan, J. The role of urethral dilation in managing pediatric neurogenic bladder dysfunction. Curr Urol
Rep, 2009. 10: 153.
https://pubmed.ncbi.nlm.nih.gov/19239821
674. Blocksom, B.H., Jr. Bladder pouch for prolonged tubeless cystostomy. J Urol, 1957. 78: 398.
https://pubmed.ncbi.nlm.nih.gov/13476506
675. Lee, M.W., et al. Intractable high-pressure bladder in female infants with spina bifida: clinical
characteristics and use of vesicostomy. Urology, 2005. 65: 568.
https://pubmed.ncbi.nlm.nih.gov/15780378
676. Mosiello, G., et al. Button Cystostomy: Is it really a Safe and Effective Therapeutic Option in
Paediatric Patients with Neurogenic Bladder? Urology, 2016. 29: 29.
https://pubmed.ncbi.nlm.nih.gov/27693876
677. Ausili, E., et al. Transanal irrigation in myelomeningocele children: an alternative, safe and valid
approach for neurogenic constipation. Spinal Cord, 2010. 48: 560.
https://pubmed.ncbi.nlm.nih.gov/20084075
678. Christensen, P., et al. Long-term outcome and safety of transanal irrigation for constipation and
fecal incontinence. Dis Colon Rectum, 2009. 52: 286.
https://pubmed.ncbi.nlm.nih.gov/19279425
679. Malone, P.S., et al. Preliminary report: the antegrade continence enema. Lancet, 1990. 336: 1217.
https://pubmed.ncbi.nlm.nih.gov/1978072
680. Anselmo, C.B., et al. Left-colon antegrade enema (LACE): Long-term experience with the Macedo-
Malone approach. Neurourol Urodyn, 2017. 36: 111.
https://pubmed.ncbi.nlm.nih.gov/26417710
681. Siddiqui, A.A., et al. Long-term follow-up of patients after antegrade continence enema procedure.
J Pediatr Gastroenterol Nutr, 2011. 52: 574.
https://pubmed.ncbi.nlm.nih.gov/21502828

146 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


682. Zegers, B.S.H.J., et al. Urinary tract infections in children with spina bifida: An inventory of 41
European centers. Pediatr Nephrol, 2009. 24: 783.
https://pubmed.ncbi.nlm.nih.gov/50357955
683. Hansson, S., et al. Untreated bacteriuria in asymptomatic girls with renal scarring. Pediatrics, 1989.
84: 964.
https://pubmed.ncbi.nlm.nih.gov/2587151
684. Hansson, S., et al. Untreated asymptomatic bacteriuria in girls: I--Stability of urinary isolates. BMJ,
1989. 298: 853.
https://pubmed.ncbi.nlm.nih.gov/2497822
685. Zegers, S.H., et al. The influence of antibiotic prophylaxis on bacterial resistance in urinary tract
infections in children with spina bifida. BMC Infect Dis, 2017. 17: 63.
https://pubmed.ncbi.nlm.nih.gov/28081719
686. Akil, I., et al. Do patients with neurogenic bladder treated with clean intermittent catheterization
need antibacterial prophylaxis? Turkish J Med Sci, 2016. 46: 1151.
https://pubmed.ncbi.nlm.nih.gov/27513418
687. Mutlu, H., et al. Urinary tract infection prophylaxis in children with neurogenic bladder with cranberry
capsules: randomized controlled trial. Isrn Pediat Print, 2012. 2012: 317280.
https://pubmed.ncbi.nlm.nih.gov/22811926
688. Johnson, H.W., et al. A short-term study of nitrofurantoin prophylaxis in children managed with
clean intermittent catheterization. Pediatrics, 1994. 93: 752.
https://pubmed.ncbi.nlm.nih.gov/8165073
689. Schlager, T.A., et al. Nitrofurantoin prophylaxis for bacteriuria and urinary tract infection in children
with neurogenic bladder on intermittent catheterization. J Pediatr, 1998. 132: 704.
https://pubmed.ncbi.nlm.nih.gov/9580774
690. Schlager, T.A., et al. Effect of a single-use sterile catheter for each void on the frequency of
bacteriuria in children with neurogenic bladder on intermittent catheterization for bladder emptying.
Pediatrics, 2001. 108: E71.
https://pubmed.ncbi.nlm.nih.gov/11581479
691. Kanaheswari, Y., et al. Urinary tract infection and bacteriuria in children performing clean intermittent
catheterization with reused catheters. Spinal Cord, 2014. 25: 25.
https://pubmed.ncbi.nlm.nih.gov/25420498
692. Defoor, W., et al. Safety of gentamicin bladder irrigations in complex urological cases. J Urol, 2006.
175: 1861.
https://pubmed.ncbi.nlm.nih.gov/16600780
693. Wan, J., et al. Intravesical instillation of gentamicin sulfate: in vitro, rat, canine, and human studies.
Urology, 1994. 43: 531.
https://pubmed.ncbi.nlm.nih.gov/8154077
694. Misseri, R., et al. Reflux in cystoplasties. Arch Esp Urol, 2008. 61: 213.
https://pubmed.ncbi.nlm.nih.gov/18491737
695. Soygur, T., et al. The need for ureteric re-implantation during augmentation cystoplasty: video-
urodynamic evaluation. BJU Int, 2010. 105: 530.
https://pubmed.ncbi.nlm.nih.gov/19583716
696. Helmy, T.E., et al. Vesicouretral reflux with neuropathic bladder: Studying the resolution rate after
ileocystoplasty. Urology, 2013. 82: 425.
https://pubmed.ncbi.nlm.nih.gov/52558103
697. Polackwich, A.S., et al. Long-term followup after endoscopic treatment of vesicoureteral reflux with
dextranomer/hyaluronic acid copolymer in patients with neurogenic bladder. J Urol, 2012. 188: 1511.
https://pubmed.ncbi.nlm.nih.gov/52167584
698. Engel, J.D., et al. Surgical versus endoscopic correction of vesicoureteral reflux in children with
neurogenic bladder dysfunction. J Urol, 1997. 157: 2291.
https://pubmed.ncbi.nlm.nih.gov/9146655
699. Verhoef, M., et al. Sex education, relationships, and sexuality in young adults with spina bifida. Arch
Phys Med Rehabil, 2005. 86: 979.
https://pubmed.ncbi.nlm.nih.gov/15895345
700. Elias, E.R., et al. Precocious puberty in girls with myelodysplasia. Pediatrics, 1994. 93: 521.
https://pubmed.ncbi.nlm.nih.gov/8115222
701. Cardenas, D.D., et al. Sexual Functioning in Adolescents and Young Adults With Spina Bifida. Arch
Phys Med Rehabil, 2008. 89: 31.
https://pubmed.ncbi.nlm.nih.gov/18164327

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 147


702. Gatti, C., et al. Predictors of successful sexual partnering of adults with spina bifida. J Urol, 2009.
182: 1911.
https://pubmed.ncbi.nlm.nih.gov/19695634
703. Lassmann, J., et al. Sexual function in adult patients with spina bifida and its impact on quality of
life. J Urol, 2007. 178: 1611.
https://pubmed.ncbi.nlm.nih.gov/17707040
704. Palmer, J.S., et al. Erectile dysfunction in patients with spina bifida is a treatable condition. J Urol,
2000. 164: 958.
https://pubmed.ncbi.nlm.nih.gov/10958716
705. Bong, G.W., et al. Sexual health in adult men with spina bifida. Sci World J, 2007. 7: 1466.
https://pubmed.ncbi.nlm.nih.gov/17767363
706. Overgoor, M.L.E., et al. Increased sexual health after restored genital sensation in male patients with
spina bifida or a spinal cord injury: The TOMAX procedure. J Urol, 2013. 189: 626.
https://pubmed.ncbi.nlm.nih.gov/52359089
707. Stein, R., et al. Bladder augmentation and urinary diversion in patients with neurogenic bladder:
Surgical considerations. J Pediatr Urol, 2012. 8: 153.
https://pubmed.ncbi.nlm.nih.gov/22264521
708. Castellan, M., et al. Complications after use of gastric segments for lower urinary tract
reconstruction. J Urol, 2012. 187: 1823.
https://pubmed.ncbi.nlm.nih.gov/51915753
709. Nguyen, D.H., et al. The syndrome of dysuria and hematuria in pediatric urinary reconstruction with
stomach. J. Urol., 1993. 150: 707
https://pubmed.ncbi.nlm.nih.gov/8326629
710. Boissier, R., et al. What is the outcome of paediatric gastrocystoplasty when the patients reach
adulthood? BJU Int, 2016. 118: 980.
https://pubmed.ncbi.nlm.nih.gov/27322857
711. Bogaert, G.A., et al. The physiology of gastrocystoplasty: once a stomach, always a stomach.
J Urol, 1995. 153: 1977.
https://pubmed.ncbi.nlm.nih.gov/7752376
712. Herschorn, S., et al. Patient perspective of long-term outcome of augmentation cystoplasty for
neurogenic bladder. Urology, 1998. 52: 672.
https://pubmed.ncbi.nlm.nih.gov/9763092
713. Medel, R., et al. Urinary continence outcome after augmentation ileocystoplasty as a single surgical
procedure in patients with myelodysplasia. J Urol, 2002. 168: 1849.
https://pubmed.ncbi.nlm.nih.gov/12352374
714. McNamara, E.R., et al. 30-Day morbidity after augmentation enterocystoplasty and
appendicovesicostomy: A NSQIP pediatric analysis. J Pediatr Surg, 2015. 11: 209.e1.
https://pubmed.ncbi.nlm.nih.gov/26049255
715. Du, K., et al. Enterocystoplasty 30-day outcomes from National Surgical Quality Improvement
Program Pediatric 2012. J Pediatric Surg, 2015. 50: 1535.
https://pubmed.ncbi.nlm.nih.gov/25957024
716. Scales, C.D., Jr., et al. Evaluating outcomes of enterocystoplasty in patients with spina bifida: a
review of the literature. J Urol, 2008. 180: 2323.
https://pubmed.ncbi.nlm.nih.gov/18930285
717. Metcalfe, P.D., et al. Bladder augmentation: complications in the pediatric population. Curr Urol Rep,
2007. 8: 152.
https://pubmed.ncbi.nlm.nih.gov/17303021
718. Schlomer, B.J., et al. Cumulative incidence of outcomes and urologic procedures after
augmentation cystoplasty. J Pediatr Surg, 2014. 10: 1043.
https://pubmed.ncbi.nlm.nih.gov/53106110
719. Roth, J., et al. Long-Term Sequela of Pediatric Bladder Reconstruction. Curr Bladder Dysf Rep,
2015. 10: 419.
https://pubmed.ncbi.nlm.nih.gov/30314731
720. Casperson, K.J., et al. Ventriculoperitoneal shunt infections after bladder surgery: is mechanical
bowel preparation necessary? J Urol, 2011. 186: 1571.
https://pubmed.ncbi.nlm.nih.gov/21855924
721. Stein, R., et al. Bladder augmentation and urinary diversion in patients with neurogenic bladder:
non-surgical considerations. J Pediatr Urol, 2012. 8: 145.
https://pubmed.ncbi.nlm.nih.gov/21493159

148 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


722. Biardeau, X., et al. Risk of malignancy after augmentation cystoplasty: A systematic review.
Neurourol Urodyn, 2016. 35: 675.
https://pubmed.ncbi.nlm.nih.gov/25867054
723. Higuchi, T.T., et al. Augmentation cystoplasty and risk of neoplasia: fact, fiction and controversy.
J Urol, 2010. 184: 2492.
https://pubmed.ncbi.nlm.nih.gov/20961577
724. Higuchi, T.T., et al. Annual endoscopy and urine cytology for the surveillance of bladder tumors after
enterocystoplasty for congenital bladder anomalies. J Urol, 2011. 186: 1791.
https://pubmed.ncbi.nlm.nih.gov/21944100
725. Kokorowski, P.J., et al. Screening for malignancy after augmentation cystoplasty in children with
spina bifida: A decision analysis. J Urol, 2011. 186: 1437.
https://pubmed.ncbi.nlm.nih.gov/51577929
726. Hamid, R., et al. Routine surveillance cystoscopy for patients with augmentation and substitution
cystoplasty for benign urological conditions: is it necessary? BJU Int, 2009. 104: 392.
https://pubmed.ncbi.nlm.nih.gov/19239457
727. Lopez Pereira, P., et al. Are urodynamic studies really needed during bladder augmentation follow-
up? J Pediatr Surg, 2009. 5: 30.
https://pubmed.ncbi.nlm.nih.gov/50262453
728. Eckstein, H.B., et al. Uretero-Cystoplastik. Akt. Urol, 1973. 4: 255. [No abstract available].
729. Youssif, M., et al. Augmentation ureterocystoplasty in boys with valve bladder syndrome. J Pediatr
Urol, 2007. 3: 433.
https://pubmed.ncbi.nlm.nih.gov/18947790
730. Husmann, D.A., et al. Ureterocystoplasty: indications for a successful augmentation. J Urol, 2004.
171: 376.
https://pubmed.ncbi.nlm.nih.gov/14665935
731. Marte, A., et al. A long-term follow-up of autoaugmentation in myelodysplastic children. BJU Int,
2002. 89: 928.
https://pubmed.ncbi.nlm.nih.gov/12010242
732. Cartwright, P.C., et al. Bladder autoaugmentation: early clinical experience. J Urol, 1989. 142: 505.
https://pubmed.ncbi.nlm.nih.gov/2746767
733. Chrzan, R., et al. Detrusorectomy reduces the need for augmentation and use of antimuscarinics in
children with neuropathic bladders. J Pediatr Surg, 2013. 9: 193.
https://pubmed.ncbi.nlm.nih.gov/51877943
734. Hansen, E.L., et al. Promising long-term outcome of bladder autoaugmentation in children with
neurogenic bladder dysfunction. J Urol, 2013. 190: 1869.
https://pubmed.ncbi.nlm.nih.gov/23707450
735. Cartwright, P.C. Bladder autoaugmentation (partial detrusor myectomy)--where does it stand after
2 decades? J Urol, 2013. 190: 1643.
https://pubmed.ncbi.nlm.nih.gov/23954194
736. Dik, P., et al. Detrusorectomy for neuropathic bladder in patients with spinal dysraphism. J Urol,
2003. 170: 1351.
https://pubmed.ncbi.nlm.nih.gov/14501768
737. Bandi, G., et al. Comparison of traditional enterocystoplasty and seromuscular colocystoplasty lined
with urothelium. J Pediatr Urol, 2007. 3: 484.
https://pubmed.ncbi.nlm.nih.gov/18947800
738. Joseph, D.B., et al. Autologous cell seeded biodegradable scaffold for augmentation cystoplasty:
Phase II study in children and adolescents with spina bifida. J Urol, 2014. 191: 1389.
https://pubmed.ncbi.nlm.nih.gov/53074764
739. Atala, A., et al. Tissue-engineered autologous bladders for patients needing cystoplasty. Lancet,
2006. 367: 1241.
https://pubmed.ncbi.nlm.nih.gov/16631879
740. Austin, P.F., et al. Advantages of rectus fascial slings for urinary incontinence in children with
neuropathic bladders. J Urol, 2001. 165: 2369.
https://pubmed.ncbi.nlm.nih.gov/11398778
741. Guys, J.M., et al. Endoscopic treatment of urinary incontinence: long-term evaluation of the results.
J Urol, 2001. 165: 2389.
https://pubmed.ncbi.nlm.nih.gov/11371983
742. Holmes, N.M., et al. Placement of artificial urinary sphincter in children and simultaneous
gastrocystoplasty. J Urol, 2001. 165: 2366.
https://pubmed.ncbi.nlm.nih.gov/11371944

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 149


743. Kassouf, W., et al. Collagen injection for treatment of urinary incontinence in children. J Urol, 2001.
165: 1666.
https://pubmed.ncbi.nlm.nih.gov/11342951
744. Kryger, J.V., et al. Long-term results of artificial urinary sphincters in children are independent of age
at implantation. J Urol, 2001. 165: 2377.
https://pubmed.ncbi.nlm.nih.gov/11371981
745. Naglo, A.S. Continence training of children with neurogenic bladder and detrusor hyperactivity:
effect of atropine. Scand J Urol Nephrol, 1982. 16: 211.
https://pubmed.ncbi.nlm.nih.gov/7163785
746. Castellan, M., et al. Bladder neck sling for treatment of neurogenic incontinence in children with
augmentation cystoplasty: long-term followup. J Urol, 2005. 173: 2128.
https://pubmed.ncbi.nlm.nih.gov/15879865
747. Chrzan, R., et al. Sling suspension of the bladder neck for pediatric urinary incontinence. J Pediatr
Urol, 2009. 5: 82.
https://pubmed.ncbi.nlm.nih.gov/18976960
748. Pannek, J., et al. Clinical usefulness of the transobturator sub-urethral tape in the treatment of
stress urinary incontinence in female patients with spinal cord lesion. J Spinal Cord Med, 2012.
35: 102.
https://pubmed.ncbi.nlm.nih.gov/22525323
749. Groen, L.A., et al. The advance male sling as a minimally invasive treatment for intrinsic sphincter
deficiency in patients with neurogenic bladder sphincter dysfunction: A pilot study. Neurourol
Urodyn, 2012. 31: 1284.
https://pubmed.ncbi.nlm.nih.gov/52138240
750. Scott, F.B., et al. Treatment of incontinence secondary to myelodysplasia by an implantable
prosthetic urinary sphincter. South Med J, 1973. 66: 987.
https://pubmed.ncbi.nlm.nih.gov/4582131
751. Catti, M., et al. Artificial Urinary Sphincter in Children-Voiding or Emptying? An Evaluation of
Functional Results in 44 Patients. J Urol, 2008. 180: 690.
https://pubmed.ncbi.nlm.nih.gov/50175127
752. Gonzalez, R., et al. Seromuscular colocystoplasty lined with urothelium: experience with 16 patients.
Urology, 1995. 45: 124.
https://pubmed.ncbi.nlm.nih.gov/7817464
753. Kryger, J.V., et al. The outcome of artificial urinary sphincter placement after a mean 15-year follow-
up in a paediatric population. BJU Int, 1999. 83: 1026.
https://pubmed.ncbi.nlm.nih.gov/10368250
754. Herndon, C.D., et al. The Indiana experience with artificial urinary sphincters in children and young
adults. J Urol, 2003. 169: 650.
https://pubmed.ncbi.nlm.nih.gov/12544336
755. Simeoni, J., et al. Artificial urinary sphincter implantation for neurogenic bladder: a multi-institutional
study in 107 children. Br J Urol, 1996. 78: 287.
https://pubmed.ncbi.nlm.nih.gov/8813930
756. Kryger, J.V., et al. Surgical management of urinary incontinence in children with neurogenic
sphincteric incompetence. J Urol, 2000. 163: 256.
https://pubmed.ncbi.nlm.nih.gov/10604371
757. Grimsby, G.M., et al. Long-Term Outcomes of Bladder Neck Reconstruction without Augmentation
Cystoplasty in Children. J Urol, 2016. 195: 155.
https://pubmed.ncbi.nlm.nih.gov/26173106
758. Whittam, B., et al. Long-term fate of the bladder after isolated bladder neck procedure. J Pediatr
Surg, 2014. 10: 886.
https://pubmed.ncbi.nlm.nih.gov/24517903
759. Hayes, M.C., et al. The Pippi Salle urethral lengthening procedure; experience and outcome from
three United Kingdom centres. BJU Int, 1999. 84: 701.
https://pubmed.ncbi.nlm.nih.gov/10510119
760. Szymanski, K.M., et al. Long-term outcomes of the Kropp and Salle urethral lengthening bladder
neck reconstruction procedures. J Pediatr Surg, 2016. 12: 403.e1.
https://pubmed.ncbi.nlm.nih.gov/27687531
761. Churchill, B.M., et al. Improved continence in patients with neurogenic sphincteric incompetence
with combination tubularized posterior urethroplasty and fascial wrap: The lengthening, narrowing
and tightening procedure. J Urol, 2010. 184: 1763.
https://pubmed.ncbi.nlm.nih.gov/51040323

150 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


762. Alova, I., et al. Long-term effects of endoscopic injection of dextranomer/hyaluronic acid based
implants for treatment of urinary incontinence in children with neurogenic bladder. J Urol, 2012.
188: 1905.
https://pubmed.ncbi.nlm.nih.gov/52218046
763. Guys, J.M., et al. Endoscopic injection with polydimethylsiloxane for the treatment of pediatric
urinary incontinence in the neurogenic bladder: long-term results. J Urol, 2006. 175: 1106.
https://pubmed.ncbi.nlm.nih.gov/16469633
764. De Vocht, T.F., et al. Long-Term Results of Bulking Agent Injection for Persistent Incontinence in
Cases of Neurogenic Bladder Dysfunction. J Urol, 2010. 183: 719.
https://pubmed.ncbi.nlm.nih.gov/50737549
765. De Troyer, B., et al. A comparative study between continent diversion and bladder neck closure
versus continent diversion and bladder neck reconstruction in children. J Pediatr Surg, 2011. 7: 209.
https://pubmed.ncbi.nlm.nih.gov/50916701
766. Kavanagh, A., et al. Bladder neck closure in conjunction with enterocystoplasty and mitrofanoff
diversion for complex incontinence: Closing the door for good. J Urol, 2012. 188: 1561.
https://pubmed.ncbi.nlm.nih.gov/52167570
767. Shpall, A.I., et al. Bladder neck closure with lower urinary tract reconstruction: technique and long-
term followup. J Urol, 2004. 172: 2296.
https://pubmed.ncbi.nlm.nih.gov/15538252
768. Landau, E.H., et al. Bladder neck closure in children: a decade of followup. J Urol, 2009. 182: 1797.
https://pubmed.ncbi.nlm.nih.gov/19692069
769. Deuker, M., et al. Long-term outcome after urinary diversion using the ileocecal segment in children
and adolescents: Complications of the efferent segment. J Pediatr Surg, 2016. 12: 247.e1.
https://pubmed.ncbi.nlm.nih.gov/27282550
770. Faure, A., et al. Bladder continent catheterizable conduit (the Mitrofanoff procedure): Long-term
issues that should not be underestimated. J Pediatr Surg, 2016. 11.
https://pubmed.ncbi.nlm.nih.gov/27707652
771. Landau, E.H., et al. Superiority of the VQZ over the tubularized skin flap and the umbilicus for
continent abdominal stoma in children. J Urol, 2008. 180: 1761.
https://pubmed.ncbi.nlm.nih.gov/18721990
772. Stein, R., et al. Urinary diversion in children and adolescents with neurogenic bladder: the Mainz
experience Part III: Colonic conduit. Pediatr Nephrol, 2005.
https://pubmed.ncbi.nlm.nih.gov/15864655
773. Cass, A.S., et al. A 22-year followup of ileal conduits in children with a neurogenic bladder. J Urol,
1984. 132: 529.
https://pubmed.ncbi.nlm.nih.gov/6471190
774. Dunn, M., et al. The long-term results of ileal conduit urinary diversion in children. Br J Urol, 1979.
51: 458.
https://pubmed.ncbi.nlm.nih.gov/534825
775. Middleton, A.W., Jr., et al. Ileal conduits in children at the Massachusetts General Hospital from
1955 to 1970. J Urol, 1976. 115: 591.
https://pubmed.ncbi.nlm.nih.gov/1271557
776. Mitchell, M.E., et al. Intestinocystoplasty and total bladder replacement in children and young
adults: followup in 129 cases. J Urol, 1987. 138: 579.
https://pubmed.ncbi.nlm.nih.gov/3625861
777. Shekarriz, B., et al. Surgical complications of bladder augmentation: comparison between various
enterocystoplasties in 133 patients. Urology, 2000. 55: 123.
https://pubmed.ncbi.nlm.nih.gov/10654908
778. Husmann, D.A., et al. Long-term follow up of enteric bladder augmentations: The risk for
malignancy. J Pediatr Surg, 2008. 4: 381.
https://pubmed.ncbi.nlm.nih.gov/50215681
779. Balachandra, B., et al. Adenocarcinoma arising in a gastrocystoplasty. J Clin Pathol, 2007. 60: 85.
https://pubmed.ncbi.nlm.nih.gov/17213351
780. Castellan, M., et al. Tumor in bladder reservoir after gastrocystoplasty. J Urol, 2007. 178: 1771.
https://pubmed.ncbi.nlm.nih.gov/17707009
781. Soergel, T.M., et al. Transitional cell carcinoma of the bladder following augmentation cystoplasty for
the neuropathic bladder. J Urol, 2004. 172: 1649.
https://pubmed.ncbi.nlm.nih.gov/15371782

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 151


782. Sung, M.T., et al. Urothelial carcinoma following augmentation cystoplasty: an aggressive variant
with distinct clinicopathological characteristics and molecular genetic alterations. Histopathology,
2009. 55: 161.
https://pubmed.ncbi.nlm.nih.gov/19694823
783. Vemulakonda, V.M., et al. Metastatic adenocarcinoma after augmentation gastrocystoplasty. J Urol,
2008. 179: 1094.
https://pubmed.ncbi.nlm.nih.gov/18206936
784. Austin, J.C., et al. Patients With Spina Bifida and Bladder Cancer: Atypical Presentation, Advanced
Stage and Poor Survival. J Urol, 2007. 178: 798.
https://pubmed.ncbi.nlm.nih.gov/17631349
785. Sammer, U., et al. Do we need surveillance urethro-cystoscopy in patients with neurogenic lower
urinary tract dysfunction? PLoS ONE, 2015. 10: e0140970.
https://pubmed.ncbi.nlm.nih.gov/26513149
786. Lebowitz, R.L., et al. Neonatal hydronephrosis: 146 cases. Radiol Clin North Am, 1977. 15: 49.
https://pubmed.ncbi.nlm.nih.gov/139634
787. Brown, T., et al. Neonatal hydronephrosis in the era of sonography. AJR Am J Roentgenol, 1987.
148: 959.
https://pubmed.ncbi.nlm.nih.gov/3034009
788. Koff, S.A. Problematic ureteropelvic junction obstruction. J Urol, 1987. 138: 390.
https://pubmed.ncbi.nlm.nih.gov/3599261
789. Gunn, T.R., et al. Antenatal diagnosis of urinary tract abnormalities by ultrasonography after 28
weeks’ gestation: incidence and outcome. Am J Obstet Gynecol, 1995. 172: 479.
https://pubmed.ncbi.nlm.nih.gov/7856673
790. Grignon, A., et al. Ureteropelvic junction stenosis: antenatal ultrasonographic diagnosis, postnatal
investigation, and follow-up. Radiology, 1986. 160: 649.
https://pubmed.ncbi.nlm.nih.gov/3526403
791. Flashner, S.C., et al., Ureteropelvic junction, in Clinical Pediatric Urology. 1976, WB Saunders:
Philadelphia.
792. Thomas, D.F. Prenatally detected uropathy: epidemiological considerations. Br J Urol, 1998.
81 Suppl 2: 8.
https://pubmed.ncbi.nlm.nih.gov/9602790
793. Ebel, K.D. Uroradiology in the fetus and newborn: diagnosis and follow-up of congenital obstruction
of the urinary tract. Pediatr Radiol, 1998. 28: 630.
https://pubmed.ncbi.nlm.nih.gov/9716640
794. O’Reilly, P., et al. Consensus on diuresis renography for investigating the dilated upper urinary tract.
Radionuclides in Nephrourology Group. Consensus Committee on Diuresis Renography. J Nucl
Med, 1996. 37: 1872.
https://pubmed.ncbi.nlm.nih.gov/8917195
795. Choong, K.K., et al. Volume expanded diuretic renography in the postnatal assessment of
suspected uretero-pelvic junction obstruction. J Nucl Med, 1992. 33: 2094.
https://pubmed.ncbi.nlm.nih.gov/1460498
796. Reddy, P.P., et al. Prenatal diagnosis. Therapeutic implications. Urol Clin North Am, 1998. 25: 171.
https://pubmed.ncbi.nlm.nih.gov/9633572
797. Braga, L.H., et al. Pilot randomized, placebo controlled trial to investigate the effect of antibiotic
prophylaxis on the rate of urinary tract infection in infants with prenatal hydronephrosis. J Urol,
2014. 191: 1501. 24679865
https://pubmed.ncbi.nlm.nih.gov/24679865
798. Craig, J., et al., Long-term antibiotics to prevent urinary tract infection in children with isolated
vesicoureteric reflux: a placebo-controlled randomized trial., In: Australian and New Zeland Society
of Nephrology 38th Annual Scientific Meeting 2002: Sydney.
799. Silay, M.S., et al. The role of antibiotic prophylaxis in antenatal hydronephrosis: A systematic review.
J Ped Urol, 2017. prior to print
https://pubmed.ncbi.nlm.nih.gov/28462806
800. Weitz, M., et al. Surgery versus non-surgical management for unilateral ureteric-pelvic junction
obstruction in newborns and infants less than two years of age. Cochrane Database Syst Rev, 2016.
7: Cd010716.
https://pubmed.ncbi.nlm.nih.gov/27416073
801. Novick, A.C., et al., Surgery of the kidney, in Campell’s Urology. 1998, WB Saunders: Philadelphia.

152 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


802. Nasser, F.M., et al. Dismembered Pyeloplasty in Infants 6 Months Old or Younger With and Without
External Trans-anastomotic Nephrostent: A Prospective Randomized Study. Urology, 2017. 101: 38.
https://pubmed.ncbi.nlm.nih.gov/27693478
803. Reddy, M.N., et al. The laparoscopic pyeloplasty: is there a role in the age of robotics? Urol Clin
North Am, 2015. 42: 43.
https://pubmed.ncbi.nlm.nih.gov/25455171
804. Tasian, G.E., et al. The robotic-assisted laparoscopic pyeloplasty: gateway to advanced
reconstruction. Urol Clin North Am, 2015. 42: 89.
https://pubmed.ncbi.nlm.nih.gov/25455175
805. Huang, Y., et al. An updated meta-analysis of laparoscopic versus open pyeloplasty for ureteropelvic
junction obstruction in children. Int J Clin Exp Med, 2015. 8: 4922.
https://pubmed.ncbi.nlm.nih.gov/26131065
806. Alhazmi, H.H. Redo laparoscopic pyeloplasty among children: A systematic review and meta-
analysis. Urol Ann, 2018. 10: 347.
https://pubmed.ncbi.nlm.nih.gov/30386084
807. Cundy, T.P., et al. Meta-analysis of robot-assisted vs conventional laparoscopic and open
pyeloplasty in children. BJU Int, 2014. 114: 582.
https://pubmed.ncbi.nlm.nih.gov/25383399
808. Trevisani, L.F., et al. Current controversies in pediatric urologic robotic surgery. Curr Opin Urol, 2013.
23: 72.
https://pubmed.ncbi.nlm.nih.gov/23169150
809. Silay, M.S., et al. Laparoscopy versus robotic-assisted pyeloplasty in children: preliminary results of
a pilot prospective randomized controlled trial. World J Urol, 2020. 38: 1841.
https://pubmed.ncbi.nlm.nih.gov/31435732
810. Arena, F., et al. Conservative treatment in primary neonatal megaureter. Eur J Pediatr Surg, 1998.
8: 347.
https://pubmed.ncbi.nlm.nih.gov/9926303
811. Peters, C.A., et al. Congenital obstructed megaureters in early infancy: diagnosis and treatment.
J Urol, 1989. 142: 641.
https://pubmed.ncbi.nlm.nih.gov/2746792
812. Onen, A., et al. Long-term followup of prenatally detected severe bilateral newborn hydronephrosis
initially managed nonoperatively. J Urol, 2002. 168: 1118.
https://pubmed.ncbi.nlm.nih.gov/12187248
813. Shukla, A.R., et al. Prenatally detected primary megaureter: a role for extended followup. J Urol,
2005. 173: 1353.
https://pubmed.ncbi.nlm.nih.gov/15758800
814. Sripathi, V., et al. Primary obstructive megaureter. J Pediatr Surg, 1991. 26: 826.
https://pubmed.ncbi.nlm.nih.gov/1895193
815. Doudt, A.D., et al. Endoscopic Management of Primary Obstructive Megaureter: A Systematic
Review. J Endourol, 2018. 32: 482.
https://pubmed.ncbi.nlm.nih.gov/29676162
816. Lee, T., et al. Impact of Clinical Guidelines on Voiding Cystourethrogram Use and Vesicoureteral
Reflux Incidence. J Urol, 2018. 199: 831.
https://pubmed.ncbi.nlm.nih.gov/28866466
817. Fanos, V., et al. Antibiotics or surgery for vesicoureteric reflux in children. Lancet, 2004. 364: 1720.
https://pubmed.ncbi.nlm.nih.gov/15530633
818. Sargent, M.A. What is the normal prevalence of vesicoureteral reflux? Pediatr Radiol, 2000. 30: 587.
https://pubmed.ncbi.nlm.nih.gov/11009294
819. Skoog, S.J., et al. Pediatric Vesicoureteral Reflux Guidelines Panel Summary Report: Clinical
Practice Guidelines for Screening Siblings of Children With Vesicoureteral Reflux and Neonates/
Infants With Prenatal Hydronephrosis. J Urol, 2010. 184: 1145.
https://pubmed.ncbi.nlm.nih.gov/20650494
820. Estrada, C.R., Jr., et al. Nomograms for predicting annual resolution rate of primary vesicoureteral
reflux: results from 2,462 children. J Urol, 2009. 182: 1535.
https://pubmed.ncbi.nlm.nih.gov/19683762
821. Pirker, M.E., et al. Renal scarring in familial vesicoureteral reflux: is prevention possible? J Urol,
2006. 176: 1842.
https://pubmed.ncbi.nlm.nih.gov/16945668

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 153


822. Alsaywid, B.S., et al. High grade primary vesicoureteral reflux in boys: long-term results of a
prospective cohort study. J Urol, 2010. 184: 1598.
https://pubmed.ncbi.nlm.nih.gov/20728178
823. Hannula, A., et al. Vesicoureteral reflux in children with suspected and proven urinary tract infection.
Pediatr Nephrol, 2010. 25: 1463.
https://pubmed.ncbi.nlm.nih.gov/20467791
824. Menezes, M., et al. Familial vesicoureteral reflux--is screening beneficial? J Urol, 2009. 182: 1673.
https://pubmed.ncbi.nlm.nih.gov/19692047
825. Noe, H.N. The long-term results of prospective sibling reflux screening. J Urol, 1992. 148: 1739.
https://pubmed.ncbi.nlm.nih.gov/1433599
826. Koff, S.A., et al. The relationship among dysfunctional elimination syndromes, primary vesicoureteral
reflux and urinary tract infections in children. J Urol, 1998. 160: 1019.
https://pubmed.ncbi.nlm.nih.gov/9719268
827. Ural, Z., et al. Bladder dynamics and vesicoureteral reflux: factors associated with idiopathic lower
urinary tract dysfunction in children. J Urol, 2008. 179: 1564.
https://pubmed.ncbi.nlm.nih.gov/18295262
828. Sillen, U., et al. The Swedish reflux trial in children: v. Bladder dysfunction. J Urol, 2010. 184: 298.
https://pubmed.ncbi.nlm.nih.gov/20488486
829. Esbjorner, E., et al. Management of children with dilating vesico-ureteric reflux in Sweden. Acta
Paediatr, 2004. 93: 37.
https://pubmed.ncbi.nlm.nih.gov/14989437
830. Sjostrom, S., et al. Spontaneous resolution of high grade infantile vesicoureteral reflux. J Urol, 2004.
172: 694.
https://pubmed.ncbi.nlm.nih.gov/15247764
831. Knudson, M.J., et al. Predictive factors of early spontaneous resolution in children with primary
vesicoureteral reflux. J Urol, 2007. 178: 1684.
https://pubmed.ncbi.nlm.nih.gov/17707023
832. Sjostrom, S., et al. Predictive factors for resolution of congenital high grade vesicoureteral reflux in
infants: results of univariate and multivariate analyses. J Urol, 2010. 183: 1177.
https://pubmed.ncbi.nlm.nih.gov/20096864
833. Yeung, C.K., et al. Renal and bladder functional status at diagnosis as predictive factors for the
outcome of primary vesicoureteral reflux in children. J Urol, 2006. 176: 1152.
https://pubmed.ncbi.nlm.nih.gov/16890714
834. Mohanan, N., et al. Renal parenchymal damage in intermediate and high grade infantile
vesicoureteral reflux. J Urol, 2008. 180: 1635.
https://pubmed.ncbi.nlm.nih.gov/18708232
835. Olbing, H., et al. New renal scars in children with severe VUR: a 10-year study of randomized
treatment. Pediatr Nephrol, 2003. 18: 1128.
https://pubmed.ncbi.nlm.nih.gov/14523634
836. Peters, C., et al. Vesicoureteral reflux associated renal damage: congenital reflux nephropathy and
acquired renal scarring. J Urol, 2010. 184: 265.
https://pubmed.ncbi.nlm.nih.gov/20483150
837. Coplen, D.E., et al. Correlation of prenatal and postnatal ultrasound findings with the incidence of
vesicoureteral reflux in children with fetal renal pelvic dilatation. J Urol, 2008. 180: 1631.
https://pubmed.ncbi.nlm.nih.gov/18718617
838. Estrada, C.R., et al. Vesicoureteral reflux and urinary tract infection in children with a history of
prenatal hydronephrosis--should voiding cystourethrography be performed in cases of postnatally
persistent grade II hydronephrosis? J Urol, 2009. 181: 801.
https://pubmed.ncbi.nlm.nih.gov/19095265
839. Lee, R.S., et al. Antenatal hydronephrosis as a predictor of postnatal outcome: a meta-analysis.
Pediatrics, 2006. 118: 586.
https://pubmed.ncbi.nlm.nih.gov/16882811
840. Mallik, M., et al. Antenatally detected urinary tract abnormalities: more detection but less action.
Pediatr Nephrol, 2008. 23: 897.
https://pubmed.ncbi.nlm.nih.gov/18278521
841. Phan, V., et al. Vesicoureteral reflux in infants with isolated antenatal hydronephrosis. Pediatr
Nephrol, 2003. 18: 1224.
https://pubmed.ncbi.nlm.nih.gov/14586679

154 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


842. Ylinen, E., et al. Risk of renal scarring in vesicoureteral reflux detected either antenatally or during
the neonatal period. Urology, 2003. 61: 1238.
https://pubmed.ncbi.nlm.nih.gov/12809909
843. Naseer, S.R., et al. New renal scars in children with urinary tract infections, vesicoureteral reflux and
voiding dysfunction: a prospective evaluation. J Urol, 1997. 158: 566.
https://pubmed.ncbi.nlm.nih.gov/9224361
844. Blumenthal, I. Vesicoureteric reflux and urinary tract infection in children. Postgrad Med J, 2006. 82: 31.
https://pubmed.ncbi.nlm.nih.gov/16397077
845. Darge, K., et al. Current status of vesicoureteral reflux diagnosis. World J Urol, 2004. 22: 88.
https://pubmed.ncbi.nlm.nih.gov/15173954
846. Lebowitz, R.L., et al. International system of radiographic grading of vesicoureteric reflux.
International Reflux Study in Children. Pediatr Radiol, 1985. 15: 105.
https://pubmed.ncbi.nlm.nih.gov/3975102
847. Westwood, M.E., et al. Further investigation of confirmed urinary tract infection (UTI) in children
under five years: a systematic review. BMC Pediatr, 2005. 5: 2.
https://pubmed.ncbi.nlm.nih.gov/15769296
848. Snow, B.W., et al. Non-invasive vesicoureteral reflux imaging. J Pediatr Urol, 2010. 6: 543.
https://pubmed.ncbi.nlm.nih.gov/20488755
849. Darge, K. Voiding urosonography with US contrast agents for the diagnosis of vesicoureteric reflux
in children. II. Comparison with radiological examinations. Pediatr Radiol, 2008. 38: 54.
https://pubmed.ncbi.nlm.nih.gov/17639371
850. Papadopoulou, F., et al. Harmonic voiding urosonography with a second-generation contrast agent
for the diagnosis of vesicoureteral reflux. Pediatr Radiol, 2009. 39: 239.
https://pubmed.ncbi.nlm.nih.gov/19096835
851. Takazakura, R., et al. Magnetic resonance voiding cystourethrography for vesicoureteral reflux.
J Magn Reson Imaging, 2007. 25: 170.
https://pubmed.ncbi.nlm.nih.gov/17154372
852. Duran, C., et al. Contrast-enhanced Voiding Urosonography for Vesicoureteral Reflux Diagnosis in
Children. Radiographics, 2017. 37: 1854.
https://pubmed.ncbi.nlm.nih.gov/29019761
853. Medical versus surgical treatment of primary vesicoureteral reflux: report of the International Reflux
Study Committee, Pediatrics, 1981. 67: 392.
https://pubmed.ncbi.nlm.nih.gov/7017578
854. Scherz, H.C., et al. The selective use of dimercaptosuccinic acid renal scans in children with
vesicoureteral reflux. J Urol, 1994. 152: 628.
https://pubmed.ncbi.nlm.nih.gov/8021985
855. Hoberman, A., et al. Imaging studies after a first febrile urinary tract infection in young children.
N Engl J Med, 2003. 348: 195.
https://pubmed.ncbi.nlm.nih.gov/12529459
856. Hong, I.K., et al. Prediction of vesicoureteral reflux in children with febrile urinary tract infection
using relative uptake and cortical defect in DMSA scan. Pediatr Neonatol, 2018. 59: 618.
https://pubmed.ncbi.nlm.nih.gov/29576374
857. Grazioli, S., et al. Antenatal and postnatal ultrasound in the evaluation of the risk of vesicoureteral
reflux. Pediatr Nephrol, 2010. 25: 1687.
https://pubmed.ncbi.nlm.nih.gov/20524012
858. Lidefelt, K.J., et al. Antenatal hydronephrosis: infants with minor postnatal dilatation do not need
prophylaxis. Pediatr Nephrol, 2008. 23: 2021.
https://pubmed.ncbi.nlm.nih.gov/18560902
859. Hafez, A.T., et al. Analysis of trends on serial ultrasound for high grade neonatal hydronephrosis.
J Urol, 2002. 168: 1518.
https://pubmed.ncbi.nlm.nih.gov/12352447
860. Lee, J.H., et al. Nonrefluxing neonatal hydronephrosis and the risk of urinary tract infection. J Urol,
2008. 179: 1524.
https://pubmed.ncbi.nlm.nih.gov/18295269
861. Sidhu, G., et al. Outcome of isolated antenatal hydronephrosis: a systematic review and meta-
analysis. Pediatr Nephrol, 2006. 21: 218.
https://pubmed.ncbi.nlm.nih.gov/16362721
862. Visuri, S., et al. Postnatal imaging of prenatally detected hydronephrosis-when is voiding
cystourethrogram necessary? Pediatr Nephrol, 2018. 33: 1751.
https://pubmed.ncbi.nlm.nih.gov/29626243

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 155


863. Houle, A.M., et al. Impact of early screening for reflux in siblings on the detection of renal damage.
BJU Int, 2004. 94: 123.
https://pubmed.ncbi.nlm.nih.gov/15217445
864. Puri, P., et al. Urinary tract infection and renal damage in sibling vesicoureteral reflux. J Urol, 1998.
160: 1028.
https://pubmed.ncbi.nlm.nih.gov/9719271
865. Hansson, S., et al. Dimercapto-succinic acid scintigraphy instead of voiding cystourethrography for
infants with urinary tract infection. J Urol, 2004. 172: 1071.
https://pubmed.ncbi.nlm.nih.gov/15311040
866. Herz, D., et al. 5-year prospective results of dimercapto-succinic acid imaging in children with febrile
urinary tract infection: proof that the top-down approach works. J Urol, 2010. 184: 1703.
https://pubmed.ncbi.nlm.nih.gov/20728131
867. Preda, I., et al. Normal dimercaptosuccinic acid scintigraphy makes voiding cystourethrography
unnecessary after urinary tract infection. J Pediatr, 2007. 151: 581.
https://pubmed.ncbi.nlm.nih.gov/18035134
868. Colen, J., et al. Dysfunctional elimination syndrome is a negative predictor for vesicoureteral reflux.
J Pediatr Urol, 2006. 2: 312.
https://pubmed.ncbi.nlm.nih.gov/18947628
869. Elder, J.S., et al. Pediatric Vesicoureteral Reflux Guidelines Panel summary report on the
management of primary vesicoureteral reflux in children. J Urol, 1997. 157: 1846.
https://pubmed.ncbi.nlm.nih.gov/9112544
870. Dias, C.S., et al. Risk factors for recurrent urinary tract infections in a cohort of patients with primary
vesicoureteral reflux. Pediatr Infect Dis J, 2010. 29: 139.
https://pubmed.ncbi.nlm.nih.gov/20135833
871. Wheeler, D.M., et al. Interventions for primary vesicoureteric reflux. Cochrane Database Syst Rev,
2004: Cd001532.
https://pubmed.ncbi.nlm.nih.gov/15266449
872. Williams, G.J., et al. Long-term antibiotics for preventing recurrent urinary tract infection in children.
Cochrane Database Syst Rev, 2006: CD001534.
https://pubmed.ncbi.nlm.nih.gov/16855971
873. Singh-Grewal, D., et al. Circumcision for the prevention of urinary tract infection in boys: a
systematic review of randomised trials and observational studies. Arch Dis Child, 2005. 90: 853.
https://pubmed.ncbi.nlm.nih.gov/15890696
874. Greenfield, S.P. Antibiotic prophylaxis in pediatric urology: an update. Curr Urol Rep, 2011. 12: 126.
https://pubmed.ncbi.nlm.nih.gov/21229337
875. Greenfield, S.P., et al. Vesicoureteral reflux: the RIVUR study and the way forward. J Urol, 2008.
179: 405.
https://pubmed.ncbi.nlm.nih.gov/18076937
876. Brandstrom, P., et al. The Swedish reflux trial in children: IV. Renal damage. J Urol, 2010. 184: 292.
https://pubmed.ncbi.nlm.nih.gov/20494369
877. Garin, E.H., et al. Clinical significance of primary vesicoureteral reflux and urinary antibiotic
prophylaxis after acute pyelonephritis: a multicenter, randomized, controlled study. Pediatrics, 2006.
117: 626.
https://pubmed.ncbi.nlm.nih.gov/16510640
878. Montini, G., et al. Prophylaxis after first febrile urinary tract infection in children? A multicenter,
randomized, controlled, noninferiority trial. Pediatrics, 2008. 122: 1064.
https://pubmed.ncbi.nlm.nih.gov/18977988
879. Pennesi, M., et al. Is antibiotic prophylaxis in children with vesicoureteral reflux effective in preventing
pyelonephritis and renal scars? A randomized, controlled trial. Pediatrics, 2008. 121: e1489.
https://pubmed.ncbi.nlm.nih.gov/18490378
880. Roussey-Kesler, G., et al. Antibiotic prophylaxis for the prevention of recurrent urinary tract infection
in children with low grade vesicoureteral reflux: results from a prospective randomized study. J Urol,
2008. 179: 674.
https://pubmed.ncbi.nlm.nih.gov/18082208
881. Hoberman, A., et al. Antimicrobial prophylaxis for children with vesicoureteral reflux. N Engl J Med,
2014. 370: 2367.
https://pubmed.ncbi.nlm.nih.gov/24795142
882. Wang, H.H., et al. Efficacy of antibiotic prophylaxis in children with vesicoureteral reflux: systematic
review and meta-analysis. J Urol, 2015. 193: 963.
https://pubmed.ncbi.nlm.nih.gov/25196653

156 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


883. de Bessa, J., Jr., et al. Antibiotic prophylaxis for prevention of febrile urinary tract infections in
children with vesicoureteral reflux: a meta-analysis of randomized, controlled trials comparing
dilated to nondilated vesicoureteral reflux. J Urol, 2015. 193: 1772.
https://pubmed.ncbi.nlm.nih.gov/25817142
884. Hidas, G., et al. Predicting the Risk of Breakthrough Urinary Tract Infections: Primary Vesicoureteral
Reflux. J Urol, 2015. 194: 1396.
https://pubmed.ncbi.nlm.nih.gov/26066405
885. Mathews, R., et al. The role of antimicrobial prophylaxis in the management of children with
vesicoureteral reflux--the RIVUR study outcomes. Adv Chronic Kidney Dis, 2015. 22: 325.
https://pubmed.ncbi.nlm.nih.gov/26088078
886. Wang, Z.T., et al. A Reanalysis of the RIVUR Trial Using a Risk Classification System. J Urol, 2018.
199: 1608.
https://pubmed.ncbi.nlm.nih.gov/29198997
887. Dogan, H.S., et al. Factors affecting the success of endoscopic treatment of vesicoureteral reflux
and comparison of two dextranomer based bulking agents: does bulking substance matter?
J Pediatr Urol, 2015. 11: 90.e1.
https://pubmed.ncbi.nlm.nih.gov/25791422
888. Kocherov, S., et al. Multicenter survey of endoscopic treatment of vesicoureteral reflux using
polyacrylate-polyalcohol bulking copolymer (Vantris). Urology, 2014. 84: 689.
https://pubmed.ncbi.nlm.nih.gov/25168553
889. Puri, P., et al. Multicenter survey of endoscopic treatment of vesicoureteral reflux using
polytetrafluoroethylene. J Urol, 1998. 160: 1007.
https://pubmed.ncbi.nlm.nih.gov/9719265
890. Steyaert, H., et al. Migration of PTFE paste particles to the kidney after treatment for vesico-ureteric
reflux. BJU Int, 2000. 85: 168.
https://pubmed.ncbi.nlm.nih.gov/10619969
891. Lightner, D.J. Review of the available urethral bulking agents. Curr Opin Urol, 2002. 12: 333.
https://pubmed.ncbi.nlm.nih.gov/12072655
892. Elder, J.S., et al. Endoscopic therapy for vesicoureteral reflux: a meta-analysis. I. Reflux resolution
and urinary tract infection. J Urol, 2006. 175: 716.
https://pubmed.ncbi.nlm.nih.gov/16407037
893. Ben-Meir, D., et al. Late-onset Uretero-vesical Junction Obstruction Following Endoscopic Injection
of Bulking Material for the Treatment of Vesico-ureteral Reflux. Urology, 2017. 101: 60.
https://pubmed.ncbi.nlm.nih.gov/27993711
894. Warchol, S., et al. Endoscopic correction of vesicoureteral reflux in children using polyacrylate-
polyalcohol copolymer (Vantris): 5-years of prospective follow-up. Centr Eur J Urol, 2017. 70: 314.
https://pubmed.ncbi.nlm.nih.gov/29104797
895. Okawada, M., et al. Incidence of ureterovesical obstruction and Cohen antireflux surgery after
Deflux(R) treatment for vesicoureteric reflux. J Pediatr Surg, 2018. 53: 310.
https://pubmed.ncbi.nlm.nih.gov/29217322
896. Holmdahl, G., et al. The Swedish reflux trial in children: II. Vesicoureteral reflux outcome. J Urol,
2010. 184: 280.
https://pubmed.ncbi.nlm.nih.gov/20488469
897. Duckett, J.W., et al. Surgical results: International Reflux Study in Children--United States branch.
J Urol, 1992. 148: 1674.
https://pubmed.ncbi.nlm.nih.gov/1433586
898. Lipski, B.A., et al. Voiding dysfunction after bilateral extravesical ureteral reimplantation. J Urol,
1998. 159: 1019.
https://pubmed.ncbi.nlm.nih.gov/9474222
899. Kurtz, M.P., et al. Robotic versus open pediatric ureteral reimplantation: Costs and complications
from a nationwide sample. J Pediatr Urol, 2016. 12: 408.e1.
https://pubmed.ncbi.nlm.nih.gov/27593917
900. Esposito, C., et al. Robot-assisted extravesical ureteral reimplantation (revur) for unilateral vesico-
ureteral reflux in children: results of a multicentric international survey. World J Urol, 2018. 36: 481.
https://pubmed.ncbi.nlm.nih.gov/29248949
901. Deng, T., et al. Robot-assisted laparoscopic versus open ureteral reimplantation for pediatric
vesicoureteral reflux: a systematic review and meta-analysis. World J Urol, 2018. 36: 819.
https://pubmed.ncbi.nlm.nih.gov/29374841

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 157


902. Boysen, W.R., et al. Prospective multicenter study on robot-assisted laparoscopic extravesical
ureteral reimplantation (RALUR-EV): Outcomes and complications. J Pediatr Urol, 2018. 14: 262.e1.
https://pubmed.ncbi.nlm.nih.gov/29503220
903. Austin, J.C., et al. Vesicoureteral reflux: who benefits from correction. Urol Clin North Am, 2010. 37: 243.
https://pubmed.ncbi.nlm.nih.gov/20569802
904. Canon, S.J., et al. Vesicoscopic cross-trigonal ureteral reimplantation: a minimally invasive option
for repair of vesicoureteral reflux. J Urol, 2007. 178: 269.
https://pubmed.ncbi.nlm.nih.gov/17499791
905. Chung, P.H., et al. Comparing open and pneumovesical approach for ureteric reimplantation in
pediatric patients--a preliminary review. J Pediatr Surg, 2008. 43: 2246.
https://pubmed.ncbi.nlm.nih.gov/19040945
906. El-Ghoneimi, A. Paediatric laparoscopic surgery. Curr Opin Urol, 2003. 13: 329.
https://pubmed.ncbi.nlm.nih.gov/12811298
907. Grimsby, G.M., et al. Multi-institutional review of outcomes of robot-assisted laparoscopic
extravesical ureteral reimplantation. J Urol, 2015. 193: 1791.
https://pubmed.ncbi.nlm.nih.gov/25301094
908. Janetschek, G., et al. Laparoscopic ureteral anti-reflux plasty reimplantation. First clinical
experience. Ann Urol (Paris), 1995. 29: 101.
https://pubmed.ncbi.nlm.nih.gov/7645993
909. Jayanthi, V., et al. Vesicoscopic ureteral reimplantation: a minimally invasive technique for the
definitive repair of vesicoureteral reflux. Adv Urol, 2008: 973616.
https://pubmed.ncbi.nlm.nih.gov/19009038
910. Marchini, G.S., et al. Robotic assisted laparoscopic ureteral reimplantation in children: case
matched comparative study with open surgical approach. J Urol, 2011. 185: 1870.
https://pubmed.ncbi.nlm.nih.gov/21421223
911. Riquelme, M., et al. Laparoscopic extravesical transperitoneal approach for vesicoureteral reflux.
J Laparoendosc Adv Surg Tech A, 2006. 16: 312.
https://pubmed.ncbi.nlm.nih.gov/16796449
912. Straub, M., et al. Diagnosis and metaphylaxis of stone disease. Consensus concept of the National
Working Committee on Stone Disease for the upcoming German Urolithiasis Guideline. World J Urol,
2005. 23: 309.
https://pubmed.ncbi.nlm.nih.gov/16315051
913. Metcalfe, P.D., et al. What is the need for additional bladder surgery after bladder augmentation in
childhood? J Urol, 2006. 176: 1801.
https://pubmed.ncbi.nlm.nih.gov/16945653
914. Bush, N.C., et al. Hospitalizations for pediatric stone disease in United States, 2002-2007. J Urol,
2010. 183: 1151.
https://pubmed.ncbi.nlm.nih.gov/20096871
915. Novak, T.E., et al. Sex prevalence of pediatric kidney stone disease in the United States: an
epidemiologic investigation. Urology, 2009. 74: 104.
https://pubmed.ncbi.nlm.nih.gov/19428065
916. Tasian, G.E., et al. Annual Incidence of Nephrolithiasis among Children and Adults in South Carolina
from 1997 to 2012. Clin J Am Soc Nephrol, 2016. 11: 488.
https://pubmed.ncbi.nlm.nih.gov/26769765
917. Sas, D.J., et al. Increasing incidence of kidney stones in children evaluated in the emergency
department. J Pediatr, 2010. 157: 132.
https://pubmed.ncbi.nlm.nih.gov/20362300
918. Kirejczyk, J.K., et al. An association between kidney stone composition and urinary metabolic
disturbances in children. J Pediatr Urol, 2014. 10: 130.
https://pubmed.ncbi.nlm.nih.gov/23953243
919. Saitz, T.R., et al. 24 Hour urine metabolic differences between solitary and multiple stone formers:
Results of the Collaboration on Urolithiasis in Pediatrics (CUP) working group. J Pediatr Urol, 2017.
13: 506.e1.
https://pubmed.ncbi.nlm.nih.gov/28526618
920. Kruse, K., et al. Reference values for urinary calcium excretion and screening for hypercalciuria in
children and adolescents. Eur J Pediatr, 1984. 143: 25.
https://pubmed.ncbi.nlm.nih.gov/6510426
921. Sargent, J.D., et al. Normal values for random urinary calcium to creatinine ratios in infancy.
J Pediatr, 1993. 123: 393.
https://pubmed.ncbi.nlm.nih.gov/8355114

158 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


922. Stapleton, F.B., et al. Urinary excretion of calcium following an oral calcium loading test in healthy
children. Pediatrics, 1982. 69: 594.
https://pubmed.ncbi.nlm.nih.gov/7079015
923. Borghi, L., et al. Comparison of two diets for the prevention of recurrent stones in idiopathic
hypercalciuria. N Engl J Med, 2002. 346: 77.
https://pubmed.ncbi.nlm.nih.gov/11784873
924. Curhan, G.C., et al. A prospective study of dietary calcium and other nutrients and the risk of
symptomatic kidney stones. N Engl J Med, 1993. 328: 833.
https://pubmed.ncbi.nlm.nih.gov/8441427
925. Bartosh, S.M. Medical management of pediatric stone disease. Urol Clin North Am, 2004. 31: 575.
https://pubmed.ncbi.nlm.nih.gov/15313066
926. Choi, J.N., et al. Low-dose thiazide diuretics in children with idiopathic renal hypercalciuria. Acta
Paediatr, 2011. 100: e71.
https://pubmed.ncbi.nlm.nih.gov/21284722
927. Naseri, M., et al. Role of high-dose hydrochlorothiazide in idiopathic hypercalciuric urolithiasis of
childhood. Iran J Kidney Dis, 2011. 5: 162.
https://pubmed.ncbi.nlm.nih.gov/21525575
928. Preminger, G.M., et al. Eventual attenuation of hypocalciuric response to hydrochlorothiazide in
absorptive hypercalciuria. J Urol, 1987. 137: 1104.
https://pubmed.ncbi.nlm.nih.gov/3586136
929. Tekin, A., et al. Oral potassium citrate treatment for idiopathic hypocitruria in children with calcium
urolithiasis. J Urol, 2002. 168: 2572.
https://pubmed.ncbi.nlm.nih.gov/12441986
930. Hoppe, B., et al. Urinary calcium oxalate saturation in healthy infants and children. J Urol, 1997.
158: 557.
https://pubmed.ncbi.nlm.nih.gov/9224359
931. Neuhaus, T.J., et al. Urinary oxalate excretion in urolithiasis and nephrocalcinosis. Arch Dis Child,
2000. 82: 322.
https://pubmed.ncbi.nlm.nih.gov/10735843
932. Turudic, D., et al. Calcium oxalate urolithiasis in children: urinary promoters/inhibitors and role of
their ratios. Eur J Pediatr, 2016. 175: 1959.
https://pubmed.ncbi.nlm.nih.gov/27730307
933. Morgenstern, B.Z., et al. Urinary oxalate and glycolate excretion patterns in the first year of life: a
longitudinal study. J Pediatr, 1993. 123: 248.
https://pubmed.ncbi.nlm.nih.gov/8345420
934. Defoor, W., et al. Results of a prospective trial to compare normal urine supersaturation in children
and adults. J Urol, 2005. 174: 1708.
https://pubmed.ncbi.nlm.nih.gov/16148687
935. Kovacevic, L., et al. From hypercalciuria to hypocitraturia--a shifting trend in pediatric urolithiasis?
J Urol, 2012. 188: 1623.
https://pubmed.ncbi.nlm.nih.gov/22910255
936. Tekin, A., et al. A study of the etiology of idiopathic calcium urolithiasis in children: hypocitruria is the
most important risk factor. J Urol, 2000. 164: 162.
https://pubmed.ncbi.nlm.nih.gov/10840454
937. Celiksoy, M.H., et al. Metabolic disorders in Turkish children with urolithiasis. Urology, 2015. 85: 909.
https://pubmed.ncbi.nlm.nih.gov/25817115
938. DeFoor, W., et al. Calcium-to-Citrate Ratio Distinguishes Solitary and Recurrent Urinary Stone
Forming Children. J Urol, 2017. 198: 416.
https://pubmed.ncbi.nlm.nih.gov/28365270
939. Zu’bi, F., et al. Stone growth patterns and risk for surgery among children presenting with
hypercalciuria, hypocitraturia and cystinuria as underlying metabolic causes of urolithiasis. J Pediatr
Urol, 2017. 13: 357.e1.
https://pubmed.ncbi.nlm.nih.gov/28865885
940. Tekin, A., et al. Cystine calculi in children: the results of a metabolic evaluation and response to
medical therapy. J Urol, 2001. 165: 2328.
https://pubmed.ncbi.nlm.nih.gov/11371943
941. Gabrielsen, J.S., et al. Pediatric urinary stone composition in the United States. J Urol, 2012.
187: 2182.
https://pubmed.ncbi.nlm.nih.gov/22503021

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 159


942. Rellum, D.M., et al. Pediatric urolithiasis in a non-endemic country: a single center experience from
The Netherlands. J Pediatr Urol, 2014. 10: 155.
https://pubmed.ncbi.nlm.nih.gov/23981680
943. Bove, P., et al. Reexamining the value of hematuria testing in patients with acute flank pain. J Urol,
1999. 162: 685.
https://pubmed.ncbi.nlm.nih.gov/10458342
944. Sternberg, K., et al. Pediatric stone disease: an evolving experience. J Urol, 2005. 174: 1711.
https://pubmed.ncbi.nlm.nih.gov/16148688
945. Memarsadeghi, M., et al. Unenhanced multi-detector row CT in patients suspected of having urinary
stone disease: effect of section width on diagnosis. Radiology, 2005. 235: 530.
https://pubmed.ncbi.nlm.nih.gov/15758192
946. Oner, S., et al. Comparison of spiral CT and US in the evaluation of pediatric urolithiasis. Jbr-btr,
2004. 87: 219.
https://pubmed.ncbi.nlm.nih.gov/15587558
947. Strouse, P.J., et al. Non-contrast thin-section helical CT of urinary tract calculi in children. Pediatr
Radiol, 2002. 32: 326.
https://pubmed.ncbi.nlm.nih.gov/11956719
948. Kwon, J.K., et al. Usefulness of low-dose nonenhanced computed tomography with iterative
reconstruction for evaluation of urolithiasis: diagnostic performance and agreement between the
urologist and the radiologist. Urology, 2015. 85: 531.
https://pubmed.ncbi.nlm.nih.gov/25733262
949. Alpay, H., et al. Clinical and metabolic features of urolithiasis and microlithiasis in children. Pediatr
Nephrol, 2009. 24: 2203.
https://pubmed.ncbi.nlm.nih.gov/19603196
950. Skolarikos, A., et al. Metabolic evaluation and recurrence prevention for urinary stone patients: EAU
guidelines. Eur Urol, 2015. 67: 750.
https://pubmed.ncbi.nlm.nih.gov/25454613
951. Chan, K.H., et al. The ability of a limited metabolic assessment to identify pediatric stone formers
with metabolic abnormalities. J Pediatr Urol, 2018. 14: 331.e1.
https://pubmed.ncbi.nlm.nih.gov/30177386
952. Chan, K.H., et al. Initial collection of an inadequate 24-hour urine sample in children does not
predict subsequent inadequate collections. J Pediatr Urol, 2019. 15: 74.e1.
https://pubmed.ncbi.nlm.nih.gov/30467015
953. Raza, A., et al. Pediatric urolithiasis: 15 years of local experience with minimally invasive
endourological management of pediatric calculi. J Urol, 2005. 174: 682.
https://pubmed.ncbi.nlm.nih.gov/16006948
954. Rizvi, S.A., et al. Pediatric urolithiasis: developing nation perspectives. J Urol, 2002. 168: 1522.
https://pubmed.ncbi.nlm.nih.gov/12352448
955. Shahat, A., et al. Is Tamsulosin Effective after Shock Wave Lithotripsy for Pediatric Renal Stones?
A Randomized, Controlled Study. J Urol, 2016. 195: 1284.
https://pubmed.ncbi.nlm.nih.gov/26926538
956. Velazquez, N., et al. Medical expulsive therapy for pediatric urolithiasis: Systematic review and
meta-analysis. J Pediatr Urol, 2015. 11: 321.
https://pubmed.ncbi.nlm.nih.gov/26165192
957. Dincel, N., et al. Are small residual stone fragments really insignificant in children? J Pediatr Surg,
2013. 48: 840.
https://pubmed.ncbi.nlm.nih.gov/23583144
958. El-Assmy, A., et al. Clinically Insignificant Residual Fragments: Is It an Appropriate Term in Children?
Urology, 2015. 86: 593.
https://pubmed.ncbi.nlm.nih.gov/26126693
959. Akin, Y., et al. Long-term effects of pediatric extracorporeal shockwave lithotripsy on renal function.
Res Rep Urol, 2014. 6: 21.
https://pubmed.ncbi.nlm.nih.gov/24892029
960. Aksoy, Y., et al. Extracorporeal shock wave lithotripsy in children: experience using a mpl-9000
lithotriptor. World J Urol, 2004. 22: 115.
https://pubmed.ncbi.nlm.nih.gov/14740160
961. Aldridge, R.D., et al. Anesthesia for pediatric lithotripsy. Paediatr Anaesth, 2006. 16: 236.
https://pubmed.ncbi.nlm.nih.gov/16490086

160 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


962. McLorie, G.A., et al. Safety and efficacy of extracorporeal shock wave lithotripsy in infants. Can
J Urol, 2003. 10: 2051.
https://pubmed.ncbi.nlm.nih.gov/14704109
963. Reisiger, K., et al. Pediatric nephrolithiasis: does treatment affect renal growth? Urology, 2007.
69: 1190.
https://pubmed.ncbi.nlm.nih.gov/17572213
964. Villanyi, K.K., et al. Short-term changes in renal function after extracorporeal shock wave lithotripsy
in children. J Urol, 2001. 166: 222.
https://pubmed.ncbi.nlm.nih.gov/11435873
965. Vlajkovic, M., et al. Long-term functional outcome of kidneys in children with urolithiasis after ESWL
treatment. Eur J Pediatr Surg, 2002. 12: 118.
https://pubmed.ncbi.nlm.nih.gov/12015657
966. Willis, L.R., et al. Relationship between kidney size, renal injury, and renal impairment induced by
shock wave lithotripsy. J Am Soc Nephrol, 1999. 10: 1753.
https://pubmed.ncbi.nlm.nih.gov/10446943
967. Ather, M.H., et al. Does size and site matter for renal stones up to 30-mm in size in children treated
by extracorporeal lithotripsy? Urology, 2003. 61: 212.
https://pubmed.ncbi.nlm.nih.gov/12559298
968. Muslumanoglu, A.Y., et al. Extracorporeal shock wave lithotripsy as first line treatment alternative for
urinary tract stones in children: a large scale retrospective analysis. J Urol, 2003. 170: 2405.
https://pubmed.ncbi.nlm.nih.gov/14634438
969. Ugur, G., et al. Anaesthetic/analgesic management of extracorporeal shock wave lithotripsy in
paediatric patients. Paediatr Anaesth, 2003. 13: 85.
https://pubmed.ncbi.nlm.nih.gov/12535048
970. Afshar, K., et al. Outcome of small residual stone fragments following shock wave lithotripsy in
children. J Urol, 2004. 172: 1600.
https://pubmed.ncbi.nlm.nih.gov/15371769
971. Al-Busaidy, S.S., et al. Pediatric staghorn calculi: the role of extracorporeal shock wave lithotripsy
monotherapy with special reference to ureteral stenting. J Urol, 2003. 169: 629.
https://pubmed.ncbi.nlm.nih.gov/12544330
972. Lottmann, H.B., et al. Monotherapy extracorporeal shock wave lithotripsy for the treatment of
staghorn calculi in children. J Urol, 2001. 165: 2324.
https://pubmed.ncbi.nlm.nih.gov/11371942
973. Rodrigues Netto, N., Jr., et al. Extracorporeal shock wave lithotripsy in children. J Urol, 2002.
167: 2164.
https://pubmed.ncbi.nlm.nih.gov/11956471
974. Tan, A.H., et al. Results of shockwave lithotripsy for pediatric urolithiasis. J Endourol, 2004. 18: 527.
https://pubmed.ncbi.nlm.nih.gov/15333214
975. Demirkesen, O., et al. Efficacy of extracorporeal shock wave lithotripsy for isolated lower caliceal
stones in children compared with stones in other renal locations. Urology, 2006. 67: 170.
https://pubmed.ncbi.nlm.nih.gov/16413356
976. Onal, B., et al. The impact of caliceal pelvic anatomy on stone clearance after shock wave lithotripsy
for pediatric lower pole stones. J Urol, 2004. 172: 1082.
https://pubmed.ncbi.nlm.nih.gov/15311043
977. Ozgur Tan, M., et al. The impact of radiological anatomy in clearance of lower calyceal stones after
shock wave lithotripsy in paediatric patients. Eur Urol, 2003. 43: 188.
https://pubmed.ncbi.nlm.nih.gov/12565778
978. Hochreiter, W.W., et al. Extracorporeal shock wave lithotripsy for distal ureteral calculi: what a
powerful machine can achieve. J Urol, 2003. 169: 878.
https://pubmed.ncbi.nlm.nih.gov/12576804
979. Landau, E.H., et al. Extracorporeal shock wave lithotripsy is highly effective for ureteral calculi in
children. J Urol, 2001. 165: 2316.
https://pubmed.ncbi.nlm.nih.gov/11371970
980. McAdams, S., et al. Preoperative Stone Attenuation Value Predicts Success After Shock Wave
Lithotripsy in Children. J Urol, 2010. 184: 1804.
https://pubmed.ncbi.nlm.nih.gov/20728112
981. Dogan, H.S., et al. A new nomogram for prediction of outcome of pediatric shock-wave lithotripsy.
J Pediatr Urol, 2015. 11: 84 e1.
https://pubmed.ncbi.nlm.nih.gov/25812469

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 161


982. Onal, B., et al. Nomogram and scoring system for predicting stone-free status after extracorporeal
shock wave lithotripsy in children with urolithiasis. BJU Int, 2013. 111: 344.
https://pubmed.ncbi.nlm.nih.gov/22672514
983. Yanaral, F., et al. Shock-wave Lithotripsy for Pediatric Patients: Which Nomogram Can Better
Predict Postoperative Outcomes? Urology, 2018. 117: 126.
https://pubmed.ncbi.nlm.nih.gov/29630952
984. Ergin, G., et al. Shock wave lithotripsy or retrograde intrarenal surgery: which one is more effective
for 10-20-mm renal stones in children. Ir J Med Sci, 2018. 187: 1121.
https://pubmed.ncbi.nlm.nih.gov/29502272
985. Marchetti, K.A., et al. Extracorporeal shock wave lithotripsy versus ureteroscopy for management
of pediatric nephrolithiasis in upper urinary tract stones: multi-institutional outcomes of efficacy and
morbidity. J Pediatr Urol, 2019.
https://pubmed.ncbi.nlm.nih.gov/31326329
986. Wu, H.Y., et al. Surgical management of children with urolithiasis. Urol Clin North Am, 2004. 31: 589.
https://pubmed.ncbi.nlm.nih.gov/15313067
987. Bassiri, A., et al. Transureteral lithotripsy in pediatric practice. J Endourol, 2002. 16: 257.
https://pubmed.ncbi.nlm.nih.gov/12042111
988. Caione, P., et al. Endoscopic manipulation of ureteral calculi in children by rigid operative
ureterorenoscopy. J Urol, 1990. 144: 484.
https://pubmed.ncbi.nlm.nih.gov/2374225
989. De Dominicis, M., et al. Retrograde ureteroscopy for distal ureteric stone removal in children. BJU
Int, 2005. 95: 1049.
https://pubmed.ncbi.nlm.nih.gov/15839930
990. Desai, M.R., et al. Percutaneous nephrolithotomy for complex pediatric renal calculus disease.
J Endourol, 2004. 18: 23.
https://pubmed.ncbi.nlm.nih.gov/15006048
991. Dogan, H.S., et al. Use of the holmium:YAG laser for ureterolithotripsy in children. BJU Int, 2004.
94: 131.
https://pubmed.ncbi.nlm.nih.gov/15217447
992. Raza, A., et al. Ureteroscopy in the management of pediatric urinary tract calculi. J Endourol, 2005.
19: 151.
https://pubmed.ncbi.nlm.nih.gov/15798409
993. Satar, N., et al. Rigid ureteroscopy for the treatment of ureteral calculi in children. J Urol, 2004. 172: 298.
https://pubmed.ncbi.nlm.nih.gov/15201799
994. Soygur, T., et al. Hydrodilation of the ureteral orifice in children renders ureteroscopic access
possible without any further active dilation. J Urol, 2006. 176: 285.
https://pubmed.ncbi.nlm.nih.gov/16753421
995. Thomas, J.C., et al. Pediatric ureteroscopic stone management. J Urol, 2005. 174: 1072.
https://pubmed.ncbi.nlm.nih.gov/16094060
996. Van Savage, J.G., et al. Treatment of distal ureteral stones in children: similarities to the american
urological association guidelines in adults. J Urol, 2000. 164: 1089.
https://pubmed.ncbi.nlm.nih.gov/10958749
997. ElSheemy, M.S., et al. Lower calyceal and renal pelvic stones in preschool children: A comparative
study of mini-percutaneous nephrolithotomy versus extracorporeal shockwave lithotripsy. Int J Urol,
2016. 23: 564.
https://pubmed.ncbi.nlm.nih.gov/27173126
998. Jackman, S.V., et al. Percutaneous nephrolithotomy in infants and preschool age children:
experience with a new technique. Urology, 1998. 52: 697.
https://pubmed.ncbi.nlm.nih.gov/9763096
999. Badawy, H., et al. Percutaneous management of renal calculi: experience with percutaneous
nephrolithotomy in 60 children. J Urol, 1999. 162: 1710.
https://pubmed.ncbi.nlm.nih.gov/10524919
1000. Boormans, J.L., et al. Percutaneous nephrolithotomy for treating renal calculi in children. BJU Int,
2005. 95: 631.
https://pubmed.ncbi.nlm.nih.gov/15705093
1001. Dawaba, M.S., et al. Percutaneous nephrolithotomy in children: early and late anatomical and
functional results. J Urol, 2004. 172: 1078.
https://pubmed.ncbi.nlm.nih.gov/15311042
1002. Sahin, A., et al. Percutaneous nephrolithotomy in older children. J Pediatr Surg, 2000. 35: 1336.
https://pubmed.ncbi.nlm.nih.gov/10999692

162 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


1003. Shokeir, A.A., et al. Percutaneous nephrolithotomy in treatment of large stones within horseshoe
kidneys. Urology, 2004. 64: 426.
https://pubmed.ncbi.nlm.nih.gov/15351557
1004. Dogan, H.S., et al. Percutaneous nephrolithotomy in children: does age matter? World J Urol, 2011.
29: 725.
https://pubmed.ncbi.nlm.nih.gov/21590468
1005. Guven, S., et al. Successful percutaneous nephrolithotomy in children: multicenter study on current
status of its use, efficacy and complications using Clavien classification. J Urol, 2011. 185: 1419.
https://pubmed.ncbi.nlm.nih.gov/21334653
1006. Khairy Salem, H., et al. Tubeless percutaneous nephrolithotomy in children. J Pediatr Urol, 2007.
3: 235.
https://pubmed.ncbi.nlm.nih.gov/18947742
1007. Nouralizadeh, A., et al. Experience of percutaneous nephrolithotomy using adult-size instruments in
children less than 5 years old. J Pediatr Urol, 2009. 5: 351.
https://pubmed.ncbi.nlm.nih.gov/19230776
1008. Ozden, E., et al. Modified Clavien classification in percutaneous nephrolithotomy: assessment of
complications in children. J Urol, 2011. 185: 264.
https://pubmed.ncbi.nlm.nih.gov/21074805
1009. Unsal, A., et al. Safety and efficacy of percutaneous nephrolithotomy in infants, preschool age, and
older children with different sizes of instruments. Urology, 2010. 76: 247.
https://pubmed.ncbi.nlm.nih.gov/20022089
1010. Onal, B., et al. Factors affecting complication rates of percutaneous nephrolithotomy in children:
results of a multi-institutional retrospective analysis by the Turkish pediatric urology society. J Urol,
2014. 191: 777.
https://pubmed.ncbi.nlm.nih.gov/24095906
1011. Ozden, E., et al. Percutaneous renal surgery in children with complex stones. J Pediatr Urol, 2008.
4: 295.
https://pubmed.ncbi.nlm.nih.gov/18644533
1012. Bilen, C.Y., et al. Percutaneous nephrolithotomy in children: lessons learned in 5 years at a single
institution. J Urol, 2007. 177: 1867.
https://pubmed.ncbi.nlm.nih.gov/17437838
1013. Jackman, S.V., et al. The “mini-perc” technique: a less invasive alternative to percutaneous
nephrolithotomy. World J Urol, 1998. 16: 371.
https://pubmed.ncbi.nlm.nih.gov/9870281
1014. Dede, O., et al. Ultra-mini-percutaneous nephrolithotomy in pediatric nephrolithiasis: Both low
pressure and high efficiency. J Pediatr Urol, 2015. 11: 253 e1.
https://pubmed.ncbi.nlm.nih.gov/25964199
1015. Farouk, A., et al. Is mini-percutaneous nephrolithotomy a safe alternative to extracorporeal
shockwave lithotripsy in pediatric age group in borderline stones? a randomized prospective study.
World J Urol, 2018. 36: 1139.
https://pubmed.ncbi.nlm.nih.gov/29450731
1016. Sarica, K., et al. Super-mini percutaneous nephrolithotomy for renal stone less than 25mm in
pediatric patients: Could it be an alternative to shockwave lithotripsy? Actas Urol Esp, 2017.
https://pubmed.ncbi.nlm.nih.gov/29273258
1017. Yuan, D., et al. Super-Mini Percutaneous Nephrolithotomy Reduces the Incidence of Postoperative
Adverse Events in Pediatric Patients: A Retrospective Cohort Study. Urol Int, 2019. 103: 81.
https://pubmed.ncbi.nlm.nih.gov/31039558
1018. Liu, Y., et al. Comparison of super-mini PCNL (SMP) versus Miniperc for stones larger than 2 cm: a
propensity score-matching study. World J Urol, 2018. 36: 955.
https://pubmed.ncbi.nlm.nih.gov/29387932
1019. Desai, M.R., et al. Single-step percutaneous nephrolithotomy (microperc): the initial clinical report.
J Urol, 2011. 186: 140.
https://pubmed.ncbi.nlm.nih.gov/21575966
1020. Hatipoglu, N.K., et al. Comparison of shockwave lithotripsy and microperc for treatment of kidney
stones in children. J Endourol, 2013. 27: 1141.
https://pubmed.ncbi.nlm.nih.gov/23713511
1021. Karatag, T., et al. A Comparison of 2 Percutaneous Nephrolithotomy Techniques for the Treatment of
Pediatric Kidney Stones of Sizes 10-20 mm: Microperc vs Miniperc. Urology, 2015. 85: 1015.
https://pubmed.ncbi.nlm.nih.gov/25917724

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 163


1022. Wang, W., et al. Comparing micropercutaneous nephrolithotomy and retrograde intrarenal surgery in
treating 1-2 cm solitary renal stones in pediatric patients younger than 3 years. J Pediatr Urol, 2019.
https://pubmed.ncbi.nlm.nih.gov/31301976
1023. Bilen, C.Y., et al. Tubeless mini percutaneous nephrolithotomy in infants and preschool children: a
preliminary report. J Urol, 2010. 184: 2498.
https://pubmed.ncbi.nlm.nih.gov/20961572
1024. Aghamir, S.M., et al. Feasibility of totally tubeless percutaneous nephrolithotomy under the age of
14 years: a randomized clinical trial. J Endourol, 2012. 26: 621.
https://pubmed.ncbi.nlm.nih.gov/22192104
1025. Nouralizadeh, A., et al. Fluoroscopy-free ultrasonography-guided percutaneous nephrolithotomy in
pediatric patients: a single-center experience. World J Urol, 2018. 36: 667.
https://pubmed.ncbi.nlm.nih.gov/29349571
1026. Simayi, A., et al. Clinical application of super-mini PCNL (SMP) in the treatment of upper urinary
tract stones under ultrasound guidance. World J Urol, 2019. 37: 943.
https://pubmed.ncbi.nlm.nih.gov/30167833
1027. Gamal, W., et al. Supine pediatric percutaneous nephrolithotomy (PCNL). J Pediatr Urol, 2015.
11: 78 e1.
https://pubmed.ncbi.nlm.nih.gov/25819602
1028. Bodakci, M.N., et al. Ultrasound-guided micropercutaneous nephrolithotomy in pediatric patients
with kidney stones. Int J Urol, 2015. 22: 773.
https://pubmed.ncbi.nlm.nih.gov/25975519
1029. al Busaidy, S.S., et al. Paediatric ureteroscopy for ureteric calculi: a 4-year experience. Br J Urol,
1997. 80: 797.
https://pubmed.ncbi.nlm.nih.gov/9393306
1030. Hill, D.E., et al. Ureteroscopy in children. J Urol, 1990. 144: 481.
https://pubmed.ncbi.nlm.nih.gov/2374224
1031. Richter, S., et al. Early postureteroscopy vesicoureteral reflux--a temporary and infrequent
complication: prospective study. J Endourol, 1999. 13: 365.
https://pubmed.ncbi.nlm.nih.gov/10446797
1032. Schuster, T.G., et al. Ureteroscopy for the treatment of urolithiasis in children. J Urol, 2002. 167: 1813.
https://pubmed.ncbi.nlm.nih.gov/11912438
1033. Gokce, M.I., et al. Evaluation of Postoperative Hydronephrosis Following Ureteroscopy in Pediatric
Population: Incidence and Predictors. Urology, 2016. 93: 164.
https://pubmed.ncbi.nlm.nih.gov/26972147
1034. Dogan, H.S., et al. Factors affecting complication rates of ureteroscopic lithotripsy in children:
results of multi-institutional retrospective analysis by Pediatric Stone Disease Study Group of
Turkish Pediatric Urology Society. J Urol, 2011. 186: 1035.
https://pubmed.ncbi.nlm.nih.gov/21784482
1035. Citamak, B., et al. Semi-Rigid Ureteroscopy Should Not Be the First Option for Proximal Ureteral
Stones in Children. J Endourol, 2018. 32: 1028.
https://pubmed.ncbi.nlm.nih.gov/30226405
1036. Abu Ghazaleh, L.A., et al. Retrograde intrarenal lithotripsy for small renal stones in prepubertal
children. Saudi J Kidney Dis Transpl, 2011. 22: 492.
https://pubmed.ncbi.nlm.nih.gov/21566306
1037. Corcoran, A.T., et al. When is prior ureteral stent placement necessary to access the upper urinary
tract in prepubertal children? J Urol, 2008. 180: 1861.
https://pubmed.ncbi.nlm.nih.gov/18721946
1038. Dave, S., et al. Single-institutional study on role of ureteroscopy and retrograde intrarenal surgery in
treatment of pediatric renal calculi. Urology, 2008. 72: 1018.
https://pubmed.ncbi.nlm.nih.gov/18585764
1039. Kim, S.S., et al. Pediatric flexible ureteroscopic lithotripsy: the children’s hospital of Philadelphia
experience. J Urol, 2008. 180: 2616.
https://pubmed.ncbi.nlm.nih.gov/18950810
1040. Tanaka, S.T., et al. Pediatric ureteroscopic management of intrarenal calculi. J Urol, 2008. 180: 2150.
https://pubmed.ncbi.nlm.nih.gov/18804225
1041. Li, J., et al. Application of flexible ureteroscopy combined with holmium laser lithotripsy and their
therapeutic efficacy in the treatment of upper urinary stones in children and infants. Urol J, 2019.
16: 343.
https://pubmed.ncbi.nlm.nih.gov/30784036

164 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


1042. Erkurt, B., et al. Treatment of renal stones with flexible ureteroscopy in preschool age children.
Urolithiasis, 2014. 42: 241.
https://pubmed.ncbi.nlm.nih.gov/24374900
1043. Mokhless, I.A., et al. Retrograde intrarenal surgery monotherapy versus shock wave lithotripsy for
stones 10 to 20 mm in preschool children: a prospective, randomized study. J Urol, 2014. 191: 1496.
https://pubmed.ncbi.nlm.nih.gov/24679882
1044. Saad, K.S., et al. Percutaneous Nephrolithotomy vs Retrograde Intrarenal Surgery for Large Renal
Stones in Pediatric Patients: A Randomized Controlled Trial. J Urol, 2015. 194: 1716.
https://pubmed.ncbi.nlm.nih.gov/26165587
1045. Bas, O., et al. Comparison of Retrograde Intrarenal Surgery and Micro-Percutaneous
Nephrolithotomy in Moderately Sized Pediatric Kidney Stones. J Endourol, 2016. 30: 765.
https://pubmed.ncbi.nlm.nih.gov/26983791
1046. He, Q., et al. Which is the best treatment of pediatric upper urinary tract stones among
extracorporeal shockwave lithotripsy, percutaneous nephrolithotomy and retrograde intrarenal
surgery: a systematic review. BMC Urol, 2019. 19: 98.
https://pubmed.ncbi.nlm.nih.gov/31640693
1047. Casale, P., et al. Transperitoneal laparoscopic pyelolithotomy after failed percutaneous access in the
pediatric patient. J Urol, 2004. 172: 680.
https://pubmed.ncbi.nlm.nih.gov/15247760
1048. Ghani, K.R., et al. Robotic nephrolithotomy and pyelolithotomy with utilization of the robotic
ultrasound probe. Int Braz J Urol, 2014. 40: 125.
https://pubmed.ncbi.nlm.nih.gov/24642160
1049. Lee, R.S., et al. Early results of robot assisted laparoscopic lithotomy in adolescents. J Urol, 2007.
177: 2306.
https://pubmed.ncbi.nlm.nih.gov/17509345
1050. Srivastava, A., et al. Laparoscopic Ureterolithotomy in Children: With and Without Stent - Initial
Tertiary Care Center Experience with More Than 1-Year Follow-Up. Eur J Pediatr Surg, 2016.
https://pubmed.ncbi.nlm.nih.gov/26878339
1051. Uson, A.C., et al. Ureteroceles in infants and children: a report based on 44 cases. Pediatrics, 1961.
27: 971.
https://pubmed.ncbi.nlm.nih.gov/13779382
1052. Prewitt, L.H., Jr., et al. The single ectopic ureter. AJR Am J Roentgenol, 1976. 127: 941.
https://pubmed.ncbi.nlm.nih.gov/998831
1053. Ahmed, S., et al. Single-system ectopic ureters: a review of 12 cases. J Pediatr Surg, 1992. 27: 491.
https://pubmed.ncbi.nlm.nih.gov/1522464
1054. Chwalla, R. The process of formation of cystic dilatation of the vesical end of the ureter and of
diverticula at the ureteral ostium. Urol Cutan Ren 1927. 31: 499. [No abstract available].
1055. Stephens, D. Caecoureterocele and concepts on the embryology and aetiology of ureteroceles. Aust
N Z J Surg, 1971. 40: 239.
https://pubmed.ncbi.nlm.nih.gov/5279434
1056. Tokunaka, S., et al. Muscle dysplasia in megaureters. J Urol, 1984. 131: 383.
https://pubmed.ncbi.nlm.nih.gov/6699978
1057. Zerin, J.M., et al. Single-system ureteroceles in infants and children: imaging features. Pediatr
Radiol, 2000. 30: 139.
https://pubmed.ncbi.nlm.nih.gov/10755749
1058. Monfort, G., et al. Surgical management of duplex ureteroceles. J Pediatr Surg, 1992. 27: 634.
https://pubmed.ncbi.nlm.nih.gov/1625138
1059. Upadhyay, J., et al. Impact of prenatal diagnosis on the morbidity associated with ureterocele
management. J Urol, 2002. 167: 2560.
https://pubmed.ncbi.nlm.nih.gov/11992089
1060. Ellerker, A.G. The extravesical ectopic ureter. Br J Surg, 1958. 45: 344.
https://pubmed.ncbi.nlm.nih.gov/13536326
1061. Pfister, C., et al. The value of endoscopic treatment for ureteroceles during the neonatal period.
J Urol, 1998. 159: 1006.
https://pubmed.ncbi.nlm.nih.gov/9474217
1062. Kwatra, N., et al. Scintigraphic features of duplex kidneys on DMSA renal cortical scans. Pediatr
Radiol, 2013. 43: 1204.
https://pubmed.ncbi.nlm.nih.gov/23385361

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 165


1063. Meneghesso, D., et al. Clinico-pathological correlation in duplex system ectopic ureters and
ureteroceles: can preoperative work-up predict renal histology? Pediatr Surg Int, 2012. 28: 309.
https://pubmed.ncbi.nlm.nih.gov/22127487
1064. Kocyigit, A., et al. Efficacy of magnetic resonance urography in detecting renal scars in children with
vesicoureteral reflux. Pediatr Nephrol, 2014. 29: 1215.
https://pubmed.ncbi.nlm.nih.gov/24500707
1065. Khrichenko, D., et al., Intra- and inter-observer variability of functional MR urography (fMRU)
assessment in children, In: Pediatr Radiol. 2016. p. 666.
1066. Bellah, R.D., et al. Ureterocele eversion with vesicoureteral reflux in duplex kidneys: findings at
voiding cystourethrography. AJR Am J Roentgenol, 1995. 165: 409.
https://pubmed.ncbi.nlm.nih.gov/7618568
1067. Carrico, C., et al. Incontinence due to an infrasphincteric ectopic ureter: why the delay in diagnosis
and what the radiologist can do about it. Pediatr Radiol, 1998. 28: 942.
https://pubmed.ncbi.nlm.nih.gov/9880638
1068. Ehammer, T., et al. High resolution MR for evaluation of lower urogenital tract malformations in
infants and children: feasibility and preliminary experiences. Eur J Radiol, 2011. 78: 388.
https://pubmed.ncbi.nlm.nih.gov/20138451
1069. Sumfest, J.M., et al. Pseudoureterocele: potential for misdiagnosis of an ectopic ureter as a
ureterocele. Br J Urol, 1995. 75: 401.
https://pubmed.ncbi.nlm.nih.gov/7735809
1070. Figueroa, V.H., et al. Utility of MR urography in children suspected of having ectopic ureter. Pediatr
Radiol, 2014. 44: 956.
https://pubmed.ncbi.nlm.nih.gov/24535117
1071. Beganovic, A., et al. Ectopic ureterocele: long-term results of open surgical therapy in 54 patients.
J Urol, 2007. 178: 251.
https://pubmed.ncbi.nlm.nih.gov/17499769
1072. Byun, E., et al. A meta-analysis of surgical practice patterns in the endoscopic management of
ureteroceles. J Urol, 2006. 176: 1871.
https://pubmed.ncbi.nlm.nih.gov/16945677
1073. Chertin, B., et al. Endoscopic treatment of vesicoureteral reflux associated with ureterocele. J Urol,
2007. 178: 1594.
https://pubmed.ncbi.nlm.nih.gov/17707044
1074. Decter, R.M., et al. Individualized treatment of ureteroceles. J Urol, 1989. 142: 535.
https://pubmed.ncbi.nlm.nih.gov/2746775
1075. Husmann, D., et al. Management of ectopic ureterocele associated with renal duplication: a
comparison of partial nephrectomy and endoscopic decompression. J Urol, 1999. 162: 1406.
https://pubmed.ncbi.nlm.nih.gov/10492225
1076. Castagnetti, M., et al. Management of duplex system ureteroceles in neonates and infants. Nat Rev
Urol, 2009. 6: 307.
https://pubmed.ncbi.nlm.nih.gov/19498409
1077. Monfort, G., et al. [Simplified treatment of ureteroceles]. Chir Pediatr, 1985. 26: 26.
https://pubmed.ncbi.nlm.nih.gov/3995671
1078. Sander, J.C., et al. Outcomes of endoscopic incision for the treatment of ureterocele in children at a
single institution. J Urol, 2015. 193: 662.
https://pubmed.ncbi.nlm.nih.gov/25167992
1079. Han, M.Y., et al. Indications for nonoperative management of ureteroceles. J Urol, 2005. 174: 1652.
https://pubmed.ncbi.nlm.nih.gov/16148674
1080. Mariyappa, B., et al. Management of duplex-system ureterocele. J Paediatr Child Health, 2014.
50: 96.
https://pubmed.ncbi.nlm.nih.gov/24372828
1081. Adorisio, O., et al. Effectiveness of primary endoscopic incision in treatment of ectopic ureterocele
associated with duplex system. Urology, 2011. 77: 191.
https://pubmed.ncbi.nlm.nih.gov/21168903
1082. DeFoor, W., et al. Ectopic ureterocele: clinical application of classification based on renal unit
jeopardy. J Urol, 2003. 169: 1092.
https://pubmed.ncbi.nlm.nih.gov/12576859
1083. Jesus, L.E., et al. Clinical evolution of vesicoureteral reflux following endoscopic puncture in children
with duplex system ureteroceles. J Urol, 2011. 186: 1455.
https://pubmed.ncbi.nlm.nih.gov/21862045

166 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


1084. Husmann, D.A., et al. Ureterocele associated with ureteral duplication and a nonfunctioning upper
pole segment: management by partial nephroureterectomy alone. J Urol, 1995. 154: 723.
https://pubmed.ncbi.nlm.nih.gov/7609163
1085. Gran, C.D., et al. Primary lower urinary tract reconstruction for nonfunctioning renal moieties
associated with obstructing ureteroceles. J Urol, 2005. 173: 198.
https://pubmed.ncbi.nlm.nih.gov/15592074
1086. Gander, R., et al. Evaluation of the Initial Treatment of Ureteroceles. Urology, 2016. 89: 113.
https://pubmed.ncbi.nlm.nih.gov/26674749
1087. Pohl, H.G. Recent advances in the management of ureteroceles in infants and children: Why less
may be more. Curr Opin Urol, 2011. 21: 322.
https://pubmed.ncbi.nlm.nih.gov/21519275
1088. Biles, M.J., et al. Innovation in Robotics and Pediatric Urology: Robotic Ureteroureterostomy for
Duplex Systems with Ureteral Ectopia. J Endourol, 2016. 30: 1041.
https://pubmed.ncbi.nlm.nih.gov/27542552
1089. Castagnetti, M., et al. Dismembered extravesical reimplantation of dilated upper pole ectopic
ureters in duplex systems. J Pediatr Surg, 2013. 48: 459.
https://pubmed.ncbi.nlm.nih.gov/23414887
1090. Esposito, C., et al. A comparison between laparoscopic and retroperitoneoscopic approach for
partial nephrectomy in children with duplex kidney: a multicentric survey. World J Urol, 2016. 34: 939.
https://pubmed.ncbi.nlm.nih.gov/26577623
1091. Herz, D., et al. Robot-assisted laparoscopic management of duplex renal anomaly: Comparison of
surgical outcomes to traditional pure laparoscopic and open surgery. J Pediatr Urol, 2016. 12: 44.e1.
https://pubmed.ncbi.nlm.nih.gov/26443241
1092. Cohen, S.A., et al. Examining trends in the treatment of ureterocele yields no definitive solution.
J Pediatr Surg, 2015. 11: 29.e1.
https://pubmed.ncbi.nlm.nih.gov/25459387
1093. Roy Choudhury, S., et al. Spectrum of ectopic ureters in children. Pediatr Surg Int, 2008. 24: 819.
https://pubmed.ncbi.nlm.nih.gov/18463883
1094. Lee, P.A., et al. Consensus statement on management of intersex disorders. International
Consensus Conference on Intersex. Pediatrics, 2006. 118: e488.
https://pubmed.ncbi.nlm.nih.gov/16882788
1095. Parliamentary Assembly, Council of Europe. Promoting the human rights of and eliminating
discrimination against intersex people. 2017.
http://assembly.coe.int/nw/xml/XRef/Xref-DocDetails-en.asp?FileId=24027
1096. Wolffenbuttel, K.P., et al. Gonadal dysgenesis in disorders of sex development: Diagnosis and
surgical management. J Pediatr Urol, 2016. 12: 411.
https://pubmed.ncbi.nlm.nih.gov/27769830
1097. Maggi, M., et al. Standard operating procedures: pubertas tarda/delayed puberty--male. J Sex Med,
2013. 10: 285.
https://pubmed.ncbi.nlm.nih.gov/22376050
1098. Wales, J.K. Disordered pubertal development. Arch Dis Child Educ Pract Ed, 2012. 97: 9.
https://pubmed.ncbi.nlm.nih.gov/21278425
1099. Lee, P.A., et al. Global Disorders of Sex Development Update since 2006: Perceptions, Approach
and Care. Horm Res Paediatr, 2016. 85: 158.
https://pubmed.ncbi.nlm.nih.gov/26820577
1100. Feldman, K.W., et al. Fetal phallic growth and penile standards for newborn male infants. J Pediatr,
1975. 86: 395.
https://pubmed.ncbi.nlm.nih.gov/1113226
1101. Creighton, S., et al. Medical photography: ethics, consent and the intersex patient. BJU Int, 2002.
89: 67.
https://pubmed.ncbi.nlm.nih.gov/11849163
1102. Biswas, K., et al. Imaging in intersex disorders. J Pediatr Endocrinol Metab, 2004. 17: 841.
https://pubmed.ncbi.nlm.nih.gov/15270401
1103. Wright, N.B., et al. Imaging children with ambiguous genitalia and intersex states. Clin Radiol, 1995.
50: 823.
https://pubmed.ncbi.nlm.nih.gov/8536391
1104. Chertin, B., et al. The use of laparoscopy in intersex patients. Pediatr Surg Int, 2006. 22: 405.
https://pubmed.ncbi.nlm.nih.gov/16521001
1105. Denes, F.T., et al. Laparoscopic management of intersexual states. Urol Clin North Am, 2001. 28: 31.
https://pubmed.ncbi.nlm.nih.gov/11277066

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 167


1106. Timing of elective surgery on the genitalia of male children with particular reference to the risks,
benefits, and psychological effects of surgery and anesthesia. American Academy of Pediatrics.
Pediatrics, 1996. 97: 590.
https://pubmed.ncbi.nlm.nih.gov/8632952
1107. Mouriquand, P., et al. The ESPU/SPU standpoint on the surgical management of Disorders of Sex
Development (DSD). J Pediatr Urol, 2014. 10: 8.
https://pubmed.ncbi.nlm.nih.gov/24528671
1108. Urology., E.S.f.P. Open letter to the Council of Europe. 2018.
Wolffenbuttel K.P. et al. Open letter to the Council of Europe. J Pediatr Urol, 2018. 14: 4.
https://www.jpurol.com/article/S1477-5131(18)30060-3/abstract
1109. van der Zwan, Y.G., et al. Gonadal maldevelopment as risk factor for germ cell cancer: towards a
clinical decision model. Eur Urol, 2015. 67: 692.
https://pubmed.ncbi.nlm.nih.gov/25240975
1110. Cools, M., et al. Germ cell tumors in the intersex gonad: old paths, new directions, moving frontiers.
Endocr Rev, 2006. 27: 468.
https://pubmed.ncbi.nlm.nih.gov/16735607
1111. Berte, N., et al. Long-term renal outcome in infants with congenital lower urinary tract obstruction.
Prog Urol, 2018. 28: 596.
https://pubmed.ncbi.nlm.nih.gov/29980359
1112. Malin, G., et al. Congenital lower urinary tract obstruction: a population-based epidemiological
study. Bjog, 2012. 119: 1455.
https://pubmed.ncbi.nlm.nih.gov/22925164
1113. Ruano, R., et al. Lower urinary tract obstruction: fetal intervention based on prenatal staging. Pediatr
Nephrol, 2017. 32: 1871.
https://pubmed.ncbi.nlm.nih.gov/28730376
1114. Johnson, M.P., et al. Natural History of Fetal Lower Urinary Tract Obstruction with Normal Amniotic
Fluid Volume at Initial Diagnosis. Fetal Diagn Ther, 2017.
https://pubmed.ncbi.nlm.nih.gov/224239
1115. Fontanella, F., et al. Fetal megacystis: a lot more than LUTO. Ultrasound Obstet Gynecol, 2019.
53: 779.
https://pubmed.ncbi.nlm.nih.gov/30043466
1116. Taghavi, K., et al. Fetal megacystis: A systematic review. J Pediatr Urol, 2017. 13: 7.
https://pubmed.ncbi.nlm.nih.gov/27889224
1117. Fontanella, F., et al. Antenatal staging of congenital lower urinary tract obstruction. Ultrasound
Obstet Gynecol, 2019. 53: 520.
https://pubmed.ncbi.nlm.nih.gov/29978555
1118. Chen, L., et al. Outcomes in fetuses diagnosed with megacystis: Systematic review and meta-
analysis. Eur J Obstet Gynecol Reprod Biol, 2019. 233: 120.
https://pubmed.ncbi.nlm.nih.gov/30594021
1119. Hennus, P.M., et al. A systematic review on the accuracy of diagnostic procedures for infravesical
obstruction in boys. PLoS One, 2014. 9: e85474.
https://pubmed.ncbi.nlm.nih.gov/24586242
1120. Hodges, S.J., et al. Posterior urethral valves. Sci World J, 2009. 9: 1119.
https://pubmed.ncbi.nlm.nih.gov/19838598
1121. Thakkar, D., et al. Epidemiology and demography of recently diagnosed cases of posterior urethral
valves. Pediatr Res, 2014. 76: 560.
https://pubmed.ncbi.nlm.nih.gov/25198372
1122. Young, H.H., et al. Congenital obstruction of the posterior urethra. J Urol, 3: 289-365, 1919. J Urol,
2002. 167: 265.
https://pubmed.ncbi.nlm.nih.gov/11743334
1123. Rosenfeld, B., et al. Type III posterior urethral valves: presentation and management. J Pediatr Surg,
1994. 29: 81.
https://pubmed.ncbi.nlm.nih.gov/8120770
1124. Stephens, F.D., et al. Pathogenesis of the prune belly syndrome. J Urol, 1994. 152: 2328.
https://pubmed.ncbi.nlm.nih.gov/7966734
1125. Roy, S., et al. [Contribution of ultrasound signs for the prenatal diagnosis of posterior urethral
valves: Experience of 3years at the maternity of the Bicetre Hospital]. J Gynecol Obstet Biol Reprod
(Paris), 2016. 45: 478.
https://pubmed.ncbi.nlm.nih.gov/25980903

168 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


1126. Cheung, K.W., et al. Congenital urinary tract obstruction. Best Pract Res Clin Obstet Gynaecol,
2019. 58: 78.
https://pubmed.ncbi.nlm.nih.gov/30819578
1127. Kajbafzadeh, A.M., et al. Comparison of magnetic resonance urography with ultrasound studies in
detection of fetal urogenital anomalies. J Pediatr Urol, 2008. 4: 32.
https://pubmed.ncbi.nlm.nih.gov/18631889
1128. Pico, H., et al. Contribution of the foetal uro-MRI in the prenatal diagnosis of uronephropathies.
Diagn Interv Imaging, 2014. 95: 573.
https://pubmed.ncbi.nlm.nih.gov/24637205
1129. Calvo-Garcia, M.A. Imaging Evaluation of Fetal Megacystis: How Can Magnetic Resonance Imaging
Help? Semin Ultrasound CT MR, 2015. 36: 537.
https://pubmed.ncbi.nlm.nih.gov/26614135
1130. Churchill, B.M., et al. Emergency treatment and long-term follow-up of posterior urethral valves. Urol
Clin North Am, 1990. 17: 343.
https://pubmed.ncbi.nlm.nih.gov/2186540
1131. Hoover, D.L., et al. Posterior urethral valves, unilateral reflux and renal dysplasia: a syndrome. J Urol,
1982. 128: 994.
https://pubmed.ncbi.nlm.nih.gov/7176067
1132. Rittenberg, M.H., et al. Protective factors in posterior urethral valves. J Urol, 1988. 140: 993.
https://pubmed.ncbi.nlm.nih.gov/3139895
1133. Cuckow, P.M., et al. Long-term renal function in the posterior urethral valves, unilateral reflux and
renal dysplasia syndrome. J Urol, 1997. 158: 1004.
https://pubmed.ncbi.nlm.nih.gov/9258130
1134. Kleppe, S., et al. Impact of prenatal urinomas in patients with posterior urethral valves and postnatal
renal function. J Perinat Med, 2006. 34: 425.
https://pubmed.ncbi.nlm.nih.gov/16965232
1135. Heikkila, J., et al. Posterior Urethral Valves are Often Associated With Cryptorchidism and Inguinal
Hernias. J Urol, 2008. 180: 715.
https://pubmed.ncbi.nlm.nih.gov/18554641
1136. Wong, J., et al. Why do undescended testes and posterior urethral valve occur together? Pediatr
Surg Int, 2016. 32: 509.
https://pubmed.ncbi.nlm.nih.gov/27072813
1137. Fontanella, F., et al. Fetal megacystis: prediction of spontaneous resolution and outcome.
Ultrasound Obstet Gynecol, 2017. 50: 458.
https://pubmed.ncbi.nlm.nih.gov/28133847
1138. Kagan, K.O., et al. The 11-13-week scan: diagnosis and outcome of holoprosencephaly,
exomphalos and megacystis. Ultrasound Obstet Gynecol, 2010. 36: 10.
https://pubmed.ncbi.nlm.nih.gov/20564304
1139. Brennan, S., et al. Evaluation of fetal kidney growth using ultrasound: A systematic review. Eur
J Radiol, 2017. 96: 55.
https://pubmed.ncbi.nlm.nih.gov/29103476
1140. Freedman, A.L., et al. Fetal therapy for obstructive uropathy: past, present.future? Pediatr Nephrol,
2000. 14: 167.
https://pubmed.ncbi.nlm.nih.gov/10684370
1141. Abdennadher, W., et al. Fetal urine biochemistry at 13-23 weeks of gestation in lower urinary tract
obstruction: criteria for in-utero treatment. Ultrasound Obstet Gynecol, 2015. 46: 306.
https://pubmed.ncbi.nlm.nih.gov/25412852
1142. Ibirogba, E.R., et al. Fetal lower urinary tract obstruction: What should we tell the prospective
parents? Prenat Diagn, 2020. 40: 661.
https://pubmed.ncbi.nlm.nih.gov/32065667
1143. Dębska, M., et al. Early vesico-amniotic shunting - does it change the prognosis in fetal lower
urinary tract obstruction diagnosed in the first trimester? Ginekol Pol, 2017. 88: 486.
https://pubmed.ncbi.nlm.nih.gov/29057434
1144. Morris, R.K., et al. Percutaneous vesicoamniotic shunting versus conservative management for fetal
lower urinary tract obstruction (PLUTO): a randomised trial. Lancet, 2013. 382: 1496.
https://pubmed.ncbi.nlm.nih.gov/23953766
1145. Nassr, A.A., et al. Effectiveness of vesicoamniotic shunt in fetuses with congenital lower urinary tract
obstruction: an updated systematic review and meta-analysis. Ultrasound Obstet Gynecol, 2017.
49: 696.
https://pubmed.ncbi.nlm.nih.gov/27270578

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 169


1146. Sananes, N., et al. Urological fistulas after fetal cystoscopic laser ablation of posterior urethral
valves: surgical technical aspects. Ultrasound Obstet Gynecol, 2015. 45: 183.
https://pubmed.ncbi.nlm.nih.gov/24817027
1147. Dębska, M., et al. Balloon catheterization in fetal lower urinary tract obstruction: an observational
study of 10 fetuses. Ultrasound Obstet Gynecol, 2019.
https://pubmed.ncbi.nlm.nih.gov/31763721
1148. Babu, R., et al. Early outcome following diathermy versus cold knife ablation of posterior urethral
valves. J Pediatr Surg, 2013. 9: 7.
https://pubmed.ncbi.nlm.nih.gov/22417679
1149. Sarhan, O., et al. Surgical complications of posterior urethral valve ablation: 20 years experience.
J Pediatric Surg, 2010. 45: 2222.
https://pubmed.ncbi.nlm.nih.gov/21034948
1150. Shirazi, M., et al. Which patients are at higher risk for residual valves after posterior urethral valve
ablation? Korean J Urol, 2014. 55: 64.
https://pubmed.ncbi.nlm.nih.gov/24466400
1151. Nawaz, G., et al. Justification For Re-Look Cystoscopy After Posterior Urethral Valve Fulguration.
J Ayub Med Coll Abbottabad : JAMC, 2017. 29: 30.
https://pubmed.ncbi.nlm.nih.gov/28712168
1152. Smeulders, N., et al. The predictive value of a repeat micturating cystourethrogram for remnant
leaflets after primary endoscopic ablation of posterior urethral valves. J Pediatr Urol, 2011. 7: 203.
https://pubmed.ncbi.nlm.nih.gov/20537589
1153. Krahn, C.G., et al. Cutaneous vesicostomy in the young child: indications and results. Urology, 1993.
41: 558.
https://pubmed.ncbi.nlm.nih.gov/8516992
1154. Sharifiaghdas, F., et al. Can transient resting of the bladder with vesicostomy reduce the need for a
major surgery in some patients? J Pediatr Urol, 2019. 15: 379.e1.
https://pubmed.ncbi.nlm.nih.gov/31060966
1155. Kim, Y.H., et al. Comparative urodynamic findings after primary valve ablation, vesicostomy or
proximal diversion. J Urol, 1996. 156: 673.
https://pubmed.ncbi.nlm.nih.gov/8683757
1156. Podesta, M., et al. Bladder function associated with posterior urethral valves after primary valve
ablation or proximal urinary diversion in children and adolescents. J Urol, 2002. 168: 1830.
https://pubmed.ncbi.nlm.nih.gov/12352370
1157. Chua, M.E., et al. Impact of Adjuvant Urinary Diversion versus Valve Ablation Alone on Progression
from Chronic to End Stage Renal Disease in Posterior Urethral Valves: A Single Institution 15-Year
Time-to-Event Analysis. J Urol, 2018.
https://pubmed.ncbi.nlm.nih.gov/29061539
1158. Novak, M.E., et al. Single-stage reconstruction of urinary tract after loop cutaneous ureterostomy.
Urology, 1978. 11: 134.
https://pubmed.ncbi.nlm.nih.gov/628990
1159. Sober, I. Pelvioureterostomy-en-Y. J Urol, 1972. 107: 473.
https://pubmed.ncbi.nlm.nih.gov/5010719
1160. Williams, D.I., et al. Ring ureterostomy. Br J Urol, 1975. 47: 789.
https://pubmed.ncbi.nlm.nih.gov/1222345
1161. Ghanem, M.A., et al. Long-term followup of bilateral high (sober) urinary diversion in patients with
posterior urethral valves and its effect on bladder function. J Urol, 2005. 173: 1721.
https://pubmed.ncbi.nlm.nih.gov/15821568
1162. Scott, J.E. Management of congenital posterior urethral valves. Br J Urol, 1985. 57: 71.
https://pubmed.ncbi.nlm.nih.gov/3971107
1163. Mukherjee, S., et al. What is the effect of circumcision on risk of urinary tract infection in boys with
posterior urethral valves? J Pediatr Surg, 2009. 44: 417.
https://pubmed.ncbi.nlm.nih.gov/19231547
1164. Casey, J.T., et al. Early administration of oxybutynin improves bladder function and clinical
outcomes in newborns with posterior urethral valves. J Urol, 2012. 188: 1516.
https://pubmed.ncbi.nlm.nih.gov/22910256
1165. Cozzi, D.A., et al. Posterior urethral valves: relationship between vesicoureteral reflux and renal
function. Urology, 2011. 77: 1209.
https://pubmed.ncbi.nlm.nih.gov/21109298

170 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


1166. Heikkila, J., et al. Long-term risk of end stage renal disease in patients with posterior urethral valves.
J Urol, 2011. 186: 2392.
https://pubmed.ncbi.nlm.nih.gov/22014822
1167. Bellinger, M.F. Ureterocystoplasty: a unique method for vesical augmentation in children. J Urol,
1993. 149: 811.
https://pubmed.ncbi.nlm.nih.gov/8455246
1168. Koff, S.A., et al. The valve bladder syndrome: pathophysiology and treatment with nocturnal bladder
emptying. J Urol, 2002. 167: 291.
https://pubmed.ncbi.nlm.nih.gov/11743343
1169. Nguyen, M.T., et al. Overnight catheter drainage in children with poorly compliant bladders improves
post-obstructive diuresis and urinary incontinence. J Urol, 2005. 174: 1633.
https://pubmed.ncbi.nlm.nih.gov/16148670
1170. Holmdahl, G. Bladder dysfunction in boys with posterior urethral valves. Scand J Urol Nephrol
Suppl, 1997. 188: 1.
https://pubmed.ncbi.nlm.nih.gov/9458522
1171. Neel, K.F. Feasibility and outcome of clean intermittent catheterization for children with sensate
urethra. Can Urol Assoc J, 2010. 4: 403.
https://pubmed.ncbi.nlm.nih.gov/21191500
1172. King, T., et al. Mitrofanoff for valve bladder syndrome: Effect on urinary tract and renal function.
J Urol, 2014. 191: 1517.
https://pubmed.ncbi.nlm.nih.gov/24679888
1173. Coleman, R., et al. Nadir creatinine in posterior urethral valves: How high is low enough? J Pediatr
Surg, 2015. 11: 356.
https://pubmed.ncbi.nlm.nih.gov/26292912
1174. Sarhan, O., et al. Prognostic value of serum creatinine levels in children with posterior urethral
valves treated by primary valve ablation. Urology, 2009. 74: S267.
https://pubmed.ncbi.nlm.nih.gov/19581129
1175. Akdogan, B., et al. Significance of age-specific creatinine levels at presentation in posterior urethral
valve patients. J Pediatr Urol, 2006. 2: 446.
https://pubmed.ncbi.nlm.nih.gov/18947654
1176. Lemmens, A.S., et al. Population-specific serum creatinine centiles in neonates with posterior
urethral valves already predict long-term renal outcome. J Matern Fetal Neonatal Med, 2015.
28: 1026.
https://pubmed.ncbi.nlm.nih.gov/25000449
1177. Odeh, R., et al. Predicting Risk of Chronic Kidney Disease in Infants and Young Children at
Diagnosis of Posterior Urethral Valves: Initial Ultrasound Kidney Characteristics and Validation of
Parenchymal Area as Forecasters of Renal Reserve. J Urol, 2016. 196: 862.
https://pubmed.ncbi.nlm.nih.gov/27017936
1178. Jalkanen, J., et al. Controlled Outcomes for Achievement of Urinary Continence among Boys
Treated for Posterior Urethral Valves. J Urol, 2016. 196: 213.
https://pubmed.ncbi.nlm.nih.gov/26964916
1179. Smith, G.H., et al. The long-term outcome of posterior urethral valves treated with primary valve
ablation and observation. J Urol, 1996. 155: 1730.
https://pubmed.ncbi.nlm.nih.gov/8627873
1180. Concodora, C.W., et al. The Role of Video Urodynamics in the Management of the Valve Bladder.
Curr Urol Rep, 2017. 18: 24.
https://pubmed.ncbi.nlm.nih.gov/28233231
1181. Capitanucci, M.L., et al. Long-term bladder function followup in boys with posterior urethral valves:
Comparison of noninvasive vs invasive urodynamic studies. J Urol, 2012. 188: 953.
https://pubmed.ncbi.nlm.nih.gov/22819111
1182. Kim, Y.H., et al. Management of posterior urethral valves on the basis of urodynamic findings. J Urol,
1997. 158: 1011.
https://pubmed.ncbi.nlm.nih.gov/9258132
1183. Misseri, R., et al. Myogenic failure in posterior urethral valve disease: real or imagined? J Urol, 2002.
168: 1844.
https://pubmed.ncbi.nlm.nih.gov/12352373
1184. Abraham, M.K., et al. Role of alpha adrenergic blocker in the management of posterior urethral
valves. Pediatr Surg Int, 2009. 25: 1113.
https://pubmed.ncbi.nlm.nih.gov/19727771

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 171


1185. Skenazy, J., et al. 1618 Alpha adrenergic blockade in neonates with posterior urethral valves. J Urol,
2012. 187: e654.
https://www.sciencedirect.com/science/article/abs/pii/S0022534712017752
1186. DeFoor, W., et al. Risk Factors for End Stage Renal Disease in Children With Posterior Urethral
Valves. J Urol, 2008. 180: 1705.
https://pubmed.ncbi.nlm.nih.gov/18708224
1187. Ansari, M.S., et al. Risk factors for progression to end-stage renal disease in children with posterior
urethral valves. J Pediatr Surg, 2010. 6: 261.
https://pubmed.ncbi.nlm.nih.gov/19833558
1188. Fine, M.S., et al. Posterior urethral valve treatments and outcomes in children receiving kidney
transplants. J Urol, 2011. 185: 2507.
https://pubmed.ncbi.nlm.nih.gov/21527196
1189. Kamal, M.M., et al. Impact of posterior urethral valves on pediatric renal transplantation: a single-
center comparative study of 297 cases. Pediatr Transplant, 2011. 15: 482.
https://pubmed.ncbi.nlm.nih.gov/21599816
1190. Lopez Pereira, P., et al. Long-term bladder function, fertility and sexual function in patients with
posterior urethral valves treated in infancy. J Pediatr Surg, 2013. 9: 38.
https://pubmed.ncbi.nlm.nih.gov/22154080
1191. Woodhouse, C.R., et al. Sexual function and fertility in patients treated for posterior urethral valves.
J Urol, 1989. 142: 586.
https://pubmed.ncbi.nlm.nih.gov/2746783
1192. Taskinen, S., et al. Effects of posterior urethral valves on long-term bladder and sexual function. Nat
Rev Urol, 2012. 9: 699.
https://pubmed.ncbi.nlm.nih.gov/23147930
1193. Arena, S., et al. Anterior urethral valves in children: an uncommon multipathogenic cause of
obstructive uropathy. Pediatr Surg Int, 2009. 25: 613.
https://pubmed.ncbi.nlm.nih.gov/19517125
1194. Firlit, R.S., et al. Obstructing anterior urethral valves in children. J Urol, 1978. 119: 819.
https://pubmed.ncbi.nlm.nih.gov/566334
1195. Zia-ul-Miraj, M. Anterior urethral valves: a rare cause of infravesical obstruction in children. J Pediatr
Surg, 2000. 35: 556.
https://pubmed.ncbi.nlm.nih.gov/10770380
1196. Routh, J.C., et al. Predicting renal outcomes in children with anterior urethral valves: a systematic
review. J Urol, 2010. 184: 1615.
https://pubmed.ncbi.nlm.nih.gov/20728183
1197. Adam, A., et al. Congenital anterior urethral diverticulum: antenatal diagnosis with subsequent
neonatal endoscopic management. Urology, 2015. 85: 914.
https://pubmed.ncbi.nlm.nih.gov/25704997
1198. Gupta, D.K., et al. Congenital anterior urethral diverticulum in children. Pediatr Surg Int, 2000. 16: 565.
https://pubmed.ncbi.nlm.nih.gov/11149395
1199. Rawat, J., et al. Congenital anterior urethral valves and diverticula: diagnosis and management in six
cases. Afr J Paediatr Surg, 2009. 6: 102.
https://pubmed.ncbi.nlm.nih.gov/19661640
1200. Cruz-Diaz, O., et al. Anterior urethral valves: not such a benign condition. Front Pediatr, 2013. 1: 35.
https://pubmed.ncbi.nlm.nih.gov/24400281
1201. Quoraishi, S.H., et al. Congenital anterior urethral diverticulum in a male teenager: a case report and
review of the literature. Case Rep Urol, 2011. 2011: 738638.
https://pubmed.ncbi.nlm.nih.gov/22606624
1202. Maizels, M., et al. Cowper’s syringocele: a classification of dilatations of Cowper’s gland duct based
upon clinical characteristics of 8 boys. J Urol, 1983. 129: 111.
https://pubmed.ncbi.nlm.nih.gov/6827661
1203. Melquist, J., et al. Current diagnosis and management of syringocele: a review. Int Braz J Urol,
2010. 36: 3.
https://pubmed.ncbi.nlm.nih.gov/20202229
1204. Campobasso, P., et al. Cowper’s syringocele: an analysis of 15 consecutive cases. Arch Dis Child,
1996. 75: 71.
https://pubmed.ncbi.nlm.nih.gov/8813875
1205. Bevers, R.F., et al. Cowper’s syringocele: symptoms, classification and treatment of an
unappreciated problem. J Urol, 2000. 163: 782.
https://pubmed.ncbi.nlm.nih.gov/10687976

172 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


1206. Dewan, P.A., et al. Congenital urethral obstruction: Cobb’s collar or prolapsed congenital obstructive
posterior urethral membrane (COPUM). Br J Urol, 1994. 73: 91.
https://pubmed.ncbi.nlm.nih.gov/8298906
1207. Nonomura, K., et al. Impact of congenital narrowing of the bulbar urethra (Cobb’s collar) and its
transurethral incision in children. Eur Urol, 1999. 36: 144.
https://pubmed.ncbi.nlm.nih.gov/10420036
1208. Gonzalez, R., et al. Urethral atresia: long-term outcome in 6 children who survived the neonatal
period. J Urol, 2001. 165: 2241.
https://pubmed.ncbi.nlm.nih.gov/11371953
1209. Passerini-Glazel, G., et al. The P.A.D.U.A. (progressive augmentation by dilating the urethra anterior)
procedure for the treatment of severe urethral hypoplasia. J Urol, 1988. 140: 1247.
https://pubmed.ncbi.nlm.nih.gov/2972844
1210. Freedman, A.L., et al. Long-term outcome in children after antenatal intervention for obstructive
uropathies. Lancet, 1999. 354: 374.
https://pubmed.ncbi.nlm.nih.gov/10437866
1211. Downs, R.A. Congenital polyps of the prostatic urethra. A review of the literature and report of two
cases. Br J Urol, 1970. 42: 76.
https://pubmed.ncbi.nlm.nih.gov/5435705
1212. Natsheh, A., et al. Fibroepithelial polyp of the bladder neck in children. Pediatr Surg Int, 2008. 24: 613.
https://pubmed.ncbi.nlm.nih.gov/18097674
1213. Akbarzadeh, A., et al. Congenital urethral polyps in children: report of 18 patients and review of
literature. J Pediatr Surg, 2014. 49: 835.
https://pubmed.ncbi.nlm.nih.gov/24851781
1214. Parrott, T.S., et al., The bladder and urethra, in Embryology for surgeons : the embryological basis
for the treatment of congenital anomalies 2nd ed., J.E. Skandalakis & S.W. Gray, Editors. 1994,
Williams & Wilkins: Baltimore.
1215. Atala, A., et al. Patent urachus and urachal cysts. Gellis & Kagan’s current pediatric therapy.
Philadelphia: WB Saunders, 1993: 386.
1216. Gearhart JP, J.R., Urachal abnormalities, in Campbell’s urology 7th edn., P.C. Walsh., A.B. Retik &
E.D. Vaughan, Editors. 1998, WB Saunders: Philadelphia.
1217. Moore, K.L., The urogenital system, in The Developing Human 3rd edn., K.L. Moore, Editor. 1982,
Elsevier Health Sciences: Philadelphia.
1218. Berman, S.M., et al. Urachal remnants in adults. Urology, 1988. 31: 17.
https://pubmed.ncbi.nlm.nih.gov/3122397
1219. Metwalli, Z.A., et al. Imaging features of intravesical urachal cysts in children. Pediatr Radiol, 2013.
43: 978.
https://pubmed.ncbi.nlm.nih.gov/23370693
1220. Yohannes, P., et al. Laparoscopic radical excision of urachal sinus. J Endourol, 2003. 17: 475.
https://pubmed.ncbi.nlm.nih.gov/14565877
1221. Robert, Y., et al. Urachal remnants: sonographic assessment. J Clin Ultrasound, 1996. 24: 339.
https://pubmed.ncbi.nlm.nih.gov/8873855
1222. Stopak, J.K., et al. Trends in surgical management of urachal anomalies. J Pediatr Surg, 2015.
50: 1334.
https://pubmed.ncbi.nlm.nih.gov/26227313
1223. Rubin., A.A., Handbook of Congenital Malformations. 1967, Philadelphia. PA.
1224. Nix, J.T., et al. Congenital patent urachus. J Urol, 1958. 79: 264.
https://pubmed.ncbi.nlm.nih.gov/13514877
1225. Blichert-Toft, M., et al. Congenital patient urachus and acquired variants. Diagnosis and treatment.
Review of the literature and report of five cases. Acta Chir Scand, 1971. 137: 807.
https://pubmed.ncbi.nlm.nih.gov/5148862
1226. Holten, I., et al. The ultrasonic diagnosis of urachal anomalies. Australas Radiol, 1996. 40: 2.
https://pubmed.ncbi.nlm.nih.gov/8838878
1227. Copp, H.L., et al. Clinical presentation and urachal remnant pathology: implications for treatment.
J Urol, 2009. 182: 1921.
https://pubmed.ncbi.nlm.nih.gov/19695622
1228. Galati, V., et al. Management of urachal remnants in early childhood. J Urol, 2008. 180: 1824.
https://pubmed.ncbi.nlm.nih.gov/18721938
1229. Lipskar, A.M., et al. Nonoperative management of symptomatic urachal anomalies. J Pediatr Surg,
2010. 45: 1016.
https://pubmed.ncbi.nlm.nih.gov/20438945

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 173


1230. Dethlefs, C.R., et al. Conservative management of urachal anomalies. J Pediatr Surg, 2019. 54: 1054.
https://pubmed.ncbi.nlm.nih.gov/30867097
1231. McCollum, M.O., et al. Surgical implications of urachal remnants: Presentation and management.
J Pediatr Surg, 2003. 38: 798.
https://pubmed.ncbi.nlm.nih.gov/12720197
1232. Yiee, J.H., et al. A diagnostic algorithm for urachal anomalies. J Pediatr Urol, 2007. 3: 500.
https://pubmed.ncbi.nlm.nih.gov/18947803
1233. Naiditch, J.A., et al. Current diagnosis and management of urachal remnants. J Pediatr Surg, 2013.
48: 2148.
https://pubmed.ncbi.nlm.nih.gov/24094971
1234. Ward, T.T., et al. Infected urachal remnants in the adult: case report and review. Clin Infect Dis, 1993.
16: 26.
https://pubmed.ncbi.nlm.nih.gov/8448315
1235. al-Hindawi, M.K., et al. Benign non-infected urachal cyst in an adult: review of the literature and a
case report. Br J Radiol, 1992. 65: 313.
https://pubmed.ncbi.nlm.nih.gov/1581788
1236. Choi, Y.J., et al. Urachal anomalies in children: a single center experience. Yonsei Med J, 2006.
47: 782.
https://pubmed.ncbi.nlm.nih.gov/17191305
1237. Ashley, R.A., et al. Urachal anomalies: a longitudinal study of urachal remnants in children and
adults. J Urol, 2007. 178: 1615.
https://pubmed.ncbi.nlm.nih.gov/17707039
1238. Lane, V. Congenital patent urachus associated with complete (hypospadiac( duplication of the
urethra and solitary crossed renal ectopia. J Urol, 1982. 127: 990.
https://pubmed.ncbi.nlm.nih.gov/7087008
1239. Suita, S., et al. Urachal remnants. Semin Pediatr Surg, 1996. 5: 107.
https://pubmed.ncbi.nlm.nih.gov/9138709
1240. Newman, B.M., et al. Advances in the management of infected urachal cysts. J Pediatr Surg, 1986.
21: 1051.
https://pubmed.ncbi.nlm.nih.gov/3794969
1241. Little, D.C., et al. Urachal anomalies in children: the vanishing relevance of the preoperative voiding
cystourethrogram. J Pediatr Surg, 2005. 40: 1874.
https://pubmed.ncbi.nlm.nih.gov/16338309
1242. Herr, H.W., et al. Urachal carcinoma: contemporary surgical outcomes. J Urol, 2007. 178: 74.
https://pubmed.ncbi.nlm.nih.gov/17499279
1243. Nogueras-Ocaña, M., et al. Urachal anomalies in children: surgical or conservative treatment?
J Pediatr Urol, 2014. 10: 522.
https://pubmed.ncbi.nlm.nih.gov/24321777
1244. Zieger, B., et al. Sonomorphology and involution of the normal urachus in asymptomatic newborns.
Pediatr Radiol, 1998. 28: 156.
https://pubmed.ncbi.nlm.nih.gov/9561533
1245. Sato, H., et al. The current strategy for urachal remnants. Pediatr Surg Int, 2015. 31: 581.
https://pubmed.ncbi.nlm.nih.gov/25896294
1246. Gregory, G.C., et al. Laparoscopic management of urachal cyst associated with umbilical hernia.
Hernia, 2011. 15: 93.
https://pubmed.ncbi.nlm.nih.gov/20069440
1247. Block, R.I., et al. Are anesthesia and surgery during infancy associated with altered academic
performance during childhood? Anesthesiology, 2012. 117: 494.
https://pubmed.ncbi.nlm.nih.gov/22801049
1248. DiMaggio, C., et al. Early childhood exposure to anesthesia and risk of developmental and
behavioral disorders in a sibling birth cohort. Anesth Analg, 2011. 113: 1143.
https://pubmed.ncbi.nlm.nih.gov/21415431
1249. Gleason, J.M., et al. A comprehensive review of pediatric urachal anomalies and predictive analysis
for adult urachal adenocarcinoma. J Urol, 2015. 193: 632.
https://pubmed.ncbi.nlm.nih.gov/25219697
1250. Sheldon, C.A., et al. Malignant urachal lesions. J Urol, 1984. 131: 1.
https://pubmed.ncbi.nlm.nih.gov/6361280
1251. Ueno, T., et al. Urachal anomalies: ultrasonography and management. J Pediatr Surg, 2003.
38: 1203.
https://pubmed.ncbi.nlm.nih.gov/12891493

174 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


1252. Molina, J.R., et al. Predictors of survival from urachal cancer: a Mayo Clinic study of 49 cases.
Cancer, 2007. 110: 2434.
https://pubmed.ncbi.nlm.nih.gov/17932892
1253. Ashley, R.A., et al. Urachal carcinoma: clinicopathologic features and long-term outcomes of an
aggressive malignancy. Cancer, 2006. 107: 712.
https://pubmed.ncbi.nlm.nih.gov/16826585
1254. Arora, H., et al. Diagnosis and Management of Urachal Anomalies in Children. Current Bladder
Dysfunction Reports, 2015. 10: 256.
https://link.springer.com/article/10.1007/s11884-015-0310-y
1255. Heuga, B., et al. [Management of urachal remnants in children: Is surgical excision mandatory?].
Prog Urol, 2015. 25: 603.
https://pubmed.ncbi.nlm.nih.gov/26094100
1256. Saltsman, J.A., et al. Urothelial neoplasms in pediatric and young adult patients: A large single-
center series. J Pediatr Surg, 2018. 53: 306.
https://pubmed.ncbi.nlm.nih.gov/29221636
1257. Chu, S., et al. Transitional Cell Carcinoma in the Pediatric Patient: A Review of the Literature.
Urology, 2016. 91: 175.
https://pubmed.ncbi.nlm.nih.gov/26802795
1258. Caione, P., et al. Nonmuscular Invasive Urothelial Carcinoma of the Bladder in Pediatric and Young
Adult Patients: Age-related Outcomes. Urology, 2017. 99: 215.
https://pubmed.ncbi.nlm.nih.gov/27450943
1259. Hoenig, D.M., et al. Transitional cell carcinoma of the bladder in the pediatric patient. J Urol, 1996.
156: 203.
https://pubmed.ncbi.nlm.nih.gov/8648805
1260. Kutarski, P.W., et al. Transitional cell carcinoma of the bladder in young adults. Br J Urol, 1993.
72: 749.
https://pubmed.ncbi.nlm.nih.gov/8281408
1261. Jaworski, D., et al. Diagnostic difficulties in cases of papillary urothelial neoplasm of low malignant
potential, urothelial proliferation of uncertain malignant potential, urothelial dysplasia and urothelial
papilloma: A review of current literature. Ann Diagn Pathol, 2019. 40: 182.
https://pubmed.ncbi.nlm.nih.gov/29395466
1262. Song, D., et al. Inflammatory myofibroblastic tumor of urinary bladder with severe hematuria: A Case
report and literature review. Medicine (Baltimore), 2019. 98: e13987.
https://pubmed.ncbi.nlm.nih.gov/30608442
1263. Teoh, J.Y., et al. Inflammatory myofibroblastic tumors of the urinary bladder: a systematic review.
Urology, 2014. 84: 503.
https://pubmed.ncbi.nlm.nih.gov/25168523
1264. Collin, M., et al. Inflammatory myofibroblastic tumour of the bladder in children: a review. J Pediatr
Urol, 2015. 11: 239.
https://pubmed.ncbi.nlm.nih.gov/25982020
1265. Wang, X., et al. Malignant Inflammatory Myofibroblastic Tumor of the Urinary Bladder in a 14-Year-
Old Boy. J Pediatr Hematol Oncol, 2015. 37: e402.
https://pubmed.ncbi.nlm.nih.gov/26207771
1266. Houben, C.H., et al. Inflammatory myofibroblastic tumor of the bladder in children: what can be
expected? Pediatr Surg Int, 2007. 23: 815.
https://pubmed.ncbi.nlm.nih.gov/17443333
1267. Alezra, E., et al. [Complete resolution of inflammatory myofibroblastic tumor of the bladder after
antibiotic therapy]. Arch Pediatr, 2016. 23: 612.
https://pubmed.ncbi.nlm.nih.gov/27102996
1268. Papali, A.C., et al. A review of pediatric glans malformations: a handy clinical reference. J Urol,
2008. 180: 1737.
https://pubmed.ncbi.nlm.nih.gov/18721953
1269. Eisner, B.H., et al. Pediatric penile tumors of mesenchymal origin. Urology, 2006. 68: 1327.
https://pubmed.ncbi.nlm.nih.gov/17169655
1270. Ealai, P.A., et al. Penile epidermal inclusion cyst: a rare location. BMJ Case Rep, 2015. 2015.
https://pubmed.ncbi.nlm.nih.gov/26290567
1271. De Mendonça, R.R., et al. Mucoid cyst of the penis: Case report and literature review. Can Urol
Assoc J, 2010. 4: E155.
https://pubmed.ncbi.nlm.nih.gov/21749810

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 175


1272. Syed, M.M.A., et al. Median raphe cyst of the penis: a case report and review of the literature. J Med
Case Rep, 2019. 13: 214.
https://pubmed.ncbi.nlm.nih.gov/31301740
1273. Shao, I.H., et al. Male median raphe cysts: serial retrospective analysis and histopathological
classification. Diagn Pathol, 2012. 7: 121.
https://pubmed.ncbi.nlm.nih.gov/22978603
1274. Sonthalia, S., et al. Smegma Pearls in Young Uncircumcised Boys. Pediatr Dermatol, 2016. 33: e186.
https://pubmed.ncbi.nlm.nih.gov/27071486
1275. Ramos, L.M., et al. Venous malformation of the glans penis: efficacy of treatment with
neodymium:yttruim-aluminum-garnet laser. Urology, 1999. 53: 779.
https://pubmed.ncbi.nlm.nih.gov/10197856
1276. Shah, S.D., et al. Rebound Growth of Infantile Hemangiomas After Propranolol Therapy. Pediatrics,
2016. 137.
https://pubmed.ncbi.nlm.nih.gov/26952504
1277. Dagur, G., et al. Unusual Glomus Tumor of the Penis. Curr Urol, 2016. 9: 113.
https://pubmed.ncbi.nlm.nih.gov/27867327
1278. Saito, T. Glomus tumor of the penis. Int J Urol, 2000. 7: 115.
https://pubmed.ncbi.nlm.nih.gov/10750892
1279. Dwosh, J., et al. Neurofibroma involving the penis in a child. J Urol, 1984. 132: 988.
https://pubmed.ncbi.nlm.nih.gov/6436512
1280. Taib, F., et al. Infantile fibrosarcoma of the penis in a 2-year-old boy. Urology, 2012. 80: 931.
https://pubmed.ncbi.nlm.nih.gov/22854139
1281. Hu, J., et al. Congenital primary penile teratoma in a child. Urology, 2014. 83: 1404.
https://pubmed.ncbi.nlm.nih.gov/24767514
1282. Smeltzer, D.M., et al. Primary lymphedema in children and adolescents: a follow-up study and
review. Pediatrics, 1985. 76: 206.
https://pubmed.ncbi.nlm.nih.gov/4022694
1283. Brorson, H. Adipose tissue in lymphedema: the ignorance of adipose tissue in lymphedema.
Lymphology, 2004. 37: 175.
https://pubmed.ncbi.nlm.nih.gov/15693531
1284. Schook, C.C., et al. Male genital lymphedema: clinical features and management in 25 pediatric
patients. J Pediatr Surg, 2014. 49: 1647.
https://pubmed.ncbi.nlm.nih.gov/25475811
1285. Vricella, G.J., et al. Granulomatous lymphangitis. J Urol, 2013. 190: 1052.
https://pubmed.ncbi.nlm.nih.gov/23773564
1286. Sackett, D.D., et al. Isolated penile lymphedema in an adolescent male: a case of metastatic
Crohn’s disease. J Pediatr Urol, 2012. 8: e55.
https://pubmed.ncbi.nlm.nih.gov/22507210
1287. Bolt, R.J., et al. Congenital lymphoedema of the genitalia. Eur J Pediatr, 1998. 157: 943.
https://pubmed.ncbi.nlm.nih.gov/9835443
1288. Dandapat, M.C., et al. Elephantiasis of the penis and scrotum. A review of 350 cases. Am J Surg,
1985. 149: 686.
https://pubmed.ncbi.nlm.nih.gov/3993854
1289. McAninch, J.W., et al. Renal reconstruction after injury. J Urol, 1991. 145: 932.
https://pubmed.ncbi.nlm.nih.gov/2016804
1290. McAleer, I.M., et al. Genitourinary trauma in the pediatric patient. Urology, 1993. 42: 563.
https://pubmed.ncbi.nlm.nih.gov/8236601
1291. Miller, R.C., et al. The incidental discovery of occult abdominal tumors in children following blunt
abdominal trauma. J Trauma, 1966. 6: 99.
https://pubmed.ncbi.nlm.nih.gov/5901856
1292. Moore, E.E., et al. Organ injury scaling: spleen, liver, and kidney. J Trauma, 1989. 29: 1664.
https://pubmed.ncbi.nlm.nih.gov/2593197
1293. Stalker, H.P., et al. The significance of hematuria in children after blunt abdominal trauma. AJR Am
J Roentgenol, 1990. 154: 569.
https://pubmed.ncbi.nlm.nih.gov/2106223
1294. Mee, S.L., et al. Radiographic assessment of renal trauma: a 10-year prospective study of patient
selection. J Urol, 1989. 141: 1095.
https://pubmed.ncbi.nlm.nih.gov/2709493

176 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


1295. Stein, J.P., et al. Blunt renal trauma in the pediatric population: indications for radiographic
evaluation. Urology, 1994. 44: 406.
https://pubmed.ncbi.nlm.nih.gov/8073555
1296. Gaither, T.W., et al. Missed Opportunities to Decrease Radiation Exposure in Children with Renal
Trauma. J Urol, 2018. 199: 552.
https://pubmed.ncbi.nlm.nih.gov/28899768
1297. LeeVan, E., et al. Management of pediatric blunt renal trauma: A systematic review. J Trauma Acute
Care Surg, 2016. 80: 519.
https://pubmed.ncbi.nlm.nih.gov/26713980
1298. Radmayr, C., et al. Blunt renal trauma in children: 26 years clinical experience in an alpine region.
Eur Urol, 2002. 42: 297.
https://pubmed.ncbi.nlm.nih.gov/12234516
1299. Hagedorn, J.C., et al. Pediatric blunt renal trauma practice management guidelines: Collaboration
between the Eastern Association for the Surgery of Trauma and the Pediatric Trauma Society.
J Trauma Acute Care Surg, 2019. 86: 916.
https://pubmed.ncbi.nlm.nih.gov/30741880
1300. Presti, J.C., Jr., et al. Ureteral and renal pelvic injuries from external trauma: diagnosis and
management. J Trauma, 1989. 29: 370.
https://pubmed.ncbi.nlm.nih.gov/2926851
1301. Mulligan, J.M., et al. Ureteropelvic junction disruption secondary to blunt trauma: excretory phase
imaging (delayed films) should help prevent a missed diagnosis. J Urol, 1998. 159: 67.
https://pubmed.ncbi.nlm.nih.gov/9400439
1302. al-Ali, M., et al. The late treatment of 63 overlooked or complicated ureteral missile injuries: the
promise of nephrostomy and role of autotransplantation. J Urol, 1996. 156: 1918.
https://pubmed.ncbi.nlm.nih.gov/8911355
1303. Fernandez Fernandez, A., et al. Blunt traumatic rupture of the high right ureter, repaired with
appendix interposition. Urol Int, 1994. 53: 97.
https://pubmed.ncbi.nlm.nih.gov/7801425
1304. Sivit, C.J., et al. CT diagnosis and localization of rupture of the bladder in children with blunt
abdominal trauma: significance of contrast material extravasation in the pelvis. AJR Am
J Roentgenol, 1995. 164: 1243.
https://pubmed.ncbi.nlm.nih.gov/7717239
1305. Hochberg, E., et al. Bladder rupture associated with pelvic fracture due to blunt trauma. Urology,
1993. 41: 531.
https://pubmed.ncbi.nlm.nih.gov/8516988
1306. Haas, C.A., et al. Limitations of routine spiral computerized tomography in the evaluation of bladder
trauma. J Urol, 1999. 162: 51.
https://pubmed.ncbi.nlm.nih.gov/10379738
1307. Volpe, M.A., et al. Is there a difference in outcome when treating traumatic intraperitoneal bladder
rupture with or without a suprapubic tube? J Urol, 1999. 161: 1103.
https://pubmed.ncbi.nlm.nih.gov/10081847
1308. Richardson, J.R., Jr., et al. Non-operative treatment of the ruptured bladder. J Urol, 1975. 114: 213.
https://pubmed.ncbi.nlm.nih.gov/1159910
1309. Cass, A.S., et al. Urethral injury due to external trauma. Urology, 1978. 11: 607.
https://pubmed.ncbi.nlm.nih.gov/675928
1310. Pokorny, M., et al. Urological injuries associated with pelvic trauma. J Urol, 1979. 121: 455.
https://pubmed.ncbi.nlm.nih.gov/439217
1311. Elliott, D.S., et al. Long-term followup and evaluation of primary realignment of posterior urethral
disruptions. J Urol, 1997. 157: 814.
https://pubmed.ncbi.nlm.nih.gov/9072573
1312. Boone, T.B., et al. Postpubertal genitourinary function following posterior urethral disruptions in
children. J Urol, 1992. 148: 1232.
https://pubmed.ncbi.nlm.nih.gov/1404642
1313. Koraitim, M.M. Posttraumatic posterior urethral strictures in children: a 20-year experience. J Urol,
1997. 157: 641.
https://pubmed.ncbi.nlm.nih.gov/8996388
1314. Baradaran, N., et al. Long-term follow-up of urethral reconstruction for blunt urethral injury at a
young age: urinary and sexual quality of life outcomes. J Pediatr Urol, 2019. 15: 224.e1.
https://pubmed.ncbi.nlm.nih.gov/30967356

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 177


1315. Avanoglu, A., et al. Posterior urethral injuries in children. Br J Urol, 1996. 77: 597.
https://pubmed.ncbi.nlm.nih.gov/8777627
1316. Nair, S.G., et al. Perioperative fluid and electrolyte management in pediatric patients. Indian
J Anaesth, 2004. 48: 355.
https://www.researchgate.net/publication/228596506
1317. Imura, K., et al. Perioperative nutrition and metabolism in pediatric patients. World J Surg, 2000.
24: 1498.
https://pubmed.ncbi.nlm.nih.gov/11193714
1318. Ward Platt, M.P., et al. The effects of anesthesia and surgery on metabolic homeostasis in infancy
and childhood. J Pediatr Surg, 1990. 25: 472.
https://pubmed.ncbi.nlm.nih.gov/2191106
1319. Andersson, H., et al. Introducing the 6-4-0 fasting regimen and the incidence of prolonged
preoperative fasting in children. Paediatr Anaesth, 2018. 28: 46.
https://pubmed.ncbi.nlm.nih.gov/29168341
1320. Frykholm, P., et al. Preoperative fasting in children: review of existing guidelines and recent
developments. Br J Anaesth, 2018. 120: 469.
https://pubmed.ncbi.nlm.nih.gov/29452803
1321. Andersson, H., et al. Low incidence of pulmonary aspiration in children allowed intake of clear fluids
until called to the operating suite. Paediatr Anaesth, 2015. 25: 770.
https://pubmed.ncbi.nlm.nih.gov/25940831
1322. Fawcett, W.J., et al. Pre-operative fasting in adults and children: clinical practice and guidelines.
Anaesthesia, 2019. 74: 83.
https://pubmed.ncbi.nlm.nih.gov/30500064
1323. Rove, K.O., et al. Enhanced recovery after surgery in children: Promising, evidence-based
multidisciplinary care. Paediatr Anaesth, 2018. 28: 482.
https://pubmed.ncbi.nlm.nih.gov/29752858
1324. Feld, L.G., et al. Clinical Practice Guideline: Maintenance Intravenous Fluids in Children. Pediatrics,
2018. 142.
https://pubmed.ncbi.nlm.nih.gov/30478247
1325. Sumpelmann, R., et al. Perioperative intravenous fluid therapy in children: guidelines from the
Association of the Scientific Medical Societies in Germany. Paediatr Anaesth, 2017. 27: 10.
https://pubmed.ncbi.nlm.nih.gov/27747968
1326. Pedraza Bermeo, A.M., et al. Risk factors for postobstructive diuresis in pediatric patients with
ureteropelvic junction obstruction, following open pyeloplasty in three high complexity institutions.
J Pediatr Urol, 2018. 14: 260.e1.
https://pubmed.ncbi.nlm.nih.gov/29501380
1327. Cheng, W., et al. Electrogastrographic changes in children who undergo day-surgery anesthesia.
J Pediatr Surg, 1999. 34: 1336.
https://pubmed.ncbi.nlm.nih.gov/10507424
1328. Chauvin, C., et al. Early postoperative oral fluid intake in paediatric day case surgery influences the
need for opioids and postoperative vomiting: a controlled randomized trialdagger. Br J Anaesth,
2017. 118: 407.
https://pubmed.ncbi.nlm.nih.gov/28203729
1329. Haid, B., et al. Enhanced Recovery after Surgery Protocol for Pediatric Urological Augmentation and
Diversion Surgery Using Small Bowel. J Urol, 2018. 200: 1100.
https://pubmed.ncbi.nlm.nih.gov/29886091
1330. Shinnick, J.K., et al. Enhancing recovery in pediatric surgery: a review of the literature. J Surg Res,
2016. 202: 165.
https://pubmed.ncbi.nlm.nih.gov/27083963
1331. Ivani, G., et al. Postoperative analgesia in infants and children: new developments. Minerva
Anestesiol, 2004. 70: 399.
https://pubmed.ncbi.nlm.nih.gov/15181422
1332. Prevention and management of pain and stress in the neonate. American Academy of Pediatrics.
Committee on Fetus and Newborn. Committee on Drugs. Section on Anesthesiology. Section on
Surgery. Canadian Paediatric Society. Fetus and Newborn Committee. Pediatrics, 2000. 105: 454.
https://pubmed.ncbi.nlm.nih.gov/10654977
1333. Anand, K.J. Consensus statement for the prevention and management of pain in the newborn. Arch
Pediatr Adolesc Med, 2001. 155: 173.
https://pubmed.ncbi.nlm.nih.gov/11177093

178 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


1334. Kain, Z.N., et al. Preoperative anxiety, postoperative pain, and behavioral recovery in young children
undergoing surgery. Pediatrics, 2006. 118: 651.
https://pubmed.ncbi.nlm.nih.gov/16882820
1335. Taddio, A., et al. Effect of neonatal circumcision on pain response during subsequent routine
vaccination. Lancet, 1997. 349: 599.
https://pubmed.ncbi.nlm.nih.gov/9057731
1336. Stapelkamp, C., et al. Assessment of acute pain in children: development of evidence-based
guidelines. Int J Evid Based Healthc, 2011. 9: 39.
https://pubmed.ncbi.nlm.nih.gov/21332662
1337. Cravero, J.P., et al. The Society for Pediatric Anesthesia recommendations for the use of opioids in
children during the perioperative period. Paediatr Anaesth, 2019. 29: 547.
https://pubmed.ncbi.nlm.nih.gov/30929307
1338. Jonas, D.A. Parent’s management of their child’s pain in the home following day surgery. J Child
Health Care, 2003. 7: 150.
https://pubmed.ncbi.nlm.nih.gov/14516009
1339. Woolf, C.J., et al. Preemptive analgesia--treating postoperative pain by preventing the establishment
of central sensitization. Anesth Analg, 1993. 77: 362.
https://pubmed.ncbi.nlm.nih.gov/8346839
1340. Kendall, M.C., et al. Regional anesthesia to ameliorate postoperative analgesia outcomes in
pediatric surgical patients: an updated systematic review of randomized controlled trials. Local Reg
Anesth, 2018. 11: 91.
https://pubmed.ncbi.nlm.nih.gov/30532585
1341. Cancer Pain Relief and Palliative Care in Children. 1998, World Health Organization: Geneva.
https://apps.who.int/iris/bitstream/10665/42001/1/9241545127.pdf
1342. Harbaugh, C.M., et al. Pediatric postoperative opioid prescribing and the opioid crisis. Curr Opin
Pediatr, 2019. 31: 378.
https://pubmed.ncbi.nlm.nih.gov/31090580
1343. Wong, I., et al. Opioid-sparing effects of perioperative paracetamol and nonsteroidal anti-
inflammatory drugs (NSAIDs) in children. Paediatr Anaesth, 2013. 23: 475.
https://pubmed.ncbi.nlm.nih.gov/23570544
1344. Martin, L.D., et al. A review of perioperative anesthesia and analgesia for infants: updates and
trends to watch. F1000Res, 2017. 6: 120.
https://pubmed.ncbi.nlm.nih.gov/28232869
1345. Cardona-Grau, D., et al. Reducing Opioid Prescriptions in Outpatient Pediatric Urological Surgery.
J Urol, 2019. 201: 1012.
https://pubmed.ncbi.nlm.nih.gov/30688774
1346. Vittinghoff, M., et al. Postoperative pain management in children: Guidance from the pain committee
of the European Society for Paediatric Anaesthesiology (ESPA Pain Management Ladder Initiative).
Paediatr Anaesth, 2018. 28: 493.
https://pubmed.ncbi.nlm.nih.gov/29635764
1347. Paix, B.R., et al. Circumcision of neonates and children without appropriate anaesthesia is
unacceptable practice. Anaesth Intensive Care, 2012. 40: 511.
https://pubmed.ncbi.nlm.nih.gov/22577918
1348. Grunau, R.E., et al. Demographic and therapeutic determinants of pain reactivity in very low birth
weight neonates at 32 Weeks’ postconceptional Age. Pediatrics, 2001. 107: 105.
https://pubmed.ncbi.nlm.nih.gov/11134442
1349. Zhu, C., et al. Analgesic efficacy and impact of caudal block on surgical complications of
hypospadias repair: a systematic review and meta-analysis. Reg Anesth Pain Med, 2019. 44: 259.
https://pubmed.ncbi.nlm.nih.gov/30700621
1350. Chan, K.H., et al. Comparison of Intraoperative and Early Postoperative Outcomes of Caudal vs
Dorsal Penile Nerve Blocks for Outpatient Penile Surgeries. Urology, 2018. 118: 164.
https://pubmed.ncbi.nlm.nih.gov/29122625
1351. Hermansson, O., et al. Local delivery of bupivacaine in the wound reduces opioid requirements after
intraabdominal surgery in children. Pediatr Surg Int, 2013. 29: 451.
https://pubmed.ncbi.nlm.nih.gov/23483343
1352. Hidas, G., et al. Application of continuous incisional infusion of local anesthetic after major pediatric
urological surgery: prospective randomized controlled trial. J Pediatr Surg, 2015. 50: 481.
https://pubmed.ncbi.nlm.nih.gov/25746712

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 179


1353. Chalmers, D.J., et al. Continuous local anesthetic infusion for children with spina bifida undergoing
major reconstruction of the lower urinary tract. J Pediatr Urol, 2015. 11: 72.e1.
https://pubmed.ncbi.nlm.nih.gov/25819374
1354. Hong, J.Y., et al. Fentanyl-sparing effect of acetaminophen as a mixture of fentanyl in intravenous
parent-/nurse-controlled analgesia after pediatric ureteroneocystostomy. Anesthesiology, 2010.
113: 672.
https://pubmed.ncbi.nlm.nih.gov/20693884
1355. Routh, J.C., et al. Ketorolac is underutilized after ureteral reimplantation despite reduced hospital
cost and reduced length of stay. Urology, 2010. 76: 9.
https://pubmed.ncbi.nlm.nih.gov/20138342
1356. Jo, Y.Y., et al. Ketorolac or fentanyl continuous infusion for post-operative analgesia in children
undergoing ureteroneocystostomy. Acta Anaesthesiol Scand, 2011. 55: 54.
https://pubmed.ncbi.nlm.nih.gov/21083540
1357. Kumar, R., et al. Dorsal lumbotomy incision for pediatric pyeloplasty--a good alternative. Pediatr
Surg Int, 1999. 15: 562.
https://pubmed.ncbi.nlm.nih.gov/10631734
1358. Hamill, J.K., et al. Rectus sheath and transversus abdominis plane blocks in children: a systematic
review and meta-analysis of randomized trials. Paediatr Anaesth, 2016. 26: 363.
https://pubmed.ncbi.nlm.nih.gov/26846889
1359. Narasimhan, P., et al. Comparison of caudal epidural block with paravertebral block for renal
surgeries in pediatric patients: A prospective randomised, blinded clinical trial. J Clin Anesth, 2019.
52: 105.
https://pubmed.ncbi.nlm.nih.gov/30243061
1360. Martin, L.D., et al. Comparison between epidural and opioid analgesia for infants undergoing major
abdominal surgery. Paediatr Anaesth, 2019. 29: 835.
https://pubmed.ncbi.nlm.nih.gov/31140664
1361. Freilich, D.A., et al. The effectiveness of aerosolized intraperitoneal bupivacaine in reducing
postoperative pain in children undergoing robotic-assisted laparoscopic pyeloplasty. J Pediatr Urol,
2008. 4: 337.
https://pubmed.ncbi.nlm.nih.gov/18790415
1362. Spinelli, G., et al. Pediatric anesthesia for minimally invasive surgery in pediatric urology. Transl
Pediatr, 2016. 5: 214.
https://pubmed.ncbi.nlm.nih.gov/27867842
1363. Menes, T., et al. Laparoscopy: searching for the proper insufflation gas. Surg Endosc, 2000. 14: 1050.
https://pubmed.ncbi.nlm.nih.gov/11116418
1364. McHoney, M., et al. Carbon dioxide elimination during laparoscopy in children is age dependent.
J Pediatr Surg, 2003. 38: 105.
https://pubmed.ncbi.nlm.nih.gov/12592630
1365. Peters, C.A. Complications in pediatric urological laparoscopy: results of a survey. J Urol, 1996. 155:
1070.
https://pubmed.ncbi.nlm.nih.gov/8583567
1366. Passerotti, C.C., et al. Patterns and predictors of laparoscopic complications in pediatric urology:
the role of ongoing surgical volume and access techniques. J Urol, 2008. 180: 681.
https://pubmed.ncbi.nlm.nih.gov/18554647
1367. Zhou, R., et al. Abdominal Wall Elasticity of Children during Pneumoperitoneum. J Pediatr Surg, 2019.
https://pubmed.ncbi.nlm.nih.gov/31307782
1368. Sureka, S.K., et al. Safe and optimal pneumoperitoneal pressure for transperitoneal laparoscopic
renal surgery in infant less than 10 kg, looked beyond intraoperative period: A prospective
randomized study. J Pediatr Urol, 2016. 12: 281.e1.
https://pubmed.ncbi.nlm.nih.gov/27751832
1369. Streich, B., et al. Increased carbon dioxide absorption during retroperitoneal laparoscopy. Br
J Anaesth, 2003. 91: 793.
https://pubmed.ncbi.nlm.nih.gov/14633746
1370. Kalfa, N., et al. Tolerance of laparoscopy and thoracoscopy in neonates. Pediatrics, 2005. 116: e785.
https://pubmed.ncbi.nlm.nih.gov/16322135
1371. Meininger, D., et al. Effects of posture and prolonged pneumoperitoneum on hemodynamic
parameters during laparoscopy. World J Surg, 2008. 32: 1400.
https://pubmed.ncbi.nlm.nih.gov/18224479

180 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021


1372. Gueugniaud, P.Y., et al. The hemodynamic effects of pneumoperitoneum during laparoscopic
surgery in healthy infants: assessment by continuous esophageal aortic blood flow echo-Doppler.
Anesth Analg, 1998. 86: 290.
https://pubmed.ncbi.nlm.nih.gov/9459234
1373. Sakka, S.G., et al. Transoesophageal echocardiographic assessment of haemodynamic changes
during laparoscopic herniorrhaphy in small children. Br J Anaesth, 2000. 84: 330.
https://pubmed.ncbi.nlm.nih.gov/10793591
1374. De Waal, E.E., et al. Haemodynamic changes during low-pressure carbon dioxide
pneumoperitoneum in young children. Paediatr Anaesth, 2003. 13: 18.
https://pubmed.ncbi.nlm.nih.gov/12535034
1375. Demyttenaere, S., et al. Effect of pneumoperitoneum on renal perfusion and function: a systematic
review. Surg Endosc, 2007. 21: 152.
https://pubmed.ncbi.nlm.nih.gov/17160650
1376. Gomez Dammeier, B.H., et al. Anuria during pneumoperitoneum in infants and children: a
prospective study. J Pediatr Surg, 2005. 40: 1454.
https://pubmed.ncbi.nlm.nih.gov/16150348
1377. Halverson, A., et al. Evaluation of mechanism of increased intracranial pressure with insufflation.
Surg Endosc, 1998. 12: 266.
https://pubmed.ncbi.nlm.nih.gov/9502709
1378. Mobbs, R.J., et al. The dangers of diagnostic laparoscopy in the head injured patient. J Clin
Neurosci, 2002. 9: 592.
https://pubmed.ncbi.nlm.nih.gov/12383425
1379. Al-Mufarrej, F., et al. Laparoscopic procedures in adults with ventriculoperitoneal shunts. Surg
Laparosc Endosc Percutan Tech, 2005. 15: 28.
https://pubmed.ncbi.nlm.nih.gov/15714153

5. CONFLICT OF INTEREST
All members of the Paediatric Urology Guidelines Panel have provided disclosure statements on all
relationships that they have that might be perceived to be a potential source of a conflict of interest. This
information is publically accessible through the European Association of Urology website: http://www.uroweb.
org/guidelines/.

This Guidelines document was developed with the financial support of the EAU. No external sources of funding
and support have been involved. The EAU is a non-profit organisation, and funding is limited to administrative
assistance and travel and meeting expenses. No honoraria or other reimbursements have
been provided.

6. CITATION INFORMATION
The format in which to cite the EAU Guidelines will vary depending on the style guide of the journal in which the
citation appears. Accordingly, the number of authors or whether, for instance, to include the publisher, location,
or an ISBN number may vary.

The compilation of the complete Guidelines should be referenced as:


EAU Guidelines. Edn. presented at the EAU Annual Congress Milan 2021. ISBN 978-94-92671-13-4.

If a publisher and/or location is required, include:


EAU Guidelines Office, Arnhem, The Netherlands. http://uroweb.org/guidelines/compilations-of-all-guidelines/
References to individual guidelines should be structured in the following way:
Contributors’ names. Title of resource. Publication type. ISBN. Publisher and publisher location, year.

PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021 181


182 PAEDIATRIC UROLOGY - LIMITED UPDATE MARCH 2021
EAU Guidelines on
Penile Cancer
O.W. Hakenberg (Chair), E. Compérat, S. Minhas,
A. Necchi, C. Protzel, N. Watkin (Vice-chair)
Guidelines Associate: R. Robinson

© European Association of Urology 2020


TABLE OF CONTENTS PAGE
1. INTRODUCTION 4
1.1 Aim and objectives 4
1.2 Panel composition 4
1.3 Available publications 4
1.4 Publication history 4
1.5 Summary of changes 4

2. METHODS 5
2.1 Data identification 5

3. EPIDEMIOLOGY, AETIOLOGY AND PATHOLOGY 6


3.1 Definition of penile cancer 6
3.2 Epidemiology 6
3.3 Risk factors and prevention 7
3.4 Pathology 8
3.4.1 Gross handling of pathology specimens 9
3.4.2 Pathology report 9
3.4.3 Grading 9
3.4.4 Pathological prognostic factors 10
3.4.5 Penile cancer and HPV 10
3.4.6 Penile biopsy 10
3.4.7 Intra-operative frozen sections and surgical margins 11
3.4.8 Guidelines for the pathological assessment of tumour specimens 11

4. STAGING AND CLASSIFICATION SYSTEMS 11


4.1 TNM classification 11
4.2 Guidelines on staging and classification 12

5. DIAGNOSTIC EVALUATION AND STAGING 12


5.1 Primary lesion 12
5.2 Regional lymph nodes 12
5.2.1 Non-palpable inguinal nodes 12
5.2.2 Palpable inguinal nodes 13
5.3 Distant metastases 13
5.4 Guidelines for the diagnosis and staging of penile cancer 13

6. DISEASE MANAGEMENT 13
6.1 Treatment of the primary tumour 13
6.1.1 Treatment of superficial non-invasive disease (PeIN) 14
6.1.2 Treatment of invasive disease confined to the glans (category T1/T2) 14
6.1.2.1 Intra-operative frozen section 14
6.1.2.2 Width of negative surgical margins 14
6.1.3 Results of different surgical organ-preserving treatments 14
6.1.3.1 Laser therapy 14
6.1.3.2 Moh’s micrographic surgery 15
6.1.3.3 Glans resurfacing 15
6.1.3.4 Glansectomy 15
6.1.3.5 Partial penectomy 15
6.1.3.6 Summary of results of surgical techniques 15
6.1.4 Summary of results of radiotherapy for T1 and T2 disease 15
6.1.5 Treatment recommendations for invasive penile cancer (T2-T4) 16
6.1.5.1 Treatment of invasive disease confined to the glans with or
without urethral involvement (T2) 16
6.1.5.2 Treatment of disease invading the corpora cavernosa and/or
urethra (T3) 16
6.1.5.3 Treatment of locally advanced disease invading adjacent
structures (T4) 16
6.1.5.4 Local recurrence after organ-conserving surgery 16

2 PENILE CANCER - MARCH 2018


6.1.6 Guidelines for stage-dependent local treatment of penile carcinoma 17
6.2 Management of regional lymph nodes 17
6.2.1 Management of patients with clinically normal inguinal lymph nodes (cN0) 17
6.2.1.1 Surveillance 18
6.2.1.2 Invasive nodal staging 18
6.2.2 Management of patients with palpable inguinal nodes (cN1/cN2) 18
6.2.2.1 Radical inguinal lymphadenectomy 18
6.2.2.2 Pelvic lymphadenectomy 19
6.2.2.3 Adjuvant treatment 19
6.2.3 Management of patients with fixed inguinal nodes (cN3) 19
6.2.4 Management of lymph node recurrence 19
6.2.5 The role of radiotherapy in lymph node disease 19
6.2.6 Guidelines for treatment strategies for nodal metastases 20
6.3 Chemotherapy 20
6.3.1 Adjuvant chemotherapy in node-positive patients after radical inguinal
lymphadenectomy 20
6.3.2 Neoadjuvant chemotherapy in patients with fixed or relapsed inguinal nodes 20
6.3.3 Palliative chemotherapy in advanced and relapsed disease 21
6.3.4 Intra-arterial chemotherapy 21
6.3.5 Targeted therapy 21
6.3.6 Guidelines for chemotherapy 21

7. FOLLOW-UP 22
7.1 Rationale for follow-up 22
7.1.1 When and how to follow-up 22
7.1.2 Recurrence of the primary tumour 22
7.1.3 Regional recurrence 22
7.1.4 Guidelines for follow-up in penile cancer 23
7.2 Quality of life 23
7.2.1 Consequences after penile cancer treatment 23
7.2.2 Sexual activity and quality of life after laser treatment 23
7.2.3 Sexual activity after glans resurfacing 24
7.2.4 Sexual activity after glansectomy 24
7.2.5 Sexual function after partial penectomy 24
7.2.6 Quality of life and sexual function after total penectomy 24
7.2.7 Quality of life after partial penectomy 24
7.3 Total phallic reconstruction 24
7.4 Specialised care 24

8. REFERENCES 25

9. CONFLICT OF INTEREST 36

10. CITATION INFORMATION 36

PENILE CANCER - MARCH 2018 3


1. INTRODUCTION
1.1 Aim and objectives
The European Association of Urology (EAU) Guidelines on Penile Cancer provides up-to-date information on
the diagnosis and management of penile squamous cell carcinoma (SCC).
It must be emphasised that clinical guidelines present the best evidence available to the experts
but following guideline recommendations will not necessarily result in the best outcome. Guidelines can never
replace clinical expertise when making treatment decisions for individual patients, but rather help to focus
decisions - also taking personal values and preferences/individual circumstances of patients into account.
Guidelines are not mandates and do not purport to be a legal standard of care.

1.2 Panel composition


The EAU Penile Cancer Guidelines Panel consists of an international multi-disciplinary group of clinicians,
including a pathologist and an oncologist. Members of this panel have been selected based on their expertise
and to represent the professionals treating patients suspected of having penile cancer. All experts involved in
the production of this document have submitted potential conflict of interest statements, which can be viewed
on the EAU website Uroweb: http://uroweb.org/guideline/penile-cancer/.

1.3 Available publications


A quick reference document (Pocket guidelines) is available, both in print and as an app for iOS and Android
devices. These are abridged versions which may require consultation together with the full text version. Several
scientific publications are available, the most recent dating back to 2014 [1], as are a number of translations
of all versions of the Penile Cancer Guidelines. All documents are available through the EAU website Uroweb:
http://uroweb.org/guideline/penile-cancer/.

1.4 Publication history


The EAU Penile Cancer Guidelines were first published in 2000; the current publication presents a limited
update of the 2017 print.

1.5 Summary of changes


Key changes for the 2018 print:

Chapter 3 - Epidemiology, aetiology and pathology. New information has been added on the various
histological subtypes of penile carcinomas, risk factors and human papilloma virus (HPV) association.

New and changed recommendations can be found in sections:

3.4.8 Guidelines for the pathological assessment of tumour specimens

Recommendations Strength rating


The pathological evaluation of penile carcinoma specimens must include an Strong
assessment of the human papilloma virus status.
The pathological evaluation of penile carcinoma specimens must include a Strong
diagnosis of the squamous cell carcinoma subtype.
The pathological evaluation of penile carcinoma surgical specimens must include an Strong
assessment of surgical margins including the width of the surgical margin.

4.2 Guidelines on staging and classification

Recommendation Strength rating


The pathological evaluation of penile carcinoma specimens must include the pTNM Strong
stage and an assessment of tumour grade.

5.4 Guidelines for the diagnosis and staging of penile cancer

Recommendations Strength rating


Primary tumour
Perform a physical examination, record morphology, extent and invasion of penile Strong
structures.

4 PENILE CANCER - MARCH 2018


Obtain a penile Doppler ultrasound or MRI with artificial erection in cases with Weak
intended organ-sparing surgery.
Inguinal lymph nodes
Perform a physical examination of both groins, record the number, laterality and Strong
characteristics of inguinal nodes and:
• If nodes are not palpable, offer invasive lymph node staging in
intermediate- and high-risk patients;
• If nodes are palpable, stage with a pelvic computed tomography (CT) or
positron emission tomography (PET)/CT.
Distant metastases
In N+ patients, obtain an abdominopelvic CT scan and chest X-ray/thoracic CT for Strong
systemic staging. Alternatively, stage with a PET/CT scan.
In patients with systemic disease or with relevant symptoms, obtain a bone scan.

6.2.6 Guidelines for treatment strategies for nodal metastases

Regional lymph nodes Management of regional lymph nodes is Strength rating


fundamental in the treatment of penile cancer
Radiotherapy Not recommended for nodal disease except as a Strong
Radiotherapy palliative option.
> T1G2: invasive lymph node staging by either bilateral Strong
modified inguinal lymphadenectomy or dynamic
sentinel node biopsy.
Palpable inguinal nodes Radical inguinal lymphadenectomy. Strong
(cN1/cN2)
Fixed inguinal lymph nodes Neoadjuvant chemotherapy followed by radical Weak
(cN3) inguinal lymphade-nectomy in responders.
Pelvic Ipsilateral pelvic lymphadenectomy if two or more Strong
Lymph nodes inguinal nodes are involved on one side (pN2) or if
extracapsular nodal metastasis (pN3) reported
Adjuvant chemotherapy In pN2/pN3 patients after radical lymphadenectomy. Strong
Radiotherapy Not recommended for nodal disease except as a Strong
palliative option.

6.3.6 Guidelines for chemotherapy

Recommendations Strength rating


Offer patients with pN2-3 tumours adjuvant chemotherapy after radical Strong
lymphadenectomy (three to four cycles of cisplatin, a taxane and 5-fluorouracil or
ifosfamide).
Offer palliative chemotherapy to patients with systemic disease. Weak

A systematic review (SR) was performed by the Panel on ‘Risks and benefits of adjuvant radiotherapy after
inguinal lymphadenectomy in node-positive penile cancer’ [2]. Even though not fully published, the review
findings support the information presented in Section 6.2.2.3 Adjuvant treatment.
This review was performed using standard Cochrane SR methodology: http://www.cochranelibrary.
com/about/about-cochrane-systematic-reviews.html

2. METHODS
2.1 Data identification
For the 2018 Penile Cancer Guidelines, new and relevant evidence has been identified, collated and appraised
through a structured assessment of the literature. A broad and comprehensive literature search, covering all
sections of the Penile Cancer Guidelines, was performed. Databases searched included Medline, EMBASE and
the Cochrane Libraries, covering the period between November 1st 2013 and September 20th 2016. All articles
relating to penile cancer (n = 838) in the relevant literature databases were reviewed resulting in the inclusion of
29 new publication in this print.

PENILE CANCER - MARCH 2018 5


Fully revised Guidelines were produced using the updated research base, together with several
national and international guidelines on penile cancer (National Comprehensive Cancer Network [3], French
Association of Urology [4] and the European Society of Medical Oncology [5]).

For the 2018 edition of the EAU Guidelines the Guidelines Office have transitioned to a modified GRADE
methodology across all 20 guidelines [6, 7]. For each recommendation within the guidelines there is an
accompanying online strength rating form which addresses a number of key elements namely:
1. the overall quality of the evidence which exists for the recommendation, references used in
this text are graded according to a classification system modified from the Oxford Centre for
Evidence-Based Medicine Levels of Evidence [8];
2. the magnitude of the effect (individual or combined effects);
3. the certainty of the results (precision, consistency, heterogeneity and other statistical or
study related factors);
4. the balance between desirable and undesirable outcomes;
5. the impact of patient values and preferences on the intervention;
6. the certainty of those patient values and preferences.

These key elements are the basis which panels use to define the strength rating of each recommendation
which is represented by the words ‘strong’ or ‘weak’ [9]. The strength of each recommendation is determined
by the balance between desirable and undesirable consequences of alternative management strategies,
the quality of the evidence (including certainty of estimates), and nature and variability of patient values and
preferences. The strength rating forms will be available online.
Additional information can be found in the general Methodology section of this print, and online at
the EAU website: http://www.uroweb.org/guideline/.
A list of associations endorsing the EAU Guidelines can also be viewed online at the above address.

3. EPIDEMIOLOGY, AETIOLOGY AND


PATHOLOGY
3.1 Definition of penile cancer
Penile carcinoma is usually a SCC and there are several recognised subtypes of penile SCC with different
clinical features and natural history (see Table 1). Penile SCC usually arises from the epithelium of the inner
prepuce or the glans.

Table 1: Histological subtypes of penile carcinomas, their frequency and outcome

Subtype Frequency Prognosis


(% of cases)
Common squamous cell carcinoma 48-65 Depends on location, stage and grade
(SCC)
Basaloid carcinoma 4-10 Poor prognosis, frequently early inguinal nodal metastasis
[10]
Warty carcinoma 7-10 Good prognosis, metastasis rare
Verrucous carcinoma 3-8 Good prognosis, no metastasis
Papillary carcinoma 5-15 Good prognosis, metastasis rare
Sarcomatoid carcinoma 1-3 Very poor prognosis, early vascular metastasis
Mixed carcinoma 9-10 Heterogeneous group
Pseudohyperplastic carcinoma <1 Foreskin, related to lichen sclerosus, good prognosis,
metastasis not reported
Carcinoma cuniculatum <1 Variant of verrucous carcinoma, good prognosis,
metastasis not reported
Pseudoglandular carcinoma <1 High-grade carcinoma, early metastasis, poor prognosis
Warty-basaloid carcinoma 9-14 Poor prognosis, high metastatic potential [11] (higher than
in warty, lower than in basaloid SCC)

6 PENILE CANCER - MARCH 2018


Adenosquamous carcinoma <1 Central and peri-meatal glans, high-grade carcinoma,
high metastatic potential but low mortality
Mucoepidermoid carcinoma <1 Highly aggressive, poor prognosis
Clear cell variant of penile 1-2 Exceedingly rare, associated with human papilloma virus,
carcinoma aggressive, early metastasis, poor prognosis, outcome is
lesion-dependent, frequent lymphatic metastasis [12]

3.2 Epidemiology
In industrialised countries, penile cancer is uncommon, with an overall incidence of around 1/100,000 males
in Europe and the USA [13, 14]. There are several areas in Europe with a higher incidence (Figure 1) [15].
Recent data from Scandinavia report an incidence of around 2/100,000 men. In the USA, the incidence of
penile cancer is affected by race and ethnicity, with the highest incidence in white Hispanics (1.01), followed
by Alaskans and Native American Indians (0.77), African Americans (0.62) and white non-Hispanics (0.51), per
100,000, respectively. In contrast, in some other parts of the world such as South America, South East Asia
and parts of Africa the incidence is much higher and can account for 1-2% of malignant diseases in men [15].
The annual age-adjusted incidence is 0.7-3.0 in India, 8.3 in Brazil (per 100,000, respectively) and even higher
in Uganda, where it is the most commonly diagnosed male cancer [15, 16].
In the USA, the overall age-adjusted incidence rate decreased from 1973 to 2002 from 0.84 in 1973-
1982 to 0.69 in 1983-1992, and to 0.58 in 1993-2002, per 100,000, respectively [13]. In Europe, the overall
incidence has been stable from the 1980s until 2013 [14]. An increased incidence was observed in Denmark
[17] and the UK (21% between 1979 and 2009) [18].
The incidence of penile cancer increases with age [14], with a peak in the sixth decade but it does
occur in younger men [19]. Penile cancer is common in regions with a high prevalence of HPV and this may
account for the worldwide variation in incidence [13]. About one third of cases are attributed to HPV-related
carcinogenesis [20]. Penile cancer is not linked to HIV or AIDS.

Figure 1: Annual incidence rate (world standardised) by European region/country*

Penis: ASR (World) (per 100,000) (All ages)

Spain, Albacete
Malta
Switzerland, Neuchatel
France, Haut-Rhin
Italy, Ragusa Province
UK, Scotland
Denmark
Austria, Tyrol
Norway
Spain, Asturias
France, Bas-Rhin
UK, England
Estonia
Slovakia
Switzerland, Ticino
The Netherlands
Belgium Flanders (excl. Limburg)
Italy, Torino
Poland, Warsaw city
Germany, Saarland
Portugal, Vila Nova de Gaia
Slovenia
Italy, Sassari
0 0.5 1.0 1.5 2.0

*Adapted from [15].

PENILE CANCER - MARCH 2018 7


3.3 Risk factors and prevention
Several risk factors for penile cancer have been identified (Table 2) [21] (LE: 2a).

Table 1: Recognised aetiological and epidemiological risk factors for penile cancer

Risk factors Relevance Ref


Phimosis Odds ratio 11-16 vs. no phimosis [22-24]
Chronic penile inflammation (balanoposthitis Risk [25]
related to phimosis), lichen sclerosus
Sporalene and ultraviolet A phototherapy for Incidence rate ratio 9.51 with > 250 treatments [26]
various dermatological conditions such as
psoriasis
Smoking Five-fold increased risk (95% Confidence [22, 23,
interval (CI): 2.0-10.1) vs. non-smokers 27]
HPV infection, condylomata acuminate 22.4% in verrucous squamous cell carcinoma [13, 28]
36-66.3% in basaloid-warty
Rural areas, low socio-economic status, [29-32]
unmarried
Multiple sexual partners, early age of first Three to five-fold increased risk of penile cancer [21, 23,
intercourse 33]

Human papilloma virus infection is a risk factor for penile cancer [34]. Human papilloma virus DNA has been
identified in 70-100% of intra-epithelial neoplasia and in 30-40% of invasive penile cancer tissue samples
(LE: 2a). The HPV virus interacts with oncogenes and tumour suppressor genes (p16, P53, Rb genes) [28, 35].
The rate of HPV-positivity differs between different histological subtypes of penile SCC. Human papilloma virus
is a cofactor in the carcinogenesis of some variants of penile SCC, while others are not related to HPV. The
commonest HPV subtypes in penile cancer are types 16 and 18 [36]. The risk of penile cancer is increased in
patients with condyloma acuminata [37] (LE: 2b).
A significantly better five-year disease-specific survival has been reported for HPV-positive vs. HPV-
negative cases (93% vs. 78%) in one study [38], while no difference in lymph node metastases and ten-year
survival was reported in another study [39] (Table 3). There is no association between the incidence of penile
and cervical cancer, although both are linked to HPV [40, 41]. Female sexual partners of patients with penile
cancer do not have an increased incidence of cervical cancer [42].

Table 3: Outcomes for HPV and non-HPV penile carcinomas

Non HPV related Prognosis HPV related Prognosis


SCC usual type/NOS 30% DOD Basaloid SCC > 50% DOD
Pseudohyperplastic 0% Papillary basaloid carcinoma
carcinoma
Pseudoglandular carcinoma > 50% Warty carcinoma Mortality low
Verrucous carcinoma Good Warty-basaloid carcinoma 30% DOD
Carcinoma cuniculatum Good Clear-cell carcinoma 20%
Papillary carcinoma NOS Good Lymphoepithelioma-like Not known
carcinoma
Adenosquamous carcinoma Good
Sarcomatoid carcinoma 75% DOD
DOD = died of disease; HPV = human papillomavirus; SCC = squamous cell carcinoma.

At present, except for a few countries, there is no general recommendation for HPV vaccination in males
because of the different HPV-associated risk patterns in penile- and cervical cancer. Furthermore, the
epidemiological effects of HPV vaccination in girls still have to be assessed [43, 44].
Phimosis is strongly associated with invasive penile cancer [23, 29, 45, 46], due to associated
chronic infection. However, smegma is not a carcinogen [45]. The incidence of lichen sclerosus is relatively high
in penile cancer but is not associated with adverse histopathological features, including penile intraepithelial
neoplasia (PeIN). Other epidemiological risk factors are cigarette smoking, low socioeconomic status and a low
level of education [29, 46].

8 PENILE CANCER - MARCH 2018


Neonatal circumcision reduces the incidence of penile cancer; however, it does not seem to reduce
the risk of PeIN [23]. The lowest incidence of penile cancer is reported for Israeli Jews (0.3/100,000/year).
One matched-pair, case-control study reported that the protective effect of neonatal circumcision against
invasive penile cancer (OR 0.41) was much weaker when the analysis was restricted to men without a history of
phimosis (OR 0.79, 95% CI: 0.29-2) [23]. Circumcision in adult life does not have any protective effect.
The controversial discussion about neonatal circumcision should take into account that
circumcision removes approximately half the tissue that can develop into penile cancer.

3.4 Pathology
Squamous cell carcinoma accounts for over 95% of penile malignancies (see Table 1). It is not known how
often SCC is preceded by premalignant lesions (see Table 4) [47-50].
Different histological types of penile SCC with different growth patterns, clinical aggressiveness and
HPV associations have been identified (see Table 5). Numerous mixed forms exist such as the warty-basaloid
form, with 50-60% the most common mixed form, the usual-verrucous (hybrid), usual-warty, usual-basaloid
and the usual-papillary, as well as other rarer combinations.
Other malignant lesions of the penis, all much less common than penile SCC, are melanocytic
lesions, mesenchymal tumours, lymphomas and metastases. Penile metastases are frequently of prostatic or
colorectal origin. Different types of penile sarcoma have been reported.

Table 4: Premalignant penile lesions (precursor lesions)

Lesions sporadically associated with squamous cell carcinoma (SCC) of the penis:
• Bowenoid papulosis of the penis (HPV related)
• Lichen sclerosis
Premalignant lesions (up to one-third transform to invasive SCC):
• Penile intraepithelial lesions
• Giant condylomata (Buschke-Löwenstein)
• Bowen’s disease
• Paget’s disease (intradermal ADK)

Table 5: Classification of intra-epithelial neoplasia (PeIN)

• Non-HPV-related PeIN
• o Differentiated PeIN
• HPV-related PeIN
• o Basaloid PeIN
• Warty PeIN
• Warty-basaloid PeIN
• Other rare patterns of PeIN (pleomorphic, spindle, clear cell, pagetoid)

3.4.1 Gross handling of pathology specimens


Tissue sections determine the accuracy of histological diagnosis. Small lesions should be fully included, bigger
lesions should have at least 3-4 blocks. Lymph nodes must be included in their entirety after having been
inked, in order to detect metastases. After having been inked, surgical margins have to be completely included
[51]. Second-opinion pathology review is highly desirable for this rare tumour entity [52].

3.4.2 Pathology report


The pathology report must include the anatomical site of the primary tumour, the histological type of SCC,
grade, perineural invasion, depth of invasion, vascular invasion (venous/lymphatic), irregular growth and front
of invasion, urethral invasion, invasion of corpus spongiosum/cavernosum, surgical margins and the p16/HPV
status (Table 6) [53-56].

PENILE CANCER - MARCH 2018 9


Table 6: Outcomes for HPV and non-HPV penile carcinomas

Information to include in the pathology report Recommended required


Clinical information x
• Prior treatments (topic, radiotherapy, chemotherapy)
Surgical procedure x
Tumour localisation x
Macroscopic tumour dimension x
• Depth of invasion
• Millimetres from basement membrane to deepest point of invasion
• Maximum thickness
• Size of tumour
Block identification x
Histological tumour type x
Histological grade x
Microscopic maximum dimensions x
• Combination of gross and microscopic if large tumours
Extent of invasion x
LVI [58, 59] x
Perineural invasion x
Margin status in mm x
Lymph node (LN) status x
• Size of largest nodal tumour deposit (not LN size)
• Number of LN+, extracapsular spread (ECS), inguinal or pelvic, to
be reported in every site separately
TNM Stage x
p16/HPV status x
* See also www.ICCR-cancer.org. database.

3.4.3 Grading
The TNM classification for penile cancer includes tumour grade, due to its prognostic relevance (Table 9).
Tumour grading in penile cancer has been shown to be highly observer-dependent and can be problematic,
especially in heterogeneous tumours. Grading should use the categories specified by the WHO for penile
cancer (Table 7).

Table 7: Grading recommendations for penile SCC

Feature Grade 1 Grade 2 Grade 3 Sarcomatoid


Cytological atypia Mild Moderate Anaplasia Sarcomatoid
Keratinisation Usually abundant Less prominent May be present Absent
Intercellular bridges Prominent Occasional Few Absent
Mitotic activity Rare Increased Abundant Abundant
Tumour margin Pushing/well Infiltrative/ill defined Infiltrative/ill defined Infiltrative/ill defined

3.4.4 Pathological prognostic factors


Pathological subtype, perineural invasion, lymphovascular invasion [58], depth of invasion and grade in the
primary tumour are strong predictors of poor prognosis and high cancer-specific mortality [60]. Tumour grade
is a predictor of metastatic spread, and lymphatic invasion is a predictor of metastasis. Venous embolism is
often seen in advanced stages. The extent of lymph node metastasis and extracapsular spread are also strong
predictors of prognosis.
The variants of penile SCC can be divided into three prognostically different groups (Table 8).

10 PENILE CANCER - MARCH 2018


Table 8: Prognosis of the variants of penile SCC

Penile SCC Good prognosis Intermediate prognosis Poor prognosis


Local growth Destructive Destructive Destructive
Metastasis Rare Intermediate Common
Risk of cancer-related Very low Intermediate High
mortality
SCC variants • Verrucous • Usual SCC • Basaloid,
• Papillary • Mixed forms • Sarcomatoid
• Warty • Pleomorphic form of adenosquamous
Pseudohyperplastic warty carcinoma
carcinoma cuniculatum

There is discussion as to whether cases that show invasion of the distal urethra have a worse prognosis;
however, there is no evidence to support this [61]. Nevertheless, invasion of the more proximal urethra signifies
a highly aggressive SCC with a poor prognosis (see Table 9). pT3 denotes a worse prognosis than pT2 [62, 63]
(LE: 2b). Capsular extension in even one single lymph node carries a poor prognosis and is denoted as pN3
[64-66].
Chaux et al. suggested a prognostic index which incorporates grade, anatomical level of infiltration
and perineural invasion to predict the likelihood of inguinal lymph node metastases and 5-year survival [67].

3.4.5 Penile cancer and HPV


The association between penile cancer and HPV is different for the different variants of penile SCC. A high
prevalence of HPV infection is found in basaloid (76%), mixed warty-basaloid (82%) and warty penile (39%)
SCCs. Verrucous and papillary penile SCCs are HPV-negative. The commonest HPV-types in penile SCC are
HPV-16 (72%), HPV-6 (9%) and HPV-18 (6%). Overall, only one-third of penile SCCs show HPV infection, but
those that do are usually infected by several HPV strains.

3.4.6 Penile biopsy


Any doubtful penile lesion should be biopsied and, even in clinically obvious cases, histological verification
must be obtained before local treatment. Before definitive surgical treatment, confirmatory frozen section
excisional biopsy can be done. Histological confirmation is necessary to guide management when:
• there is doubt about the exact nature of the lesion (e.g. PeIN, metastasis or melanoma);
• treatment is planned with topical agents, radiotherapy or laser surgery.

The size of a biopsy is important. In one study, in biopsies with an average size of 0.1 cm it was difficult to
evaluate the depth of invasion in 91% of cases. The grade at biopsy, and in the final specimen, may differ in
up to 30% of cases, with failure to detect cancer in 3.5% of cases [47]. Furthermore, vascular and lymphatic
tumour emboli were detected in only 9-11% of cases. Although a punch biopsy may be sufficient for superficial
lesions, an excisional biopsy which is deep enough to properly assess the degree of invasion and stage is
preferable.

3.4.7 Intra-operative frozen sections and surgical margins


Surgical treatment must completely remove the penile carcinoma with negative surgical margins, which may
be confirmed by intra-operative frozen section [68]. The width of negative surgical margins should follow a risk-
adapted strategy based on tumour grade. Only 3 mm of tumour-free tissue is sufficient to consider the surgical
margins to be negative [69].

3.4.8 Guidelines for the pathological assessment of tumour specimens

Recommendations Strength rating


The pathological evaluation of penile carcinoma specimens must include an assessment of Strong
the human papilloma virus status.
The pathological evaluation of penile carcinoma specimens must include a diagnosis of the Strong
squamous cell carcinoma subtype.
The pathological evaluation of penile carcinoma surgical specimens must include an Strong
assessment of surgical margins including the width of the surgical margin.

PENILE CANCER - MARCH 2018 11


4. STAGING AND CLASSIFICATION SYSTEMS
4.1 TNM classification
The 2016 UICC TNM classification for penile cancer [51] introduced some changes in comparison to previous
editions. The T1 category is stratified into two prognostically different risk groups, depending on the presence
or absence of lymphovascular invasion and grading (Table 9). The classification T2 denotes invasion of the
corpus spongiosum, while T3 is defined as invasion of the corpora cavernosa, due to the different prognosis
of these two patterns [62, 63]. For penile cancer, unlike in other neoplasms, tumour grade is used for the TNM
classification in the subdivision of the T1 stage (Table 9).
The current pN1 group consists of one or two ipsilateral inguinal lymph node metastases, pN2 is
defined as more than two uni- or bilateral metastatic nodes and pN3 any pelvic nodes, uni- or bilateral, or any
extranodal extension regardless of the number of lymph node metastases [51]. Retroperitoneal lymph node
metastases are classified as extra-regional nodal and, therefore, distant metastases.

Table 9: 2016 TNM clinical and pathological classification of penile cancer [51]

Clinical classification
T - Primary Tumour
TX Primary tumour cannot be assessed
T0 No evidence of primary tumour
Tis Carcinoma in situ
Ta Non-invasive verrucous carcinoma*
T1 Tumour invades subepithelial connective tissue
T1a Tumour invades subepithelial connective tissue without lymphovascular invasion and is not
poorly differentiated
T1b Tumour invades subepithelial connective tissue with lymphovascular invasion or is poorly
differentiated
T2 Tumour invades corpus spongiosum with or without invasion of the urethra
T3 Tumour invades corpus cavernosum with or without invasion of the urethra
T4 Tumour invades other adjacent structures
N - Regional Lymph Nodes
NX Regional lymph nodes cannot be assessed
N0 No palpable or visibly enlarged inguinal lymph nodes
N1 Palpable mobile unilateral inguinal lymph node
N2 Palpable mobile multiple or bilateral inguinal lymph nodes
N3 Fixed inguinal nodal mass or pelvic lymphadenopathy, unilateral or bilateral
M - Distant Metastasis
M0 No distant metastasis
M1 Distant metastasis
Pathological classification
The pT categories correspond to the clinical T categories.
The pN categories are based upon biopsy or surgical excision
pN - Regional Lymph Nodes
pNX Regional lymph nodes cannot be assessed
pN0 No regional lymph node metastasis
pN1 Metastasis in one or two inguinal lymph nodes
pN2 Metastasis in more than two unilateral inguinal nodes or bilateral inguinal lymph nodes
pN3 Metastasis in pelvic lymph node(s), unilateral or bilateral extranodal or extension of regional lymph
node metastasis
pM - Distant Metastasis
pM1 Distant metastasis microscopically confirmed
G - Histopathological Grading
GX Grade of differentiation cannot be assessed
G1 Well differentiated
G2 Moderately differentiated
G3 Poorly differentiated
G4 Undifferentiated

*Verrucous carcinoma not associated with destructive invasion.

12 PENILE CANCER - MARCH 2018


4.2 Guidelines on staging and classification

Recommendation Strength rating


The pathological evaluation of penile carcinoma specimens must include the pTNM stage Strong
and an assessment of tumour grade.

5. DIAGNOSTIC EVALUATION AND STAGING


Penile cancer can be cured in over 80% of cases if diagnosed early, but is a life-threatening disease when
lymphatic metastasis occurs. Local treatment can be mutilating, and devastating for the patient’s psychological
well-being.

5.1 Primary lesion


Penile carcinoma is usually a clinically obvious lesion but it may be hidden under a phimosis [24]. Physical
examination should include palpation of the penis to assess the extent of local invasion and palpation of both
groins to assess the lymph node status.
Ultrasound (US) can provide information about infiltration of the corpora [70, 71]. Magnetic
resonance imaging (MRI) with an artificially induced erection can be used to exclude corporal invasion but is very
unpleasant for the patient [72, 73]. The sensitivity and specificity of MRI in predicting corporal or urethral invasion
was reported as 82.1% and 73.6%, and 62.5% and 82.1%, respectively [74]. Penile Doppler US has been
reported to have a higher staging accuracy than an MRI in detecting corporal infiltration [75].

5.2 Regional lymph nodes


Careful palpation of both groins for enlarged inguinal lymph nodes must be part of the initial physical
examination of patients suspected of having penile cancer.

5.2.1 Non-palpable inguinal nodes


If there are no palpable lymph nodes, the likelihood of micro-metastatic disease is about 25%. Imaging studies
are not helpful in staging clinically normal inguinal regions, although may be used in obese patients in whom
palpation is unreliable:
• Inguinal US (7.5 MHz) can detect abnormal, enlarged nodes. The longitudinal/transverse diameter ratio
and absence of the lymph node hilum are findings with relatively high specificity [76].
• Conventional computed tomography (CT) or MRI cannot detect micro-metastases reliably [77].
• 18FDG-positron emission tomography (PET/CT) does not detect lymph node metastases < 10 mm [78, 79].

Further management of patients with normal inguinal nodes should be guided by pathological risk factors of
the primary tumour. Lymphovascular invasion, local stage and grade are predictive of lymphatic metastasis
[80, 81]. Existing nomograms are not accurate. Invasive lymph node staging is required in patients at
intermediate- or high risk of lymphatic spread (see Section 6.2).

5.2.2 Palpable inguinal nodes


Palpably enlarged lymph nodes are highly indicative of lymph node metastases. Physical examination should
note the number of palpable nodes on each side and whether these are fixed or mobile. Additional imaging
does not alter management and is not required (see Section 6).
A pelvic CT scan can be used to assess the pelvic lymph nodes. Imaging with 18FDG-PET/CT has
shown high sensitivity (88-100%) and specificity (98-100%) for confirming metastatic nodes in patients with
palpable inguinal lymph nodes [79, 82].

5.3 Distant metastases


Staging for systemic metastases should be performed in patients with positive inguinal nodes [83-85] (LE: 2b).
Abdominal and pelvic CT should be done plus a chest X-ray, although a thoracic CT is more sensitive. PET/CT
is an option [81].
There is no tumour marker for penile cancer. The SCC antigen (SCC Ag) is increased in less than
25% of penile cancer patients. One study reported that SCC Ag did not predict occult metastatic disease, but
was an indicator of disease-free survival (DFS) in lymph-node-positive patients [86].

PENILE CANCER - MARCH 2018 13


5.4 Guidelines for the diagnosis and staging of penile cancer

Recommendations Strength rating


Primary tumour
Perform a physical examination, record morphology, extent and invasion of penile Strong
structures.
Obtain a penile Doppler ultrasound or MRI with artificial erection in cases with intended Weak
organ-sparing surgery.
Inguinal lymph nodes
Perform a physical examination of both groins, record the number, laterality and Strong
characteristics of inguinal nodes and:
• If nodes are not palpable, offer invasive lymph node staging in intermediate- and
high-risk patients;
• If nodes are palpable, stage with a pelvic computed tomography (CT) or positron
emission tomography (PET)/CT.
Distant metastases
In N+ patients, obtain an abdominopelvic CT scan and chest X-ray/thoracic CT for systemic Strong
staging. Alternatively, stage with a PET/CT scan.
In patients with systemic disease or with relevant symptoms, obtain a bone scan.

6. DISEASE MANAGEMENT
6.1 Treatment of the primary tumour
The aims of the treatment of the primary tumour are complete tumour removal with as much organ preservation
as possible, without compromising oncological control. Local recurrence has little influence on long-term
survival, so organ preservation strategies are justified [87].
There are no randomised controlled trials (RCTs) or observational comparative studies for any of
the treatment options for localised penile cancer. Penile preservation appears to be superior in functional
and cosmetic outcomes to partial or total penectomy, and is considered to be the primary treatment method
for localised penile cancer. However, there are no RCTs comparing organ-preserving and ablative treatment
strategies.
Histological diagnosis with local staging must be obtained before using non-surgical treatments.
With surgical treatment, negative surgical margins must be obtained. Treatment of the primary tumour and of
the regional nodes can be staged.
Local treatment modalities for small and localised penile cancer include excisional surgery, external
beam radiotherapy (EBRT), brachytherapy and laser ablation. Patients should be counselled about all relevant
treatment options.

6.1.1 Treatment of superficial non-invasive disease (PeIN)


Topical chemotherapy with imiquimod or 5-fluorouracil (5-FU) is an effective first-line treatment. Circumcision
is advisable prior to the use of topical agents. Due to high persistence/recurrence rates, treatment must be
assessed by biopsy and long-term surveillance is warranted. An insufficient response may signify underlying
invasive disease. Significant inflammatory responses may occur [88, 89]. Complete responses have been
reported in up to 57% of PeIN cases [90] and in 74% of cases treated by circumcision and 5-FU without
relapse. If topical treatment fails, it should not be repeated.
Laser treatment with a neodymium:yttrium-aluminium-garnet (Nd:YAG) or Carbon dioxide (CO2)
laser is an effective treatment option [91-96]. Visualisation may be improved by photodynamic diagnosis with
the CO2 laser [97]. Rebiopsy for treatment control is mandatory.
Glans resurfacing, total or partial, can be a primary treatment for PeIN or a secondary option in
case of failure of topical chemotherapy or laser therapy. Glans resurfacing consists of complete removal of
the glandular epithelium followed by reconstruction with a graft (split skin or buccal mucosa). However, in one
study in cases of glans resurfacing for presumed PeIN, up to 20% of patients were found to have invasive
disease on histopathological examination [88].

14 PENILE CANCER - MARCH 2018


6.1.2 Treatment of invasive disease confined to the glans (category T1/T2)
Small and localised invasive lesions should receive organ-sparing treatment. Additional circumcision is
advisable for glandular tumours. Foreskin tumours are treated by ‘radical circumcision’. Local excision, partial
glansectomy or total glansectomy with reconstruction are surgical options. External beam radiotherapy or
brachytherapy are radiotherapeutic options. Small lesions can also be treated by laser therapy but the risk of
more invasive disease must be recognised.
Treatment choice depends on tumour size, histology, stage and grade, localisation (especially
relative to the meatus) and patient preference.

6.1.2.1 Intra-operative frozen section


Many authors recommend intraoperative frozen sections to assess surgical margins. Others have suggested
that frozen sections are only needed if there are definite concerns [98]. For glans resurfacing, some advocate
the use of acetic acid staining to delineate abnormal areas [99]. Data from one multi-centre study suggests that
differentiated penile intraepithelial neoplasia, squamous hyperplasia and lichen sclerosis present at the surgical
margins are frequent findings and are not relevant for cancer-specific survival [65].

6.1.2.2 Width of negative surgical margins


There is no clear evidence as to the required width of negative surgical margins. With organ-sparing these can
be minimal. For a general recommendation, 3-5 mm can be considered a safe maximum [100, 101]. A grade-
based differentiated approach can also be used, with 3 mm for grade one, 5 mm for grade two and 8 mm for
grade three. This approach has its limitations due to the difficulties with penile cancer grading.

6.1.3 Results of different surgical organ-preserving treatments


6.1.3.1 Laser therapy
The results of CO2 laser treatment have been reported by three retrospective studies from the same institution
with a median follow-up of five years and a total of 195 patients [91-93]. Laser treatment was given in
combination with radiotherapy or chemotherapy for PeIN or T1 penile cancers. No cancer-specific deaths
were reported. Local recurrence ranged from 14% for PeIN [93] to 23% for T1 tumours [92], with an estimated
cumulative risk of local recurrence at five years of 10% for PeIN (n = 106) and 16% for T1 (n = 78) tumours [91].
The reported rate of inguinal nodal recurrence was between 0% [93] and 4% [92]. The rate of secondary partial
penectomy at ten years was 3% for PeIN and 10% for T1 tumours [91].
Four studies, three from the same institution, reported results of Nd:YAG laser treatment for a total
of 150 patients with a follow-up of at least four years [94-96, 102]. Local recurrence rates ranged from 10% to
48% [94, 95]. One study [96] reported recurrence-free survival rates of 100%, 95% and 89% at one, two and
five years, respectively. Inguinal nodal recurrence was reported in 21% of patients [94] and cancer-specific
mortality was reported as 2% [102] and 9% [95]. The three studies from the same institution reported overall
survival (OS) rates of 100% at four years [94] and 85-95% [96, 103] at seven years. The rate of secondary
partial penectomy was highly divergent, with 4% in one study [96] and 45% in another [95]. One study reported
that no complications and no adverse effects on urinary or sexual function were observed [94].
Other studies have presented data on a variety of laser treatments with either a CO2 or a Nd:YAG
laser, a combination of both, or a potassium titanyl phosphate (KTP) laser [104-107], with a mean follow-up of
32-60 months with stages PeIN to T3 included. These studies reported on a total of 138 patients, with local
recurrence rates of 11% [92], 19% [105] and 26% [107]. In one study, recurrence-free survival at five years was
88% [105]. The cancer-specific survival (CSS) probability at five years was 95% in one study [105], and 2% at
five years in another [105].

6.1.3.2 Moh’s micrographic surgery


Moh’s micrographic surgery is a historical technique by which histological margins are taken in a geometrical
fashion around a conus of excision. The original description [108] consisted of 33 consecutive patients treated
between 1936 and 1986 with 79% cured at five years [108]. The second study reported 68% recurrence-free
survival at three years, 32% local recurrences and 8% inguinal nodal recurrence [109]. In both studies, one
partial amputation and one cancer-specific death occurred. In a contemporary series of 48 cases, there were
no recurrences among 10 primary invasive SCCs with a cure rate of 100% (mean follow-up, 161 months,
median follow-up, 177 months), but one recurrence in 19 cases of penile intraepithelial neoplasia (cure rate
94.7%) [110].

6.1.3.3 Glans resurfacing


Three studies have reported results of glans resurfacing in a total of 71 patients with PeIN or T1 with a
median follow-up of 21-30 months [88, 111, 112]. No cancer-specific deaths were reported, the rates of local
recurrence were 0% [111] and 6% [112], without reports of nodal recurrence or complications.

PENILE CANCER - MARCH 2018 15


6.1.3.4 Glansectomy
Results of glansectomy were reported in three studies [100, 113, 114], while a fourth also reported on glans-
preserving surgery [114]. One study reported 87 patients with six local (6.9%), eleven regional (12.6%) and
two systemic recurrences (2.3%) with a mean follow-up of 42 months [100]. The other two studies reported on
a total of 68 patients with a follow-up of 63 [114] and 114 months [113], respectively, in which there was one
patient (8%) with local recurrence [113], six (9%) with inguinal nodal recurrence, and no cancer-specific deaths.

6.1.3.5 Partial penectomy


Results of partial penectomy were reported in rather heterogeneous studies with a total of 184 patients with
T1-T3 tumours and a follow-up of 40-194 months [93, 114-119]. The rate of local recurrence ranged from
4-50% and cancer-specific mortality from 0-27%. The reported five-year OS ranged from 59-89% [117, 119,
120].

6.1.3.6 Summary of results of surgical techniques


Although conservative, organ-sparing surgery may improve quality of life (QoL), local recurrence is more likely
than after amputation surgery for penile cancer. In one study the local recurrence rate after organ-sparing
surgery was 18%, most of these occurred within 36 months [121], and amputation was necessary in 17% of
the recurrences. Compared to this, the local recurrence rate after amputation surgery (partial or radical) was
lower (4%). Glansectomy with circumcision for the treatment of small penile lesions has a very low rate of local
recurrence (2%) [100].
In one large cohort of patients undergoing organ-sparing surgery, isolated local recurrence
was 8.9% and five-year disease-specific survival (DSS) 91.7%. Tumour grade, stage and lymphovascular
invasion were predictors of local recurrence. In the largest cohort of penile surgery, the 5-year cumulative
incidence of local recurrence after organ-sparing (including laser treatment) was 27% while it was only 3.8%
in the amputation group [98]. Of the 451 patients treated by organ-sparing surgery, 16% eventually underwent
amputation. However, there was no significant difference in survival between the organ-sparing and the
amputation groups. These results suggest that the local recurrence rates following penile preserving surgery
are higher than with partial penectomy, although survival appears to be unaffected.

6.1.4 Summary of results of radiotherapy for T1 and T2 disease


Radiotherapy is an organ-preserving approach with good results in selected patients with T1-2 lesions
< 4 cm in diameter [122-127] (LE: 2b). It can be given as external radiotherapy with a minimum dose of 60 Gy
combined with a brachytherapy boost or as brachytherapy alone [123, 125]. Reported results are best with
brachytherapy with local control rates ranging from 70-90% [123, 125]. The American Brachytherapy Society
and the Groupe Européen de Curiethérapie-European Society of Therapeutic Radiation Oncology consensus
statement for penile brachytherapy also reported good tumour control rates, acceptable morbidity and
functional organ preservation for penile brachytherapy for stages T1 and T2 [128]. Penile preservation rates of
70-88% have been reported [129], with overall penile conservation rates of 87% and 70% at five and ten years.
Pulsed-dose-rate brachytherapy has been introduced as a new modality and 15% local recurrences have been
reported in one series [130].
In the few comparisons of surgical treatment and radiotherapy, results of surgery were slightly
better. In a meta-analysis comparing surgery and brachytherapy, 5-year OS and local control rates with surgery
were 76%/84% for surgery and 73%/79% for brachytherapy, respectively [131]. The organ preservation rate
for brachytherapy was 74% and there was no difference in survival. Local recurrence after radiotherapy can be
salvaged by surgery [132].
Specific complications of radiotherapy for penile cancer are urethral stenosis (20-35%), glans
necrosis (10-20%) and late fibrosis of the corpora cavernosa [133] (LE: 3). With brachytherapy, meatal stenosis
has been reported to occur in 40% of cases, but was much lower in a contemporary series of 73 patients with
only 6.6%. In that series, 2.6% reported pain with sexual intercourse and 5.3% dysuria over a follow-up of
5 years. Penile amputation for necrosis was necessary in 6.8 % of patients [134].
Functional outcome after radiotherapy has not often been reported. In one report, 17/18 patients
with normal erections before treatment maintained these after treatment [129].
Table 10 provides an overview of the complications and outcomes of primary local treatments.

16 PENILE CANCER - MARCH 2018


Table 10: Summary of reported complications and oncological outcomes of local treatments*

Treatment Complications Local Nodal Cancer-specific References


recurrence recurrence deaths
ND:YAG laser n.r. 10-48% 21% 2-9% [94-96, 102]
C02 laser Bleeding, meatal 14-23% 2-4% n.r. [91-93]
stenosis, both < 1%
Lasers (unspecified) Bleeding 8%, 11-26% 2% 2-3% [104-107]
local infection 2%
Moh’s micrographic Local infection 3%, 32% 8% 3-4% [108-110]
surgery Meatal stenosis 6%
Glans resurfacing n.r. 4-6% n.r. n.r. [88, 111, 112,
135]
Glansectomy n.r. 8% 9% n.r. [113, 114]
Partial penectomy n.r. 4-13% 14-19% 11-27% [93, 117, 119,
120]
Brachytherapy Meatal stenosis > 10-30% n.r. n.r. [122, 123, 125]
40%
External beam Urethral stenosis n.r. n.r. n.r. [123, 127, 128,
radiotherapy 20-35%, Glans 132, 133]
necrosis 10-20%
*The ranges are the lowest and highest number of occurrences reported in different series.

6.1.5 Treatment recommendations for invasive penile cancer (T2-T4)


6.1.5.1 Treatment of invasive disease confined to the glans with or without urethral involvement (T2)
Total glansectomy, with or without resurfacing of the corporeal heads, is recommended [115] (LE: 3).
Radiotherapy is an option (see Section 6.1.6). Partial amputation should be considered in patients unfit for
reconstructive surgery [132].

6.1.5.2 Treatment of disease invading the corpora cavernosa and/or urethra (T3)
Glansectomy with distal corporectomy and reconstruction or partial amputation with reconstruction are
standard [100, 101, 126]. Radiation therapy is an option.

6.1.5.3 Treatment of locally advanced disease invading adjacent structures (T4)


Extensive partial amputation or total penectomy with perineal urethrostomy is the standard advisable treatment
[101]. For locally advanced and ulcerated cases, neoadjuvant chemotherapy may be an option. Otherwise,
adjuvant chemotherapy or palliative radiotherapy are options (see Sections 6.2.4 and 6.1.6).

6.1.5.4 Local recurrence after organ-conserving surgery


A second organ-conserving procedure can be performed if there is no corpus cavernosum invasion [97, 101,
121, 126, 136]. For large or high-stage recurrence, partial or total amputation is required [133]. A total phallic
reconstruction may be offered to patients undergoing total/subtotal amputation [137, 138].

6.1.6 Guidelines for stage-dependent local treatment of penile carcinoma

Primary tumour Use organ-preserving treatment whenever possible Strength rating


Tis Topical treatment with 5-fluorouracil (5-FU) or imiquimod for Strong
superficial lesions with or without photodynamic control.
Laser ablation with carbon dioxide (CO2) or
neodymium:yttrium-aluminium-garnet (Nd:YAG) laser.
Glans resurfacing.
Ta, T1a (G1, G2) Wide local excision with circumcision, CO2 or Nd:YAG Strong
laser with circumcision.
Laser ablation with CO2 or Nd:YAG laser.
Glans resurfacing.
Glansectomy with reconstruction.
Radiotherapy for lesions < 4 cm.

PENILE CANCER - MARCH 2018 17


T1b (G3) and T2 Wide local excision plus reconstruction. Strong
Glansectomy with circumcision and reconstruction.
Radiotherapy for lesions < 4 cm in diameter.
T3 Partial amputation with reconstruction or radiotherapy for Strong
lesions < 4 cm in diameter.
T3 with invasion of the urethra Partial penectomy or total penectomy with perineal Strong
urethrostomy.
T4 Neoadjuvant chemotherapy followed by surgery in Weak
responders or palliative radiotherapy.
Local recurrence Salvage surgery with penis-sparing in small recurrences Weak
or partial amputation.
Large or high-stage recurrence: partial or total amputation.

6.2 Management of regional lymph nodes


The development of lymphatic metastases in penile cancer follows the route of anatomical drainage. The
inguinal lymph nodes, followed by the pelvic lymph nodes, provide the regional drainage system of penis. The
superficial and deep inguinal lymph nodes are the first regional node group to be affected, which can be uni- or
bilateral [87].
The ‘sentinel’ inguinal nodes, i.e. those first affected by lymphatic spread, appear to be located in
the medial superior zone followed by the central inguinal zones [90]. No solitary lymphatic spread has been
observed from the penis to the two inferior groin regions and no direct drainage to the pelvic nodes, either
[88, 97]. These findings confirm earlier studies.
Pelvic nodal disease does not occur without ipsilateral inguinal lymph node metastasis. Also,
crossover metastatic spread, from one groin to the contralateral pelvis, has never been reported. Further
lymphatic spread from the pelvic nodes to retroperitoneal nodes (para-aortic, para-caval) is classified as
systemic metastatic disease.
The management of regional lymph nodes is decisive for patient survival. Cure can be achieved in
limited lymph node disease confined to the regional lymph nodes. Radical lymphadenectomy is the treatment
of choice. Multimodal treatment combining surgery and chemotherapy is often indicated.
The management of regional lymph nodes is dependent on the clinical inguinal lymph node status.
There are three possible scenarios. First, the clinical lymph nodes appear normal on palpation and are not
enlarged. Secondly, the inguinal lymph nodes are palpably enlarged, either uni- or bilaterally. Thirdly, there are
grossly enlarged and sometimes ulcerated inguinal lymph nodes, uni- or bilaterally.
In clinically node-negative patients (cN0), micro-metastatic disease occurs in up to 25% of cases
and invasive lymph node staging is required since no imaging technique can reliably detect or exclude micro-
metastatic disease. In clinically positive lymph nodes (cN1/cN2), metastatic disease is highly likely and lymph
node surgery with histology is required. Enlarged fixed inguinal lymph nodes (cN3) require multimodal treatment
by (neoadjuvant) chemotherapy and surgery. Even if present in only one node, capsular penetration/extra-nodal
extension in lymph node metastasis carries a high risk of progression and is classified as pN3, which also
requires multimodal treatment.

6.2.1 Management of patients with clinically normal inguinal lymph nodes (cN0)
Risk stratification for the micro-metastatic inguinal lymph node disease depends on stage, grade and the
presence/absence of lymphovascular invasion in the primary tumour [100]. pTa/pTis tumours and those
with low grade have a comparatively low risk of lymphatic spread. Well-differentiated G1 pT1 tumours are
considered low risk, pT1G2 intermediate risk and pT1G3 and all higher stage tumours are considered high risk
for lymphatic spread [101].
For these patients, three management strategies are possible: surveillance, invasive nodal staging
or radical lymphadenectomy. Early inguinal lymphadenectomy in clinically node-negative patients is superior
for long-term patient survival compared to later lymphadenectomy with regional nodal recurrence [91, 92].
One prospective study comparing bilateral lymphadenectomy, radiotherapy and surveillance in such patients
reported significantly better five-year OS lymphadenectomy vs. inguinal radiotherapy or surveillance (74% vs.
66% and 63%, respectively) [93].

6.2.1.1 Surveillance
Surveillance of regional lymph nodes carries the risk of regional recurrence arising later from existing
micro-metastatic disease. Patient survival is over 90% with early lymphadenectomy and below 40% with
lymphadenectomy for regional recurrence [94, 95]. Patients considering surveillance must be informed about
this risk. Surveillance is only recommended in patients with pTis/pTa tumours and with the appropriate caveats
in low risk G1 pT1 tumours [94-96]. Compliance is required for surveillance.

18 PENILE CANCER - MARCH 2018


6.2.1.2 Invasive nodal staging
Since no imaging technique can detect micro-metastatic disease, invasive lymph node staging is
recommended for pT1 tumours of intermediate and high risk, as well as for T2-T4 tumours [92, 105]
(LE: 2b). Fine-needle aspiration cytology also does not reliably exclude micro-metastatic disease and is not
recommended.
Invasive nodal staging can be done by either dynamic sentinel-node biopsy (DSNB) or by modified
inguinal lymphadenectomy (mILND), both of which are standard techniques [139]. Dynamic sentinel-node
biopsy aims to detect affected sentinel nodes in both groins. Technetium-99m (99mTc) nanocolloid is injected
around the penile cancer site on the day before surgery often combined with patent blue. A gamma-ray probe
is used intra-operatively to detect the sentinel nodes, which is possible in 97% of cases. The protocol has
been standardised for routine use [107]. Dynamic sentinel-node biopsy has a reported high sensitivity in some
centres (90-94%) [107, 108] (LE: 2b). In a meta-analysis of eighteen studies, the pooled sensitivity was 88%,
which improved to 90% with the addition of patent blue [109].
Modified ILND is an alternative option, whereby the medial superficial inguinal lymph nodes and those
from the central zone are removed bilaterally [87, 106] (LE: 3), leaving the greater saphenous vein untouched.
Both methods of invasive lymph node staging may miss micro-metastatic disease leading to
regional recurrence [91]. The false-negative rate may be as high as 12-15% for DSNB, even in experienced
centres [95, 96]. The false-negative rate of mILND is unknown. If lymph node metastasis is found, ipsilateral
radical inguinal lymphadenectomy is indicated.

6.2.2 Management of patients with palpable inguinal nodes (cN1/cN2)


With uni- or bilateral palpable inguinal lymph nodes (cN1/cN2), metastatic lymph node disease is highly likely.
The notion that these may be inflammatory and that antibiotic treatment should first be used is unfounded and
dangerous as it delays curative treatment.
Palpably enlarged groin lymph nodes should be surgically removed, pathologically assessed (by
frozen section) and, if positive, a radical inguinal lymphadenectomy should be performed. In clinically doubtful
cases, US-guided fine needle aspiration cytology is an option [140].
In such cases, CT or MRI can provide staging information about the pelvic nodal status and
18F-FDGPET/CT can identify additional metastases [141]. Dynamic sentinel-node biopsy is not indicated in

patients with palpably enlarged lymph nodes [142] (LE: 3).

6.2.2.1 Radical inguinal lymphadenectomy


Radical inguinal lymphadenectomy carries a significant morbidity due to impaired lymph drainage from the
legs and scrotum. Morbidity can be as high as 50% [143] in the presence of significant risk factors such as
increased body mass index (BMI). Recent series have reported lower morbidity in about 25% of cases [144,
145] (LE: 2b). Therapeutic radical inguinal lymphadenectomy can be life-saving and should not be underused
for fear of associated morbidity [146].
Tissue handling must be meticulous in order to minimise post-operative morbidity. Lymphatic vessel
walls do not contain smooth muscle and are therefore not reliably closed by electrocautery. Numerous metal
clips may also cause post-operative problems so that ligation of all lymphatic vessels is advisable [147, 148].
Post-operative morbidity may be reduced by preserving the saphenous vein and post-operative measures
to improve drainage, such as stockings, bandaging, inguinal pressure dressings or vacuum suction and
prophylactic antibiotics [149]. Transposition of the Sartorius muscle is not recommended. There is no benefit
from using fibrin glue intraoperatively [150]. Advanced cases may require reconstructive surgery for wound
closure. The most commonly reported complications in recent series were wound infections (1.2-1.4%), skin
necrosis (0.6-4.7%), lymphoedema (5-13.9%) and lymphocele formation (2.1-4%) [144, 145].
Minimally-invasive surgical techniques (laparoscopic, robot-assisted) for inguinal lymphadenectomy
are technically feasible and, in small series, have been reported to significantly reduce post-operative morbidity
except for the rate of lymphoceles [144, 150-153].

6.2.2.2 Pelvic lymphadenectomy


Patients with two or more inguinal lymph node metastases on one side and/or extracapsular lymph node
extension need to undergo ipsilateral pelvic lymphadenectomy. This recommendation is based on a study in
which the rate of positive pelvic nodes was found to be 23% in cases with more than two positive inguinal nodes
and 56% in those with more than three positive inguinal nodes or extracapsular extension [101, 154] (LE: 2b).
Positive pelvic nodes carry a worse prognosis than only inguinal nodal metastasis (five-year CSS
71.0% vs. 33.2%) [155]. In a study of 142 groin node-positive patients, significant risk factors for pelvic nodal
metastasis were the number of positive inguinal nodes (cut-off three), the diameter of inguinal metastatic nodes
(cut-off 30 mm) and extra-nodal extension. The percentage of pelvic nodal metastases was 0% without any of
these risk factors and 57.1% with all three risk factors present [155].

PENILE CANCER - MARCH 2018 19


Pelvic lymphadenectomy may be performed simultaneously with inguinal lymphadenectomy or
as a secondary procedure. If bilateral pelvic dissection is indicated, it can be performed through a midline
suprapubic extraperitoneal incision. It is important to avoid unnecessary delay if these procedures are indicated
[156].

6.2.2.3 Adjuvant treatment


In patients with pN2/pN3 disease, adjuvant chemotherapy is recommended after lymphadenectomy [157]
(see Section 6.3.1). One retrospective study reported long-term DFS of 84% in node-positive patients with
adjuvant chemotherapy after radical lymph node surgery vs. 39% in historical controls without adjuvant
chemotherapy after lymphadenectomy [157]. More recent studies have confirmed the survival benefit of
adjuvant chemotherapy after radical inguinal lymphadenectomy [158-160].
Although adjuvant radiotherapy has been used after inguinal lymphadenectomy, there are no
data showing definite patient benefit. Adjuvant radiotherapy after inguinal lymphadenectomy should not be
administered outside of clinical studies.

6.2.3 Management of patients with fixed inguinal nodes (cN3)


Patients with large and bulky, sometimes ulcerated, inguinal lymph nodes require staging by thoracic,
abdominal and pelvic CT for pelvic nodes and systemic disease. In clinically unequivocal cases, histological
verification by biopsy is not required.
These patients have a poor prognosis. Multimodal treatment with neoadjuvant chemotherapy
followed by radical lymphadenectomy in responders is recommended [161-163]. Responders to neoadjuvant
chemotherapy with post-chemotherapy surgery have been reported to achieve long-term survival in 37% of
cases [161]. Contemporary studies have confirmed this patient benefit [162, 164, 165].

6.2.4 Management of lymph node recurrence


Patients with regional recurrence should be treated in the same way as patients with primary cN1/cN2 disease.
However, patients with regional lymph node recurrence after DSNB or modified inguinal lymphadenectomy
already have disordered inguinal lymphatic drainage and are at a high risk of irregular metastatic progression.
Inguinal nodal recurrence after radical inguinal lymphadenectomy has a five-year CSS rate of 16% [166].
There is no evidence for the best management in such cases. Multimodal treatment with
neoadjuvant and/or adjuvant chemotherapy after radical lymph node surgery is recommended.

6.2.5 The role of radiotherapy in lymph node disease


Radiotherapy is used in some institutions for the treatment of inguinal lymph nodes. However, this
is not evidence-based. One of the rare prospective trials in penile cancer found that inguinal radical
lymphadenectomy is superior to inguinal radiotherapy for lymph-node positive penile cancer patients [167].
There is no evidence that adjuvant radiotherapy after radical inguinal lymphadenectomy improves
oncological outcome [168]. One study reported poor long-term survival in patients with adjuvant inguinal and
pelvic radiotherapy [169]. Other studies have likewise not demonstrated a patient benefit [168-174].
In one comparative retrospective study, adjuvant chemotherapy was far superior to adjuvant
radiotherapy after radical inguinal lymphadenectomy in node-positive patients [157]. A large retrospective
analysis of the SEER database (National Cancer Institute Surveillance, Epidemiology and End Results Program)
of 2,458 penile cancer patients treated with either surgery alone or surgery plus EBRT concluded that the
addition of adjuvant EBRT ‘had neither a harmful nor a beneficial effect on CSS’ [175].
Due to this lack of positive evidence, radiotherapy cannot be recommended outside of controlled
trials for the treatment of lymph node disease in penile cancer. Prophylactic radiotherapy for cN0 disease is not
indicated. Radiotherapy for advanced lymph node disease remains a palliative option.

6.2.6 Guidelines for treatment strategies for nodal metastases

Regional lymph nodes Management of regional lymph nodes is fundamental in Strength rating
the treatment of penile cancer
No palpable inguinal nodes Tis, Ta G1, T1G1: surveillance. Strong
(cN0) > T1G2: invasive lymph node staging by either bilateral Strong
modified inguinal lymphadenectomy or dynamic sentinel
node biopsy.
Palpable inguinal nodes Radical inguinal lymphadenectomy. Strong
(cN1/cN2)
Fixed inguinal lymph nodes Neoadjuvant chemotherapy followed by radical inguinal Weak
(cN3) lymphadenectomy in responders.

20 PENILE CANCER - MARCH 2018


Pelvic lymph nodes Ipsilateral pelvic lymphadenectomy if two or more inguinal Strong
nodes are involved on one side (pN2) or if extracapsular
nodal metastasis (pN3) reported.
Adjuvant chemotherapy In pN2/pN3 patients after radical lymphadenectomy. Strong
Radiotherapy Not recommended for nodal disease except as a palliative Strong
option.

6.3 Chemotherapy
6.3.1 Adjuvant chemotherapy in node-positive patients after radical inguinal lymphadenectomy
Multimodal treatment can improve patient outcome. Adjuvant chemotherapy after radical lymphadenectomy
in node-positive patients has been reported in a few small and heterogeneous series [158-160]. Comparing
different small-scale clinical studies is fraught with difficulty.
The value of adjuvant chemotherapy after radical inguinal lymphadenectomy in node-positive penile
cancer was first demonstrated by a study which reported long-term (DFS) of 84% in 25 consecutive patients
treated with twelve adjuvant weekly courses of vincristine, bleomycin, and methotrexate (VBM) during 1979-
1990 and compared this to a historical control group of 38 consecutive node-positive patients with radical
lymphadenectomy (with- or without adjuvant inguinal radiotherapy) who had achieved a DFS rate of only 39%
[161].
The same group also published results of an adjuvant chemotherapy regimen with three courses
of cisplatin and 5-FU with lower toxicity and even better results compared to VBM [176] (LE: 2b). The same
group has published results of adjuvant chemotherapy with cisplatin, 5-FU plus paclitaxel or docetaxel (TPF),
with three to four cycles after resection of pN2-3 disease [177]. Of 19 patients, 52.6% were disease-free after
a median follow up of 42 months and tolerability was good. Results of adjuvant treatment with paclitaxel and
cisplatin also improved outcome [178].
Therefore, the use of adjuvant chemotherapy is recommended, in particular when the administration
of the triple combination chemotherapy is feasible and there is curative intent (LE: 2b). There are no data
concerning adjuvant chemotherapy in stage pN1 patients. Adjuvant chemotherapy in pN1 disease is, therefore,
recommended only in clinical trials.

6.3.2 Neoadjuvant chemotherapy in patients with fixed or relapsed inguinal nodes


Bulky inguinal lymph node enlargement (cN3) indicates extensive lymphatic metastatic disease. Primary
lymph node surgery is not generally recommended since complete surgical resection is unlikely and only a few
patients will benefit from surgery alone.
Limited data is available on neoadjuvant chemotherapy before inguinal lymph node surgery.
However, it allows for early treatment of systemic disease and down-sizing of the inguinal lymph node
metastases. In responders, complete surgical treatment is possible with a good clinical response.
Results of neoadjuvant chemotherapy for bulky inguinal lymph node metastases were modest in
retrospective studies including five to twenty patients treated with bleomycin-vincristine-methotrexate (BVM)
or bleomycin-methotrexate-cisplatin (BMP) regimens [162, 163, 179], as well as in the confirmatory BMP trial of
the Southwest Oncology Group [180]. However, treatment-related toxicity was unacceptable due to bleomycin-
related mortality.
Cisplatin/5-FU (PF) chemotherapy achieved a response rate of 25-50% with more acceptable
toxicity [181, 182]. Over a period of 30 years, five different neoadjuvant chemotherapy regimens were used in
twenty patients [87], with long-term survival in 37% of responders who underwent radical lymph node surgery
after neoadjuvant chemotherapy. In the EORTC cancer study 30992, 26 patients with locally advanced or
metastatic disease received irinotecan and cisplatin chemotherapy. Although the study did not meet its primary
endpoint (response rate), there were three cases of pathologically complete remissions [183].
A phase II trial evaluated treatment with four cycles of neoadjuvant paclitaxel, cisplatin, and
ifosfamide (TIP). An objective response was reported in 15/30 patients, including three pathologically complete
remissions (pCRs). The estimated median time to progression (TTP) was 8.1 months and the median OS was
17.1 months [164] (LE: 2a).
Hypothetical similarities between penile SCC and head and neck SCC led to the evaluation, in
penile cancer, of chemotherapy regimens with an efficacy in head and neck SCC, including taxanes. The
combination of cisplatin and 5-FU plus a taxane has been used in neoadjuvant and adjuvant settings [177].
An overall objective response rate of 44% was reported in 28 patients treated neoadjuvantly, including 14%
pCR (LE: 2b). Similarly, a phase II trial with TPF using docetaxel instead of paclitaxel reported an objective
response of 38.5% in 29 locally advanced or metastatic patients, although the study did not meet its primary
endpoint. However, there was significant toxicity [184] (LE: 2a). Further evidence of the benefit of neoadjuvant
chemotherapy was published recently [165].

PENILE CANCER - MARCH 2018 21


Overall, these results support the recommendation that neoadjuvant chemotherapy using a
cisplatin- and taxane-based triple combination should be used in patients with fixed, unresectable, nodal
disease (LE: 2a).
There are hardly any data concerning the potential benefit of radiochemotherapy together with
lymph node surgery in penile cancer. It should therefore only be used in controlled clinical trials [185].

6.3.3 Palliative chemotherapy in advanced and relapsed disease


A recent retrospective study of 140 patients with advanced penile SCC reported that visceral metastases and
an > 1 ECOG-performance status were independent prognostic factors, and that cisplatin-based regimens had
better outcomes than non-cisplatin-based regimens after adjusting for prognostic factors [186] (LE: 3).
Before taxanes were introduced, chemotherapy data in penile cancer were limited by small
numbers, patient heterogeneity and retrospective design (except for the EORTC trial [183]). Initial response
rates ranged from 25% to 100%, with very few sustained responses and very few long-term survivors.
The introduction of taxanes into penile cancer chemotherapy has enhanced the activity and efficacy of the
regimens used [87, 162-164, 178-184, 187].
There are virtually no data on second-line chemotherapy in penile cancer. One report using
second-line paclitaxel monotherapy reported a response rate of < 30% and no patient survived [188] (LE: 2a).
Anecdotally, a benefit of second-line cisplatin with gemcitabine has been observed [189] (LE: 4).

6.3.4 Intra-arterial chemotherapy


Intra-arterial chemotherapy which refers to intra-aortic application has been trialled in locally advanced cases,
especially of cisplatin and gemcitabine in small case series [190-193]. Apart from a limited clinical response,
the outcome was not significantly improved.

6.3.5 Targeted therapy


Targeted drugs have been used as second-line treatment and they could be considered as single-agent
treatment in refractory cases. Anti-epidermal growth factor receptor (EGFR) targeted monotherapy has been
trialled [194], as EGFR is expressed in penile SCC [190, 191] and there are assumed similarities with head and
neck SCC [190, 191]. There have been other studies, particularly with the anti-EGFR monoclonal antibodies
panitumumab and cetuximab, without long-term response, however [195]. Some activity of tyrosine kinase
inhibitors has been reported as well [193]. Further clinical studies are needed (LE: 4).

6.3.6 Guidelines for chemotherapy

Recommendations Strength rating


Offer patients with pN2-3 tumours adjuvant chemotherapy after radical lymphadenectomy Strong
(three to four cycles of cisplatin, a taxane and 5-fluorouracil or ifosfamide).
Offer patients with non-resectable or recurrent lymph node metastases neoadjuvant Weak
chemotherapy (four cycles of a cisplatin- and taxane-based regimen) followed by radical
surgery.
Offer palliative chemotherapy to patients with systemic disease. Weak

7. FOLLOW-UP
7.1 Rationale for follow-up
Early detection of recurrence increases the likelihood of curative treatment since local recurrence does not
significantly reduce long-term survival if successfully treated [87, 196]. In contrast, disease that has spread
to the inguinal lymph nodes greatly reduces the rate of long-term DSS. Follow-up is also important in the
detection and management of treatment-related complications.
Local or regional nodal recurrences usually occur within two years of primary treatment [87]. After
five years, all recurrences were either local or new primary lesions [87]. This supports an intensive follow-up
regimen during the first two years, with a less intensive follow up later for a total of at least five years. Follow-
up after five years may be omitted in motivated patients who will undertake regular self-examination reliably
[87].

22 PENILE CANCER - MARCH 2018


7.1.1 When and how to follow-up
After local treatment with negative inguinal nodes, follow-up should include physical examination of the penis
and groins for local and/or regional recurrence. Additional imaging has no proven benefit. Follow-up also
depends on the primary treatment modality. Histology from the glans should be obtained to confirm disease-
free status following laser ablation or topical chemotherapy.
After potentially curative treatment for inguinal nodal metastases, CT or MRI imaging for the
detection of systemic disease should be performed at three-monthly intervals for the first two years.
Although rare, late local recurrence may occur, with life-threatening metastases becoming very
unusual after five years. Therefore, regular follow up can be stopped after five years, provided the patient
understands the need to report any local changes immediately [197]. In patients unlikely to self-examine, long-
term follow up may be necessary.

7.1.2 Recurrence of the primary tumour


Local recurrence is more likely with all types of local organ-sparing treatment but does not influence the rate
of cancer-specific survival in contrast to regional lymph node recurrence [87, 196]. Local recurrence occurred
during the first two years in up to 27% of patients treated with penis-preserving modalities [98]. After partial
penectomy, the risk of local recurrence is about 4-5% [87, 98, 196].
Local recurrence is easily detected by physical examination, by the patient himself or his physician.
Patient education is an essential part of follow-up and the patient should be urged to visit a specialist if any
changes are seen.

7.1.3 Regional recurrence


Most regional recurrences occur during the first two years after treatment, irrespective of whether surveillance
or invasive nodal staging were used. Although unlikely, regional recurrence can occur later than two years
after treatment. It is therefore advisable to continue follow up in these patients [197]. The highest rate of
regional recurrence (9%) occurs in patients managed by surveillance, while the lowest is in patients who have
undergone invasive nodal staging by modified inguinal lymphadenectomy or DSNB and whose lymph nodes
were negative (2.3%).
The use of US and fine needle aspiration cytology (FNAC) in suspicious cases has improved the
early detection rate of regional recurrence [76, 198, 199]. There are no data to support the routine use of CT or
MRI for the follow-up of inguinal nodes.
Patients who have had surgery for lymph node metastases without adjuvant treatment have an
increased risk of regional recurrence of 19% [87]. Regional recurrence requires timely treatment by radical
inguinal lymphadenectomy and adjuvant chemotherapy (see Section 6).

7.1.4 Guidelines for follow-up in penile cancer

Interval of follow-up Examinations and Minimum duration Strength rating


investigations of follow-up
Years Years
one to two three to five
Recommendations for follow-up of the primary tumour
Penile-preserving Three Six months Regular physician Five years Strong
treatment months or self-examination.
Repeat biopsy after
topical or laser
treatment for penile
intraepithelial neoplasia.
Amputation Three One year Regular physician or ­ Five years Strong
months self-examination.
Recommendations for follow-up of the inguinal lymph nodes
Surveillance Three Six months Regular physician or Five years Strong
months self-examination.
pN0 at initial Three One year Regular physician Five years Strong
treatment months or self-examination.
Ultrasound with fine-
needle aspiration biopsy
optional.

PENILE CANCER - MARCH 2018 23


pN+ at initial Three Six months Regular physician Five years Strong
treatment months or self-examination.
Ultrasound with fine-
needle aspiration
cytology optional,
computed tomography/
magnetic resonance
imaging optional.

7.2 Quality of life


7.2.1 Consequences after penile cancer treatment
In patients with long-term survival after penile cancer treatment, sexual dysfunction, voiding problems and
cosmetic penile appearance may adversely affect the patient’s QoL [200]. However, there is very little data
on sexual function and QoL after treatment for penile cancer. In particular, there is heterogeneity of the
psychometric tools used to assess QoL outcomes and further research is needed to develop disease-specific
patient reported outcome measures for penile cancer.

Comparative studies
There are only two comparative studies in the literature reporting on the health-related quality of life (HRQoL)
outcomes following surgery for localised penile cancer. One study compared wide local excision with
glansectomy [201]. Among 41 patients there was reduction in post-operative International Index of Erectile
Function (IIEF) and the authors concluded that local excision led to better sexual outcomes than glansectomy.
In another study of 147 patients, the IIEF-15, the SF36 Health Survey and the Impact of Cancer questionnaire
were used [202].
Compared to an age-matched population sample, men after partial penectomy reported
significantly more problems with orgasm, cosmesis, life interference and urinary function than those who had
undergone penile-sparing surgery (83% vs. 43%, p < 0.0001). Interestingly, there were no differences in erectile
function, sexual desire, intercourse satisfaction or overall sexual satisfaction.

7.2.2 Sexual activity and quality of life after laser treatment


A retrospective interview-based Swedish study after laser treatment for penile PeIN [104] in 58 out of 67
surviving patients with a mean age of 63 years, of whom 46 participated, reported a marked decrease in some
sexual practices, such as manual stimulation, caressing and fellatio, but a general satisfaction rate with life
overall and sexuality which was similar to that of the general Swedish population.
A large study on CO2 laser treatment of penile cancer in 224 patients reported no problems with
erectile or sexual function following treatment [91]. In another study [107], no sexual dysfunction occurred in
nineteen patients treated.

7.2.3 Sexual activity after glans resurfacing


In one study with ten patients [111], 7/10 completed questionnaires (IIEF-5 and a non-validated 9-item
questionnaire) at six months. The median IIEF-5 score was 24 (no erectile dysfunction). All patients who were
sexually active before treatment were active after three to five months, 7/7 stated that the sensation at the tip
of their penis was either no different or better after surgery, and 5/7 patients felt that their sex life had improved.
Overall patient satisfaction with glans resurfacing was high.

7.2.4 Sexual activity after glansectomy


Two studies reported sexual function after glansectomy [112, 113]. In one (n = 68) with unclear methodology
[113], 79% did not report any decline in spontaneous erection, rigidity or penetrative capacity after surgery, and
75% reported recovery of orgasm. In the other study [114], all twelve patients had returned to ‘normal’ sexual
activity one month after surgery.

7.2.5 Sexual function after partial penectomy


Sexual function after partial penectomy was reported by three studies [203-205]. In one with 18 patients with a
mean age of 52 years, the IIEF scores were significantly worse for all domains of sexual function after surgery
[203] and 55.6% of patients had erectile function that allowed sexual intercourse. In patients who did not
resume sexual activity, 50% were ashamed of their small penis and missing glans, while another third blamed
surgical complications. Of those who had resumed sexual intercourse, 66.7% reported the same frequency
and level of sexual activity as before surgery, while 72.2% continued to have ejaculation and orgasm every time
with sexual activity. Overall, only 33.3% maintained their pre-operative frequency of sexual intercourse and
were satisfied with their sex life.

24 PENILE CANCER - MARCH 2018


In another study, an ‘Overall Sexual Functioning Questionnaire’ was used in 14 patients with a
median time of 11.5 months after surgery (range 6-72) [204]. Prior to surgery, all patients had had normal
erectile function and intercourse at least once a month. In 9/14 patients, sexual function was ‘normal’
or ‘slightly decreased’, while 3/14 had had no sexual intercourse since surgery. Alei et al. reported an
improvement in erectile function over time [205]. In a report of 25 patients after partial penectomy and
neoglans formation, the IIEF-5, Quality of Erection Questionnaire (QEQ), Erectile Dysfunction Inventory of
Treatment Satisfaction (EDITS) and Self-Esteem and Relationship (SEAR) were used. This study reported a high
percentage of patient and partner satisfaction with surgical treatment and recovery of sexual function, self-
esteem, and overall relationship satisfaction [206].

7.2.6 Quality of life and sexual function after total penectomy


In ten patients with penile cancer evaluated after total amputation of the penis, there were significant effects on
sexual life and overall QoL, although there were no negative implications in terms of partner relationships, self-
assessment or the evaluation of masculinity [207].

7.2.7 Quality of life after partial penectomy


Several qualitative and quantitative instruments have been used to assess ‘psychological behaviour and
adjustment’ and ‘social activity’ as QoL indicators [204, 208]. Patient-reported fears were those of mutilation,
loss of sexual pleasure and of cancer death and what this would mean for their families. The study reported no
significant levels of anxiety and depression on the General Health Questionnaire-12 and the Hospital Anxiety
and Depression Scale. ‘Social activity’ remained the same after surgery in terms of living conditions, family life
and social interactions.

7.3 Total phallic reconstruction


There is very limited data about total phallic reconstruction following full or near-total penile amputation
[137, 209, 210]. Although it is not possible to restore function without a penile prosthesis, cosmetically
acceptable results can be obtained.

7.4 Specialised care


Since penile cancer is rare, patients should be referred to a centre with experience and expertise in local
treatment, pathological diagnosis, chemotherapy and psychological support for penile cancer patients. Some
countries have centralised the care of penile cancer patients (Sweden, Denmark, the Netherlands, the UK).

8. REFERENCES
1. Hakenberg, O.W., et al. EAU guidelines on penile cancer: 2014 update. Eur Urol, 2015. 67: 142.
https://www.ncbi.nlm.nih.gov/pubmed/25457021
2. Robinson, R.N., et al. What are the risks and benefits of adjuvant radiotherapy after inguinal
lymphadenectomy for penile cancer? PROSPERO, 2015.
http://www.crd.york.ac.uk/PROSPERO/display_record.php?ID=CRD42015024904
3. Clark, P.E., et al. Penile cancer: Clinical Practice Guidelines in Oncology. J Natl Compr Canc Netw,
2013. 11: 594.
https://www.ncbi.nlm.nih.gov/pubmed/23667209
4. Souillac, I., et al. [Penile cancer in 2010: update from the Oncology Committee of the French
Association of Urology: external genital organs group (CCAFU-OGE)]. Prog Urol, 2011. 21: 909.
https://www.ncbi.nlm.nih.gov/pubmed/22118355
5. Van Poppel, H., et al. Penile cancer: ESMO Clinical Practice Guidelines for diagnosis, treatment and
follow-up. Ann Oncol, 2013. 24 Suppl 6: vi115.
https://www.ncbi.nlm.nih.gov/pubmed/23975666
6. Guyatt, G.H., et al. GRADE: an emerging consensus on rating quality of evidence and strength of
recommendations. Bmj, 2008. 336: 924.
https://www.ncbi.nlm.nih.gov/pubmed/18436948
7. Guyatt, G.H., et al. What is “quality of evidence” and why is it important to clinicians? Bmj, 2008.
336: 995.
https://www.ncbi.nlm.nih.gov/pubmed/18456631

PENILE CANCER - MARCH 2018 25


8. Phillips, B., et al. Oxford Centre for Evidence-based Medicine Levels of Evidence. Updated by
Jeremy Howick March 2009.
http://www.cebm.net/oxford-centre-evidence-based-medicine-levels-evidence-march-2009/
9. Guyatt, G.H., et al. Going from evidence to recommendations. Bmj, 2008. 336: 1049.
https://www.ncbi.nlm.nih.gov/pubmed/18467413
10. Cubilla, A.L., et al. Pathologic features of epidermoid carcinoma of the penis. A prospective study of
66 cases. Am J Surg Pathol, 1993. 17: 753.
https://www.ncbi.nlm.nih.gov/pubmed/8338190
11. Chaux, A., et al. Papillary squamous cell carcinoma, not otherwise specified (NOS) of the penis:
clinicopathologic features, differential diagnosis, and outcome of 35 cases. Am J Surg Pathol, 2010.
34: 223.
https://www.ncbi.nlm.nih.gov/pubmed/22116602
12. Mannweiler, S., et al. Clear-cell differentiation and lymphatic invasion, but not the revised TNM
classification, predict lymph node metastases in pT1 penile cancer: a clinicopathologic study of 76
patients from a low incidence area. Urol Oncol, 2013. 31: 1378.
https://www.ncbi.nlm.nih.gov/pubmed/22421354
13. Backes, D.M., et al. Systematic review of human papillomavirus prevalence in invasive penile
cancer. Cancer Causes Control, 2009. 20: 449.
https://www.ncbi.nlm.nih.gov/pubmed/19082746
14. Chaux, A., et al. Epidemiologic profile, sexual history, pathologic features, and human
papillomavirus status of 103 patients with penile carcinoma. World J Urol, 2013. 31: 861.
https://www.ncbi.nlm.nih.gov/pubmed/22116602
15. Cancer Incidence in Five Continents Vol. VIII. IARC Scientific Publication No. 155. Vol. Vol III. 2002,
The International Agency for Research on Cancer, 150 cours Albert Thomas, 69372 Lyon CEDEX 08,
France.
http://www.iarc.fr/en/publications/pdfs-online/epi/sp155/
16. Parkin, D.M., et al. Chapter 2: The burden of HPV-related cancers. Vaccine, 2006. 24 Suppl 3: S3/11.
https://www.ncbi.nlm.nih.gov/pubmed/16949997
17. Baldur-Felskov, B., et al. Increased incidence of penile cancer and high-grade penile intraepithelial
neoplasia in Denmark 1978-2008: a nationwide population-based study. Cancer Causes Control,
2012. 23: 273.
https://www.ncbi.nlm.nih.gov/pubmed/22101453
18. Arya, M., et al. Long-term trends in incidence, survival and mortality of primary penile cancer in
England. Cancer Causes Control, 2013. 24: 2169.
https://www.ncbi.nlm.nih.gov/pubmed/24101363
19. Barnholtz-Sloan, J.S., et al. Incidence trends in primary malignant penile cancer. Urol Oncol, 2007.
25: 361.
https://www.ncbi.nlm.nih.gov/pubmed/17826651
20. Hartwig, S., et al. Estimation of the epidemiological burden of human papillomavirus-related cancers
and non-malignant diseases in men in Europe: a review. BMC Cancer, 2012. 12: 30.
https://www.ncbi.nlm.nih.gov/pubmed/22260541
21. Dillner, J., et al. Etiology of squamous cell carcinoma of the penis. Scand J Urol Nephrol Suppl,
2000: 189.
https://www.ncbi.nlm.nih.gov/pubmed/11144896
22. Maden, C., et al. History of circumcision, medical conditions, and sexual activity and risk of penile
cancer. J Natl Cancer Inst, 1993. 85: 19.
https://www.ncbi.nlm.nih.gov/pubmed/8380060
23. Tsen, H.F., et al. Risk factors for penile cancer: results of a population-based case-control study in
Los Angeles County (United States). Cancer Causes Control, 2001. 12: 267.
https://www.ncbi.nlm.nih.gov/pubmed/11405332
24. Afonso, L.A., et al. High Risk Human Papillomavirus Infection of the Foreskin in Asymptomatic Men
and Patients with Phimosis. J Urol, 2016. 195: 1784.
https://www.ncbi.nlm.nih.gov/pubmed/26796413
25. Archier, E., et al. Carcinogenic risks of psoralen UV-A therapy and narrowband UV-B therapy in
chronic plaque psoriasis: a systematic literature review. J Eur Acad Dermatol Venereol, 2012. 26
Suppl 3: 22.
https://www.ncbi.nlm.nih.gov/pubmed/22512677
26. Stern, R.S. The risk of squamous cell and basal cell cancer associated with psoralen and ultraviolet
A therapy: a 30-year prospective study. J Am Acad Dermatol, 2012. 66: 553.
https://www.ncbi.nlm.nih.gov/pubmed/22264671

26 PENILE CANCER - MARCH 2018


27. Daling, J.R., et al. Cigarette smoking and the risk of anogenital cancer. Am J Epidemiol, 1992. 135:
180.
https://www.ncbi.nlm.nih.gov/pubmed/1311142
28. Stankiewicz, E., et al. HPV infection and immunochemical detection of cell-cycle markers in
verrucous carcinoma of the penis. Mod Pathol, 2009. 22: 1160.
https://www.ncbi.nlm.nih.gov/pubmed/19465901
29. Koifman, L., et al. Epidemiological aspects of penile cancer in Rio de Janeiro: evaluation of 230
cases. Int Braz J Urol, 2011. 37: 231.
https://www.ncbi.nlm.nih.gov/pubmed/21557840
30. Thuret, R., et al. A population-based analysis of the effect of marital status on overall and cancer-
specific mortality in patients with squamous cell carcinoma of the penis. Cancer Causes Control,
2013. 24: 71.
https://www.ncbi.nlm.nih.gov/pubmed/23109172
31. McIntyre, M., et al. Penile cancer: an analysis of socioeconomic factors at a southeastern tertiary
referral center. Can J Urol, 2011. 18: 5524.
https://www.ncbi.nlm.nih.gov/pubmed/21333043
32. Benard, V.B., et al. Examining the association between socioeconomic status and potential human
papillomavirus-associated cancers. Cancer, 2008. 113: 2910.
https://www.ncbi.nlm.nih.gov/pubmed/18980274
33. Ulff-Moller, C.J., et al. Marriage, cohabitation and incidence trends of invasive penile squamous cell
carcinoma in Denmark 1978-2010. Int J Cancer, 2013. 133: 1173.
https://www.ncbi.nlm.nih.gov/pubmed/23404289
34. Lebelo, R.L., et al. Diversity of HPV types in cancerous and pre-cancerous penile lesions of South
African men: implications for future HPV vaccination strategies. J Med Virol, 2014. 86: 257.
https://www.ncbi.nlm.nih.gov/pubmed/24155172
35. Kayes, O., et al. Molecular and genetic pathways in penile cancer. Lancet Oncol, 2007. 8: 420.
https://www.ncbi.nlm.nih.gov/pubmed/17466899
36. Munoz, N., et al. Chapter 1: HPV in the etiology of human cancer. Vaccine, 2006. 24 Suppl 3: S3/1.
https://www.ncbi.nlm.nih.gov/pubmed/16949995
37. Nordenvall, C., et al. Cancer risk among patients with condylomata acuminata. Int J Cancer, 2006.
119: 888.
https://www.ncbi.nlm.nih.gov/pubmed/16557590
38. Lont, A.P., et al. Presence of high-risk human papillomavirus DNA in penile carcinoma predicts
favorable outcome in survival. Int J Cancer, 2006. 119: 1078.
https://www.ncbi.nlm.nih.gov/pubmed/16570278
39. Bezerra, A.L., et al. Human papillomavirus as a prognostic factor in carcinoma of the penis: analysis
of 82 patients treated with amputation and bilateral lymphadenectomy. Cancer, 2001. 91: 2315.
https://www.ncbi.nlm.nih.gov/pubmed/11413520
40. Philippou, P., et al. Genital lichen sclerosus/balanitis xerotica obliterans in men with penile
carcinoma: a critical analysis. BJU Int, 2013. 111: 970.
https://www.ncbi.nlm.nih.gov/pubmed/23356463
41. D’Hauwers, K.W., et al. Human papillomavirus, lichen sclerosus and penile cancer: a study in
Belgium. Vaccine, 2012. 30: 6573.
https://www.ncbi.nlm.nih.gov/pubmed/22939906
42. de Bruijn, R.E., et al. Patients with penile cancer and the risk of (pre)malignant cervical lesions in
female partners: a retrospective cohort analysis. BJU Int, 2013. 112: 905.
https://www.ncbi.nlm.nih.gov/pubmed/23905914
43. Newman, P.A., et al. HPV vaccine acceptability among men: a systematic review and meta-analysis.
Sex Transm Infect, 2013. 89: 568.
https://www.ncbi.nlm.nih.gov/pubmed/23828943
44. Fisher, H., et al. Inequalities in the uptake of human papillomavirus vaccination: a systematic review
and meta-analysis. Int J Epidemiol, 2013. 42: 896.
https://www.ncbi.nlm.nih.gov/pubmed/23620381
45. Van Howe, R.S., et al. The carcinogenicity of smegma: debunking a myth. J Eur Acad Dermatol
Venereol, 2006. 20: 1046.
https://www.ncbi.nlm.nih.gov/pubmed/16987256
46. Daling, J.R., et al. Penile cancer: importance of circumcision, human papillomavirus and smoking in
in situ and invasive disease. Int J Cancer, 2005. 116: 606.
https://www.ncbi.nlm.nih.gov/pubmed/15825185

PENILE CANCER - MARCH 2018 27


47. Velazquez, E.F., et al. Limitations in the interpretation of biopsies in patients with penile squamous
cell carcinoma. Int J Surg Pathol, 2004. 12: 139.
https://www.ncbi.nlm.nih.gov/pubmed/15173919
48. Velazquez, E.F., et al. Lichen sclerosus in 68 patients with squamous cell carcinoma of the penis:
frequent atypias and correlation with special carcinoma variants suggests a precancerous role.
Am J Surg Pathol, 2003. 27: 1448.
https://www.ncbi.nlm.nih.gov/pubmed/14576478
49. Teichman, J.M., et al. Noninfectious penile lesions. Am Fam Physician, 2010. 81: 167.
https://www.ncbi.nlm.nih.gov/pubmed/20082512
50. Renaud-Vilmer, C., et al. Analysis of alterations adjacent to invasive squamous cell carcinoma of the
penis and their relationship with associated carcinoma. J Am Acad Dermatol, 2010. 62: 284.
https://www.ncbi.nlm.nih.gov/pubmed/20115951
51. Brierley, J., et al., TNM Classification of Malignant Tumours, 8th Edn. 2016.
https://www.uicc.org/8th-edition-uicc-tnm-classification-malignant-tumors-published
52. Tang, V., et al. Should centralized histopathological review in penile cancer be the global standard?
BJU Int, 2014. 114: 340.
https://www.ncbi.nlm.nih.gov/pubmed/24053106
53. Aumayr, K., et al. P16INK4A immunohistochemistry for detection of human papilloma virus-
associated penile squamous cell carcinoma is superior to in-situ hybridization. Int J Immunopathol
Pharmacol, 2013. 26: 611.
https://www.ncbi.nlm.nih.gov/pubmed/24067458
54. Bezerra, S.M., et al. Human papillomavirus infection and immunohistochemical p16(INK4a)
expression as predictors of outcome in penile squamous cell carcinomas. Hum Pathol, 2015. 46:
532.
https://www.ncbi.nlm.nih.gov/pubmed/25661481
55. Mannweiler, S., et al. Two major pathways of penile carcinogenesis: HPV-induced penile cancers
overexpress p16ink4a, HPV-negative cancers associated with dermatoses express p53, but lack
p16ink4a overexpression. J Am Acad Dermatol, 2013. 69: 73.
https://www.ncbi.nlm.nih.gov/pubmed/23474228
56. Corbishley C., et al. Carcinoma of the Penis and Distal Urethra Histopathology Reporting Guide 1st
edition. International Collaboration on Cancer Reporting. 2017. 2018.
http://www.iccr-cancer.org/datasets/published-datasets/urinary-male-genital/carcinoma-of-the-
penis-tnm8
57. Erbersdobler, A. Pathologic Evaluation and Reporting of Carcinoma of the Penis. Clin Genitourin
Cancer, 2017. 15: 192.
https://www.ncbi.nlm.nih.gov/pubmed/27594553
58. Winters, B.R., et al. Predictors of Nodal Upstaging in Clinical Node Negative Patients With Penile
Carcinoma: A National Cancer Database Analysis. Urology, 2016. 96: 29.
https://www.ncbi.nlm.nih.gov/pubmed/27450944
59. Feng, M.A., et al. Concordance of lymphovascular invasion diagnosed in penile carcinoma with and
without the immunohistochemical markers ERG and CD31. Histol Histopathol, 2016. 31: 293.
https://www.ncbi.nlm.nih.gov/pubmed/26452171
60. Cubilla, A.L. The role of pathologic prognostic factors in squamous cell carcinoma of the penis.
World J Urol, 2009. 27: 169.
https://www.ncbi.nlm.nih.gov/pubmed/8338190
61. Velazquez, E.F., et al. Epithelial abnormalities and precancerous lesions of anterior urethra in
patients with penile carcinoma: a report of 89 cases. Mod Pathol, 2005. 18: 917.
https://www.ncbi.nlm.nih.gov/pubmed/15920559
62. Rees, R.W., et al. pT2 penile squamous cell carcinomas: cavernosus vs. spongiosus invasion.
Eur Urol Suppl, 2008. 7: 111 (abstract #163).
https://www.eusupplements.europeanurology.com/article/S1569-9056(08)60162-1/fulltext
63. Leijte, J.A., et al. Evaluation of current TNM classification of penile carcinoma. J Urol, 2008. 180: 933.
https://www.ncbi.nlm.nih.gov/pubmed/18635216
64. Zhang, Z.L., et al. The importance of extranodal extension in penile cancer: a meta-analysis.
BMC Cancer, 2015. 15: 815.
https://www.ncbi.nlm.nih.gov/pubmed/26510975
65. Gunia, S., et al. Does the width of the surgical margin of safety or premalignant dermatoses at the
negative surgical margin affect outcome in surgically treated penile cancer? J Clin Pathol, 2014. 67:
268.
https://www.ncbi.nlm.nih.gov/pubmed/24100380

28 PENILE CANCER - MARCH 2018


66. Wang, J.Y., et al. Prognostic significance of the degree of extranodal extension in patients with
penile carcinoma. Asian J Androl, 2014. 16: 437.
https://www.ncbi.nlm.nih.gov/pubmed/24480925
67. Chaux, A., et al. The prognostic index: a useful pathologic guide for prediction of nodal metastases
and survival in penile squamous cell carcinoma. Am J Surg Pathol, 2009. 33: 1049.
https://www.ncbi.nlm.nih.gov/pubmed/19384188
68. Velazquez, E.F., et al. Positive resection margins in partial penectomies: sites of involvement and
proposal of local routes of spread of penile squamous cell carcinoma. Am J Surg Pathol, 2004. 28:
384.
https://www.ncbi.nlm.nih.gov/pubmed/15104302
69. Mahesan, T., et al. Advances in Penile-Preserving Surgical Approaches in the Management of Penile
Tumors. Urol Clin North Am, 2016. 43: 427.
https://www.ncbi.nlm.nih.gov/pubmed/27717429
70. Bertolotto, M., et al. Primary and secondary malignancies of the penis: ultrasound features.
Abdom Imaging, 2005. 30: 108.
https://www.ncbi.nlm.nih.gov/pubmed/15759326
71. Lont, A.P., et al. A comparison of physical examination and imaging in determining the extent of
primary penile carcinoma. BJU Int, 2003. 91: 493.
https://www.ncbi.nlm.nih.gov/pubmed/12656901
72. Kayes, O., et al. The role of magnetic resonance imaging in the local staging of penile cancer.
Eur Urol, 2007. 51: 1313.
https://www.ncbi.nlm.nih.gov/pubmed/17113213
73. Petralia, G., et al. Local staging of penile cancer using magnetic resonance imaging with
pharmacologically induced penile erection. Radiol Med, 2008. 113: 517.
https://www.ncbi.nlm.nih.gov/pubmed/18478188
74. Hanchanale, V., et al. The accuracy of magnetic resonance imaging (MRI) in predicting the invasion
of the tunica albuginea and the urethra during the primary staging of penile cancer. BJU Int, 2016.
117: 439.
https://www.ncbi.nlm.nih.gov/pubmed/25600638
75. Bozzini, G., et al. Role of Penile Doppler US in the Preoperative Assessment of Penile Squamous
Cell Carcinoma Patients: Results From a Large Prospective Multicenter European Study. Urology,
2016. 90: 131.
https://www.ncbi.nlm.nih.gov/pubmed/26776562
76. Krishna, R.P., et al. Sonography: an underutilized diagnostic tool in the assessment of metastatic
groin nodes. J Clin Ultrasound, 2008. 36: 212.
https://www.ncbi.nlm.nih.gov/pubmed/17960822
77. Mueller-Lisse, U.G., et al. Functional imaging in penile cancer: PET/computed tomography, MRI,
and sentinel lymph node biopsy. Curr Opin Urol, 2008. 18: 105.
https://www.ncbi.nlm.nih.gov/pubmed/18090498
78. Leijte, J.A., et al. Prospective evaluation of hybrid 18F-fluorodeoxyglucose positron emission
tomography/computed tomography in staging clinically node-negative patients with penile
carcinoma. BJU Int, 2009. 104: 640.
https://www.ncbi.nlm.nih.gov/pubmed/19281465
79. Schlenker, B., et al. Detection of inguinal lymph node involvement in penile squamous cell
carcinoma by 18F-fluorodeoxyglucose PET/CT: a prospective single-center study. Urol Oncol, 2012.
30: 55.
https://www.ncbi.nlm.nih.gov/pubmed/20022269
80. Alkatout, I., et al. Squamous cell carcinoma of the penis: predicting nodal metastases by histologic
grade, pattern of invasion and clinical examination. Urol Oncol, 2011. 29: 774.
https://www.ncbi.nlm.nih.gov/pubmed/20060332
81. Graafland, N.M., et al. Prognostic factors for occult inguinal lymph node involvement in penile
carcinoma and assessment of the high-risk EAU subgroup: a two-institution analysis of 342
clinically node-negative patients. Eur Urol, 2010. 58: 742.
https://www.ncbi.nlm.nih.gov/pubmed/20800339
82. Souillac, I., et al. Prospective evaluation of (18)F-fluorodeoxyglucose positron emission tomography-
computerized tomography to assess inguinal lymph node status in invasive squamous cell
carcinoma of the penis. J Urol, 2012. 187: 493.
https://www.ncbi.nlm.nih.gov/pubmed/22177157

PENILE CANCER - MARCH 2018 29


83. Horenblas, S., et al. Squamous cell carcinoma of the penis. III. Treatment of regional lymph nodes.
J Urol, 1993. 149: 492.
https://www.ncbi.nlm.nih.gov/pubmed/8437253
84. Ornellas, A.A., et al. Surgical treatment of invasive squamous cell carcinoma of the penis:
retrospective analysis of 350 cases. J Urol, 1994. 151: 1244.
https://www.ncbi.nlm.nih.gov/pubmed/7512656
85. Zhu, Y., et al. Predicting pelvic lymph node metastases in penile cancer patients: a comparison of
computed tomography, Cloquet’s node, and disease burden of inguinal lymph nodes. Onkologie,
2008. 31: 37.
https://www.ncbi.nlm.nih.gov/pubmed/18268397
86. Zhu, Y., et al. The value of squamous cell carcinoma antigen in the prognostic evaluation, treatment
monitoring and followup of patients with penile cancer. J Urol, 2008. 180: 2019.
https://www.ncbi.nlm.nih.gov/pubmed/18801542
87. Leijte, J.A., et al. Recurrence patterns of squamous cell carcinoma of the penis: recommendations
for follow-up based on a two-centre analysis of 700 patients. Eur Urol, 2008. 54: 161.
https://www.ncbi.nlm.nih.gov/pubmed/18440124
88. Shabbir, M., et al. Glans resurfacing for the treatment of carcinoma in situ of the penis: surgical
technique and outcomes. Eur Urol, 2011. 59: 142.
https://www.ncbi.nlm.nih.gov/pubmed/21050658
89. Manjunath, A., et al. Topical Therapy for non-invasive penile cancer (Tis)-updated results and
toxicity. Transl Androl Urol, 2017. 6: 803.
https://www.ncbi.nlm.nih.gov/pubmed/29184776
90. Alnajjar, H.M., et al. Treatment of carcinoma in situ of the glans penis with topical chemotherapy
agents. Eur Urol, 2012. 62: 923.
https://www.ncbi.nlm.nih.gov/pubmed/22421082
91. Bandieramonte, G., et al. Peniscopically controlled CO2 laser excision for conservative treatment of
in situ and T1 penile carcinoma: report on 224 patients. Eur Urol, 2008. 54: 875.
https://www.ncbi.nlm.nih.gov/pubmed/18243513
92. Colecchia, M., et al. pT1 penile squamous cell carcinoma: a clinicopathologic study of 56 cases
treated by CO2 laser therapy. Anal Quant Cytol Histol, 2009. 31: 153.
https://www.ncbi.nlm.nih.gov/pubmed/19639702
93. Piva, L., et al. [Therapeutic alternatives in the treatment of class T1N0 squamous cell carcinoma of
the penis: indications and limitations]. Arch Ital Urol Androl, 1996. 68: 157.
https://www.ncbi.nlm.nih.gov/pubmed/8767503
94. Frimberger, D., et al. Penile carcinoma. Is Nd:YAG laser therapy radical enough? J Urol, 2002. 168:
2418.
https://www.ncbi.nlm.nih.gov/pubmed/12441930
95. Meijer, R.P., et al. Long-term follow-up after laser therapy for penile carcinoma. Urology, 2007. 69: 759.
https://www.ncbi.nlm.nih.gov/pubmed/17445665
96. Rothenberger, K.H., et al. [Laser therapy of penile carcinoma]. Urologe A, 1994. 33: 291.
https://www.ncbi.nlm.nih.gov/pubmed/7941174
97. Paoli, J., et al. Penile intraepithelial neoplasia: results of photodynamic therapy. Acta Derm Venereol,
2006. 86: 418.
https://www.ncbi.nlm.nih.gov/pubmed/16955186
98. Djajadiningrat, R.S., et al. Penile sparing surgery for penile cancer-does it affect survival? J Urol,
2014. 192: 120.
https://www.ncbi.nlm.nih.gov/pubmed/24373799
99. Corbishley, C.M., et al. Glans resurfacing for precancerous and superficially invasive carcinomas of
the glans penis: Pathological specimen handling and reporting. Semin Diagn Pathol, 2015. 32: 232.
https://www.ncbi.nlm.nih.gov/pubmed/25662797
100. Philippou, P., et al. Conservative surgery for squamous cell carcinoma of the penis: resection
margins and long-term oncological control. J Urol, 2012. 188: 803.
https://www.ncbi.nlm.nih.gov/pubmed/22818137
101. Ornellas, A.A., et al. Surgical treatment of invasive squamous cell carcinoma of the penis: Brazilian
National Cancer Institute long-term experience. J Surg Oncol, 2008. 97: 487.
https://www.ncbi.nlm.nih.gov/pubmed/18425779
102. Schlenker, B., et al. Organ-preserving neodymium-yttrium-aluminium-garnet laser therapy for penile
carcinoma: a long-term follow-up. BJU Int, 2010. 106: 786.
https://www.ncbi.nlm.nih.gov/pubmed/20089106

30 PENILE CANCER - MARCH 2018


103. Schlenker, B., et al. Intermediate-differentiated invasive (pT1 G2) penile cancer--oncological
outcome and follow-up. Urol Oncol, 2011. 29: 782.
https://www.ncbi.nlm.nih.gov/pubmed/19945307
104. Skeppner, E., et al. Treatment-seeking, aspects of sexual activity and life satisfaction in men with
laser-treated penile carcinoma. Eur Urol, 2008. 54: 631.
https://www.ncbi.nlm.nih.gov/pubmed/18788122
105. Windahl, T., et al. Combined laser treatment for penile carcinoma: results after long-term followup.
J Urol, 2003. 169: 2118.
https://www.ncbi.nlm.nih.gov/pubmed/12771731
106. Tietjen, D.N., et al. Laser therapy of squamous cell dysplasia and carcinoma of the penis. Urology,
1998. 52: 559.
https://www.ncbi.nlm.nih.gov/pubmed/9763071
107. van Bezooijen, B.P., et al. Laser therapy for carcinoma in situ of the penis. J Urol, 2001. 166: 1670.
https://www.ncbi.nlm.nih.gov/pubmed/11586199
108. Mohs, F.E., et al. Mohs micrographic surgery for penile tumors. Urol Clin North Am, 1992. 19: 291.
https://www.ncbi.nlm.nih.gov/pubmed/1574820
109. Shindel, A.W., et al. Mohs micrographic surgery for penile cancer: management and long-term
followup. J Urol, 2007. 178: 1980.
https://www.ncbi.nlm.nih.gov/pubmed/17869306
110. Machan, M., et al. Penile Squamous Cell Carcinoma: Penis-Preserving Treatment With Mohs
Micrographic Surgery. Dermatol Surg, 2016. 42: 936.
https://www.ncbi.nlm.nih.gov/pubmed/27467227
111. Hadway, P., et al. Total glans resurfacing for premalignant lesions of the penis: initial outcome data.
BJU Int, 2006. 98: 532.
https://www.ncbi.nlm.nih.gov/pubmed/16925748
112. Ayres, B., et al., Glans resurfacing – a new penile preserving option for superficially invasive penile
cancer. Eur Urol Suppl, 2011. 10: 340.
http://www.eusupplements.europeanurology.com/article/S1569-9056(11)61084-1/abstract
113. Austoni E., et al. Reconstructive surgery for penile cancer with preservation of sexual function.
Eur Urol Suppl, 2008. 7: 116 (Abstract #183).
https://www.eusupplements.europeanurology.com/article/S1569-9056(08)60182-7/pdf
114. Li, J., et al. Organ-sparing surgery for penile cancer: complications and outcomes. Urology, 2011.
78: 1121.
https://www.ncbi.nlm.nih.gov/pubmed/22054385
115. Smith, Y., et al. Reconstructive surgery for invasive squamous carcinoma of the glans penis.
Eur Urol, 2007. 52: 1179.
https://www.ncbi.nlm.nih.gov/pubmed/17349734
116. Morelli, G., et al. Glansectomy with split-thickness skin graft for the treatment of penile carcinoma.
Int J Impot Res, 2009. 21: 311.
https://www.ncbi.nlm.nih.gov/pubmed/19458620
117. Modig, H., et al. Carcinoma of the penis. Treatment by surgery or combined bleomycin and radiation
therapy. Acta Oncol, 1993. 32: 653.
https://www.ncbi.nlm.nih.gov/pubmed/7505090
118. Persky, L., et al. Carcinoma of the penis. CA Cancer J Clin, 1986. 36: 258.
https://www.ncbi.nlm.nih.gov/pubmed/3093013
119. Lummen, G., et al. [Treatment and follow-up of patients with squamous epithelial carcinoma of the
penis]. Urologe A, 1997. 36: 157.
https://www.ncbi.nlm.nih.gov/pubmed/9199044
120. Khezri, A.A., et al. Carcinoma of the penis. Br J Urol, 1978. 50: 275.
https://www.ncbi.nlm.nih.gov/pubmed/753475
121. Veeratterapillay, R., et al. Oncologic Outcomes of Penile Cancer Treatment at a UK Supraregional
Center. Urology, 2015. 85: 1097.
https://www.ncbi.nlm.nih.gov/pubmed/25769781
122. Crook, J., et al. MP-21.03: Penile brachytherapy: results for 60 patients. Urology, 2007. 70: 161.
https://www.goldjournal.net/article/S0090-4295(07)00764-9/abstract
123. Crook, J., et al. Penile brachytherapy: technical aspects and postimplant issues. Brachytherapy,
2010. 9: 151.
https://www.ncbi.nlm.nih.gov/pubmed/19854685

PENILE CANCER - MARCH 2018 31


124. Crook, J., et al. Radiation therapy in the management of the primary penile tumor: an update.
World J Urol, 2009. 27: 189.
https://www.ncbi.nlm.nih.gov/pubmed/18636264
125. de Crevoisier, R., et al. Long-term results of brachytherapy for carcinoma of the penis confined to
the glans (N- or NX). Int J Radiat Oncol Biol Phys, 2009. 74: 1150.
https://www.ncbi.nlm.nih.gov/pubmed/19395183
126. Gotsadze, D., et al. Is conservative organ-sparing treatment of penile carcinoma justified? Eur Urol,
2000. 38: 306.
https://www.ncbi.nlm.nih.gov/pubmed/10940705
127. Ozsahin, M., et al. Treatment of penile carcinoma: to cut or not to cut? Int J Radiat Oncol Biol Phys,
2006. 66: 674.
https://www.ncbi.nlm.nih.gov/pubmed/16949770
128. Crook, J.M., et al. American Brachytherapy Society-Groupe Europeen de Curietherapie-European
Society of Therapeutic Radiation Oncology (ABS-GEC-ESTRO) consensus statement for penile
brachytherapy. Brachytherapy, 2013. 12: 191.
https://www.ncbi.nlm.nih.gov/pubmed/23453681
129. Delaunay, B., et al. Brachytherapy for penile cancer: efficacy and impact on sexual function.
Brachytherapy, 2014. 13: 380.
https://www.ncbi.nlm.nih.gov/pubmed/23896397
130. Kamsu-Kom, L., et al. Clinical Experience with Pulse Dose Rate Brachytherapy for Conservative
Treatment of Penile Carcinoma and Comparison with Historical Data of Low Dose Rate
Brachytherapy. Clin Oncol (R Coll Radiol), 2015. 27: 387.
https://www.ncbi.nlm.nih.gov/pubmed/26003455
131. Hasan, S., et al. The role of brachytherapy in organ preservation for penile cancer: A meta-analysis
and review of the literature. Brachytherapy, 2015. 14: 517.
https://www.ncbi.nlm.nih.gov/pubmed/25944394
132. Azrif, M., et al. External-beam radiotherapy in T1-2 N0 penile carcinoma. Clin Oncol (R Coll Radiol),
2006. 18: 320.
https://www.ncbi.nlm.nih.gov/pubmed/16703750
133. Zouhair, A., et al. Radiation therapy alone or combined surgery and radiation therapy in squamous-
cell carcinoma of the penis? Eur J Cancer, 2001. 37: 198.
https://www.ncbi.nlm.nih.gov/pubmed/11166146
134. Cordoba, A., et al. Low-dose brachytherapy for early stage penile cancer: a 20-year single-
institution study (73 patients). Radiat Oncol, 2016. 11: 96.
https://www.ncbi.nlm.nih.gov/pubmed/27464910
135. Lucky, M., et al. The treatment of penile carcinoma in situ (CIS) within a UK supra-regional network.
BJU Int, 2015. 115: 595.
https://www.ncbi.nlm.nih.gov/pubmed/25060513
136. Minhas, S., et al. What surgical resection margins are required to achieve oncological control in men
with primary penile cancer? BJU Int, 2005. 96: 1040.
https://www.ncbi.nlm.nih.gov/pubmed/16225525
137. Garaffa, G., et al. Total phallic reconstruction after penile amputation for carcinoma. BJU Int, 2009.
104: 852.
https://www.ncbi.nlm.nih.gov/pubmed/19239449
138. Salgado, C.J., et al. Glans penis coronaplasty with palmaris longus tendon following total penile
reconstruction. Ann Plast Surg, 2009. 62: 690.
https://www.ncbi.nlm.nih.gov/pubmed/19461287
139. Zou, Z.J., et al. Radiocolloid-based dynamic sentinel lymph node biopsy in penile cancer with
clinically negative inguinal lymph node: an updated systematic review and meta-analysis. Int Urol
Nephrol, 2016. 48: 2001.
https://www.ncbi.nlm.nih.gov/pubmed/27577753
140. Saisorn, I., et al. Fine-needle aspiration cytology predicts inguinal lymph node metastasis without
antibiotic pretreatment in penile carcinoma. BJU Int, 2006. 97: 1225.
https://www.ncbi.nlm.nih.gov/pubmed/16686716
141. Rosevear, H.M., et al. Utility of (1)(8)F-FDG PET/CT in identifying penile squamous cell carcinoma
metastatic lymph nodes. Urol Oncol, 2012. 30: 723.
https://www.ncbi.nlm.nih.gov/pubmed/21396850
142. Horenblas, S. Lymphadenectomy for squamous cell carcinoma of the penis. Part 1: diagnosis of
lymph node metastasis. BJU Int, 2001. 88: 467.
https://www.ncbi.nlm.nih.gov/pubmed/11589659

32 PENILE CANCER - MARCH 2018


143. Stuiver, M.M., et al. Early wound complications after inguinal lymphadenectomy in penile cancer: a
historical cohort study and risk-factor analysis. Eur Urol, 2013. 64: 486.
https://www.ncbi.nlm.nih.gov/pubmed/23490726
144. Koifman, L., et al. Radical open inguinal lymphadenectomy for penile carcinoma: surgical technique,
early complications and late outcomes. J Urol, 2013. 190: 2086.
https://www.ncbi.nlm.nih.gov/pubmed/23770135
145. Yao, K., et al. Modified technique of radical inguinal lymphadenectomy for penile carcinoma:
morbidity and outcome. J Urol, 2010. 184: 546.
https://www.ncbi.nlm.nih.gov/pubmed/20620415
146. Hegarty, P.K., et al. Controversies in ilioinguinal lymphadenectomy. Urol Clin North Am, 2010. 37:
421.
https://www.ncbi.nlm.nih.gov/pubmed/20674697
147. Protzel, C., et al. Lymphadenectomy in the surgical management of penile cancer. Eur Urol, 2009.
55: 1075.
https://www.ncbi.nlm.nih.gov/pubmed/19264390
148. Thuret, R., et al. A contemporary population-based assessment of the rate of lymph node dissection
for penile carcinoma. Ann Surg Oncol, 2011. 18: 439.
https://www.ncbi.nlm.nih.gov/pubmed/20839061
149. La-Touche, S., et al. Trial of ligation versus coagulation of lymphatics in dynamic inguinal sentinel
lymph node biopsy for staging of squamous cell carcinoma of the penis. Ann R Coll Surg Engl,
2012. 94: 344.
https://www.ncbi.nlm.nih.gov/pubmed/22943231
150. Weldrick, C., et al. A comparison of fibrin sealant versus standard closure in the reduction of
postoperative morbidity after groin dissection: A systematic review and meta-analysis. Eur J Surg
Oncol, 2014. 40: 1391.
https://www.ncbi.nlm.nih.gov/pubmed/25125341
151. Cui, Y., et al. Saphenous vein sparing during laparoscopic bilateral inguinal lymphadenectomy for
penile carcinoma patients. Int Urol Nephrol, 2016. 48: 363.
https://www.ncbi.nlm.nih.gov/pubmed/26660956
152. Kumar, V., et al. Prospective study comparing video-endoscopic radical inguinal lymph node
dissection (VEILND) with open radical ILND (OILND) for penile cancer over an 8-year period.
BJU Int, 2017. 119: 530.
https://www.ncbi.nlm.nih.gov/pubmed/27628265
153. Tauber, R., et al. Inguinal lymph node dissection: epidermal vacuum therapy for prevention of wound
complications. J Plast Reconstr Aesthet Surg, 2013. 66: 390.
https://www.ncbi.nlm.nih.gov/pubmed/23107617
154. Lughezzani, G., et al. The relationship between characteristics of inguinal lymph nodes and pelvic
lymph node involvement in penile squamous cell carcinoma: a single institution experience.
J Urol, 2014. 191: 977.
https://www.ncbi.nlm.nih.gov/pubmed/24262497
155. Tobias-Machado, M., et al. Video endoscopic inguinal lymphadenectomy: a new minimally invasive
procedure for radical management of inguinal nodes in patients with penile squamous cell
carcinoma. J Urol, 2007. 177: 953.
https://www.ncbi.nlm.nih.gov/pubmed/17296386
156. Graafland, N.M., et al. Prognostic significance of extranodal extension in patients with pathological
node positive penile carcinoma. J Urol, 2010. 184: 1347.
https://www.ncbi.nlm.nih.gov/pubmed/20723934
157. Lucky, M.A., et al. Referrals into a dedicated British penile cancer centre and sources of possible
delay. Sex Transm Infect, 2009. 85: 527.
https://www.ncbi.nlm.nih.gov/pubmed/19584061
158. Nicolai, N., et al. A Combination of Cisplatin and 5-Fluorouracil With a Taxane in Patients Who
Underwent Lymph Node Dissection for Nodal Metastases From Squamous Cell Carcinoma of
the Penis: Treatment Outcome and Survival Analyses in Neoadjuvant and Adjuvant Settings. Clin
Genitourin Cancer, 2016. 14: 323.
https://www.ncbi.nlm.nih.gov/pubmed/26341040
159. Necchi, A., et al. Prognostic Factors of Adjuvant Taxane, Cisplatin, and 5-Fluorouracil
Chemotherapy for Patients With Penile Squamous Cell Carcinoma After Regional
Lymphadenectomy. Clin Genitourin Cancer, 2016. 14: 518.
https://www.ncbi.nlm.nih.gov/pubmed/27050716

PENILE CANCER - MARCH 2018 33


160. Sharma, P., et al. Adjuvant chemotherapy is associated with improved overall survival in pelvic
node-positive penile cancer after lymph node dissection: a multi-institutional study. Urol Oncol,
2015. 33: 496 e17.
https://www.ncbi.nlm.nih.gov/pubmed/26072110
161. Pizzocaro, G., et al. Adjuvant and neoadjuvant vincristine, bleomycin, and methotrexate for inguinal
metastases from squamous cell carcinoma of the penis. Acta Oncol, 1988. 27: 823.
https://www.ncbi.nlm.nih.gov/pubmed/2466471
162. Leijte, J.A., et al. Neoadjuvant chemotherapy in advanced penile carcinoma. Eur Urol, 2007. 52: 488.
https://www.ncbi.nlm.nih.gov/pubmed/17316964
163. Bermejo, C., et al. Neoadjuvant chemotherapy followed by aggressive surgical consolidation for
metastatic penile squamous cell carcinoma. J Urol, 2007. 177: 1335.
https://www.ncbi.nlm.nih.gov/pubmed/17382727
164. Pagliaro, L.C., et al. Neoadjuvant paclitaxel, ifosfamide, and cisplatin chemotherapy for metastatic
penile cancer: a phase II study. J Clin Oncol, 2010. 28: 3851.
https://www.ncbi.nlm.nih.gov/pubmed/20625118
165. Dickstein, R.J., et al. Prognostic factors influencing survival from regionally advanced squamous cell
carcinoma of the penis after preoperative chemotherapy. BJU Int, 2016. 117: 118.
https://www.ncbi.nlm.nih.gov/pubmed/25294319
166. Pizzocaro, G., et al. Taxanes in combination with cisplatin and fluorouracil for advanced penile
cancer: preliminary results. Eur Urol, 2009. 55: 546.
https://www.ncbi.nlm.nih.gov/pubmed/18649992
167. Kulkarni, J.N., et al. Prophylactic bilateral groin node dissection versus prophylactic radiotherapy
and surveillance in patients with N0 and N1-2A carcinoma of the penis. Eur Urol, 1994. 26: 123.
https://www.ncbi.nlm.nih.gov/pubmed/7957466
168. Graafland, N.M., et al. Inguinal recurrence following therapeutic lymphadenectomy for node positive
penile carcinoma: outcome and implications for management. J Urol, 2011. 185: 888.
https://www.ncbi.nlm.nih.gov/pubmed/21239009
169. Franks, K.N., et al. Radiotherapy for node positive penile cancer: experience of the Leeds teaching
hospitals. J Urol, 2011. 186: 524.
https://www.ncbi.nlm.nih.gov/pubmed/21700296
170. Ravi, R., et al. Role of radiation therapy in the treatment of carcinoma of the penis. Br J Urol, 1994.
74: 646.
https://www.ncbi.nlm.nih.gov/pubmed/7530129
171. Demkow, T. The treatment of penile carcinoma: experience in 64 cases. Int Urol Nephrol, 1999. 31:
525.
https://www.ncbi.nlm.nih.gov/pubmed/10668948
172. Chen, M.F., et al. Contemporary management of penile cancer including surgery and adjuvant
radiotherapy: an experience in Taiwan. World J Urol, 2004. 22: 60.
https://www.ncbi.nlm.nih.gov/pubmed/14657999
173. Djajadiningrat, R.S., et al. Contemporary management of regional nodes in penile cancer-
improvement of survival? J Urol, 2014. 191: 68.
https://www.ncbi.nlm.nih.gov/pubmed/23917166
174. Tang, D.H., et al. Adjuvant pelvic radiation is associated with improved survival and decreased
disease recurrence in pelvic node-positive penile cancer after lymph node dissection: A multi-
institutional study. Urol Oncol, 2017. 35: 605 e17.
https://www.ncbi.nlm.nih.gov/pubmed/28666722
175. Burt, L.M., et al. Stage presentation, care patterns, and treatment outcomes for squamous cell
carcinoma of the penis. Int J Radiat Oncol Biol Phys, 2014. 88: 94.
https://www.ncbi.nlm.nih.gov/pubmed/24119832
176. Pizzocaro, G., et al. Up-to-date management of carcinoma of the penis. Eur Urol, 1997. 32: 5.
https://www.ncbi.nlm.nih.gov/pubmed/9266225
177. Giannatempo P., et al. Survival analyses of adjuvant or neoadjuvant combination of a taxane plus
cisplatin and 5-fluorouracil (T-PF) in patients with bulky nodal metastases from squamous cell
carcinoma of the penis (PSCC): Results of a single high-volume center. J Clin Oncol, 2014. 32: 5.
https://meetinglibrary.asco.org/record/90280/abstract
178. Noronha, V., et al. Role of paclitaxel and platinum-based adjuvant chemotherapy in high-risk penile
cancer. Urol Ann, 2012. 4: 150.
https://www.ncbi.nlm.nih.gov/pubmed/23248520

34 PENILE CANCER - MARCH 2018


179. Hakenberg, O.W., et al. Cisplatin, methotrexate and bleomycin for treating advanced penile
carcinoma. BJU Int, 2006. 98: 1225.
https://www.ncbi.nlm.nih.gov/pubmed/17125480
180. Haas, G.P., et al. Cisplatin, methotrexate and bleomycin for the treatment of carcinoma of the penis:
a Southwest Oncology Group study. J Urol, 1999. 161: 1823.
https://www.ncbi.nlm.nih.gov/pubmed/10332445
181. Hussein, A.M., et al. Chemotherapy with cisplatin and 5-fluorouracil for penile and urethral
squamous cell carcinomas. Cancer, 1990. 65: 433.
https://www.ncbi.nlm.nih.gov/pubmed/2297633
182. Shammas, F.V., et al. Cisplatin and 5-fluorouracil in advanced cancer of the penis. J Urol, 1992. 147:
630.
https://www.ncbi.nlm.nih.gov/pubmed/1538445
183. Theodore, C., et al. A phase II multicentre study of irinotecan (CPT 11) in combination with cisplatin
(CDDP) in metastatic or locally advanced penile carcinoma (EORTC PROTOCOL 30992). Ann Oncol,
2008. 19: 1304.
https://www.ncbi.nlm.nih.gov/pubmed/18417462
184. Nicholson, S., et al. Phase II trial of docetaxel, cisplatin and 5FU chemotherapy in locally advanced
and metastatic penis cancer (CRUK/09/001). Br J Cancer, 2013. 109: 2554.
https://www.ncbi.nlm.nih.gov/pubmed/24169355
185. Eliason, M., et al. Primary treatment of verrucous carcinoma of the penis with fluorouracil, cis-
diamino-dichloro-platinum, and radiation therapy. Arch Dermatol, 2009. 145: 950.
https://www.ncbi.nlm.nih.gov/pubmed/19687438
186. Pond, G.R., et al. Prognostic risk stratification derived from individual patient level data for men with
advanced penile squamous cell carcinoma receiving first-line systemic therapy. Urol Oncol, 2014.
32: 501.
https://www.ncbi.nlm.nih.gov/pubmed/24332646
187. Di Lorenzo, G., et al. Cisplatin and 5-fluorouracil in inoperable, stage IV squamous cell carcinoma of
the penis. BJU Int, 2012. 110: E661.
https://www.ncbi.nlm.nih.gov/pubmed/22958571
188. Di Lorenzo, G., et al. Paclitaxel in pretreated metastatic penile cancer: final results of a phase 2
study. Eur Urol, 2011. 60: 1280.
https://www.ncbi.nlm.nih.gov/pubmed/21871710
189. Power, D.G., et al. Cisplatin and gemcitabine in the management of metastatic penile cancer.
Urol Oncol, 2009. 27: 187.
https://www.ncbi.nlm.nih.gov/pubmed/18367122
190. Gou, H.F., et al. Epidermal growth factor receptor (EGFR)-RAS signaling pathway in penile
squamous cell carcinoma. PLoS One, 2013. 8: e62175.
https://www.ncbi.nlm.nih.gov/pubmed/23637996
191. Necchi, A., et al. Proof of activity of anti-epidermal growth factor receptor-targeted therapy for
relapsed squamous cell carcinoma of the penis. J Clin Oncol, 2011. 29: e650.
https://www.ncbi.nlm.nih.gov/pubmed/21632506
192. Carthon, B.C., et al. Epidermal growth factor receptor-targeted therapy in locally advanced or
metastatic squamous cell carcinoma of the penis. BJU Int, 2014. 113: 871.
https://www.ncbi.nlm.nih.gov/pubmed/24053151
193. Zhu, Y., et al. Feasibility and activity of sorafenib and sunitinib in advanced penile cancer: a
preliminary report. Urol Int, 2010. 85: 334.
https://www.ncbi.nlm.nih.gov/pubmed/20980789
194. Di Lorenzo, G., et al. Cytosolic phosphorylated EGFR is predictive of recurrence in early stage penile
cancer patients: a retropective study. J Transl Med, 2013. 11: 161.
https://www.ncbi.nlm.nih.gov/pubmed/23819610
195. Necchi, A., et al. Panitumumab Treatment for Advanced Penile Squamous Cell Carcinoma When
Surgery and Chemotherapy Have Failed. Clin Genitourin Cancer, 2016. 14: 231.
https://www.ncbi.nlm.nih.gov/pubmed/26362073
196. Horenblas, S., et al. Local recurrent tumour after penis-conserving therapy. A plea for long-term
follow-up. Br J Urol, 1993. 72: 976.
https://www.ncbi.nlm.nih.gov/pubmed/8306171
197. Kroon, B.K., et al. Patients with penile carcinoma benefit from immediate resection of clinically
occult lymph node metastases. J Urol, 2005. 173: 816.
https://www.ncbi.nlm.nih.gov/pubmed/15711276

PENILE CANCER - MARCH 2018 35


198. Kroon, B.K., et al. Ultrasonography-guided fine-needle aspiration cytology before sentinel node
biopsy in patients with penile carcinoma. BJU Int, 2005. 95: 517.
https://www.ncbi.nlm.nih.gov/pubmed/15705071
199. Djajadiningrat, R.S., et al. Ultrasound examination and fine needle aspiration cytology-useful for
followup of the regional nodes in penile cancer? J Urol, 2014. 191: 652.
https://www.ncbi.nlm.nih.gov/pubmed/23994372
200. Schover, L.R. Sexuality and fertility after cancer. Hematology Am Soc Hematol Educ Program, 2005:
523.
https://www.ncbi.nlm.nih.gov/pubmed/16304430
201. Sedigh, O., et al. Sexual function after surgical treatment for penile cancer: Which organ-sparing
approach gives the best results? Can Urol Assoc J, 2015. 9: E423.
https://www.ncbi.nlm.nih.gov/pubmed/26279710
202. Kieffer, J.M., et al. Quality of life for patients treated for penile cancer. J Urol, 2014. 192: 1105.
https://www.ncbi.nlm.nih.gov/pubmed/24747092
203. Romero, F.R., et al. Sexual function after partial penectomy for penile cancer. Urology, 2005. 66:
1292.
https://www.ncbi.nlm.nih.gov/pubmed/16360459
204. D’Ancona, C.A., et al. Quality of life after partial penectomy for penile carcinoma. Urology, 1997. 50:
593.
https://www.ncbi.nlm.nih.gov/pubmed/9338738
205. Alei, G., et al. Lichen sclerosus in patients with squamous cell carcinoma. Our experience with
partial penectomy and reconstruction with ventral fenestrated flap. Ann Ital Chir, 2012. 83: 363.
https://www.ncbi.nlm.nih.gov/pubmed/22759475
206. Sansalone, S., et al. Sexual outcomes after partial penectomy for penile cancer: results from a multi-
institutional study. Asian J Androl, 2017. 19: 57.
https://www.ncbi.nlm.nih.gov/pubmed/26643562
207. Sosnowski, R., et al. Quality of life in penile carcinoma patients – post-total penectomy. Centr Eur
J Urol, 2016. 69: 204.
https://www.ncbi.nlm.nih.gov/pubmed/27551559
208. Yu, C., et al. Sexual Function after Partial Penectomy: A Prospectively Study From China. Sci Rep,
2016. 6: 21862.
https://www.ncbi.nlm.nih.gov/pubmed/26902397
209. Gerullis, H., et al. Construction of a penoid after penectomy using a transpositioned testicle.
Urol Int, 2013. 90: 240.
https://www.ncbi.nlm.nih.gov/pubmed/22922734
210. Hage, J.J. Simple, safe, and satisfactory secondary penile enhancement after near-total oncologic
amputation. Ann Plast Surg, 2009. 62: 685.
https://www.ncbi.nlm.nih.gov/pubmed/19461286

9. CONFLICT OF INTEREST
All members of the Penile Cancer Guidelines working group have provided disclosure statements of all
relationships that they have that might be perceived as a potential source of a conflict of interest. This
information is publicly accessible through the European Association of Urology website: http://uroweb.org/
guideline/penile-cancer/.
This guidelines document was developed with the financial support of the European Association of
Urology. No external sources of funding and support have been involved. The EAU is a non-profit organisation
and funding is limited to administrative assistance and travel and meeting expenses. No honoraria or other
reimbursements have been provided.

36 PENILE CANCER - MARCH 2018


10. CITATION INFORMATION
The format in which to cite the EAU Guidelines will vary depending on the style guide of the journal in which the
citation appears. Accordingly, the number of authors or whether, for instance, to include the publisher, location,
or an ISBN number may vary.

The compilation of the complete Guidelines should be referenced as:


EAU Guidelines. Edn. presented at the EAU Annual Congress Amsterdam, 2020. ISBN 978-94-92671-07-3.

If a publisher and/or location is required, include:


EAU Guidelines Office, Arnhem, The Netherlands. http://uroweb.org/guidelines/compilations-of-all-guidelines/

References to individual guidelines should be structured in the following way:


Contributors’ names. Title of resource. Publication type. ISBN. Publisher and publisher location, year.

PENILE CANCER - MARCH 2018 37


38 PENILE CANCER - MARCH 2018
EAU Guidelines on
Primary Urethral
Carcinoma
G. Gakis, J.A. Witjes (Chair), H.M. Bruins, R. Cathomas,
E. Compérat, N.C. Cowan, J.A. Efstathiou, A.G. van der Heijden,
V. Hernàndez, A. Lorch, M.I. Milowsky, M.J. Ribal (Vice-chair),
G.N. Thalmann, E. Veskimäe
Guidelines Associates: E.E. Linares Espinós,
Y. Neuzillet, M. Rouanne

© European Association of Urology 2021


TABLE OF CONTENTS PAGE
1. INTRODUCTION 4
1.1 Aims and scope 4
1.2 Panel composition 4
1.3 Available publications 4
1.4 Publication history & summary of changes 4
1.4.1 Summary of changes 4

2. METHODS 4
2.1 Data identification 4
2.2 Review 5
2.3 Future goals 5

3. EPIDEMIOLOGY, AETIOLOGY AND PATHOLOGY 5


3.1 Epidemiology 5
3.2 Aetiology 6
3.3 Histopathology 6

4. STAGING AND CLASSIFICATION SYSTEMS 6


4.1 Tumour, Node, Metastasis (UICC/TNM) staging system 6
4.2 Tumour grade 7
4.3 Handling of tumour specimens 7
4.4 Guideline for staging and classification systems 8

5. DIAGNOSTIC EVALUATION AND STAGING 8


5.1 History 8
5.2 Clinical examination 8
5.3 Urinary cytology 8
5.4 Diagnostic urethrocystoscopy and biopsy 8
5.5 Imaging for diagnosis and stagingg 8
5.6 Regional lymph nodes 9
5.7 Summary of evidence and guidelines for diagnostic evaluation and staging 9

6. PROGNOSIS 9
6.1 Long-term survival after primary urethral carcinoma 9
6.2 Predictors of survival in primary urethral carcinoma 9
6.3 Summary of evidence for prognosis 9

7. DISEASE MANAGEMENT 10
7.1 Treatment of localised primary urethral carcinoma in males 10
7.1.1 Summary of evidence and guidelines for the treatment of localised primary
urethral carcinoma in males 10
7.2 Treatment of localised urethral carcinoma in females 10
7.2.1 Urethrectomy and urethra-sparing surgery 10
7.2.2 Radiotherapy 10
7.2.3 Summary of evidence and guidelines for the treatment of localised urethral
carcinoma in females 11
7.3 Multimodal treatment in locally advanced urethral carcinoma in both genders 11
7.3.1 Introduction 11
7.3.2 Preoperative cisplatin-based chemotherapy 11
7.3.3 Chemoradiotherapy in locally advanced squamous cell carcinoma of the
urethra 11
7.3.4 Salvage treatment in recurrent primary urethral carcinoma after surgery for
primary treatment 11
7.3.5 Treatment of regional lymph nodes 11
7.3.6 Summary of evidence and guidelines for multimodal treatment in advanced
urethral carcinoma in both genders 12

2 PRIMARY URETHRAL CARCINOMA - LIMITED UPDATE MARCH 2021


7.4 Treatment of urothelial carcinoma of the prostate 12
7.4.1 Summary of evidence and guidelines for the treatment of urothelial carcinoma
of the prostate 12
7.5 Metastatic disease 12

8. FOLLOW-UP 14

9. REFERENCES 14

10. CONFLICT OF INTEREST 18

11. CITATION INFORMATION 19

PRIMARY URETHRAL CARCINOMA - LIMITED UPDATE MARCH 2021 3


1. INTRODUCTION
1.1 Aims and scope
The aim of these guidelines is to deliver current evidence-based information on the diagnosis and treatment of
patients with primary urethral carcinoma. When the first carcinoma in the urinary tract is detected in the urethra,
this is defined as primary urethral carcinoma, in contrast to secondary urethral carcinoma, which presents as
recurrent carcinoma in the urethra after prior diagnosis and treatment of carcinoma elsewhere in the urinary
tract. Most often, secondary urethral carcinoma is reported after radical cystectomy for bladder cancer [1, 2]
(see Chapter 7.4 of the European Association of Urology [EAU] Guidelines on Muscle-invasive and Metastatic
Bladder Cancer [MIBC]) [2].
It must be emphasised that clinical guidelines present the best evidence available to the experts
but following guideline recommendations will not necessarily result in the best outcome. Guidelines can never
replace clinical expertise when making treatment decisions for individual patients, but rather help to focus
decisions - also taking personal values and preferences/individual circumstances of patients into account.
Guidelines are not mandates and do not purport to be a legal standard of care.

1.2 Panel composition


The EAU Guidelines Panel on MIBC is responsible for this publication. This is an international multidisciplinary
group of clinicians, including urologists, oncologists, a pathologist, a radiotherapist and a radiologist. Members
of this panel have been selected based on their expertise to represent the professionals treating patients
suspected of suffering from urethral carcinoma. All experts involved in the production of this document
have submitted potential conflict of interest statements, which can be viewed on the EAU Website Uroweb:
https://uroweb.org/guideline/primary-urethral-carcinoma/.

1.3 Available publications


A quick reference document (Pocket guidelines) is available in print and as an app for iOS and Android devices,
presenting the main findings of the Primary Urethral Carcinoma Guidelines. These are abridged versions which
may require consultation together with the full text version. The most recent scientific summary was published
in 2020 [3].

1.4 Publication history & summary of changes


The Primary Urethral Carcinoma Guidelines were first published in 2013. This is the eight update of this
document.

1.4.1 Summary of changes


The literature for the complete document has been assessed and updated, where relevant. In particular for:
• Section 5.3 – Urinary cytology;
• Section 5.5 – Imaging for diagnosis and staging;
• Chapter 6 – Prognosis;
• Section 7.3 – Multimodal treatment in locally advanced urethral carcinoma in both genders. A new
recommendation was provided:

7.3.6 S
 ummary of evidence and guidelines for multimodal treatment in advanced urethral carcinoma in
both genders

Recommendation Strength rating


Offer inguinal lymph node (LN) dissection to patients with limited LN-positive urethral Strong
squamous cell carcinoma.

2. METHODS
2.1 Data identification
For the 2021 Primary Urethral Carcinoma Guidelines, new and relevant evidence has been identified, collated,
and appraised through a structured assessment of the literature. An updated systematic literature search was
performed to identify studies reporting data on urethral malignancies since the prior search, covering a time
frame between July 3rd, 2019 and October 20th, 2020. Databases searched included Ovid (Medline), EMBASE

4 PRIMARY URETHRAL CARCINOMA - LIMITED UPDATE MARCH 2021


and the Cochrane Central Register of Controlled Trials and Cochrane Database of Systematic Reviews. A total
of 117 unique records were identified, retrieved, and screened for relevance. Five new references were included
in this 2021 publication. A detailed search strategy is available online: https://uroweb.org/guideline/primary-
urethral-carcinoma/?type=appendices-publications.

For each recommendation within the guidelines there is an accompanying online strength rating form, based on
a modified GRADE methodology [4, 5]. These forms address a number of key elements namely:

1. the overall quality of the evidence which exists for the recommendation, references used in
this text are graded according to a classification system modified from the Oxford Centre for
Evidence-Based Medicine Levels of Evidence [6];
2. the magnitude of the effect (individual or combined effects);
3. the certainty of the results (precision, consistency, heterogeneity and other statistical or study
related factors);
4. the balance between desirable and undesirable outcomes;
5. the impact of patient values and preferences on the intervention;
6. the certainty of those patient values and preferences.

These key elements are the basis which panels use to define the strength rating of each recommendation. The
strength of each recommendation is represented by the words ‘strong’ or ‘weak’ [7]. The strength of each
recommendation is determined by the balance between desirable and undesirable consequences of alternative
management strategies, the quality of the evidence (including certainty of estimates), and nature and variability
of patient values and preferences. The strength rating forms will be available online.
Additional information can be found in the general Methodology section of this print, and online at
the EAU website; http://www.uroweb.org/guideline/.
A list of Associations endorsing the EAU Guidelines can also be viewed online at the above address.

2.2 Review
This document was peer-reviewed prior to publication in 2020.

2.3 Future goals


The MIBC Guidelines Panel aims to systematically address the following key clinical topics in future updates of
the Primary Urethral Carcinoma Guidelines:
• assessment of the accuracy of computed tomography [CT] and magnetic resonance imaging [MRI] for
local staging of primary urethral carcinoma and their predictive value on clinical decision-making;
• the (long-term) efficacy of urethral-sparing surgery and chemoradiotherapy for genital preservation in
localised and locally advanced tumours;
• the prognostic impact of neoadjuvant and adjuvant treatment modalities in locally advanced disease;
• the prognostic impact of the extent of transurethral resection of the prostate prior to bacillus Calmette-
Guérin (BCG) treatment in urothelial malignancies of the prostatic urethra and ducts;
• the therapeutic benefit and clinical safety of programmed cell death (ligand)-1 inhibitors for the treatment
of advanced primary urethral carcinoma;
• the extent and prognostic benefit of regional Lymph node (LN) dissection at primary treatment.

3. EPIDEMIOLOGY, AETIOLOGY AND


PATHOLOGY
3.1 Epidemiology
Primary urethral carcinoma is considered a rare cancer, accounting for < 1% of all genitourinary malignancies
[8] (ICD-O3 topography code: C68.0) [9]. In 2013, the prevalence of urethral carcinoma in the 28 European
Union countries was 3,986 cases with an estimated annual incidence of 1,504 new cases, with a male/female
prevalence of 2.9: 1 [10]. Likewise, in an analysis of the Surveillance, Epidemiology and End Results (SEER)
database, the incidence of primary urethral carcinoma peaked in the > 75 years age group (7.6/million). The
age-standardised rate was 4.3/million in men and 1.5/million in women and was almost negligible in those aged
< 55 years (0.2/million) [11].

PRIMARY URETHRAL CARCINOMA - LIMITED UPDATE MARCH 2021 5


3.2 Aetiology
For male primary urethral carcinoma, various predisposing factors have been reported, including urethral
strictures [12, 13], chronic irritation after intermittent catheterisation/urethroplasty [14-16], external beam
irradiation therapy (EBRT) [17], radioactive seed implantation [18], chronic urethral inflammation/urethritis
following sexually transmitted diseases (i.e., condylomata associated with human papilloma virus 16) [19, 20]
and lichen sclerosis [13]. In female urethral carcinoma, urethral diverticula [21-23] and recurrent urinary tract
infections [24] have been associated with primary urethral carcinoma. Mid-urethral sling meshes have not been
associated with an increased risk of primary urethral carcinoma [25]. Clear-cell adenocarcinoma (AC) may also
have a congenital origin [26, 27].

3.3 Histopathology
Both the Surveillance of Rare Cancers in Europe (RARECARE) project and SEER database have reported
that urothelial carcinoma (UC) of the urethra is the predominant histological type of primary urethral cancer
(54–65%), followed by squamous cell carcinoma (SCC) (16–22%) and AC (10–16%) [10, 28]. A SEER analysis
of 2,065 men with primary urethral carcinoma (mean age 73 years) found that UC was most common (78%),
and SCC (12%) and AC (5%) were significantly less frequent [29]. In women, AC is the more frequent histology
(38–46.7%) followed by SCC (25.4–28%), UC (24.9–28%) and other histological entities (6%) [30, 31].

4. STAGING AND CLASSIFICATION SYSTEMS


4.1 Tumour, Node, Metastasis (UICC/TNM) staging system
In men and women, urethral carcinoma is classified according to the 8th edition of the TNM classification [9]
(Table 4.1). It should be noted that there is a separate TNM staging system for prostatic UC [9]. Of note, for
cancers occurring in the urethral diverticulum, stage T2 is not applicable as urethral diverticula are lacking
periurethral muscle [32].

Table 4.1: TNM classification (8th edition) for urethral carcinoma [9]

T - Primary Tumour
TX Primary tumour cannot be assessed
T0 No evidence of primary tumour
Urethra (male and female)
Ta Non-invasive papillary, polypoid, or verrucous carcinoma
Tis Carcinoma in situ
T1 Tumour invades subepithelial connective tissue
T2 Tumour invades any of the following: corpus spongiosum, prostate, periurethral muscle
T3 Tumour invades any of the following: corpus cavernosum, beyond prostatic capsule, anterior vagina,
bladder neck (extraprostatic extension)
T4 Tumour invades other adjacent organs (invasion of the bladder)
Urothelial (transitional cell) carcinoma of the prostate
Tis pu Carcinoma in situ, involvement of prostatic urethra
Tis pd Carcinoma in situ, involvement of prostatic ducts
T1 Tumour invades subepithelial connective tissue (for tumours involving prostatic urethra only)
T2 Tumour invades any of the following: prostatic stroma, corpus sponsiosum, periurethral muscle
T3 Tumour invades any of the following: corpus cavernosum, beyond prostatic capsule, bladder neck
(extraprostatic extension)
T4 Tumour invades other adjacent organs (invasion of the bladder or rectum)
N - Regional Lymph Nodes
NX Regional lymph nodes cannot be assessed
N0 No regional lymph node metastasis
N1 Metastasis in a single lymph node
N2 Metastasis in multiple lymph nodes
M - Distant Metastasis
M0 No distant metastasis
M1 Distant metastasis

6 PRIMARY URETHRAL CARCINOMA - LIMITED UPDATE MARCH 2021


4.2 Tumour grade
The former World Health Organization (WHO) grading system of 1973, which differentiated urothelial
carcinomas into three different grades (G1-G3), has been replaced by the grading system in 2004 [33]. Non-
urothelial urethral carcinoma is graded by a trinomial system that differentiates between well-differentiated
(G1), moderately-differentiated (G2), and poorly-differentiated tumours (G3). Table 4.2 lists the different grading
systems according to the WHO 1973 and 2004 systems [33]. The 2004 classification corresponds to the new
2016 WHO classification [34].

Table 4.2: Histopathological grading of urothelial and non-urothelial primary urethral carcinoma [33]

Urothelial urethral carcinoma


PUNLMP Papillary urothelial neoplasm of low malignant potential
Low grade Well differentiated
High grade Poorly differentiated

Non-urothelial urethral carcinoma


Gx Tumour grade not assessable
G1 Well differentiated
G2 Moderately differentiated
G3 Poorly differentiated

4.3 Handling of tumour specimens


Specimen handling should follow the general rules as published by the International Collaboration on Cancer
Reporting [35].

Table 4.3: Required and recommended elements for pathology reporting of carcinoma of the urethra in
urethrectomy specimens [9, 35]

Required Recommended
Operative procedure Clinical information Previous history of urinary
tract disease or distant
metastasis
Additional specimens Previous therapy
submitted
Maximum tumour Cannot be assessed Other clinical information
dimension No macroscopically Tumour focality
visible tumour
Maximum tumour Other tumour dimensions
dimension (largest (than maximum
tumour) dimension) of the largest
tumour
Macroscopic tumour site Block identification key
Macroscopic extent of Associated epithelial
invasion lesions
Histological tumour type Histological subtype/ Extranodal spread for
variant (urothelial involved regional lymph
carcinoma) node(s)
Non-invasive carcinoma Coexistent pathology
Histological tumour grade Ancillary studies
Microscopic extent of
invasion
Lymphovascular invasion
Margin status
Regional lymph node No regional lymph nodes
status submitted

PRIMARY URETHRAL CARCINOMA - LIMITED UPDATE MARCH 2021 7


4.4 Guideline for staging and classification systems

Recommendation Strength rating


Use the 2017 TNM classification and 2004/2016 WHO grading systems for pathological Strong
staging and grading of primary urethral carcinoma.

5. DIAGNOSTIC EVALUATION AND STAGING


5.1 History
When becoming clinically apparent, most patients (45–57%) with primary urethral carcinoma present with
symptoms associated with locally advanced disease (T3/T4) [36]. At initial presentation visible haematuria or
bloody urethral discharge is reported in up to 62% of the cases. Further symptoms of locally advanced disease
include; an extra-urethral mass (52%), bladder outlet obstruction (48%), pelvic pain (33%), urethrocutaneous
fistula (10%), abscess formation (5%) or dyspareunia [36].

5.2 Clinical examination


In men, physical examination should comprise palpation of the external genitalia for suspicious indurations or
masses and digital rectal examination [37]. In women, further pelvic examination with careful inspection and
palpation of the urethra should be performed, especially in those with primary onset of irritative or obstructive
voiding. In addition, bimanual examination, when necessary under general anaesthesia, should be performed for
local clinical staging and to exclude the presence of colorectal or gynaecological malignancies. Bilateral inguinal
palpation should be conducted to assess the presence of enlarged LNs, describing location, size, and mobility [38].

5.3 Urinary cytology


Cytological assessment of urine specimens in suspect cases of primary urethral carcinoma should be
conducted according to the Paris system [39]. The role of urinary cytology in primary urethral carcinoma is
limited since its sensitivity ranges between 55% and 59% [40]. Detection rates depend on the underlying
histological entity. In male patients, the sensitivity for UC and SCC was reported to be 80% and 50%,
respectively, whereas in female patients, sensitivity was found to be 77% for SCC and 50% for UC [40].

5.4 Diagnostic urethrocystoscopy and biopsy


Diagnostic urethrocystoscopy and biopsy enables primary assessment of a urethral tumour in terms of
tumour extent, location, and underlying histology [37]. To enable accurate pathological assessment of surgical
margins, biopsy sites (proximal/distal end) should be marked and sent together with clinical information to
the pathologist. To obtain all relevant information, the collection, handling, and evaluation of biopsy specimen
should follow the recommendations provided by the International Collaboration on Cancer Reporting (see
Table 4.3) [35].
Careful cystoscopic examination is necessary to exclude the presence of concomitant bladder
tumours [41]. A cold-cup biopsy enables accurate tissue retrieval for histological analysis and avoids artificial
tissue damage. In patients with larger lesions, transurethral resection (optionally in men under penile blood
arrest using a tourniquet) can be performed for histological diagnosis [42]. In patients with suspected UC of
the prostatic urethra or ducts, resectoscope loop biopsy of the prostatic urethra (between the five and seven
o’clock position from the bladder neck and distally around the area of the verumontanum) can contribute to an
improved detection rate [43].

5.5 Imaging for diagnosis and stagingg


Radiological imaging of urethral carcinoma aims to assess local tumour extent and to detect lymphatic and
distant metastatic spread. In a recent multicentre study, the accuracy of CT for clinical tumour and nodal
staging predicting final pathological staging was found to be 72.9% and 70.6%, respectively [44]. Magnetic
resonance imaging can be used to evaluate local tumour extent and presence of regional LN metastases,
focusing in particular on inguinal and pelvic LNs [45-47]. Computed tomography can be used for distant
staging and concentrate on chest, abdomen, and pelvis, with CT of the thorax and abdomen in all patients
with invasive disease (> cT1N0M0) [48]. If imaging of the remainder of the urothelium is required, CT urography
should be performed [49]. For local staging, there is evidence that MRI is an accurate tool for monitoring
tumour response to neoadjuvant chemoradiotherapy and evaluating the extent of local disease prior to
exenterative surgery [50].

8 PRIMARY URETHRAL CARCINOMA - LIMITED UPDATE MARCH 2021


5.6 Regional lymph nodes
In urethral carcinoma, in contrast to penile cancer (41%) [51], enlarged LNs often represent metastatic disease
(84%) [52-54]. In men, lymphatics from the anterior urethra drain into the superficial- and deep inguinal LNs
and, subsequently, to the pelvic (external, obturator and internal iliac) LNs. Conversely, lymphatic vessels of the
posterior urethra drain into the pelvic LNs. In women, the lymph of the proximal third drains into the pelvic LN
chains, whereas the distal two-thirds initially drain into the superficial- and deep inguinal nodes [55, 56].

5.7 Summary of evidence and guidelines for diagnostic evaluation and staging

Summary of evidence LE
Patients with clinically enlarged inguinal or pelvic LNs often exhibit pathological LN metastasis. 3

Recommendations Strength rating


Use urethrocystoscopy with biopsy and urinary cytology to diagnose urethral carcinoma. Strong
Assess the presence of distant metastases by computed tomography of the thorax and Strong
abdomen/pelvis.
Use pelvic magnetic resonance imaging to assess the local extent of urethral tumour and Strong
regional lymph node enlargement.

6. PROGNOSIS
6.1 Long-term survival after primary urethral carcinoma
According to the RARECARE project, the one- and 5-year relative overall survival (OS) rates in patients with
urethral carcinoma in Europe are 71% and 54%, respectively [10]. Based on longer follow-up, an analysis of
the SEER database, comparing prognostic factors in rare pathological types of primary urethral carcinoma
(n = 257) and common pathological groups (n = 2,651), reported 10-year OS rates of 42.4% and 31.9%,
respectively [57]. Cancer-specific survival (CSS) rates at five and ten years were 68% and 60%, respectively
[11]. Age (> 60 years), race (others vs. whites), T-stage (T3/T4 vs. Ta-T2) and M-stage (M1 vs. M0) were
independent prognostic risk factors for OS and CSS in rare pathological variants [57].

6.2 Predictors of survival in primary urethral carcinoma


In Europe, 5-year OS rate does not substantially differ between the sexes [10, 31]. Prognostic factors of
decreased survival in patients with primary urethral carcinoma are:
• advanced age (> 65 years) and black race [10, 31, 58, 59];
• stage, grade, nodal involvement [53, 59] and metastasis [29];
• tumour size and proximal tumour location [29];
• extent of surgical treatment and treatment modality [29, 58, 59];
• underlying histology [10, 29, 58-60];
• presence of concomitant bladder cancer [41];
• location of recurrence (urethral vs. non-urethral) [61].

Some limitations have to be considered when interpreting these results. In the Dutch study, the numbers were
low (n = 91) [60]. .

6.3 Summary of evidence for prognosis

Summary of evidence LE
Prognostic factors for survival in primary urethral carcinoma are: age, race, tumour stage and 3
grade, nodal stage, presence of distant metastasis, histological type, tumour size, tumour location,
concomitant bladder cancer and type and modality of treatment.

PRIMARY URETHRAL CARCINOMA - LIMITED UPDATE MARCH 2021 9


7. DISEASE MANAGEMENT
7.1 Treatment of localised primary urethral carcinoma in males
Previously, treatment of male distal urethral carcinoma followed the procedure for penile cancer, with
aggressive surgical excision of the primary lesion with a wide safety margin [37]. Distal urethral tumours exhibit
significantly improved survival rates compared with proximal tumours [62]. Therefore, optimising treatment
of distal urethral carcinoma has become the focus of clinicians to improve functional outcome and quality of
life (QoL), while preserving oncological safety. A retrospective series found no evidence of local recurrence,
even with < 5 mm resection margins (median follow-up: 17–37 months), in men with pT1-3N0-2 distal urethral
carcinoma treated with well-defined penis-preserving surgery and additional iliac/inguinal lymphadenectomy
(LND) for clinically suspected LN disease [63]. Similar results for the feasibility of penile-preserving surgery have
also been reported in recent series [64, 65]. However, a series on patients treated with penis-preserving surgery
for distal urethral carcinoma reported a higher risk of progression in patients with positive proximal margins,
which was also more frequently observed in cases with lymphovascular and peri-neural invasion of the primary
tumour [66].

7.1.1 Summary of evidence and guidelines for the treatment of localised primary urethral
carcinoma in males

Summary of evidence LE
In distal urethral tumours performing a partial urethrectomy with a minimal safety margin does not 3
increase the risk of local recurrence.

Recommendations Strength rating


Offer distal urethrectomy as an alternative to penile amputation in localised distal urethral Weak
tumours, if surgical margins are negative.
Ensure complete circumferential assessment of the proximal urethral margin if penis- Strong
preserving surgery is intended.

7.2 Treatment of localised urethral carcinoma in females


7.2.1 Urethrectomy and urethra-sparing surgery
In women with localised urethral carcinoma, to provide the highest chance of local cure, primary radical
urethrectomy should remove all the peri-urethral tissue from the bulbocavernosus muscle bilaterally and
distally, with a cylinder of all adjacent soft tissue up to the pubic symphysis and bladder neck. Bladder neck
closure and appendicovesicostomy for primary distal urethral lesions has been shown to provide satisfactory
functional results in women [37].
Previous series have reported outcomes in women with mainly distal urethral tumours undergoing
primary treatment with urethra-sparing surgery with or without additional radiotherapy (RT) compared to
primary urethrectomy, with the aim of maintaining integrity and function of the lower urinary tract [67, 68]. In
longer-term series with a median follow-up of 153–175 months, local recurrence rates in women undergoing
partial urethrectomy with intraoperative frozen section analysis were 22–60%, and distal sleeve resection of
> 2 cm resulted in secondary urinary incontinence in 42% of patients who subsequently required additional
reconstructive surgery [67, 68].
Ablative surgical techniques, i.e., transurethral resection (TUR) or laser, used for small distal
urethral tumours, have also resulted in considerable local failure rates of 16%, with a CSS rate of 50%. This
emphasises the critical role of local tumour control in women with distal urethral carcinoma to prevent local and
systemic progression [67].

7.2.2 Radiotherapy
In women, RT was investigated in several older series with a medium follow up of 91–105 months [69]. With a
median cumulative dose of 65 Gy (range 40–106 Gy), the 5-year local control rate was 64% and 7-year CSS
was 49% [69]. Most local failures (95%) occurred within the first two years after primary treatment [69]. The
extent of urethral tumour involvement was found to be the only parameter independently associated with local
tumour control but the type of RT (EBRT vs. interstitial brachytherapy) was not [69]. In one study, the addition
of brachytherapy to EBRT reduced the risk of local recurrence by a factor of 4.2 [70]. Of note, pelvic toxicity
in those achieving local control was considerable (49%), including urethral stenosis, fistula, necrosis, cystitis
and/or haemorrhage, with 30% of the reported complications graded as severe [69].

10 PRIMARY URETHRAL CARCINOMA - LIMITED UPDATE MARCH 2021


7.2.3 Summary of evidence and guidelines for the treatment of localised urethral carcinoma in
females

Summary of evidence LE
In distal tumours, urethra-sparing surgery and local RT represent alternatives to primary urethrectomy 3
but are associated with increased risk of tumour recurrence and local toxicity.

Recommendations Strength rating


Offer urethra-sparing surgery, as an alternative to primary urethrectomy, to women with Weak
distal urethral tumours, if negative surgical margins can be achieved intraoperatively.
Offer local radiotherapy, as an alternative to urethral surgery, to women with localised Weak
urethral tumours, but discuss local toxicity.

7.3 Multimodal treatment in locally advanced urethral carcinoma in both genders


7.3.1 Introduction
Multimodal therapy in primary urethral carcinoma consists of definitive surgery plus chemotherapy
with additional RT [71]. Multimodal therapy was often underutilised (16%) in locally advanced disease
notwithstanding promising results [71-74]. In a recent study monotherapy was associated with decreased local
recurrence-free survival after adjusting for stage, histology, sex, and year of treatment (p = 0.017). Its use has
decreased over time [75].

7.3.2 Preoperative cisplatin-based chemotherapy


Retrospective studies have reported that modern cisplatin-based combination chemotherapy regimens can
be effective in advanced primary urethral carcinoma providing prolonged survival even in LN-positive disease.
Moreover, they have emphasised the critical role of surgery after chemotherapy to achieve long-term survival in
patients with locally advanced urethral carcinoma.
In a series of 124 patients, 39 (31%) were treated with peri-operative platinum-based chemotherapy
for advanced primary urethral carcinoma (twelve patients received neoadjuvant chemotherapy, six received
neoadjuvant chemoradiotherapy and 21 adjuvant chemotherapy). Patients who received neoadjuvant
chemotherapy or chemoradiotherapy for locally advanced primary urethral carcinoma (≥ cT3 and/or cN+)
appeared to demonstrate improved survival compared to those who underwent upfront surgery with or without
adjuvant chemotherapy [76]. Another retrospective series including 44 patients with advanced primary urethral
carcinoma, reported outcomes on 21 patients who had preoperatively received cisplatin-based combination
chemotherapy according to the underlying histologic subtype. The overall response rate for the various
regimens was 72% and the median OS 32 months [52].

7.3.3 Chemoradiotherapy in locally advanced squamous cell carcinoma of the urethra


The clinical feasibility of local RT with concurrent chemotherapy as an alternative to surgery in locally advanced
SCC has been reported in several series. This approach offers a potential for genital preservation [77-81].
The largest, and recently updated, retrospective series reported outcomes in 25 patients with primary locally
advanced SCC of the urethra treated with two cycles of 5-fluorouracil and mitomycin C with concurrent EBRT.
A complete response to primary chemoradiotherapy was observed in ~80% of patients. The 5-year OS and
disease-specific survival was 52% and 68%, respectively. In this updated series, salvage surgery, initiated only
in non-responders or in case of local failure, was not reported to be associated with improved survival [77].

A large retrospective cohort study in patients with locally advanced urethral carcinoma treated with adjuvant RT
and surgery vs. surgery alone demonstrated that the addition of RT improved OS [82].

7.3.4 Salvage treatment in recurrent primary urethral carcinoma after surgery for primary
treatment
A multicentre study reported that patients who were treated with surgery as primary therapy and underwent
surgery or RT-based salvage treatment for recurrent solitary or concomitant urethral disease, demonstrated
similar survival rates compared to patients who never developed recurrence after primary treatment [61].

7.3.5 Treatment of regional lymph nodes


Nodal control in urethral carcinoma can be achieved either by regional LN dissection [37], RT [69] or
chemotherapy [52]. Currently, there is still no clear evidence supporting prophylactic bilateral inguinal and/or
pelvic LND in all patients with urethral carcinoma [54]. However, in patients with clinically enlarged inguinal/

PRIMARY URETHRAL CARCINOMA - LIMITED UPDATE MARCH 2021 11


pelvic LNs or invasive tumours, regional LND should be considered as initial treatment since cure might still
be achievable with limited disease [37]. It was recently shown that in patients with invasive urethral SCC and
cN1-2 disease, inguinal LND conferred an OS benefit [54].

7.3.6 Summary of evidence and guidelines for multimodal treatment in advanced urethral
carcinoma in both genders

Summary of evidence LE
In locally advanced urethral carcinoma, cisplatin-based chemotherapy with curative intent prior 3
to surgery might improve survival compared to chemotherapy alone, or surgery followed by
chemotherapy.
In locally advanced SCC of the urethra, treatment with chemoradiotherapy might be an alternative to 3
surgery.

Recommendations Strength rating


Discuss treatment of patients with locally advanced urethral carcinoma within a Strong
multidisciplinary team of urologists, radio-oncologists, and oncologists.
In locally advanced urethral carcinoma, use cisplatin-based chemotherapeutic regimens Weak
with curative intent prior to surgery.
In locally advanced squamous cell carcinoma (SCC) of the urethra, offer the combination of Weak
curative radiotherapy (RT) with radiosensitising chemotherapy for definitive treatment and
genital preservation.
Offer salvage surgery or RT to patients with urethral recurrence after primary treatment. Weak
Offer inguinal lymph node (LN) dissection to patients with limited LN-positive urethral SCC. Weak

7.4 Treatment of urothelial carcinoma of the prostate


Local conservative treatment with extensive TUR and subsequent BCG instillation is effective in patients with Ta or
Tis prostatic urethral carcinoma [83]. Likewise, patients undergoing TUR of the prostate prior to BCG experience
improved complete response rates compared with those who do not (95% vs. 66%) [84]. Risk of understaging
local extension of prostatic urethral cancer at TUR is high in patients with ductal or stromal involvement [85]. In
smaller series, response rates to BCG in patients with prostatic duct involvement have been reported to vary
between 57% and 75% [83, 86]. Some earlier series have reported superior oncological results for the initial use
of radical cystoprostatectomy as a primary treatment option in patients with ductal involvement [87, 88]. In 24
patients with prostatic stromal invasion treated with radical cystoprostatectomy, a LN mapping study found that
twelve patients had positive LNs, with an increased proportion located above the iliac bifurcation [89].

7.4.1 Summary of evidence and guidelines for the treatment of urothelial carcinoma of the prostate

Summary of evidence LE
Patients undergoing TUR of the prostate for prostatic urothelial carcinoma prior to BCG treatment 3
show superior complete response rates compared to those who do not.

Recommendations Strength rating


Offer a urethra-sparing approach with transurethral resection (TUR) and bacillus-Calmette Strong
Guérin (BCG) to patients with non-invasive urethral carcinoma or carcinoma in situ of the
prostatic urethra and prostatic ducts.
In patients not responding to BCG, or in patients with extensive ductal or stromal Weak
involvement, perform a cystoprostatectomy with extended pelvic lymphadenectomy.

7.5 Metastatic disease


There is no data addressing management of metastatic disease in primary urethral carcinoma patients.
Systemic therapy in metastatic disease should be selected based on the histology of the tumour. The EAU
Guidelines on Metastatic Bladder Cancer can be followed if UC is the predominant histology [2]. Even though
urethral carcinoma patients have been included in large clinical trials on immunotherapy, so far, in terms of
response rates, no subgroup analyses are available [90].

12 PRIMARY URETHRAL CARCINOMA - LIMITED UPDATE MARCH 2021


In addition, there is an urgent clinical need to better address the role of local palliative treatment strategies in
primary urethral carcinoma including surgery, which has shown to impact positively on QoL aspects in selected
patients with advanced genital cancers [91].

Figure 7.1: Management of primary urethral carcinoma

METASTASIS

underlying
according
DISTANT

SALVAGE SURGERY/RADIOTHERAPY
histology
• Systemic
therapy

to
chemotherapy and
consolida ve

• Chemoradio-
N1-N2
T3-T4

surgery***

therapy**
• Induc ve
LOCALLY ADVANCED
(T3-T4 N0-2 M0)

or
chemotherapy and

• Chemoradio-

radiotherapy
• Neoadjuvant
• Chest/abdomen CT

T3-T4

• Surgery and
N0

therapy**
PUC STAGING

adjuvant
surgery
• Pelvic MRI

or

or

LOCAL RECURRENCE
• Urethra-sparing
urethrectomy*

• Urethrectomy

•Radiotherapy
Distal
Ta-T2

surgery*
FEMALE
• Distal
MALE

or

or
• Urethrocystoscopy
PUC DIAGNOSIS

• Urinary cytology

tomy ****
• Biopsy/TUR

complete)
• Urethrec-
Proximal

(paral *
T2

or
LOCALISED (≤ T2 N0M0)

complete)
• Urethrec-
Proximal

(paral *
Ta-T1

tomy

or
+/- Proximal
Prostac T2

lymphade-
cystopros-
tatectomy

tomy and
urethrec-

nectomy
• Radical-
Ta-T2

****
and
BCG *****
Ta-Tis-T1
Prostac

• Repeat
TUR
+

* Ensure complete circumferential assessment if penis-preserving/urethra-sparing surgery or partial


urethrectomy is intended.
** Squamous cell carcinoma.
*** Regional lymphadenectomy should be considered in clinically enlarged lymph nodes.
**** Consider neoadjuvant chemotherapy.
***** In extensive or BCG-unresponsive disease: consider (primary) cystoprostatectomy +/- urethrectomy +
lymphadenectomy.
BCG = bacillus Calmette-Guérin; CT = computed tomography; MRI = magnetic resonance imaging;
PUC = primary urethral carcinoma; TUR = transurethral resection.

PRIMARY URETHRAL CARCINOMA - LIMITED UPDATE MARCH 2021 13


8. FOLLOW-UP
Given the low incidence of primary urethral carcinoma, follow-up has not been systematically investigated.
Therefore, it seems reasonable to tailor surveillance regimens to patients’ individual risk factors (see Section 6.2). In
patients undergoing urethra-sparing surgery, it seems prudent to advocate a more extensive follow-up with urinary
cytology, urethrocystoscopy and cross-sectional imaging despite the lack of specific data.

9. REFERENCES
1. Boorjian, S.A., et al. Risk factors and outcomes of urethral recurrence following radical cystectomy.
Eur Urol, 2011. 60: 1266.
https://pubmed.ncbi.nlm.nih.gov/21871713
2. Witjes, J.A., et al., EAU Guidelines on Muscle-invasive and Metastatic Bladder Cancer. Edn.
presented at the 35th EAU Annual Congress Amsterdam, in EAU Guidelines 2020: Arnhem. The
Netherlands.
https://uroweb.org/guideline/bladder-cancer-muscle-invasive-and-metastatic/
3. Gakis, G., et al. European Association of Urology Guidelines on Primary Urethral Carcinoma-2020
Update. Eur Urol Oncol, 2020. 3: 424.
https://pubmed.ncbi.nlm.nih.gov/23582479
4. Guyatt, G.H., et al. GRADE: an emerging consensus on rating quality of evidence and strength of
recommendations. BMJ, 2008. 336: 924.
https://pubmed.ncbi.nlm.nih.gov/18436948
5. Guyatt, G.H., et al. What is “quality of evidence” and why is it important to clinicians? BMJ, 2008.
336: 995.
https://pubmed.ncbi.nlm.nih.gov/18456631
6. Phillips, B., et al. Oxford Centre for Evidence-based Medicine Levels of Evidence. Updated by
Jeremy Howick March 2009. [accessed January 2021]
https://www.cebm.net/2009/06/oxford-centre-evidence-based-medicine-levels-evidence-march-2009/
7. Guyatt, G.H., et al. Going from evidence to recommendations. BMJ, 2008. 336: 1049.
https://pubmed.ncbi.nlm.nih.gov/18467413
8. Gatta, G., et al. Rare cancers are not so rare: the rare cancer burden in Europe. Eur J Cancer, 2011.
47: 2493.
https://pubmed.ncbi.nlm.nih.gov/22033323
9. Brierley, J.D., et al., TNM classification of malignant tumors. UICC International Union Against
Cancer. 2017, Wiley/Blackwell. p. 208.
https://www.uicc.org/resources/tnm/publications-resources
10. RARECARENet. Surveillance of Rare Cancers in Europe
https://www.rarecancerseurope.org/About-Rare-Cancers/families-and-list-of-rare-cancers
11. Swartz, M.A., et al. Incidence of primary urethral carcinoma in the United States. Urology, 2006.
68: 1164.
https://pubmed.ncbi.nlm.nih.gov/17141838
12. Krukowski, J., et al. Primary urethral carcinoma - unexpected cause of urethral stricture. Case report
and review of the literature. Med Ultrason, 2019. 21: 494.
https://pubmed.ncbi.nlm.nih.gov/31765461
13. Guo, H., et al. Lichen Sclerosus Accompanied by Urethral Squamous Cell Carcinoma: A
Retrospective Study From a Urethral Referral Center. Am J Mens Health, 2018. 12: 1692.
https://pubmed.ncbi.nlm.nih.gov/29926751
14. Colapinto, V., et al. Primary carcinoma of the male urethra developing after urethroplasty for stricture.
J Urol, 1977. 118: 581.
https://pubmed.ncbi.nlm.nih.gov/916053
15. Mohanty, N.K., et al. Squamous cell carcinoma of perineal urethrostomy. Urol Int, 1995. 55: 118.
https://pubmed.ncbi.nlm.nih.gov/8533195
16. Sawczuk, I., et al. Post urethroplasty squamous cell carcinoma. N Y State J Med, 1986. 86: 261.
https://pubmed.ncbi.nlm.nih.gov/3459083
17. Mohan, H., et al. Squamous cell carcinoma of the prostate. Int J Urol, 2003. 10: 114.
https://pubmed.ncbi.nlm.nih.gov/12588611

14 PRIMARY URETHRAL CARCINOMA - LIMITED UPDATE MARCH 2021


18. Arva, N.C., et al. Diagnostic dilemmas of squamous differentiation in prostate carcinoma case report
and review of the literature. Diagn Pathol, 2011. 6: 46.
https://pubmed.ncbi.nlm.nih.gov/21627811
19. Cupp, M.R., et al. Detection of human papillomavirus DNA in primary squamous cell carcinoma of the
male urethra. Urology, 1996. 48: 551.
https://pubmed.ncbi.nlm.nih.gov/8886059
20. Wiener, J.S., et al. Oncogenic human papillomavirus type 16 is associated with squamous cell cancer
of the male urethra. Cancer Res, 1992. 52: 5018.
https://pubmed.ncbi.nlm.nih.gov/1325290
21. Ahmed, K., et al. Urethral diverticular carcinoma: an overview of current trends in diagnosis and
management. Int Urol Nephrol, 2010. 42: 331.
https://pubmed.ncbi.nlm.nih.gov/19649767
22. Chung, D.E., et al. Urethral diverticula in women: discrepancies between magnetic resonance
imaging and surgical findings. J Urol, 2010. 183: 2265.
https://pubmed.ncbi.nlm.nih.gov/20400161
23. Thomas, A.A., et al. Urethral diverticula in 90 female patients: a study with emphasis on neoplastic
alterations. J Urol, 2008. 180: 2463.
https://pubmed.ncbi.nlm.nih.gov/18930487
24. Libby, B., et al. Non-surgical treatment of primary female urethral cancer. Rare Tumors, 2010. 2: e55.
https://pubmed.ncbi.nlm.nih.gov/21139970
25. Altman, D., et al. Cancer Risk After Midurethral Sling Surgery Using Polypropylene Mesh. Obstet
Gynecol, 2018. 131: 469.
https://pubmed.ncbi.nlm.nih.gov/29420401
26. Gandhi, J.S., et al. Clear cell adenocarcinoma of the male urethral tract. Indian J Pathol Microbiol,
2012. 55: 245.
https://pubmed.ncbi.nlm.nih.gov/22771656
27. Mehra, R., et al. Primary urethral clear-cell adenocarcinoma: comprehensive analysis by surgical
pathology, cytopathology, and next-generation sequencing. Am J Pathol, 2014. 184: 584.
https://pubmed.ncbi.nlm.nih.gov/24389164
28. Visser, O., et al. Incidence and survival of rare urogenital cancers in Europe. Eur J Cancer, 2012. 48: 456.
https://pubmed.ncbi.nlm.nih.gov/22119351
29. Rabbani, F. Prognostic factors in male urethral cancer. Cancer, 2011. 117: 2426.
https://pubmed.ncbi.nlm.nih.gov/24048790
30. Aleksic, I., et al. Primary urethral carcinoma: A Surveillance, Epidemiology, and End Results data
analysis identifying predictors of cancer-specific survival. Urol Ann, 2018. 10: 170.
https://pubmed.ncbi.nlm.nih.gov/29719329
31. Sui, W., et al. Outcomes and Prognostic Factors of Primary Urethral Cancer. Urology, 2017. 100: 180.
https://pubmed.ncbi.nlm.nih.gov/27720774
32. Greiman, A.K., et al. Urethral diverticulum: A systematic review. Arab J Urol, 2019. 17: 49.
https://pubmed.ncbi.nlm.nih.gov/31258943
33. E.J. Eble J, S.I., Sauter G., WHO Classification of Tumours: Pathology and Genetics of Tumours of
the Urinary System and Male Genital Organs (IARC WHO Classification of Tumours). 2004, Lyon.
https://publications.iarc.fr/Book-And-Report-Series/Who-Classification-Of-Tumours/WHO-
Classification-Of-Tumours-Of-The-Urinary-System-And-Male-Genital-Organs-2016
34. Comperat, E., et al. Immunochemical and molecular assessment of urothelial neoplasms and aspects
of the 2016 World Health Organization classification. Histopathology, 2016. 69: 717.
https://pubmed.ncbi.nlm.nih.gov/9730466
35. Shanks, J.H., et al. Dataset for reporting of carcinoma of the urethra (in urethrectomy specimens):
recommendations from the International Collaboration on Cancer Reporting (ICCR). Histopathology,
2019. 75: 453.
https://pubmed.ncbi.nlm.nih.gov/31009090
36. Gheiler, E.L., et al. Management of primary urethral cancer. Urology, 1998. 52: 487.
https://pubmed.ncbi.nlm.nih.gov/9730466
37. Karnes, R.J., et al. Surgery for urethral cancer. Urol Clin North Am, 2010. 37: 445.
https://pubmed.ncbi.nlm.nih.gov/20674699
38. Blaivas, J.G., et al. Periurethral masses: etiology and diagnosis in a large series of women. Obstet
Gynecol, 2004. 103: 842.
https://pubmed.ncbi.nlm.nih.gov/15121554

PRIMARY URETHRAL CARCINOMA - LIMITED UPDATE MARCH 2021 15


39. Barkan, G.A., et al. The Paris System for Reporting Urinary Cytology: The Quest to Develop a
Standardized Terminology. Acta Cytol, 2016. 60: 185.
https://pubmed.ncbi.nlm.nih.gov/27318895
40. Touijer, A.K., et al. Role of voided urine cytology in diagnosing primary urethral carcinoma. Urology,
2004. 63: 33.
https://pubmed.ncbi.nlm.nih.gov/14751342
41. Gakis, G., et al. Oncological Outcomes of Patients with Concomitant Bladder and Urethral
Carcinoma. Urol Int, 2016. 97: 134.
https://pubmed.ncbi.nlm.nih.gov/27462702
42. Samm, B.J., et al. Penectomy: a technique to reduce blood loss. Urology, 1999. 53: 393.
https://pubmed.ncbi.nlm.nih.gov/9933061
43. Donat, S.M., et al. The efficacy of transurethral biopsy for predicting the long-term clinical impact of
prostatic invasive bladder cancer. J Urol, 2001. 165: 1580.
https://pubmed.ncbi.nlm.nih.gov/11342921
44. Schubert, T., et al. MP11-20 The predictive accuracy between clinical staging and pathological
staging in patients with primary urethral carcinoma. J Urol, 2019. 201: e135.
https://www.auajournals.org/doi/abs/10.1097/01.JU.0000555184.87550.3c
45. Del Gaizo, A., et al. Magnetic resonance imaging of solid urethral and peri-urethral lesions. Insights
Imaging, 2013. 4: 461.
https://pubmed.ncbi.nlm.nih.gov/23686749
46. Itani, M., et al. MRI of female urethra and periurethral pathologies. Int Urogynecol J, 2016. 27: 195.
https://pubmed.ncbi.nlm.nih.gov/26209954
47. Stewart, S.B., et al. Imaging tumors of the penis and urethra. Urol Clin North Am, 2010. 37: 353.
https://pubmed.ncbi.nlm.nih.gov/20674692
48. Kim, B., et al. Imaging of the male urethra. Semin Ultrasound CT MR, 2007. 28: 258.
https://pubmed.ncbi.nlm.nih.gov/17874650
49. Raman, S.P., et al. Upper and Lower Tract Urothelial Imaging Using Computed Tomography
Urography. Radiol Clin North Am, 2017. 55: 225.
https://pubmed.ncbi.nlm.nih.gov/28126213
50. Gourtsoyianni, S., et al. MRI at the completion of chemoradiotherapy can accurately evaluate
the extent of disease in women with advanced urethral carcinoma undergoing anterior pelvic
exenteration. Clin Radiol, 2011. 66: 1072.
https://pubmed.ncbi.nlm.nih.gov/21839430
51. Naumann, C.M., et al. Reliability of dynamic sentinel node biopsy combined with ultrasound-guided
removal of sonographically suspicious lymph nodes as a diagnostic approach in patients with penile
cancer with palpable inguinal lymph nodes. Urol Oncol, 2015. 33: 389.e9.
https://pubmed.ncbi.nlm.nih.gov/25934562
52. Dayyani, F., et al. Retrospective analysis of survival outcomes and the role of cisplatin-based
chemotherapy in patients with urethral carcinomas referred to medical oncologists. Urol Oncol, 2013.
31: 1171.
https://pubmed.ncbi.nlm.nih.gov/22534087
53. Gakis, G., et al. Prognostic factors and outcomes in primary urethral cancer: results from the
international collaboration on primary urethral carcinoma. World J Urol, 2016. 34: 97.
https://pubmed.ncbi.nlm.nih.gov/25981402
54. Werntz, R.P., et al. The role of inguinal lymph node dissection in men with urethral squamous cell
carcinoma. Urol Oncol, 2018. 36: 526 e1.
https://pubmed.ncbi.nlm.nih.gov/30446445
55. Carroll, P.R., et al. Surgical anatomy of the male and female urethra. Urol Clin North Am, 1992. 19: 339.
https://pubmed.ncbi.nlm.nih.gov/1574824
56. Sharp, D., et al., Surgery of penile and urethral carcinoma, in Campbell’s Urology, D. McDougal, A.
Wein, L. Kavoussi, A. Novick, A. Partin, C. Peters & P. Ramchandani, Editors. 212, Saunders Elsevier:
Philadelphia, PA, USA.
https://www.researchgate.net/publication/289837938_Surgery_of_Penile_and_Urethral_Carcinoma
57. Abudurexiti, M., et al. Prognosis of rare pathological primary urethral carcinoma. Cancer Manag Res,
2018. 10: 6815.
https://pubmed.ncbi.nlm.nih.gov/30584373
58. Champ, C.E., et al. Prognostic factors and outcomes after definitive treatment of female urethral
cancer: a population-based analysis. Urology, 2012. 80: 374.
https://pubmed.ncbi.nlm.nih.gov/22857759

16 PRIMARY URETHRAL CARCINOMA - LIMITED UPDATE MARCH 2021


59. Zi, H., et al. Nomograms for predicting long-term overall survival and cancer-specific survival in
patients with primary urethral carcinoma: a population-based study. Int Urol Nephrol, 2020. 52: 287.
https://pubmed.ncbi.nlm.nih.gov/31612421
60. Derksen, J.W., et al. Primary urethral carcinoma in females: an epidemiologic study on
demographical factors, histological types, tumour stage and survival. World J Urol, 2013. 31: 147.
https://pubmed.ncbi.nlm.nih.gov/22614443
61. Gakis, G., et al. The prognostic effect of salvage surgery and radiotherapy in patients with recurrent
primary urethral carcinoma. Urol Oncol, 2018. 36: 10 e7.
http://ascopubs.org/doi/abs/10.1200/jco.2015.33.15_suppl.4568
62. Dalbagni, G., et al. Male urethral carcinoma: analysis of treatment outcome. Urology, 1999. 53: 1126.
https://pubmed.ncbi.nlm.nih.gov/10367840
63. Smith, Y., et al. Penile-preserving surgery for male distal urethral carcinoma. BJU Int, 2007. 100: 82.
https://pubmed.ncbi.nlm.nih.gov/17488307
64. Pedrosa, J.A., et al. Distal urethrectomy for localized penile squamous carcinoma in situ extending
into the urethra: an updated series. Int Urol Nephrol, 2014. 46: 1551.
https://pubmed.ncbi.nlm.nih.gov/24633698
65. Kulkarni, M., et al. MP10-16 Substitution urethroplasty for treatment of distal urethral carcinoma and
carcinoma in situ. J. Urol 193: e117.
https://www.jurology.com/article/S0022-5347(15)00728-4/fulltext
66. Torbrand, C., et al. Diagnosing Distal Urethral Carcinomas in Men Might Be Only the Tip of the
Iceberg. Clin Genitourin Cancer, 2017. 15: e1131.
https://pubmed.ncbi.nlm.nih.gov/28784424
67. Dimarco, D.S., et al. Surgical treatment for local control of female urethral carcinoma. Urol Oncol,
2004. 22: 404.
https://pubmed.ncbi.nlm.nih.gov/15464921
68. DiMarco, D.S., et al. Outcome of surgical treatment for primary malignant melanoma of the female
urethra. J Urol, 2004. 171: 765.
https://pubmed.ncbi.nlm.nih.gov/14713806
69. Garden, A.S., et al. Primary carcinoma of the female urethra. Results of radiation therapy. Cancer,
1993. 71: 3102.
https://pubmed.ncbi.nlm.nih.gov/8490839
70. Milosevic, M.F., et al. Urethral carcinoma in women: results of treatment with primary radiotherapy.
Radiother Oncol, 2000. 56: 29.
https://pubmed.ncbi.nlm.nih.gov/10869752
71. Zinman, L.N., et al. Management of Proximal Primary Urethral Cancer: Should Multidisciplinary
Therapy Be the Gold Standard? Urol Clin North Am, 2016. 43: 505.
https://pubmed.ncbi.nlm.nih.gov/27717436
72. Cahn, D.B., et al. Contemporary practice patterns and survival outcomes for locally advanced
urethral malignancies: A National Cancer Database Analysis. Urol Oncol, 2017. 35: 670 e15.
https://pubmed.ncbi.nlm.nih.gov/28803701
73. Dayyani, F., et al. Management of advanced primary urethral carcinomas. BJU Int, 2014. 114: 25.
https://pubmed.ncbi.nlm.nih.gov/24447439
74. Peyton, C.C., et al. Survival Outcomes Associated With Female Primary Urethral Carcinoma: Review
of a Single Institutional Experience. Clin Genitourin Cancer, 2018. 16: e1003.
https://pubmed.ncbi.nlm.nih.gov/29859736
75. Mano, R., et al. Primary urethral cancer: treatment patterns and associated outcomes. BJU Int, 2020.
126: 359.
https://pubmed.ncbi.nlm.nih.gov/32336001
76. Gakis, G., et al. Impact of perioperative chemotherapy on survival in patients with advanced primary
urethral cancer: results of the international collaboration on primary urethral carcinoma. Ann Oncol,
2015. 26: 1754.
https://pubmed.ncbi.nlm.nih.gov/25969370
77. Kent, M., et al. Combined chemoradiation as primary treatment for invasive male urethral cancer.
J Urol, 2015. 193: 532.
https://pubmed.ncbi.nlm.nih.gov/25088950
78. Gakis, G. Editorial Comment to Docetaxel, cisplatin and 5-fluorouracil chemotherapy with concurrent
radiation for unresectable advanced urethral carcinoma. Int J Urol, 2014. 21: 424.
https://pubmed.ncbi.nlm.nih.gov/24251884

PRIMARY URETHRAL CARCINOMA - LIMITED UPDATE MARCH 2021 17


79. Itoh, J., et al. Docetaxel, cisplatin and 5-fluorouracil chemotherapy with concurrent radiation for
unresectable advanced urethral carcinoma. Int J Urol, 2014. 21: 422.
https://pubmed.ncbi.nlm.nih.gov/24251859
80. Hara, I., et al. Successful treatment for squamous cell carcinoma of the female urethra with combined
radio- and chemotherapy. Int J Urol, 2004. 11: 678.
https://pubmed.ncbi.nlm.nih.gov/15285764
81. Cohen, M.S., et al. Coordinated chemoradiation therapy with genital preservation for the treatment of
primary invasive carcinoma of the male urethra. J Urol, 2008. 179: 536.
https://pubmed.ncbi.nlm.nih.gov/18076921
82. Son, C.H., et al. Optimizing the Role of Surgery and Radiation Therapy in Urethral Cancer Based on
Histology and Disease Extent. Int J Radiat Oncol Biol Phys, 2018. 102: 304.
https://pubmed.ncbi.nlm.nih.gov/29908944
83. Palou Redorta, J., et al. Intravesical instillations with bacillus calmette-guerin for the treatment of
carcinoma in situ involving prostatic ducts. Eur Urol, 2006. 49: 834.
https://pubmed.ncbi.nlm.nih.gov/16426729
84. Gofrit, O.N., et al. Prostatic urothelial carcinoma: is transurethral prostatectomy necessary before
bacillus Calmette-Guerin immunotherapy? BJU Int, 2009. 103: 905.
https://pubmed.ncbi.nlm.nih.gov/19021623
85. Njinou Ngninkeu, B., et al. Transitional cell carcinoma involving the prostate: a clinicopathological
retrospective study of 76 cases. J Urol, 2003. 169: 149.
https://pubmed.ncbi.nlm.nih.gov/12478124
86. Palou, J., et al. Urothelial carcinoma of the prostate. Urology, 2007. 69: 50.
https://pubmed.ncbi.nlm.nih.gov/17280908
87. Hillyard, R.W., Jr., et al. Superficial transitional cell carcinoma of the bladder associated with mucosal
involvement of the prostatic urethra: results of treatment with intravesical bacillus Calmette-Guerin.
J Urol, 1988. 139: 290.
https://pubmed.ncbi.nlm.nih.gov/3339727
88. Solsona, E., et al. The prostate involvement as prognostic factor in patients with superficial bladder
tumors. J Urol, 1995. 154: 1710.
https://pubmed.ncbi.nlm.nih.gov/7563328
89. Vazina, A., et al. Stage specific lymph node metastasis mapping in radical cystectomy specimens.
J Urol, 2004. 171: 1830.
https://pubmed.ncbi.nlm.nih.gov/15076287
90. Balar, A.V., et al. First-line pembrolizumab in cisplatin-ineligible patients with locally advanced and
unresectable or metastatic urothelial cancer (KEYNOTE-052): a multicentre, single-arm, phase 2
study. Lancet Oncol, 2017. 18: 1483.
https://pubmed.ncbi.nlm.nih.gov/28967485
91. Shukla, C.J., et al. Palliation of male genital cancers. Clin Oncol (R Coll Radiol), 2010. 22: 747.
https://pubmed.ncbi.nlm.nih.gov/20800458

10. CONFLICT OF INTEREST


All members of the Muscle-invasive and Metastatic Bladder Cancer Guidelines Panel have provided disclosure
statements of all relationships that they have that might be perceived as a potential source of a conflict of
interest. This information is open access available on the European Association of Urology website:
http://www.uroweb.org/guidelines/primary-urethral-carcinoma/.
This guidelines document was developed with the financial support of the European Association of
Urology. No external sources of funding and support have been involved. The EAU is a non-profit organisation
and funding is limited to administrative assistance and travel and meeting expenses. No honoraria or other
reimbursements have been provided.

18 PRIMARY URETHRAL CARCINOMA - LIMITED UPDATE MARCH 2021


11. CITATION INFORMATION
The format in which to cite the EAU Guidelines will vary depending on the style guide of the journal in which
the citation appears. Accordingly, the number of authors or whether, for instance, to include the publisher,
location,or an ISBN number may vary.

The compilation of the complete Guidelines should be referenced as:


EAU Guidelines. Edn. presented at the EAU Annual Congress Milan 2021. ISBN 978-94-92671-13-4.

If a publisher and/or location is required, include:


EAU Guidelines Office, Arnhem, The Netherlands. http://uroweb.org/guidelines/compilations-of-all-guidelines/

References to individual guidelines should be structured in the following way:


Contributors’ names. Title of resource. Publication type. ISBN. Publisher and publisher location, year.

PRIMARY URETHRAL CARCINOMA - LIMITED UPDATE MARCH 2021 19


20 PRIMARY URETHRAL CARCINOMA - LIMITED UPDATE MARCH 2021
EAU Guidelines on
Renal Cell
Carcinoma
B. Ljungberg (Chair), L. Albiges, J. Bedke, A. Bex (Vice-chair),
U. Capitanio, R.H. Giles (Patient Advocate), M. Hora, T. Klatte
T. Lam, L. Marconi, T. Powles, A. Volpe
Guidelines Associates: Y. Abu-Ghanem, S. Dabestani,
S. Fernández-Pello Montes, F. Hofmann, T. Kuusk, R. Tahbaz

© European Association of Urology 2021


TABLE OF CONTENTS PAGE
1. INTRODUCTION 6
1.1 Aims and scope 6
1.2 Panel composition 6
1.3 Acknowledgement 6
1.4 Available publications 6
1.5 Publication history and summary of changes 6
1.5.1 Publication history 6
1.5.2 Summary of changes 6

2. METHODS 10
2.1 Data identification 10
2.2 Review 11
2.3 Future goals 11

3. EPIDEMIOLOGY, AETIOLOGY AND PATHOLOGY 12


3.1 Epidemiology 12
3.2 Aetiology 12
3.2.1 Summary of evidence and recommendation for epidemiology, aetiology and
pathology 12
3.3 Histological diagnosis 12
3.3.1 Clear-cell RCC 13
3.3.2 Papillary RCC 13
3.3.3 Chromophobe RCC 13
3.4 Other renal tumours 13
3.4.1 Renal medullary carcinoma 13
3.4.1.1 Treatment of renal medullary carcinoma 13
3.4.2 Carcinoma associated with end-stage renal disease; acquired cystic disease-
associated RCC 14
3.4.3 Papillary adenoma 14
3.4.4 Hereditary kidney tumours 14
3.4.5 Angiomyolipoma 15
3.4.5.1 Treatment 15
3.4.6 Renal oncocytoma 15
3.4.7 Cystic renal tumours 17
3.5 Summary of evidence and recommendations for the management of other renal tumours 17
3.6 Recommendations for the management of other renal tumours 17

4. STAGING AND CLASSIFICATION SYSTEMS 18


4.1 Staging 18
4.2 Anatomic classification systems 18

5. DIAGNOSTIC EVALUATION 19
5.1 Symptoms 19
5.1.1 Physical examination 19
5.1.2 Laboratory findings 19
5.2 Imaging investigations 19
5.2.1 Presence of enhancement 19
5.2.2 Computed tomography or magnetic resonance imaging 19
5.2.3 Other investigations 20
5.2.4 Radiographic investigations to evaluate RCC metastases 20
5.2.5 Bosniak classification of renal cystic masses 20
5.3 Renal tumour biopsy 21
5.3.1 Indications and rationale 21
5.3.2 Technique 21
5.3.3 Diagnostic yield and accuracy 22
5.3.4 Morbidity 22
5.4 Summary of evidence and recommendations for the diagnostic assessment of RCC 22

2 RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021


6. PROGNOSTIC FACTORS 23
6.1 Classification 23
6.2 Anatomical factors 23
6.3 Histological factors 23
6.4 Clinical factors 24
6.5 Molecular factors 24
6.6 Prognostic models 25
6.7 Summary of evidence and recommendations for prognostic factors 25

7. DISEASE MANAGEMENT 27
7.1 Treatment of localised RCC 27
7.1.1 Introduction 27
7.1.2 Surgical treatment 27
7.1.2.1 Nephron-sparing surgery versus radical nephrectomy in localised
RCC 27
7.1.2.1.1 T1 RCC 27
7.1.2.1.2 T2 renal cell carcinoma 28
7.1.2.2 Associated procedures 28
7.1.2.2.1 Adrenalectomy 28
7.1.2.2.2 Lymph node dissection for clinically negative lymph
nodes (cN0) 29
7.1.2.2.3 Embolisation 29
7.1.2.2.4 Summary of evidence and recommendations for the
treatment of localised RCC 29
7.1.3 Radical and partial nephrectomy techniques 30
7.1.3.1 Radical nephrectomy techniques 30
7.1.3.2 Partial nephrectomy techniques 30
7.1.3.2.1 Open versus laparoscopic approach 30
7.1.3.2.2 Open versus robotic approach 31
7.1.3.2.3 Open versus hand-assisted approach 31
7.1.3.2.4 Open versus laparoscopic versus robotic approaches 31
7.1.3.2.5 Laparoscopic versus robotic approach 31
7.1.3.2.6 Surgical volume 31
7.1.3.3 Positive surgical margins on histopathological specimens 31
7.1.3.4 Summary of evidence and recommendations for radical and partial
nephrectomy techniques 32
7.1.4 Therapeutic approaches as alternatives to surgery 32
7.1.4.1 Surgical versus non-surgical treatment 32
7.1.4.2 Active surveillance and watchful waiting 32
7.1.4.3 Role of renal tumour biopsy before active surveillance 33
7.1.4.4 Tumour ablation 33
7.1.4.4.1 Role of renal mass biopsy 33
7.1.4.4.2 Cryoablation 33
7.1.4.4.3 Radiofrequency ablation 34
7.1.4.4.4 Tumour ablation versus surgery 34
7.1.4.4.5 Stereotactic ablative radiotherapy 34
7.1.4.4.6 Other ablative techniques 35
7.1.4.4.7 Summary of evidence and recommendation for
therapeutic approaches as alternative to surgery 35
7.2 Treatment of locally advanced RCC 35
7.2.1 Introduction 35
7.2.2 Role of lymph node invasion in locally advanced RCC 35
7.2.2.1 Management of clinically negative lymph nodes (cN-) in locally
advanced RCC 35
7.2.2.2 Management of clinically positive lymph nodes (cN+) in locally
advanced RCC 35
7.2.3 Management of locally advanced unresectable RCC 36
7.2.4 Management of RCC with venous tumour thrombus 36
7.2.4.1 The evidence base for surgery in patients with venous tumour
thrombus 36

RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021 3


7.2.4.2 The evidence base for different surgical strategies 36
7.2.4.3 Summary of evidence and recommendations for the management of
RCC with venous tumour thrombus 36
7.2.5 Neoadjuvant and adjuvant therapy 36
7.2.5.1 Summary of evidence and recommendations for adjuvant therapy 38
7.3 Advanced/metastatic RCC 38
7.3.1 Local therapy of advanced/metastatic RCC 38
7.3.1.1 Cytoreductive nephrectomy 38
7.3.1.1.1 Embolisation of the primary tumour 39
7.3.1.1.2 Summary of evidence and recommendations for local
therapy of advanced/metastatic RCC 39
7.3.2 Local therapy of metastases in metastatic RCC 40
7.3.2.1 Complete versus no/incomplete metastasectomy 40
7.3.2.2 Local therapies for RCC bone metastases 40
7.3.2.3 Local therapies for RCC brain metastases 40
7.3.2.4 Embolisation of metastases 41
7.3.2.5 Adjuvant treatment in cM0 patients after metastasectomy 41
7.3.2.6 Summary of evidence and recommendations for local therapy of
metastases in metastatic RCC 41
7.4 Systemic therapy for advanced/metastatic RCC 41
7.4.1 Chemotherapy 41
7.4.1.1 Recommendation for systemic therapy in advanced/metastatic RCC 42
7.4.2 Immunotherapy 42
7.4.2.1 IFN-α monotherapy and combined with bevacizumab 42
7.4.2.2 Interleukin-2 42
7.4.2.3 Immune checkpoint blockade 42
7.4.2.3.1 Immuno-oncology monotherapy 42
7.4.2.4 Immunotherapy/combination therapy 42
7.4.2.5 Summary of evidence and recommendations for immunotherapy in
metastatic RCC 45
7.4.3 Targeted therapies 46
7.4.3.1 Tyrosine kinase inhibitors 47
7.4.3.1.1 Sorafenib 47
7.4.3.1.2 Sunitinib 47
7.4.3.1.3 Pazopanib 47
7.4.3.1.4 Axitinib 47
7.4.3.1.5 Cabozantinib 47
7.4.3.1.6 Lenvatinib 48
7.4.3.1.7 Tivozanib 48
7.4.4 Monoclonal antibody against circulating VEGF 48
7.4.4.1 Bevacizumab monotherapy and bevacizumab plus IFN-α 48
7.4.5 mTOR inhibitors 48
7.4.5.1 Temsirolimus 48
7.4.5.2 Everolimus 48
7.4.6 Therapeutic strategies 48
7.4.6.1 Therapy for treatment-naïve patients with clear-cell metastatic RCC 48
7.4.6.1.1 Sequencing systemic therapy in clear-cell metastatic RCC 48
7.4.6.2 Non-clear-cell metastatic RCC 49
7.4.7 Summary of evidence and recommendations for targeted therapy in metastatic
RCC 50
7.5 Locally recurrent RCC after treatment of localised disease 51
7.5.1 Summary of evidence and recommendation on locally recurrent RCC after
treatment of localised disease 52

8. FOLLOW-UP IN RCC 52
8.1 Introduction 52
8.2 Which imaging investigations for which patients, and when? 53
8.3 Summary of evidence and recommendations for surveillance following RN or PN or
ablative therapies in RCC 55
8.4 Research priorities 55

4 RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021


9. REFERENCES 55

10. CONFLICT OF INTEREST 84

11. CITATION INFORMATION 84

RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021 5


1. INTRODUCTION
1.1 Aims and scope
The European Association of Urology (EAU) Renal Cell Cancer (RCC) Guidelines Panel has compiled these
clinical guidelines to provide urologists with evidence-based information and recommendations for the
management of RCC.
It must be emphasised that clinical guidelines present the best evidence available to the experts
but following guideline recommendations will not necessarily result in the best outcome. Guidelines can never
replace clinical expertise and judgement when making treatment decisions for individual patients, but rather
help to focus decisions whilst also taking personal values and preferences/individual circumstances of patients
into account. Guidelines are not mandates and do not purport to be a legal standard of care.

1.2 Panel composition


The RCC Guidelines Panel is an international group of clinicians consisting of urological surgeons, oncologists,
methodologists, a pathologist and a radiologist, with particular expertise in the field of renal cancer care. Since
2015, the Panel has incorporated a patient advocate to provide a consumer perspective for its guidelines.
All experts involved in the production of this document have submitted potential conflict of interest
statements, which can be viewed on the EAU website Uroweb: http://uroweb.org/guideline/renalcellcarcinoma/.

1.3 Acknowledgement
The RCC Guidelines Panel is most grateful for the continued methodological and scientific support provided by
Prof.Dr. O. Hes (pathologist, Pilzen, Czech Republic) for two sections of this document: Histological diagnosis
and Other renal tumours.

1.4 Available publications


A quick reference document (Pocket Guidelines) is available, both in print and as an app for iOS and Android
devices, presenting the main findings of the RCC Guidelines. These are abridged versions which may require
consultation together with the full text version. Several scientific publications are available, as are a number of
translations of all versions of the EAU RCC Guidelines [1]. All documents can be accessed on the EAU website:
http://uroweb.org/guideline/renal-cell-carcinoma/.

1.5 Publication history and summary of changes


1.5.1 Publication history
The EAU RCC Guidelines were first published in 2000. This 2021 RCC Guidelines document presents a
substantial update of the 2020 publication.

1.5.2 Summary of changes


All chapters of the 2021 RCC Guidelines have been updated, based on the 2020 version of the Guidelines.
References have been added throughout the document.

New data have been included in the following sections, resulting in changed recommendations in:

Section 5.4 Summary of evidence and recommendations for the diagnostic assessment of RCC

Recommendations Strength rating


Omit chest CT in patients with incidentally noted cT1a disease due to the low risk of Weak
lung metastases in this cohort.
Use non-ionising modalities, including MRI and contrast-enhanced ultrasound, for Strong
further characterisation of small renal masses, tumour thrombus and differentiation
of unclear renal masses, if the results of contrast-enhanced CT are indeterminate.

6 RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021


Section 6.7 Summary of evidence and recommendations for prognostic factors

Summary of evidence LE
In RCC patients, TNM stage, tumour size, grade, and RCC subtype provide important 2
prognostic information.

Recommendations Strength rating


Use the WHO/ISUP grading system and classify renal cell carcinoma type. Strong
Use prognostic models in localised and metastatic disease. Strong
Do not routinely use molecular markers to assess prognosis. Strong

7.1.2.2.4 Summary of evidence and recommendations for the treatment of localised RCC

Summary of evidence LE
Retrospective studies suggest that oncological outcomes are similar following PN vs. RN in 3b
patients with larger (≥ 7 cm) RCC. Post-operative complication rates are higher in PN groups.
In patients with localised disease without radiographic evidence of LN metastases, a survival 2b
advantage of LND in conjunction with RN is not demonstrated in randomised trials.

Recommendations Strength rating


Offer partial nephrectomy to patients with T2 tumours and a solitary kidney or Weak
chronic kidney disease, if technically feasible.
Do not offer an extended lymph node dissection to patients with organ-confined Weak
disease.

7.1.3.4 Summary of evidence and recommendations for radical and partial nephrectomy techniques

Summary of evidence LE
Robotic-assisted and laparoscopic PN are associated with shorter length of stay and lower 2b
blood loss compared to open PN.
Hospital volume in PN might impact on surgical complications, warm ischaemia and surgical 3
margins.
Radical nephrectomy after positive surgical margins can result in over-treatment in many cases. 3

Recommendation Strength rating


Intensify follow-up in patients with a positive surgical margin. Weak

Section 7.1.4.3.7 Summary of evidence and recommendation for therapeutic approaches as alternative
to surgery

Summary of evidence LE
Low quality studies suggest high disease recurrence rates after radiofrequency ablation of 3
tumours > 3 cm and after cryoablation of tumours > 4 cm.
Low quality studies suggest a higher local recurrence rate for thermal ablation therapies 3
compared to PN, but the quality of the data does not allow definitive conclusions.

Recommendations Strength rating


Offer active surveillance (AS) or thermal ablation (TA) to frail and/or comorbid Weak
patients with small renal masses.
Perform a percutaneous renal mass biopsy prior to, and not concomitantly with TA. Strong
When TA or AS are offered, discuss with patients about the harms/benefits with Strong
regards to oncological outcomes and complications.
Do not routinely offer TA for tumours > 3 cm and cryoablation for tumours > 4 cm. Weak

RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021 7


Section 7.2.4.3 Summary of evidence and recommendations for the management of RCC with venous tumour
thrombus

Summary of evidence LE
In patients with locally advanced disease, the survival benefit of lymph node (LN) dissection is 3
unproven but LN dissection has significant staging, prognosis and follow-up implications.

Recommendations Strength rating


In patients with clinically enlarged lymph nodes (LNs), perform LN dissection to Weak
guide staging, prognosis and follow-up.
In case of metastatic disease, discuss surgery within the context of a Weak
multidisciplinary team.

Section 7.2.5.1 Summary of evidence and recommendations for adjuvant therapy

Summary of evidence LE
Adjuvant therapy does not improve survival after nephrectomy. 1b
In one single RCT, in selected high-risk patients, adjuvant sunitinib improved disease-free 1b
survival (DFS) but not overall survival (OS).
Adjuvant sorafenib, pazopanib, everolimus, girentuximab or axitinib does not improve DFS or 1b
OS after nephrectomy.
Adjuvant RCTs are ongoing to evaluate the benefit of adjuvant immunotherapy after 1b
nephrectomy in high-risk patients.

Recommendations Strength rating


Do not offer adjuvant therapy with sorafenib, pazopanib, everolimus, girentuximab Strong
or axitinib.
Do not offer adjuvant sunitinib following surgically resected high-risk clear-cell renal Weak
cell carcinoma.

Section 7.3.1.1.2 Summary of evidence and recommendations for local therapy of advanced/metastatic RCC

Summary of evidence LE
Patients with their primary tumour in place treated with ICI-based combination therapy have 2b
better PFS and OS in exploratory subgroup analyses compared to treatment with sunitinib.

Recommendation Strength rating


Discuss delayed cytoreductive nephrectomy in patients who derive clinical benefit Weak
from systemic therapy.

Section 7.3.2.6 Summary of evidence and recommendations for local therapy of metastases in metastatic RCC

Summary of evidence LE
Tyrosine kinase inhibitors treatment after metastasectomy in patients with no evidence of 1b
disease did not improve RFS when compared to placebo or observation.

Recommendation Strength rating


Do not offer tyrosine kinase inhibitor treatment to mRCC patients after Strong
metastasectomy and no evidence of disease.

8 RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021


Section 7.4.2.5 Summary of evidence and recommendations for immunotherapy in metastatic RCC

Summary of evidence LE
The combination of pembrolizumab plus axitinib, lenvatinib plus pembrolizumab and 1b
nivolumab plus cabozantinib in treatment-naive patients with cc-mRCC across all IMDC risk
group demonstrated PFS, OS and ORR benefits compared to sunitinib.
Axitinib, cabozantinib or lenvatinib can be continued if immune-related adverse events result 4
in cessation of axitinib plus pembrolizumab, cabozantinib plus nivolumab or lenvatinib plus
pembrolizumab. Re-challenge with immunotherapy requires expert support.
Nivolumab plus ipilimumab, pembrolizumab plus axitinib, nivolumab plus cabozantinib and 4
lenvatinib plus pembrolizumab should be administered in centres with experience of immune
combination therapy and appropriate supportive care within the context of a multidisciplinary
team.
The combination of nivolumab plus ipilimumab in the IMDC intermediate- and poor-risk 2b
population of treatment-naive patients with cc-mRCC leads to superior survival compared to
sunitinib while OS was higher in IMDC good-risk patients with sunitinib.

Recommendations Strength rating


Offer pembrolizumab plus axitinib, lenvatinib plus pembrolizumab or nivolumab plus Strong
cabozantinib to treatment-naive patients in clear-cell metastatic renal cell carcinoma
(cc-mRCC).
Administer nivolumab plus ipilimumab, pembrolizumab plus axitinib, lenvatinib plus Weak
pembrolizumab and nivolumab plus cabozantinib in centres with experience of
immune combination therapy and appropriate supportive care within the context of
a multidisciplinary team.
Offer axitinib, cabozantinib or lenvatinib as subsequent treatment to patients who Weak
experience treatment-limiting immune-related adverse events after treatment with
the combination of axitinib plus pembrolizumab, cabozantinib plus nivolumab or
lenvatinib plus pembrolizumab.

Figure 7.1: Updated EAU Guidelines recommendations for the first-line treatment of metastatic ccRCC

Alternative in patients who can


Standard of Care not receive or tolerate immune
checkpoint inhibitors

nivolumab/cabozantinib [1b]
sunitinib* [1b]
IMDC favourable risk pembrolizumab/axitinib [1b]
pazopanib* [1b]
pembrolizumab/lenvatinib [1b]

nivolumab/cabozantinib [1b]
cabozantinib* [2a]
IMDC intermediate and pembrolizumab/axitinib [1b]
sunitinib*[1b]
poor risk pembrolizumab/lenvatinib [1b]
pazopanib* [1b]
nivolumab/ipilimumab [1b]

Section 7.4.7 Summary of evidence and recommendations for targeted therapy in metastatic RCC

Recommendations Strength rating


Offer nivolumab or cabozantinib for immune checkpoint inhibitor-naive vascular Strong
endothelial growth factor receptor (VEGFR)-refractory clear-cell metastatic renal cell
carcinoma (cc-mRCC) after one or two lines of therapy.
Offer VEGF-tyrosine kinase inhibitors as second-line therapy to patients refractory Weak
to nivolumab plus ipilimumab or axitinib plus pembrolizumab, cabozantinib plus
nivolumab or lenvatinib plus pembrolizumab.

RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021 9


Section 8.3 S
 ummary of evidence and recommendations for surveillance following RN or PN or
ablative therapies in RCC

Summary of evidence LE
Functional follow-up after curative treatment for RCC is useful to prevent renal and 4
cardiovascular deterioration.
Oncological follow-up can detect local recurrence or metastatic disease while the patient may 4
still be surgically curable.
Prognostic models provide stratification of RCC risk of recurrence based on TNM and 3
histological features
In competing-risk models, risk of non-RCC-related death exceeds that of RCC recurrence or 3
related death in low-risk patients.
Life expectancy estimation is feasible and may support counselling of patients on duration of 4
follow-up.

Recommendations Strength rating


Perform functional follow-up (renal function assessment and prevention of Weak
cardiovascular events) both in nephron-sparing (NSS) and radical nephrectomy (RN)
patients.
Consider curtailing follow-up when the risk of dying from other causes is double that Weak
of the recurrence risk of RCC.

2. METHODS
2.1 Data identification
For the 2021 Guidelines, new and relevant evidence has been identified, collated and appraised through a
structured assessment of the literature for the chapters as listed in Table 2.1.
A broad and comprehensive scoping search was performed, which was limited to studies
representing high certainty of evidence (i.e. systematic reviews with or without meta-analysis, randomised
controlled trials (RCTs), and prospective non-randomised comparative studies only for therapeutic
interventions, and systematic reviews and prospective studies with well-defined reference standards for
diagnostic accuracy studies) published in the English language. In case no higher level data exists for a
particular topic, lower level evidence is considered for inclusion. The search was restricted to articles published
between April 1st 2019 and June 25th 2020. Databases covered included Medline, EMBASE, and the Cochrane
Library. After de-duplication, a total of 1,973 unique records were identified, retrieved and screened for relevance.
A total of 106 new references have been included in the 2021 RCC Guidelines publication. A search
strategy is published online: https://uroweb.org/guideline/renal-cell-carcinoma/?type=appendices-publications.

For each recommendation within the guidelines there is an accompanying online strength rating form, the
basis of which is a modified GRADE methodology [2]. Each strength rating form addresses a number of key
elements, namely:
1. the overall quality of the evidence which exists for the recommendation; references used in
this text are graded according to a classification system modified from the Oxford Centre for
Evidence-Based Medicine Levels of Evidence [3];
2. the magnitude of the effect (individual or combined effects);
3. the certainty of the results (precision, consistency, heterogeneity and other statistical or
study-related factors);
4. the balance between desirable and undesirable outcomes;
5. the impact of patient values and preferences on the intervention;
6. the certainty of those patient values and preferences.

These key elements are the basis which panels use to define the strength rating of each recommendation.

The strength of each recommendation is represented by the words ‘strong’ or ‘weak’ [4]. The strength of each
recommendation is determined by the balance between desirable and undesirable consequences of alternative
management strategies, the quality of the evidence (including certainty of estimates), and nature and variability
of patient values and preferences. The strength rating forms will be available online.

10 RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021


Specific chapters were updated by way of systematic reviews, commissioned and undertaken by the Panel,
based on prioritised topics or questions. These reviews were performed using standard Cochrane systematic
review methodology: http://www.cochranelibrary.com/about/aboutcochranesystematic-reviews.html.

Table 2.1: Description of update and summary of review methodology

Chapter Brief description of review methodology


1. Introduction Not applicable.
2. Methods Not applicable.
3. Epidemiology, aetiology and pathology This chapter was updated by a narrative review,
based on a structured literature assessment.
4. Staging and grading classification systems This chapter was updated by a narrative review,
based on a structured literature assessment.
Section 3.4.5.1 (Treatment of angiomyolipoma) was
updated by means of a systematic review [5].
5. Diagnostic evaluation Section 5.2 (Diagnostic imaging) was revised based
on a systematic review [6]. The remainder of the
chapter was updated by a structured literature
assessment.
6. Prognosis This chapter was updated by a narrative review,
based on a structured literature assessment.
7. Treatment (Disease management) Sections 7.1.2 and 7.2.4 (Treatment of localised and
locally advanced disease) were revised based on an
updated systematic review. Sub-section 7.1.4.3.2
(Cryoablation versus partial nephrectomy) was
updated by means of a systematic review [7].
Section 7.4.6.2 (Non-clear-cell metastatic RCC) was
updated by means of a systematic review [8].

Some aspects of Section 7.4 (Targeted therapy for


metastatic RCC) were updated by way of a Cochrane
systematic review [9].
The remainder of the chapter was updated using a
structured literature assessment. Systemic therapy
for metastatic disease: this section was updated by a
systematic review.
8. F
 ollow-up in RCC & Surveillance following radical This chapter was updated by a narrative review,
or partial nephrectomy or ablative therapies based on a structured literature assessment. The
findings of a prospective database set up by the RCC
Panel have been included [10, 11].

Additional methodology information can be found in the general Methodology section of this print, and online
at the EAU website: http://uroweb.org/guidelines/. A list of Associations endorsing the EAU Guidelines can also
be viewed online at the above address.

2.2 Review
All publications ensuing from systematic reviews have been peer reviewed. The 2021 print of the RCC
Guidelines was peer-reviewed prior to publication.

2.3 Future goals


For their future updates, the RCC Guideline Panel aims to focus on patient-reported outcomes.
The use of clinical quality indicators is an area of interest for the RCC Panel. A number of key quality
indicators for this patient group have been selected:
• thorax computed tomography (CT) for staging of pulmonary metastasis;
• proportion of patients with T1aN0M0 tumours undergoing nephron-sparing surgery (NSS) as first treatment;
• the proportion of patients treated within six weeks after diagnosis;
• the proportion of patients with metastatic RCC (mRCC) offered systemic therapy;
• the proportion of patients who undergo minimally invasive or operative treatment as first treatment who
die within 30 days.

RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021 11


The Panel have set up a database to investigate current practice in follow-up of RCC patients in a number of
European centres. Assessing patterns of recurrence and use of imaging techniques are primary outcomes for
this project.

The results of ongoing and new systematic reviews will be included in the 2022 update of the RCC Guidelines:
• What is the best treatment option for ≥ T2 tumours?
• Adjuvant targeted therapy for renal cell carcinoma at high risk for recurrence;
• Systematic review of prevalence of intraperitoneal recurrences following robotic/laparoscopic partial
nephrectomy;
• Systematic review of individual, unit and hospital surgical volume for radical and partial nephrectomy and
their impact on outcomes;
• RECUR database analysis of recurrent disease/follow-up.

3. EPIDEMIOLOGY, AETIOLOGY AND


PATHOLOGY
3.1 Epidemiology
Renal cell carcinoma represents around 3% of all cancers, with the highest incidence occurring in Western
countries [12, 13]. In Europe and worldwide the highest incidence rates are found in the Czech Republic and
Lithuania [13]. Generally, during the last two decades until recently, there has been an annual increase of about
2% in incidence both worldwide and in Europe leading to approximately 99,200 new RCC cases and 39,100
kidney cancer-related deaths within the European Union in 2018 [12, 13]. In Europe, overall mortality rates
for RCC increased until the early 1990s, with rates generally stabilising or declining thereafter [14]. There has
been a decrease in mortality since the 1980s in Scandinavian countries and since the early 1990s in France,
Germany, Austria, the Netherlands, and Italy. However, in some European countries (Croatia, Estonia, Greece,
Ireland, Slovakia), mortality rates still show an upward trend [12, 13].
Renal cell carcinoma is the most common solid lesion within the kidney and accounts
for approximately 90% of all kidney malignancies. It comprises different RCC subtypes with specific
histopathological and genetic characteristics [15]. There is a 1.5:1 predominance in men over women with a
higher incidence in the older population [13, 16].

3.2 Aetiology
Established risk factors include lifestyle factors such as smoking, obesity, and hypertension [13, 16]. In a recent
systematic review also diabetes was found to be detrimental [17]. Having a first-degree relative with kidney
cancer is also associated with an increased risk of RCC. A number of other factors have been suggested to
be associated with higher or lower risk of RCC, including specific dietary habits and occupational exposure
to specific carcinogens, but the literature is inconclusive [16]. Moderate alcohol consumption appears to have
a protective effect for reasons as yet unknown, while also any physical activity level seems to have a small
protective effect [13, 17]. The most effective prophylaxis is to avoid cigarette smoking and reduce obesity [13, 16].

3.2.1 Summary of evidence and recommendation for epidemiology, aetiology and pathology

Summary of evidence LE
Several verified risk factors have been identified including smoking, obesity and hypertension. These 2a
are considered definite risk factors for RCC.

Recommendation Strength rating


Increase physical activity, eliminate cigarette smoking and in obese patients reduce weight Strong
are the primary preventative measures to decrease risk of RCC.

3.3 Histological diagnosis


Renal cell carcinomas comprise a broad spectrum of histopathological entities described in the 2016 World
Health Organization (WHO) classification [15]. There are three main RCC types: clear cell (ccRCC), papillary
(pRCC type I and II) and chromophobe (chRCC). The RCC type classification has been confirmed by

12 RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021


cytogenetic and genetic analyses [15, 18] (LE: 2b). Collecting duct carcinoma and other rare renal tumours are
discussed in Section 3.3.
Histological diagnosis includes, besides RCC type; evaluation of nuclear grade, sarcomatoid
features, vascular invasion, tumour necrosis, and invasion of the collecting system and peri-renal fat, pT, or
even pN categories. The four-tiered WHO/ISUP (International Society of Urological Pathology) grading system
has replaced the Fuhrman grading system [15].

3.3.1 Clear-cell RCC


Overall, clear-cell RCC (ccRCC) is well circumscribed and a capsule is usually absent. The cut surface is
golden-yellow, often with haemorrhage and necrosis. Loss of chromosome 3p and mutation of the von Hippel-
Lindau (VHL) gene at chromosome 3p25 are frequently found. The loss of von Hippel-Lindau protein function
contributes to tumour initiation, progression, and metastases. The 3p locus harbours at least four additional
cRCC tumour supressor genes (UTX, JARID1C, SETD2, PBRM1) [18]. In general, ccRCC has a worse
prognosis compared to pRCC and chRCC, but this difference disappears after adjustment for stage and grade
[19, 20]. For details about prognosis, see Section 6.3 - Histological factors.

3.3.2 Papillary RCC


Papillary RCC is the second most commonly encountered morphotype of RCC. Papillary RCC has traditionally
been subdivided into two types [15]. Type I and II pRCC, which were shown to be clinically and biologically
distinct; pRCC type I is associated with activating germline mutations of MET and pRCC type II is associated
with activation of the NRF2-ARE pathway and at least three subtypes [21]. Type II pRCC presents a
heterogenous group of tumours and future substratification is expected, e.g. oncocytic pRCC [15].
A typical histology of pRCC type I (narrow papillae without any binding, and only microcapillaries
in papillae) explains its typical clinical signs. Narrow papillae without any binding and a tough pseudocapsule
explain the ideal rounded shape (Pascal’s law) and fragility (specimens have a “minced meat” structure).
Tumour growth causes necrotisation of papillae, which is a source of hyperosmotic proteins that cause
subsequent “growth” of the tumour, fluid inside the tumour, and only a serpiginous, contrast-enhancing margin.
Stagnation in the microcapillaries explain the minimal post-contrast attenuation on CT. Papillary RCC type 1
can imitate a pathologically changed cyst (Bosniak IIF or III). The typical signs of pRCC type 1 are as follows:
an ochre colour, more frequently exophytic, extrarenal growth, low grade, and low malignant potential; over
75% of these tumours can be treated by NSS surgery. A substantial risk of renal tumour biopsy tract seeding
exists (12.5%), probably due to the fragility of the tumour papillae [22]. Papillary RCC type I is more common
and generally considered to have a better prognosis than pRCC type II [15, 20, 23].

3.3.3 Chromophobe RCC


Overall, chRCC is a pale tan, relatively homogenous and tough, well-demarcated mass without a capsule.
Chromophobe RCC cannot be graded by the Fuhrman grading system because of its innate
nuclear atypia. An alternative grading system has been proposed, but has yet to be validated [15]. Loss of
chromosomes Y, 1, 2, 6, 10, 13, 17 and 21 are typical genetic changes [15]. The prognosis is relatively good,
with high 5-year recurrence-free survival (RFS), and 10-year CSS [24]. The new WHO/ISUP grading system
merges former entity ‘hybrid oncocytic chromophobe tumour’ with chRCC.

3.4 Other renal tumours


Other renal tumours constitute the remaining renal cortical tumours. These include a variety of uncommon,
sporadic, and familial carcinomas, some only recently described, as well as a group of unclassified carcinomas.
A summary of these tumours is provided in Table 3.1, but some clinically relevant tumours and extremely rare
entities are mentioned below.

3.4.1 Renal medullary carcinoma


Renal medullary carcinoma (RMC) is a very rare tumour, comprising < 0.5% of all RCCs [25], predominantly
diagnosed in young adults (median age 28 years) with sickle haemoglobinopathies (including sickle cell trait).
It is mainly centrally located with ill-defined borders. Renal medullary carcinoma is one of the most aggressive
RCCs [26, 27] and most patients (~67%) will present with metastatic disease [26, 28]. Even patients who
present with seemingly localised disease may develop macrometastases shortly thereafter, often within a few
weeks.

3.4.1.1 Treatment of renal medullary carcinoma


Despite treatment, median OS is 13 months in the most recent series [26]. Due to the infiltrative nature and
medullary epicentre of RMC, radical nephrectomy (RN) is favoured over PN even in very early-stage disease.
Retrospective data indicate that nephrectomy in localised disease results in superior OS (16.4 vs. 7 months)

RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021 13


compared with systemic chemotherapy alone, but longer survival was noted in patients who achieved an
objective response to first-line chemotherapy. [26, 29]. There is currently no established role for distant
metastasectomy or nephrectomy in the presence of metastases.
Palliative radiation therapy is an option and may achieve regression in the targeted areas but
it will not prevent progression outside the radiation field [30, 31]. Renal medullary carcinoma is refractory
to monotherapies with targeted anti-angiogenic regimens including tyrosine kinase inhibitors (TKIs) and
mammalian target of rapamycin (mTOR) inhibitors [26, 32]. The mainstay systemic treatments for RMC are
cytotoxic combination regimens which produce partial or complete responses in ~29% of patients [32]. There
are no prospective comparisons between different chemotherapy regimens but most published series used
various combinations of platinum agents, taxanes, gemcitabine, and/or anthracyclines [26, 27]. High-dose-
intensity combination of methotrexate, vinblastine, doxorubicin, and cisplatin (MVAC) has also shown efficacy
against RMC [33] although a retrospective comparison did not show superiority of MVAC over cisplatin,
paclitaxel, and gemcitabine [27]. Single-agent anti-PD-1 (monoclonal antibodies against programmed death-1)
immune checkpoint therapy has produced responses in a few case reports, although, as yet, insufficient
data are available to determine the response rate to this approach [30, 31]. Whenever possible, patients
should be enrolled in clinical trials of novel therapeutic approaches, particularly after failing first-line cytotoxic
chemotherapy.

3.4.2 Carcinoma associated with end-stage renal disease; acquired cystic disease-associated RCC
Cystic degenerative changes (acquired cystic kidney disease [ACKD]) and a higher incidence of RCC, are
typical features of end-stage renal disease (ESRD). Renal cell carcinomas of native end-stage kidneys are
found in approximately 4% of patients. Their lifetime risk of developing RCCs is at least ten times higher than in
the general population. Compared with sporadic RCCs, RCCs associated with ESRD are generally multicentric
and bilateral, found in younger patients (mostly male), and are less aggressive. Whether the relatively indolent
outcome of tumours in ESRD is due to the mode of diagnosis or a specific ACKD-related molecular pathway
still has to be determined. Although the histological spectrum of ESRD tumours is similar to that of sporadic
RCC; pRCC occur relatively more frequently [34, 35]. A specific subtype of RCC occurring only in end-stage
kidneys has been described as Acquired Cystic Disease-associated RCC (ACD-RCC) with indolent clinical
behaviour, likely due to early detection in patients with ESRD on periodic follow-up [15, 18, 36].

3.4.3 Papillary adenoma


These tumours have a papillary or tubular architecture of low nuclear grade and may be up to 15 mm in
diameter, or smaller [37], according to the WHO 2016 classification [15].

3.4.4 Hereditary kidney tumours


Five to eight percent of RCCs are hereditary; to date there are ten hereditary RCC syndromes associated
with specific germline mutations, RCC histology, and comorbidities. Hereditary RCC syndromes are often
suggested by family history, age of onset and presence of other lesions typical for the respective syndromes.
Median age for hereditary RCC is 37 years; 70% of hereditary RCC tumours are found in the lowest decile
(46 years old) of all RCC tumours [38]. Hereditary kidney tumours are found in the following entities:
VHL syndrome, hereditary pRCC, Birt-Hogg-Dube syndrome, hereditary leiomyomatosis and RCC (HLRCC),
tuberous sclerosis, germline succinate dehydrogenase (SDH) mutation, non-polyposis colorectal cancer
syndrome, hyperparathyroidism-jaw tumour syndrome, phosphatase and tensin homolog (PTEN) hamartoma
syndrome (PHTS), constitutional chromosome 3 translocation, and familial non-syndromic ccRCC. Renal
medullary carcinoma can be included because of its association with hereditary haemoglobinopathies
[37, 39-41].

Patients with hereditary kidney cancer syndromes may require repeated surgical intervention [42, 43]. In
most hereditary RCCs nephron-sparing approaches are recommended. The exceptions are HLRCC and SDH
syndromes for which immediate surgical intervention is recommended due to the aggressive nature of this
lesion. For other hereditary syndromes such as VHL, surveillance is recommended until the largest tumour
reaches 3 cm in diameter, to reduce interventions [44]. Active surveillance (AS) for VHL, BDH and HPRCC
should, in individual patients, follow the growth kinetics, size and location of the tumours, rather than apply
a standardised follow-up interval. Regular screening for both renal and extra-renal lesions should follow
international guidelines for these syndromes. Multidisciplinary and co-ordinated care should be offered, where
appropriate [45].
Although not hereditary, somatic fusion translocations of TFE3 and TFEB may affect 15% of
patients with RCC younger than 45 years and 20-45% of children and young adults diagnosed with RCC [46].
A recent phase II trial demonstrated clinical activity of an oral HIF-2α (hypoxia-inducible factor)
inhibitor MK-6482 in VHL patients. In VHL-associated RCC the objective response rate (ORR) was 28%

14 RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021


and stable disease rate was observed in 71% of 61 evaluated patients with a median decline of the linear
growth rate by -6.4 mm (range 23.3–4.5) per year. Eighty seven per cent of patients showed a decrease from
baseline in target lesions as evaluated by independent review, which, despite a follow-up of 36 weeks only, are
promising but, as yet, unvalidated results [47].

Genetic counselling is suggested for younger patients, in case of bilateral and multiple tumours, a past family
history of RCC and uncommon morphology.

3.4.5 Angiomyolipoma
Angiomyolipoma (AML) is a benign mesenchymal tumour, which can occur sporadically or as part of tuberous
sclerosis complex [48]. Overall prevalence is 0.44%, with 0.6% in female and 0.3% in male populations.
Only 5% of these patients present with multiple AMLs [49]. Angiomyolipoma belongs to a family of so-called
PEComas (perivascular epithelioid cell tumours), characterised by the proliferation of perivascular epithelioid
cells. Some PEComas can behave aggressively and can even produce distant metastases. Classic AMLs are
completely benign [15, 37, 50]. Ultrasound (US), CT, and magnetic resonance imaging (MRI) often lead to the
diagnosis of AMLs due to the presence of adipose tissue, however in fat poor AML, diagnostic imaging cannot
reliably identify these lesions. Percutaneous biopsy is rarely useful. Renal tumours that cannot be clearly
identified as benign during the initial diagnostic work-up should be treated according to the recommendations
provided for the treatment of RCC in these Guidelines. In tuberous sclerosis, AML can be found in enlarged
lymph nodes (LNs), which does not represent metastatic spread but a multicentric spread of AMLs. In rare
cases, an extension of a non-malignant thrombus into the renal vein or inferior vena cava can be found,
associated with an angiotrophic-type growth of AML. Epithelioid AML, a very rare variant of AML, consists of at
least 80% epithelioid cells [37, 50]. Epithelioid AMLs are potentially malignant with a highly variable proportion
of cases with aggressive behaviour [51]. Criteria to predict the biological behaviour in epithelioid AML were
proposed by the WHO 2016 [37, 50]. Angiomyolipoma, in general, has a slow and consistent growth rate, and
minimal morbidity [5].
In some cases, larger AMLs can cause local pain. The main complication of AMLs is spontaneous
bleeding in the retroperitoneum or into the collecting system, which can be life threatening. Bleeding is caused
by spontaneous rupture of the tumour. Little is known about the risk factors for bleeding, but it is believed
to increase with tumour size and may be related to the angiogenic component of the tumour that includes
irregular blood vessels [5]. The major risk factors for bleeding are tumour size, grade of the angiogenic
component, and the presence of tuberous sclerosis [52, 53].

3.4.5.1 Treatment
Active surveillance is the most appropriate option for most AMLs (48%). In a group of patients on AS, only 11%
of AMLs showed growth, and spontaneous bleeding was reported in 2%, resulting in active treatment in 5%
of patients [5, 54] (LE: 3). The association between AML size and the risk of bleeding remains unclear and the
traditionally used 4-cm cut-off should not per se trigger active treatment [5]. When surgery is indicated, NSS is
the preferred option, if technically feasible. Main disadvantages of less invasive selective arterial embolisation
(SAE) are more recurrences and a need for secondary treatment (0.85% for surgery vs. 31% for SAE). For
thermal ablation only limited data are available, and this option is used less frequently [5].
Active treatment (SAE, surgery or ablation) should be instigated in case of persistent pain, ruptured
AML (acute or repeated bleeding) or in case of a very large AML. Specific patient circumstances may influence
the choice to offer active treatment; such as patients at high risk of abdominal trauma, females of childbearing
age or patients in whom follow-up or access to emergency care may be inadequate. Selective arterial
embolisation is an option in case of life-threatening AML bleeding.
In patients diagnosed with tuberous sclerosis, size reduction of often bilateral AMLs can be induced
by inhibiting the mTOR pathway using everolimus, as demonstrated in RCTs [55, 56]. In a small phase II trial
(n = 20), efficacy of everolimus was demonstrated in sporadic AML as well. A 25% or greater reduction in
tumour volume at 4 and 6 months was demonstrated in 55.6% and 71.4% of patients, respectively. Twenty per
cent of patients were withdrawn due to toxicities and 40% self-withdrew from the study due to side effects [57].

3.4.6 Renal oncocytoma


Oncocytoma is a benign tumour representing 3–7% of all solid renal tumours and its incidence increases to
18% when tumours < 4 cm are considered [15, 54]. The diagnostic accuracy of imaging modalities (CT, MRI) in
renal oncocytoma is limited and histopathology remains the only reliable diagnostic modality [15, 54]. Standard
treatment for renal oncocytoma is similar to that of other renal tumours; surgical excision by partial- or RN
with subsequent histopathological verification. However, due to the inability of modern imaging techniques
to differentiate benign from malignant renal masses, there is a renewed interest in renal mass biopsy (RMB)
prior to surgical intervention. Accuracy of the biopsy and management of advanced/progressing oncocytomas

RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021 15


need to be considered in this context since oncocytic renal neoplasms diagnosed by RMB at histological
examination after surgery showed oncocytoma in only 64.6% of cases. The remainder of the tumours were
mainly chRCC (18.7% including 6.3% hybrid oncocytic/chromophobe tumours which have now been grouped
histologically with chRCC) [15], other RCCs (12.5%), and other benign lesions (4.2%) [58]. The majority of
oncocytomas slowly progress in size with an annual growth rate < 14 mm [59-61]. Preliminary data show
that AS may be a safe option to manage oncocytoma in appropriately selected patients. Potential triggers to
change management of patients on AS are not well defined [62].

Table 3.1: Other renal cortical tumours, and recommendations for treatment (strength rating: weak) [15]

Entity Clinical relevant notes Malignant potential Treatment


Sarcomatoid variants Sign of high-grade High Surgery. Nivolumab and
of RCC transformation without being a ipilimumab. Sunitinib,
distinct histological entity. gemcitabine plus
doxorubicin is also an
option [63].
Multilocular cystic Formerly multilocular cystic Benign Nephron-sparing
renal neoplasm of low RCC surgery (NSS).
malignant potential
Carcinoma of the Rare, often presenting at an High, very aggressive. Surgery. Response to
collecting ducts of advanced stage (N+ 44% and Median survival is targeted therapies is
Bellini M1, 33% at diagnosis). The 30 months [64]. poor [65].
hazard ratio (HR) for CSS in
comparison with ccRCC is 4.49
[20].
Renal medullary Very rare. Mainly young black High, very aggressive, Surgery. Different
carcinoma men with sickle cell trait. median survival is five chemotherapy
months [64]. regimens,
radiosensitive.
Translocation RCC Rare, mainly younger patients High Surgery. Vascular
(TRCC) Xp11.2 < 40, more common in endothelial growth
females. Less commonly, TFEB factor (VEGF)-targeted
located on the short arm of therapy.
Translocation RCC chromosome 6 (6p21) [66]. Low/intermediate Surgery, NSS. VEGF-
t(6;11) targeted therapy.
Mucinous tubular and Tumour is associated with the Intermediate Surgery, NSS.
spindle cell carcinoma loop of Henle.
Acquired cystic disease- Low Surgery, NSS.
associated RCC
Clear-cell papillary RCC Also reported as renal Low Surgery, NSS.
angiomyomatous tumour (RAT).
Hereditary Rare, germline mutation of the High Surgery.
leiomyomatosis and fumarate hydratase gene [15]. No data about treatment
RCC-associated RCC 21% lifetime risk of RCC [67]. of metastatic disease.
Imaging screening is
recommended [67].
Tubulocystic RCC Mainly men, imaging can show Low (90% indolent) Surgery, NSS.
Bosniak III or IV.
Succinate Rare. Variable Surgery, NSS.
dehydrogenase-
deficient RCC
Metanephric tumours Divided into metanephric Benign Surgery, NSS.
adenoma, adenofibroma, and
metanephric stromal tumours.
Cystic nephroma/Mixed The term renal epithelial and Low/benign Surgery, NSS.
epithelial and stromal stromal tumours (REST) is used
tumour as well. Imaging – Bosniak type
III or II/IV.

16 RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021


Oncocytoma 3-7% of all renal tumours. Benign Observation (when
Imaging characteristics histologically
alone are unreliable when confirmed).
differentiating between NSS. See Section 3.4.6.
oncocytoma and RCC.
Histopathological diagnosis
remains the reference standard.
Renal cysts Simple cysts are frequently Malignant or benign Treatment or follow-
occurring, while occurring up recommendation
septa, calcifications and solid based on Bosniak
components require follow-up classification.
and/or management. See Table 5.1

3.4.7 Cystic renal tumours


Cystic renal lesions are classified according to the Bosniak classification (see Section 5.2.5). Bosniak I and II
cysts are benign lesions which do not require follow up [68]. Bosniak IV cysts are mostly malignant tumours
with pseudocystic changes only. Bosniak IIF and III cysts remain challenging for clinicians. The differentiation of
benign and malignant tumour in categories IIF/III is based on imaging, mostly CT, with an increasing role of MRI
and contrast enhanced ultrasound (CEUS). Computed tomography shows poor sensitivity (36%) and specificity
(76%; κ [kappa coefficient] = 0.11) compared with 71% sensitivity and 91% specificity (κ = 0.64) for MRI
and 100% sensitivity and 97% specificity for CEUS (κ = 0.95) [69]. Surgical and radiological cohorts pooled
estimates show a prevalence of malignancy of 0.51 (0.44–0.58) in Bosniak III and 0.89 (0.83–0.92) in Bosniak IV
cysts, respectively. In a systematic review, less than 1% of stable Bosniak IIF cysts showed malignancy during
follow-up. Twelve percent of Bosniak IIF cysts had to be reclassified to Bosniak III/IV during radiological follow-up,
with 85% of these showing malignancy, which is comparable to the malignancy rates of Bosniak IV cysts [68].
The updated Bosniak classification strengthens the classification and includes also MRI diagnostic criteria [70].

The most common histological type for Bosniak III cysts is ccRCC with pseudocystic changes and low
malignant potential [71, 72]; multilocular cystic renal neoplasm of low malignant potential ([MCRNLMP],
formerly mcRCC (see Section 3.2 and Table 3.1); pRCC type I (very low malignant potential); benign
multilocular cyst; benign group of renal epithelial and stromal tumours (REST); and other rare entities. Surgery
in Bosniak III cysts will result in over-treatment in 49% of the tumours which are lesions with a low malignant
potential. In view of the excellent outcome of these patients in general, a surveillance approach is an alternative
to surgical treatment [68, 70, 73, 74].

3.5 Summary of evidence and recommendations for the management of other renal
tumours

Summary of evidence LE
A variety of renal tumours exist of which approximately 15% are benign. 1b
Recent histological work up of Bosniak III cysts shows low risk of malignant potential. 2

3.6 Recommendations for the management of other renal tumours

Recommendations Strength rating


Manage Bosniak type III cysts the same as localised RCC, or offer active surveillance. Weak
Manage Bosniak type IV cysts the same as localised RCC. Strong
Treat angiomyolipoma (AML) with selective arterial embolisation or nephron-sparing surgery, in: Weak
• large tumours (a recommended threshold of intervention does not exist);
• females of childbearing age;
• patients in whom follow-up or access to emergency care may be inadequate;
• persistent pain or acute or repeated bleeding episodes.
Offer systemic therapy to patients in need of therapy with surgically unresectable AMLs not Weak
amendable to embolisation or surgery.
Offer active surveillance to patients with biopsy-proven oncocytomas, as an acceptable Weak
alternative to surgery or ablation.
Offer radical nephrectomy to patients with localised renal medullary carcinoma. Weak
Base systemic therapy for renal medullary carcinoma on chemotherapy regimens containing Weak
cisplatinum such as cisplatin plus gemcitabine.

RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021 17


4. STAGING AND CLASSIFICATION SYSTEMS
4.1 Staging
The Tumour Node Metastasis (TNM) classification system is recommended for clinical and scientific use
[75], but requires continuous re-assessment [15, 76]. A supplement was published in 2012, and the latter’s
prognostic value was confirmed in single and multi-institution studies [77, 78]. Tumour size, venous invasion,
renal capsular invasion, adrenal involvement, and LN and distant metastasis are included in the TNM
classification system (Table 4.1). However, some uncertainties remain:
• The sub-classification of T1 tumours using a cut-off of 4 cm might not be optimal in NSS for localised cancer.
• The value of size stratification of T2 tumours has been questioned [79].
• Renal sinus fat invasion might carry a worse prognosis than perinephric fat invasion, but, is nevertheless
included in the same pT3a stage group [80-82] (LE: 3).
• Sub T-stages (pT2b, pT3a, pT3c and pT4) may overlap [78].
• For adequate M staging, accurate pre-operative imaging (chest and abdominal CT) should be performed
[83, 84] (LE: 4).

Table 4.1: 2017 TNM classification system [75]

T - Primary tumour
TX Primary tumour cannot be assessed
T0 No evidence of primary tumour
T1 Tumour < 7 cm or less in greatest dimension, limited to the kidney
T1a Tumour < 4 cm or less
T1b Tumour > 4 cm but < 7 cm
T2 Tumour > 7 cm in greatest dimension, limited to the kidney
T2a Tumour > 7 cm but < 10 cm
T2b Tumours > 10 cm, limited to the kidney
T3 Tumour extends into major veins or perinephric tissues but not into the ipsilateral adrenal gland and
not beyond Gerota fascia
T3a Tumour extends into the renal vein or its segmental branches, or invades the pelvicalyceal
system or invades perirenal and/or renal sinus fat, but not beyond Gerota fascia*
T3b Tumour grossly extends into the vena cava below diaphragm
T3c Tumour grossly extends into vena cava above the diaphragm or invades the wall of the vena cava
T4 Tumour invades beyond Gerota fascia (including contiguous extension into the ipsilateral adrenal gland)
N - Regional Lymph Nodes
NX Regional lymph nodes cannot be assessed
N0 No regional lymph node metastasis
N1 Metastasis in regional lymph node(s)
M - Distant Metastasis
M0 No distant metastasis
M1 Distant metastasis
pTNM stage grouping
Stage I T1 N0 M0
Stage II T2 N0 M0
Stage III T3 N0 M0
T1, T2, T3 N1 M0
Stage IV T4 Any N M0
Any T Any N M1
A help desk for specific questions about TNM classification is available at http://www.uicc.org/tnm.
*Adapted based on the American Joint Committee on Cancer (AJCC), 8th Edn. 2017 [85].

4.2 Anatomic classification systems


Objective anatomic classification systems, such as the Preoperative Aspects and Dimensions Used for an
Anatomical (PADUA) classification system, the R.E.N.A.L. nephrometry score, the C-index, an Arterial Based
Complexity (ABC) Scoring System and Zonal NePhRO scoring system, have been proposed to standardise the
description of renal tumours [86-88]. These systems include assessment of tumour size, exophytic/endophytic
properties, proximity to the collecting system and renal sinus, and anterior/posterior or lower/upper pole
location.

18 RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021


The use of such a system is helpful as it allows objective prediction of potential morbidity of NSS
and tumour ablation techniques. These tools provide information for treatment planning, patient counselling,
and comparison of PN and tumour ablation series. However, when selecting the most optimal treatment option,
anatomic scores must be considered together with patient features and surgeon experience.

5. DIAGNOSTIC EVALUATION
5.1 Symptoms
Many renal masses remain asymptomatic until the late disease stages. More than 50% of RCCs are detected
incidentally by non-invasive imaging investigating various non-specific symptoms and other abdominal
diseases [78, 89] (LE: 3). The classic triad of flank pain, visible haematuria, and palpable abdominal
mass is rare (6–10%) and correlates with aggressive histology and advanced disease [90, 91] (LE: 3).
Paraneoplastic syndromes are found in approximately 30% of patients with symptomatic RCCs [92] (LE: 4).
Some symptomatic patients present with symptoms caused by metastatic disease, such as bone pain or
persistent cough [93] (LE: 3).

5.1.1 Physical examination


Physical examination has a limited role in RCC diagnosis. However, the following findings should prompt
radiological examinations:
• palpable abdominal mass;
• palpable cervical lymphadenopathy;
• non-reducing varicocele and bilateral lower extremity oedema, which suggests venous involvement.

5.1.2 Laboratory findings


Commonly assessed laboratory parameters are serum creatinine, glomerular filtration rate (GFR), complete cell
blood count, erythrocyte sedimentation rate, liver function study, alkaline phosphatase, lactate dehydrogenase
(LDH), serum corrected calcium [94], coagulation study, and urinalysis (LE: 4). For central renal masses abutting
or invading the collecting system, urinary cytology and possibly endoscopic assessment should be considered
in order to exclude urothelial cancer (LE: 4).
Split renal function should be estimated using renal scintigraphy in the following situations [95, 96]
(LE: 2b):
• when renal function is compromised, as indicated by increased serum creatinine or significantly
decreased GFR;
• when renal function is clinically important; e.g., in patients with a solitary kidney or multiple or bilateral
tumours.

Renal scintigraphy is an additional diagnostic option in patients at risk of future renal impairment due to
comorbid disorders.

5.2 Imaging investigations


Most renal tumours are diagnosed by abdominal US or CT performed for other medical reasons [89] (LE: 3).
Renal masses are classified as solid or cystic based on imaging findings.

5.2.1 Presence of enhancement


With solid renal masses, the most important criterion for differentiating malignant lesions is the presence
of enhancement [97] (LE: 3). Traditionally, US, CT and MRI are used for detecting and characterising renal
masses. Most renal masses are diagnosed accurately by imaging alone.

5.2.2 Computed tomography or magnetic resonance imaging


Computed tomography or MRI are used to characterise renal masses. Imaging must be performed
unenhanced, in an early arterial phase, and in a parchymal phase with intravenous contrast material to
demonstrate enhancement. In CT imaging, enhancement in renal masses is determined by comparing
Hounsfield units (HUs) before, and after, contrast administration. A change of fifteen HU, or more, in the solid
tumour parts demonstrates enhancement and thus vital tumour parts [98] (LE: 3). Computed tomography or
MRI allows accurate diagnosis of RCC, but cannot reliably distinguish oncocytoma and fat-free AML from

RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021 19


malignant renal neoplasms [99-102] (LE: 3). Abdominal CT provides information on [103]:
• function and morphology of the contralateral kidney [104] (LE: 3);
• primary tumour extension;
• venous involvement;
• enlargement of locoregional LNs;
• condition of the adrenal glands and other solid organs (LE: 3).

Abdominal contrast-enhanced CT angiography is useful in selected cases when detailed information on the
renal vascular supply is needed [105, 106]. If the results of CT are indeterminate, CEUS is a valuable alternative
to further characterise renal lesions [6, 107-109] (LE: 1b).

Magnetic resonance imaging may provide additional information on venous involvement if the extent of an
inferior vena cava (IVC) tumour thrombus is poorly defined on CT [110-113] (LE: 3). In MRI, especially high-
resolution T2 weighted images provide a superior delineation of the uppermost tumour thrombus, as the inflow
of the enhanced blood may be reduced due to extensive occlusive tumour thrombus growth in the inferior vena
cava. The T2 weighted images with its intrinsic contrast allows a good delineation [113].
Magnetic resonance imaging is indicated in patients who are allergic to intravenous CT contrast
medium and in pregnancy without renal failure [113, 114] (LE: 3). Magnetic resonance imaging allows the
evaluation of a dynamic enhancement without radiation exposure. Advanced MRI techniques such as
diffusion-weighted (DWI) and perfusion-weighted imaging are being explored for renal mass assessment [115].
Recently, the use of multiparametric MRI (mpMRI) to diagnose ccRCC via a clear cell likelihood score (ccLS)
in small renal masses was reported [116]. The ccLS is a 5-tier classification that denotes the likelihood of a
mass representing ccRCC, ranging from ‘very unlikely’ to ‘very likely’. The authors prospectively validated
the diagnostic performance of ccLS in 57 patients with cT1a tumours and found a high diagnostic accuracy.
The diagnostic performance of mpMRI-based ccLS was further validated in a larger retrospective cohort
(n = 434) across all tumour sizes and stages [117], and ccLS was found to be an independent prognostic factor
for identifying ccRCC. The system is promising and deserves further validation.

For the diagnosis of complex renal cysts (Bosniak IIF-III) MRI may be preferable. The accuracy of CT is limited
in these cases, with poor sensitivity (36%) and specificity (76%; κ = 0.11); MRI, due to a higher sensitivity for
enhancement, showed a 71% sensitivity and 91% specificity (κ = 0.64). Contrast-enhanced US showed high
sensitivity (100%) and specificity (97%), with a negative predictive value of 100% (κ = 0.95) [69].
In younger patients who are worried about the radiation exposure of frequent CT scans, MRI may
be offered as alternative although only limited data exist correlating diagnostic radiation exposure to the
development of secondary cancers [118].

5.2.3 Other investigations


Renal arteriography and inferior venacavography have a limited role in the work-up of selected RCC patients
(LE: 3). In patients with any sign of impaired renal function, an isotope renogram and total renal function
evaluation should be considered to optimise treatment decision-making [95, 96] (LE: 2a). Positron-emission
tomography (PET) is not recommended [6, 119] (LE: 1b).

5.2.4 Radiographic investigations to evaluate RCC metastases


Chest CT is accurate for chest staging [83, 84, 120-122] (LE: 3). Use of nomograms to calculate risk of lung
metastases have been proposed based on tumour size, clinical stage and presence of systemic symptoms
[123, 124]. These are based on large, retrospective datasets, and suggest that chest CT may be omitted in
patients with cT1a and cN0, and without systemic symptoms, anaemia or thrombocythemia, due to the low
incidence of lung metastases (< 1%) in this group of patients. There is a consensus that most bone metastases
are symptomatic at diagnosis; thus, routine bone imaging is not generally indicated [120, 125, 126] (LE: 3).
However, bone scan, brain CT, or MRI may be used in the presence of specific clinical or laboratory signs and
symptoms [125, 127, 128] (LE: 3). A recent prospective comparative blinded study involving 92 consecutive
mRCC patients treated with first-line VEGFR-TKI (median follow-up 35 months) found that whole-body DWI/
MRI detected a statistically significant higher number of bony metastases compared with conventional
thoraco-abdomino-pelvic contrast-enhanced CT, with higher number of metastases being an independent
prognostic factor for progression-free survival (PFS) and OS [129].

5.2.5 Bosniak classification of renal cystic masses


This system classifies renal cysts into five categories, based on CT imaging appearance, to predict malignancy
risk [130, 131] (LE: 3), and also advocates treatment for each category (Table 5.1). A new updated Bosniak
classification has been proposed that strengthens the classification and includes MRI diagnostic criteria

20 RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021


[70]; however, it requires further validation. The management of cystic renal tumours is also discussed in
Section 3.4.7.

Table 5.1: Bosniak classification of renal cysts [130]

Bosniak Features Work-up


category
I Simple benign cyst with a hairline-thin wall without septa, calcification, Benign
or solid components. Same density as water and does not enhance with
contrast medium.
II Benign cyst that may contain a few hairline-thin septa. Fine calcification Benign
may be present in the wall or septa. Uniformly high-attenuation lesions
< 3 cm in size, with sharp margins without enhancement.
IIF These may contain more hairline-thin septa. Minimal enhancement of Follow-up, up to
a hairline-thin septum or wall. Minimal thickening of the septa or wall. five years. Some are
The cyst may contain calcification, which may be nodular and thick, malignant.
with no contrast enhancement. No enhancing soft-tissue elements.
This category also includes totally intra-renal, non-enhancing, high
attenuation renal lesions ≥ 3 cm. Generally well-marginated.
III These are indeterminate cystic masses with thickened irregular walls or Surgery or active
septa with enhancement. surveillance – see
Chapter 7. Over 50%
are malignant.
IV Clearly malignant containing enhancing soft-tissue components. Surgery. Most are
malignant.

5.3 Renal tumour biopsy


5.3.1 Indications and rationale
Percutaneous renal tumour biopsy can reveal histology of radiologically indeterminate renal masses and can
be considered in patients who are candidates for AS of small masses, to obtain histology before ablative
treatments, and to select the most suitable medical and surgical treatment strategy in the setting of metastatic
disease [132-137] (LE: 3).
A multicentre study assessing 542 surgically removed small renal masses showed that the likelihood
of benign findings at pathology is significantly lower in centres where biopsies are performed (5% vs. 16%),
suggesting that biopsies can reduce surgery for benign tumours and the potential for short-term and long-term
morbidity associated with these procedures [138].

Renal biopsy is not indicated for comorbid and frail patients who can be considered only for conservative
management (watchful waiting) regardless of biopsy results. Due to the high diagnostic accuracy of abdominal
imaging, renal tumour biopsy is not necessary in patients with a contrast-enhancing renal mass for whom
surgery is planned (LE: 4).
Core biopsies of cystic renal masses have a lower diagnostic yield and accuracy and are not
recommended alone, unless areas with a solid pattern are present (Bosniak IV cysts) [132, 135, 139] (LE: 2b/3).

5.3.2 Technique
Percutaneous sampling can be performed under local anaesthesia with needle core biopsy and/or fine needle
aspiration (FNA). Biopsies can be performed under US or CT guidance, with a similar diagnostic yield [135,
140] (LE: 2b). Eighteen-gauge needles are ideal for core biopsies, as they result in low morbidity and provide
sufficient tissue for diagnosis [132, 136, 141] (LE: 2b). A coaxial technique allowing multiple biopsies through a
coaxial cannula should always be used to avoid potential tumour seeding [132, 136] (LE: 3).

Core biopsies are preferred for the characterisation of solid renal masses while a combination with FNA can
provide complimentary results and improve accuracy for complex cystic lesions [139, 142, 143] (LE: 2a). A
systematic review and meta-analysis of the diagnostic performance and complications of renal tumour biopsy
was performed by the Panel, including 57 publications and a total of 5,228 patients. Needle core biopsies were
found to have better accuracy for the diagnosis of malignancy compared with FNA [139]. Other studies showed
that solid pattern, larger tumour size and exophytic location are predictors of a diagnostic core biopsy [132,
135, 140] (LE: 2b).

RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021 21


5.3.3 Diagnostic yield and accuracy
In experienced centres, core biopsies have a high diagnostic yield, specificity, and sensitivity for the diagnosis
of malignancy. The above-mentioned meta-analysis showed that sensitivity and specificity of diagnostic core
biopsies for the diagnosis of malignancy are 99.1% and 99.7%, respectively [139] (LE: 2b). However, 0–22.6%
of core biopsies are non-diagnostic (8% in the meta-analysis) [133-137, 140, 141, 144] (LE: 2a). If a biopsy is
non-diagnostic, and radiologic findings are suspicious for malignancy, a further biopsy or surgical exploration
should be considered (LE: 4). Repeat biopsies have been reported to be diagnostic in a high proportion of
cases (83-100%) [132, 145-147].
Accuracy of renal tumour biopsies for the diagnosis of tumour histotype is good. The median
concordance rate between tumour histotype on renal tumour biopsy and on the surgical specimen of the
following PN or RN was 90.3% in the pooled analysis [139].
Assessment of tumour grade on core biopsies is challenging. In the pooled analysis the overall
accuracy for nuclear grading was poor (62.5%), but significantly improved (87%) using a simplified two-tier
system (high vs. low grade) [139] (LE: 2a).
The ideal number and location of core biopsies are not defined. However, at least two good quality
cores should be obtained and necrotic areas should be avoided to maximise diagnostic yield [132, 135,
148, 149] (LE: 2b). Peripheral biopsies are preferable for larger tumours, to avoid areas of central necrosis
[150] (LE: 2b). In cT2 or greater renal masses, multiple core biopsies taken from at least four separate solid
enhancing areas in the tumour were shown to achieve a higher diagnostic yield and a higher accuracy to identify
sarcomatoid features, without increasing the complication rate [151].

5.3.4 Morbidity
Overall, percutaneous biopsies have a low morbidity [139]. Tumour seeding along the needle tract has been
regarded as anecdotal in large series and pooled analyses on renal tumour biopsies. Especially the coaxial
technique has been regarded as a safe method to avoid any seeding of tumour cells. However, authors recently
reported on 7 patients in whom tumour seeding was identified on histological examination of the resection
specimen after surgical resection of RCC following diagnostic percutaneous biopsy [152]. Six of the 7 cases
were of the pRCC type. The clinical significance of these findings is still uncertain but only one of these patients
developed local tumour recurrence at the site of the previous biopsy [152].
Spontaneously resolving subcapsular/perinephric haematomas are reported in 4.3% of cases in
a pooled analysis, but clinically significant bleeding is unusual (0–1.4%; 0.7% in the pooled analysis) and
generally self-limiting [139].
Percutaneous biopsy of renal hilar masses is technically feasible with a diagnostic yield similar to that
of cortical masses, but with significantly higher post-procedural bleeding compared with cortical masses [153].

5.4 Summary of evidence and recommendations for the diagnostic assessment of RCC

Summary of evidence LE
Contrast enhanced multi-phasic CT has a high sensitivity and specificity for characterisation and 2a
detection of RCC, invasion, tumour thrombus and mRCC.
Magnetic resonance imaging has a slightly higher sensitivity and specificity for small cystic renal 2a
masses and tumour thrombi as compared to CT.
Contrast enhanced ultrasound has a high sensitivity and specificity for characterisation of renal masses. 2a
Renal mass biopsies are associated with reduced over-treatment of benign masses and offers patients 3
additional information (i.e. grade, subtype) for an informed decision regarding optimal management.
Ultrasound, power-Doppler US and positron-emission tomography CT have a low sensitivity and 2a
specificity for detection and characterisation of RCC.

Recommendations Strength rating


Use multi-phasic contrast-enhanced computed tomography (CT) of abdomen and chest for Strong
the diagnosis and staging of renal tumours.
Omit chest CT in patients with incidentally noted cT1a disease due to the low risk of lung Weak
metastases in this cohort.
Use magnetic resonance imaging (MRI) to better evaluate venous involvement, reduce Weak
radiation or avoid intravenous CT contrast medium.
Use non-ionising modalities, including MRI and contrast-enhanced ultrasound, for further Strong
characterisation of small renal masses, tumour thrombus and differentiation of unclear renal
masses, if the results of contrast-enhanced CT are indeterminate.

22 RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021


Do not routinely use bone scan and/or positron-emission tomography CT for staging of Weak
renal cell carcinoma.
Perform a renal tumour biopsy before ablative therapy and systemic therapy without Strong
previous pathology.
Perform a percutaneous biopsy in select patients who are considering active surveillance. Weak
Use a coaxial technique when performing a renal tumour biopsy. Strong
Do not perform a renal tumour biopsy of cystic renal masses. Strong
Use a core biopsy technique rather than fine needle aspiration for histological Strong
characterisation of solid renal tumours.

6. PROGNOSTIC FACTORS
6.1 Classification
Prognostic factors can be classified into: anatomical, histological, clinical, and molecular.

6.2 Anatomical factors


Tumour size, venous invasion and extension, collecting system invasion, perinephric- and sinus fat invasion,
adrenal involvement, and LN and distant metastasis are included in the TNM classification system [154, 155]
(Table 4.1).

6.3 Histological factors


Histological factors include tumour grade, RCC subtype, lymphovascular invasion, tumour necrosis, and
invasion of the collecting system [156, 157]. Tumour grade is considered one of the most important histological
prognostic factors. Fuhrman nuclear grade is based on simultaneous investigation of nuclear size, nuclear
shape and nucleolar prominence [158]. It has been the most widely accepted grading system for several
decades, but has now been largely replaced by the WHO/ISUP grading classification [159]. This relies solely
on nucleolar prominence for grade 1-3 tumours, allowing for less inter-observer variation [160]. It has been
shown that the WHO/ISUP grading provides superior prognostic information compared to Fuhrman grading,
especially for grade 2 and grade 3 tumours [161]. Rhabdoid and sarcomatoid changes can be found in all
RCC types and are equivalent to grade 4 tumours. Sarcomatoid changes are more often found in chRCC than
other subtypes [162]. The percentage of the sarcomatoid component appears to be prognostic as well, with
a larger percentage of involvement being associated with worse survival. However, there is no agreement on
the optimal prognostic cut-off for sub-classifying sarcomatoid changes [163, 164]. The WHO/ISUP grading
system is applicable to both ccRCC and pRCC. It is currently not recommended to grade chRCC. However,
a recent study suggested a two-tiered chRCC grading system (low vs. high grade) based on the presence of
sarcomatoid differentiation and/or tumour necrosis, which was statistically significant on multivariable analysis
[165]. Both the WHO/ISUP and chRCC grading systems need to be validated for prognostic systems and
nomograms [159].

Renal cell carcinoma subtype is regarded as another important prognostic factor. On univariable analysis,
patients with chRCC vs. pRCC vs. ccRCC had a better prognosis [166, 167] (Table 6.1). However, prognostic
information provided by the RCC type is lost when stratified according to tumour stage [167, 168] (LE: 3).
In a recent cohort study of 1,943 patients with ccRCC and pRCC significant survival differences were
only shown between pRCC type I and ccRCC [169]. Papillary RCC has been traditionally divided into type 1 and 2,
but a subset of tumours shows mixed features. For more details, see Section 3.2 - Histological diagnosis. Data
also suggest that type 2 pRCC is a heterogeneous entity with multiple molecular subgroups [21]. Some studies
suggest poorer survival for type 2 than type 1 [170], but this association is often lost in the multivariable analysis
[171]. A meta-analysis did not show a significant survival difference between both types [172].
Renal cell carcinoma with Xp11.2 translocation has a poor prognosis [173]. Its incidence is low, but
its presence should be systematically assessed in young patients. Renal cell carcinoma type classification has
been confirmed by cytogenetic and genetic analyses [174-176] (LE: 2b). Surgically excised malignant complex
cystic masses contain ccRCC in the majority of cases, and more than 80% are pT1. In a recent series, 5-year
cancer-specific survival (CSS) was 98% [177]. Differences in tumour stage, grade and CSS between RCC types
are illustrated in Table 6.1.

RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021 23


Table 6.1: B
 aseline characteristics and cancer-specific survival of surgically treated patients by RCC
type [162]

Survival time % RCC % Sarcomatoid % T3-4 % N1 % M1 % 10 year CSS (%)


clear-cell RCC 80 5 33 5 15 62
papillary RCC 15 1 11 5 3 86
chromophobe RCC 5 8 15 4 4 86
CSS = cancer-specific survival.

In all RCC types, prognosis worsens with stage and histopathological grade (Table 6.2). The 5-year overall
survival (OS) for all types of RCC is 49%, which has improved since 2006, probably due to an increase in
incidentally detected RCCs and new systemic treatments [178, 179]. Although not considered in the current
N classification, the number of metastatic regional LNs is an important predictor of survival in patients without
distant metastases [180].

Table 6.2: Cancer-specific survival by stage [20]

Grade HR (95% CI)


T1N0M0 Referent
T2N0M0 2.71 (2.17–3.39)
T3N0M0 5.20 (4.36–6.21)
T4N0M0 16.88 (12.40–22.98)
N+M0 16.33 (12.89–20.73)
M+ 33.23 (28.18–39.18)
CI = confidential interval. HR = hazard ratio.

6.4 Clinical factors


Clinical factors include performance status (PS), local symptoms, cachexia, anaemia, platelet count, neutrophil
count, lymphocyte count, C-reactive protein (CRP), albumin, and various indices deriving from these factors
such as the neutrophil-to-lymphocyte ratio (NLR) [93, 181-185] (LE: 3). As a marker of systemic inflammatory
response, a high pre-operative NLR has been associated with poor prognosis [186], but there is signficant
heterogeneity in the data and no agreement on the optimal prognostic cutoff. Even though obesity is an
aetiological factor for RCC, it has also been observed to provide prognostic information. A high body mass
index (BMI) appears to be associated with improved survival outcomes in both non-metastatic and metastatic
RCC [187-189]. This association is linear with regards to cancer-specific mortality, while obese RCC patients
show increasing all-cause mortality with increasing BMI [190]. There is also evolving evidence on the
prognostic value of body composition indices measured on cross-sectional imaging, such as sarcopenia and
fat accumulation [191, 192].

6.5 Molecular factors


Numerous molecular markers such as carbonic anhydrase IX (CaIX), VEGF, hypoxia-inducible factor (HIF), Ki67
(proliferation), p53, p21 [193], PTEN (phosphatase and tensin homolog) cell cycle, E-cadherin, osteopontin
[194] CD44 (cell adhesion) [195, 196], CXCR4 [197], PD-L1 [198], miRNA, SNPs, gene mutations, and
gene methylations have been investigated (LE: 3) [19]. While the majority of these markers are associated
with prognosis and many improve the discrimination of current prognostic models, there has been very
little emphasis on external validation studies. Furthermore, there is no conclusive evidence on the value of
molecular markers for treatment selection in mRCC [199]. Their routine use in clinical practice is therefore not
recommended.

Several prognostic and predictive marker signatures have been described for specific systemic treatments in
mRCC. In the JAVELIN Renal 101 trial (NCT02684006), a 26-gene immunomodulatory gene signature predicted
PFS in those treated with avelumab plus axitinib, while an angiogenesis gene signature was associated
with PFS for sunitinib. Mutational profiles and histocompatibility leukocyte antigen (HLA) types were also
associated with PFS, while PD-L1 expression and tumour mutational burden were not [200]. In IMmotion151
(NCT02420821), a T effector/IFN-γ-high or angiogenesis-low gene expression signature predicted improved
PFS for atezolizumab plus bevacizumab compared to sunitinib. The angiogenesis-high gene expression
signature correlated with longer PFS in patients treated with sunitinib [201]. In CheckMate 214 (NCT02231749),
a higher angiogenesis gene signature score was associated with better overall response rates and PFS for

24 RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021


sunitinib, while a lower angiogenesis score was associated with higher ORR in those treated with nivolumab
plus ipilimumab. Progression-free survival ≥ 18 months was more often seen in patients with higher expression
of Hallmark inflammatory response and Hallmark epithelial mesenchymal transition gene sets [202].

Urinary and plasma Kidney-Injury Molecule-1 (KIM-1) has been identified as a potential diagnostic and
prognostic marker. KIM-1 concentrations were found to predict RCC up to 5 years prior to diagnosis and were
associated with a shorter survival time [203]. KIM-1 is a glycoprotein marker of acute proximal tubular injury
and therefore mainly expressed in RCC derived from the proximal tubules such as ccRCC and pRCC [204].
While early studies are promising, more high-quality research is required. Several retrospective studies and
large molecular screening programs have identified mutated genes and chromosomal changes in ccRCC with
distinct clinical outcomes. The expression of the BAP1 and PBRM1 genes, situated on chromosome 3p in
a region that is deleted in more than 90% of ccRCCs, have shown to be independent prognostic factors for
tumour recurrence [205-207]. These published reports suggest that patients with BAP1-mutant tumours have
worse outcomes compared with patients with PBRM1-mutant tumours [206]. Loss of chromosome 9p and 14q
have been consistently shown to be associated with poorer survival [208-210]. The TRACERx renal consortium
has proposed a genetic classification based on RCC evolution (punctuated vs. branched vs. linear), which
correlates with tumour aggressiveness and survival [209]. Additionally, a 16-gene signature was shown to
predict disease-free survival (DFS) in patients with non-metastatic RCC [211]. However, these signatures have
not been validated by independent researchers yet.

6.6 Prognostic models


Prognostic models combining independent prognostic factors have been developed and externally validated
[212-218]. These models are more accurate than TNM stage or grade alone for predicting clinically relevant
oncological outcomes (LE: 3). Before being adopted, new prognostic models should be evaluated and
compared to current prognostic models with regards to discrimination, calibration and net benefit. In metastatic
disease, risk groups assigned by the Memorial Sloan Kettering Cancer Center (MSKCC) (primarily created in
the pre-targeted therapy, and validated in patients receiving targeted therapy) and the International Metastatic
Renal Cell Carcinoma Database Consortium (IMDC) (initially created in the targeted therapy era) differ in
23% of cases [219]. The IMDC model has been used in the majority of recent randomised trials, including
those with immune checkpoint inhibitors, and may therefore be the preferred model for clinical practice. The
discrimination of the IMDC model can be improved by addition of a seventh variable, namely presence of brain,
bone, and/or liver metastases [220]. For patients treated with immune checkpoint inhibitors, the monocyte-
to-lymphocyte ratio, BMI, and number and site of metastases at baseline were recently used to create a four-
tiered prediction model. This model showed greater discrimination than IMDC in predicting OS, but needs
further validation [221].
Overall, there is no conclusive evidence that one prognostic model is superior to another for both
localised and metastatic disease [19]. Tables 6.3 and 6.4 summarise the current most relevant prognostic
models.

6.7 Summary of evidence and recommendations for prognostic factors

Summary of evidence LE
In RCC patients, TNM stage, tumour size, grade, and RCC subtype provide important prognostic 2a
information.

Recommendations Strength rating


Use the current Tumour, Node, Metastasis classification system. Strong
Use the WHO/ISUP grading system and classify renal cell carcinoma type. Strong
Use prognostic models in localised and metastatic disease. Strong
Do not routinely use molecular markers to assess prognosis. Strong

RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021 25


Table 6.3: Prognostic models for localised RCC

Prognostic model Subtype* Risk factors/prognostic factors


UISS** [222] All 1. ECOG PS
2. T classification
3. N classification (N+ classified as metastatic)
4. Grade

T1N0M0G1–2, ECOG PS 0: low-risk disease


T3N0M0G2–4, ECOG PS ≥ 1 OR T4N0M0: high-risk disease
Any other N0M0: intermediate-risk disease
Leibovich CC 1. T classification (pT1a: 0 points, pT1b: 1 point, pT2: 3 points,
score/model 2003 pT3–4: 4 points)
[215] 2. N classification (pNx/N0: 0 points, pN+: 2 points)
3. Tumour size (< 10 cm: 0 points, ≥ 10 cm: 1 point)
4. Grade (G1–2: 0 points, G3: 1 point, G4: 3 points)
5. Tumour necrosis (absent: 0 points, present: 1 point)

0–2 points: low-risk disease


3–5 points: intermediate-risk disease
6 or more points: high-risk disease
Leibovich CC, P, ccRCC
score/model 2018 CHR • Progression (9 factors): constitutional symptoms, grade, tumour necrosis,
[223] sarcomatoid features, tumour size, perinephric or sinus fat invasion,
tumour thrombus level, extension beyond kidney, nodal involvement.
• Cancer-specific survival (12 factors): age, ECOG PS, constitutional
symptoms, adrenalectomy, surgical margins, grade, tumour necrosis,
saromatoid features, tumour size, perinephric or sinus fat invasion,
tumour thrombus, nodal involvement.
• No risk groups/prognostic groups.

pRCC
• Low risk (group 1): grade 1–2, no fat invasion, no tumour thrombus.
• Intermediate risk (group 2): grade 3, no fat invasion, no tumour thrombus.
• High risk (group 3): grade 4 or fat invasion or any level tumour thrombus.

chRCC
• Low risk (group 1): no fat invasion, no sarcomatoid differentiation, no
nodal involvement.
• Intermediate risk (group 2): fat invasion and no sarcomatoid
differentiation and no nodal inolvement.
• High risk (group 3): sarcomatoid differentiation or nodal involvement.
VENUSS P 1. T classification (pT1: 0 points, pT2: 1 point, pT3–4: 2 points)
score/model*** 2. N classification (pNx/pN0: 0 points, pN1: 3 points)
[171] 3. Tumour size (≤ 4 cm: 0 points, > 4 cm: 2 points)
4. Grade (G1/2: 0 points, G3/4: 2 points)
5. Tumour thrombus (absent: 0 points, present: 2 points)

0–2 points: low-risk disease


3–5 points: intermediate-risk disease
6 or more points: high-risk disease
GRANT All 1. Age > 60 years
score/model**** 2. T classification = T3b, pT3c or pT4
[224] 3. N classification = pN1
4. (Fuhrman) grade = G3 or G4

0–1 factors: favourable-risk disease


2 or more factors: unfavourable-risk disease
* ccRCC = clear-cell RCC; ECOG = Eastern Cooperative Oncology Group; pRCC = papillary RCC;
chRCC = chromophobe RCC.
** University of California Integrated Staging system. Available at https://www.mdcalc.com/ucla-integrated-
staging-system-uiss-renal-cell-carcinoma-rcc.
*** VEnous extension, NUclear grade, Size, Stage. Available at https://evidencio.com/.
**** GRade, Age, Nodes and Tumour.

26 RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021


Table 6.4: Prognostic models for metastatic RCCC

Prognostic model Subtype Risk factors/prognostic factors


MSKCC [225]** All 1. Karnofsky PS [226]* < 80%
2. Interval from diagnosis to systemic treatment < 1 year
3. Haemoglobin < Lower Limit of Normal
4. Corrected calcium >10 mg/dL/> 2.5 mmol/L
5. LDH > 1.5x Upper Limit of Normal

0 factors: favourable-risk disease


1–2 factors: intermediate-risk disease
3–5 factors: poor-risk disease
IMDC [227]*** All 1. Karnofsky PS [226]* < 80%
2. Interval from diagnosis to treatment < 1 year
3. Haemoglobin < lower limit of normal
4. Corrected calcium > upper limit of normal (i.e. > 10.2 mg/dL)
5. Neutrophil count > upper limit of normal (i.e. > 7.0×109/L)
6. Platelet count > upper limit of normal (i.e. > 400,000)

0 factors: favourable-risk disease


1–2 factors: intermediate-risk disease
3–6 factors: poor-risk disease
IMDC = International Metastatic Renal Cancer Database Consortium; LDH = lactate dehydrogenase;
MSKCC = Memorial Sloan Kettering Cancer Center; PS = performance status.
* Karnofsky performance status calculator: https://www.thecalculator.co/health/Karnofsky-Score-for-
Performance-Status-Calculator-961.html.
** MSKCC: https://www.mdcalc.com/memorial-sloan-kettering-cancer-center-mskcc-motzer-score-metastatic-
renal-cell-carcinoma-rcc.
*** IMDC: https://www.mdcalc.com/imdc-international-metastatic-rcc-database-consortium-risk-score-rcc.

7. DISEASE MANAGEMENT
7.1 Treatment of localised RCC
7.1.1 Introduction
Sections 7.1.2 and 7.2.4.2 are underpinned by a systematic review which includes all relevant published
literature comparing surgical management of localised RCC (T1-2N0M0). Randomised or quasi-RCTs were
included. However, due to the very limited number of RCTs, non-randomised studies (NRS), prospective
observational studies with controls, retrospective matched-pair studies, and comparative studies from the
databases of well-defined registries were also included. Historically, surgery has been the benchmark for the
treatment of localised RCC.

7.1.2 Surgical treatment


7.1.2.1 Nephron-sparing surgery versus radical nephrectomy in localised RCC
7.1.2.1.1 T1 RCC
Outcome 1: Cancer-specific survival
Most studies comparing the oncological outcomes of PN and RN are retrospective and include cohorts of
varied and, overall, limited size [228, 229]. There is only one, prematurely closed, prospective RCT including
patients with organ-confined RCCs of limited size (< 5 cm), showing comparable non-inferiority of CSS for PN
vs. RN (HR: 2.06 [95% CI: 0.62–6.84]) [230].

Outcomes 2 & 3: Overall mortality and renal function


Partial nephrectomy preserved kidney function better after surgery, thereby potentially lowering the risk of
development of cardiovascular disorders [228, 231-235]. When compared with a radical surgical approach,
several retrospective analyses of large databases have suggested a decreased cardiovascular-specific
mortality [232, 236] as well as improved OS for PN compared to RN. However, in some series this held true
only for younger patients and/or patients without significant comorbidity at the time of the surgical intervention

RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021 27


[237, 238]. An analysis of the U.S. Medicare database [239] could not demonstrate an OS benefit for patients
≥ 75 years of age when RN or PN were compared with non-surgical management.
Conversely, another series that addressed this question and also included Medicare patients
suggested an OS benefit in older patients (75–80 years) when subjected to surgery rather than non-surgical
management. Shuch et al. compared patients who underwent PN for RCC with a non-cancer healthy control
group via a retrospective database analysis; showing an OS benefit for the cancer cohort [240]. These
conflicting results may be an indication that unknown statistical confounders hamper the retrospective
analysis of population-based tumour registries. In the only prospectively randomised, but prematurely closed,
heavily underpowered, trial, PN seems to be less effective than RN in terms of OS in the intention to treat (ITT)
population (HR: 1.50 [95% CI: 1.03–2.16]). However, in the targeted RCC population of the only RCT, the trend
in favour of RN was no longer significant [230]. Taken together, the OS advantage suggested for PN vs. RN
remains an unresolved issue.

Patients with a normal pre-operative renal function and a decreased GFR due to surgical treatment (either RN
or PN), generally present with stable long-term renal function [235]. Adverse OS in patients with a pre-existing
GFR reduction does not seem to result from further renal function impairment following surgery, but rather from
other medical comorbidities causing pre-surgical chronic kidney disease (CKD) [241]. However, in particular in
patients with pre-existing CKD, PN is the treatment of choice to limit the risk of development of ESRD which
requires haemodialysis. Huang et al. found that 26% of patients with newly diagnosed RCC had an GFR ≤ 60
mL/min, even though their baseline serum creatinine levels were in the normal range [96].

Outcomes 4 & 5: Peri-operative outcomes and quality of life


In terms of the intra- and peri-operative morbidity/complications associated with PN vs. RN, the European
Organisation for Research and Treatment of Cancer (EORTC) randomised trial showed that PN for small, easily
resectable, incidentally discovered RCC, in the presence of a normal contralateral kidney, can be performed
safely with slightly higher complication rates than after RN [242].

Only a limited number of studies are available addressing quality of life (QoL) following PN vs. RN, irrespective
of the surgical approach used (open vs. minimally invasive). Quality of life was ranked higher following PN as
compared to RN, but in general patients’ health status deteriorated following both approaches [242, 243].

In view of the above, and since oncological safety (CSS and RFS) of PN, so far, has been found non-differing
from RN outcomes, PN is the treatment of choice for T1 RCC since it preserves kidney function better and in
the long term potentially limits the incidence of cardiovascular disorders. Whether decreased mortality from
any cause can be attributed to PN is still unresolved, but in patients with pre-existing CKD, PN is the preferred
surgical treatment option as it avoids further deterioration of kidney function; the latter being associated with
a higher risk of development of ESRD and the need for haemodialysis. Irrespective of the available data, in
frail patients, treatment decisions should be individualised, weighing the risks and benefits of PN vs. RN, the
increased risk of peri-operative complications and the risk of developing or worsening CKD post-operatively.

7.1.2.1.2 T2 renal cell carcinoma


There is very limited evidence on the optimal surgical treatment for patients with larger renal masses (T2).
Some retrospective comparative studies of PN vs. RN for T2 RCC have been published [244]. A trend for
lower tumour recurrence- and cancer-specific mortality is reported in PN groups. The estimated blood loss
is reported to be higher for PN groups, as is the likelihood of post-operative complications [244]. A recent
multicentre study compared the survival outcomes in patients with larger (≥ 7 cm) ccRCC treated with PN
vs. RN with long-term follow-up (median 102 months). Compared to the RN group, the PN group had a
significantly longer median OS (p = 0.014) and median CSS (p = 0.04) [245]. Overall the level of the evidence
is low. These studies including T2 masses all have a high risk of selection bias due to imbalance between the
PN and RN groups regarding patient’s age, comorbidities, tumour size, stage, and tumour position. These
imbalances in covariation factors may have a greater impact on patient outcome than the choice of PN or RN.
The Panel’s confidence in the results is limited and the true effects may be substantially different.
In view of the above, the risks and benefits of PN should be discussed patients with T2 RCC with
a solitary kidney, bilateral renal tumours or CKD, if technically feasible, with sufficient parenchymal volume
preserved to allow sufficient post-operative renal function, PN should be considered in these patients.

7.1.2.2 Associated procedures


7.1.2.2.1 Adrenalectomy
One prospective non-randomised study compared the outcomes of RN with or without, ipsilateral
adrenalectomy [246]. Multivariable analysis showed that upper pole location was not predictive of adrenal
involvement, but tumour size was. No difference in OS at 5 or 10 years was seen with, or without,

28 RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021


adrenalectomy. Adrenalectomy was justified using criteria based on radiographic- and intra-operative findings.
Only 48 of 2,065 patients underwent concurrent ipsilateral adrenalectomy of which 42 of the 48 interventions
were for benign lesions [246].

7.1.2.2.2 Lymph node dissection for clinically negative lymph nodes (cN0)
The indication for LN dissection (LND) together with PN or RN is still controversial [247]. The clinical
assessment of LN status is based on the detection of an enlargement of LNs either by CT/MRI or intra-
operative palpability of enlarged nodes. Less than 20% of suspected metastatic nodes (cN+) are positive for
metastatic disease at histopathological examination (pN+) [248]. Both CT and MRI are unsuitable for detecting
malignant disease in nodes of normal shape and size [249]. For clinically positive LNs (cN+) see Section 7.2.2.

Smaller retrospective studies have suggested a clinical benefit associated with a more or less extensive LND
preferably in patients at high risk for lymphogenic spread. In a large retrospective study, the outcomes of RN
with or without LND in patients with high-risk non-mRCC were compared using a propensity score analysis. In
this study LND was not significantly associated with a reduced risk of distant metastases, cancer-specific or
all-cause mortality. The extent of the LND was not associated with improved oncologic outcomes [250]. The
number of LN metastases (< / > 4) as well as the intra- and extracapsular extension of intra-nodal metastasis
correlated with the patients´ clinical prognosis in some studies [249, 251-253]. Better survival outcomes
were seen in patients with a low number of positive LNs (< 4) and no extranodal extension. On the basis of
a retrospective Surveillance, Epidemiology and End Results (SEER) database analysis of > 9,000 patients
no effects of an extended LND (eLND) on the disease-specific survival (DSS) of patients with pathologically
confined negative nodes was demonstrated [254]. However, in patients with pathologically proven lymphogenic
spread (pN+), an increase of 10 for the number of nodes dissected resulted in a 10% absolute increase in DSS.
In addition, in a larger cohort of 1,983 patients, Capitanio et al. demonstrated that eLND results in a significant
prolongation of CSS in patients with unfavourable prognostic features (e.g., sarcomatoid differentiation, large
tumour size) [255]. As to morbidity related to eLND, a recent retrospective propensity score analysis from a
large single-centre database showed that eLND is not associated with an increased risk of Clavien grade ≥ 3
complications. Furthermore, LND was not associated with length of hospital stay or estimated blood loss [256].

Only one prospective RCT evaluating the clinical value of LND combined with surgical treatment of primary
RCC has been published so far. With an incidence of LN involvement of only 4%, the risk of lymphatic
spread appears to be very low. Recognising the latter, only a staging effect was attributed to LND [248]. This
trial included a very high percentage of patients with pT2 tumours, which are not at increased risk for LN
metastases. Only 25% of patients with pT3 tumours underwent a complete LND and the LN template used by
the authors was not clearly stated.
The optimal extent of LND remains controversial. Retrospective studies suggest that an eLND
should involve the LNs surrounding the ipsilateral great vessel and the inter-aortocaval region from the
crus of the diaphragm to the common iliac artery. Involvement of inter-aortocaval LNs without regional hilar
involvement is reported in up to 35–45% of cases [249, 257, 258]. At least 15 LNs should be removed [255,
259]. Sentinel LND is an investigational technique [260, 261].

7.1.2.2.3 Embolisation
Before routine nephrectomy, tumour embolisation has no benefit [262, 263]. In patients unfit for surgery, or
with non-resectable disease, embolisation can control symptoms, including visible haematuria or flank pain
[264, 265]. These indications will be revisited in Sections 7.2 and 7.3 with cross reference to the summary of
evidence and recommendations below.

7.1.2.2.4 Summary of evidence and recommendations for the treatment of localised RCC

Summary of evidence LE
The oncological outcome in terms of OS following PN equals that of RN in patients with c/p T1 RCC. 1b
Retrospective studies suggest that oncological outcomes are similar following PN vs. RN in patients 3b
with larger (≥ 7 cm) RCC. Post-operative complication rates are higher in PN patients.
Ipsilateral adrenalectomy during RN or PN has no survival advantage in the absence of clinically 3
evident adrenal involvement.
In patients with localised disease without radiographic evidence of LN metastases, a survival 2b
advantage of LND in conjunction with RN is not demonstrated in randomised trials.
Retrospective studies suggest a clinical benefit associated with lymphadenectomy in high-risk patients. 2b
In patients unfit for surgery with massive haematuria or flank pain, embolisation can be a beneficial 3
palliative approach.

RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021 29


Recommendations Strength rating
Offer surgery to achieve cure in localised renal cell cancer. Strong
Offer partial nephrectomy (PN) to patients with T1 tumours. Strong
Offer PN to patients with T2 tumours and a solitary kidney or chronic kidney disease, if Weak
technically feasible.
Do not perform ipsilateral adrenalectomy if there is no clinical evidence of invasion of the Strong
adrenal gland.
Do not offer an extended lymph node dissection to patients with organ-confined disease. Weak
Offer embolisation to patients unfit for surgery presenting with massive haematuria or flank Weak
pain.

7.1.3 Radical and partial nephrectomy techniques


7.1.3.1 Radical nephrectomy techniques
No RCTs have assessed the oncological outcomes of laparoscopic vs. open RN. A cohort study [266] and
retrospective database reviews are available, mostly of low methodological quality, showing similar oncological
outcomes even for higher stage disease and locally more advanced tumours [267-269]. Based on a systematic
review, less morbidity was found for laparoscopic vs. open RN [228].
Data from one RCT [268] and two non-randomised studies [270, 271] showed a significantly
shorter hospital stay and lower analgesic requirement for the laparoscopic RN group as compared with
the open group. Convalescence time was also significantly shorter [271]. No difference in the number of
patients receiving blood transfusions was observed, but peri-operative blood loss was significantly less in
the laparoscopic arm in all 3 studies [268, 270, 271]. Surgical complication rates were low with very wide
confidence intervals. There was no difference in complications, but operation time was significantly shorter in
the open nephrectomy arm. Post-operative QoL scores were similar [270].
Some comparative studies focused on the peri-operative outcomes of laparoscopic vs. RN for
renal ≥ T2 tumours. Overall, patients who underwent laparoscopic RN were shown to have lower estimated
blood loss, less post-operative pain, shorter length of hospital stay and convalescence compared to those
who underwent open RN [269, 271, 272]. Intra-operative and post-operative complications were similar in the
two groups and no significant differences in CSS, PFS and OS were reported [269, 271, 272] (LE: 2b). Another
multicentre propensity matched analysis compared laparoscopic- and open surgery for pT3a RCC, showing
no significant difference in 3-year RFS between groups [273]. The best approach for laparoscopic RN was the
retroperitoneal or transperitoneal approach with similar oncological outcomes in two RTCs [274, 275] and one
quasi-randomised study [276]. Quality of life variables were similar for both approaches. Hand-assisted vs.
standard laparoscopic RN was compared in one quasi-randomised study [276] and one database review and
estimated 5-year OS, CSS, and RFS rates were comparable [277]. Duration of surgery was significantly shorter
in the hand-assisted approach, while length of hospital stay and time to non-strenuous activities were shorter
for the standard laparoscopic RN cohort [276, 277]. However, the sample size was small.
Data of a large retrospective cohort study on robot-assisted laparoscopic vs. laparoscopic
RN showed robotic-assisted laparoscopic RN was not associated with increased risk of any or major
complications but had a longer operating time and higher hospital costs compared with laparoscopic RN [278].
A systematic review reported on robot-assisted laparoscopic vs. conventional laparoscopic RN, showing no
substantial differences in local recurrence rates, nor in all-cause cancer-specific mortality [279]. Similar results
were seen in observational cohort studies comparing ‘portless’ and 3-port laparoscopic RN, with similar peri-
operative outcomes [280, 281].

7.1.3.2 Partial nephrectomy techniques


7.1.3.2.1 Open versus laparoscopic approach
Studies comparing laparoscopic and open PN found no difference in PFS [282-285] and OS [284, 285] in
centres with laparoscopic expertise. However, the oncological safety of laparoscopic vs. open PN has, so far,
only been addressed in studies with relatively limited follow-up [273]. However, the higher number of patients
treated with open surgery in this series might reflect a selection bias by offering laparoscopic surgery in case
of a less complex anatomy [273]. The mean estimated blood loss was found to be lower with the laparoscopic
approach [282, 284, 286], while post-operative mortality, deep vein thrombosis, and pulmonary embolism
events were similar [282, 284]. Operative time is generally longer with the laparoscopic approach [283-285] and
warm ischaemia time is shorter with the open approach [282, 284, 286, 287]. In a matched-pair comparison,
GFR decline was greater in the laparoscopic PN group in the immediate post-operative period [285], but
not after 3.6 years follow-up. In another comparative study, the surgical approach was not an independent
predictor for post-operative CKD [287]. Retroperitoneal and transperitoneal laparoscopic PN have similar
peri-operative outcomes [288]. Simple tumour enucleation also had similar PFS and CSS rates compared to

30 RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021


standard PN and RN in a large study [289]. The feasibility of laparo-endoscopic single-site PN has been shown
in selected patients but larger studies are needed to confirm its safety and clinical role [290].

7.1.3.2.2 Open versus robotic approach


One study prospectively compared the peri-operative outcomes of a series of robot-assisted and open PN
performed by the same experienced surgeon. Robot-assisted PN was superior to open PN in terms of lower
estimated blood loss and shorter hospital stay. Warm ischaemia time, operative time, immediate- early- and
short-term complications, variation in creatinine levels and pathologic margins were similar between groups
[291].
A multicentre French prospective database compared the outcomes of 1,800 patients who
underwent open PN and robot-assisted PN. Although the follow-up was shorter, there was a decreased
morbidity in the robotic-assisted PN group with less overall complications, less major complications, less
transfusions and a much shorter hospital stay [292].

7.1.3.2.3 Open versus hand-assisted approach


Hand-assisted laparoscopic PN (HALPN) is rarely performed. A recent comparative study of open vs. HALPN
showed no difference in OS or RFS at intermediate-term follow-up. The authors observed a lower rate of intra-
operative and all-grade post-operative 30-day complications in HALPN vs. open PN patients, but there was no
significant difference in high Clavien grade complications. Three months after the operation, GFR was lower in
the HALPN than in the open PN group [293].

7.1.3.2.4 Open versus laparoscopic versus robotic approaches


In a retrospective propensity-score-matched study, comparing open-, laparoscopic- and robot-assisted PN,
after 5-year of median follow-up, similar rates of local recurrence, distant metastasis and cancer-related death
rates were found [294].

7.1.3.2.5 Laparoscopic versus robotic approach


Another study included the 50 last patients having undergone laparoscopic and robotic PN for T1-T2 renal
tumours by two different surgeons with an experience of over 200 procedures each in laparoscopic and robotic
PN and robotic-assisted partial nephrectomy (RAPN), respectively, at the beginning of the study. Peri-operative
and short-term oncological and functional outcomes appeared broadly comparable between RAPN and LPN
when performed by highly experienced surgeons [295].

A meta-analysis, including a series of NSS with variable methodological quality compared the peri-operative
outcomes of robot-assisted- and laparoscopic PN. The robotic group had a significantly lower rate of
conversion to open surgery and to radical surgery, shorter warm ischaemia time, smaller change in estimated
GFR after surgery and shorter length of hospital stay. No significant differences were observed between the
two groups regarding complications, change of serum creatinine after surgery, operative time, estimated blood
loss and positive surgical margins [296].

7.1.3.2.6 Surgical volume


In a recent analysis of 8,753 patients who underwent PN, an inverse non-linear relationship of hospital volume
with morbidity of PN was observed, with a plateauing seen at 35 to 40 cases per year overall, and 18 to 20
cases for the robotic approach [297]. A retrospective study of a U.S. National Cancer Database looked at the
prognostic impact of hospital volume and the outcomes of robot-assisted PN, including 18,724 cases. This study
shows that undergoing RAPN at higher-volume hospitals may have better peri-operative outcomes (conversion
to open and length of hospital stay) and lower positive surgical margin rates [298]. A French study, including
1,222 RAPN patients, has shown that hospital volume is the main predictive factor of Trifecta achievement
(no complications, warm ischaemia time < 25 min, and negative surgical margins) after adjustment for other
variables, including surgeon volume [299]. The prospective REgistry of COnservative and Radical Surgery for
cortical renal tumour Disease (RECORd-2) study including 2,076 patients showed that the hospital volume
(> 60 PN/year) is an independent predictor for positive surgical margins [300].

7.1.3.3 Positive surgical margins on histopathological specimens


A positive surgical margin is encountered in about 2–8% of PNs [296]. Studies comparing surgical margins with
different surgical approaches (open, laparoscopic, robotic) are inconclusive [301, 302]. Most trials showed that
intra-operative frozen section analysis had no influence on the risk of definite positive surgical margins [303]. A
positive surgical margin status occurs more frequently in cases in which surgery is imperative (solitary kidneys
and bilateral tumours) and in patients with adverse pathological features (pT2a, pT3a, grade III-IV) [304-307].
The potential negative impact of a positive margin status on the oncologic outcome is still controversial [301].

RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021 31


The majority of retrospective analyses reported so far indicated that positive surgical margins do not translate
into a higher risk of metastases or a decreased CSS [305, 306]. On the other hand, another retrospective study
of a large single institutional series showed that positive surgical margins are an independent predictor of PFS
due to a higher incidence of distant and local relapses [308].
However, only a proportion of patients with an uncertain margin status actually harbour residual
malignancy [309]. Local tumour bed recurrences were found in 16% in patients with positive surgical margins
compared with 3% in those with negative margins [304], Therefore, RN or re-resection of margins can result in
over-treatment in many cases. Patients with positive surgical margins should be informed that they will need a
more intense surveillance (imaging) follow-up and that they are at increased risk of secondary local therapies
[305, 310]. On the other hand, protection from recurrence is not ensured by negative surgical margins [311].

7.1.3.4 Summary of evidence and recommendations for radical and partial nephrectomy techniques

Summary of evidence LE
Laparoscopic radical nephrectomy (RN) has lower morbidity than open nephrectomy. 1b
Short-term oncological outcomes for T1-T2a tumours are equivalent for laparoscopic and open RN. 2a
Partial nephrectomy can be performed, either by open-, pure laparoscopic- or robot-assisted 2b
approach, based on surgeon’s expertise and skills.
Robotic-assisted and laparoscopic PN are associated with shorter length of hospital stay and lower 2b
blood loss compared to open PN.
Partial nephrectomy is associated with a higher percentage of positive surgical margins compared to RN. 3
Hospital volume in PN might impact on surgical complications, warm ischaemia and surgical margins. 3
Radical nephrectomy after positive surgical margins can result in over-treatment in many cases. 3

Recommendations Strength rating


Offer laparoscopic radical nephrectomy (RN) to patients with T2 tumours and localised Strong
masses not treatable by partial nephrectomy (PN).
Do not perform minimally invasive RN in patients with T1 tumours for whom a PN is feasible Strong
by any approach, including open.
Do not perform minimally invasive surgery if this approach may compromise oncological-, Strong
functional- and peri-operative outcomes.
Intensify follow-up in patients with a positive surgical margin. Weak

7.1.4 Therapeutic approaches as alternatives to surgery


7.1.4.1 Surgical versus non-surgical treatment
Population-based studies compared the oncological outcomes of surgery (RN or PN) and non-surgical
management for tumours < 4 cm. The analyses showed a significantly lower cancer-specific mortality in
patients treated with surgery [239, 312, 313]. However, the patients assigned to the surveillance arm were
older and likely to be frailer and less suitable for surgery. Other-cause mortality rates in the non-surgical group
significantly exceeded that of the surgical group [312]. Analyses of older patients (> 75 years) failed to show the
same benefit in cancer-specific mortality for surgical treatment [314-316].

7.1.4.2 Active surveillance and watchful waiting


Elderly and comorbid patients with incidental small renal masses have a low RCC-specific mortality and
significant competing-cause mortality [317, 318]. Active surveillance is defined as the initial monitoring of
tumour size by serial abdominal imaging (US, CT, or MRI) with delayed intervention reserved for tumours
showing clinical progression during follow-up [319]. The concept of AS differs from the concept of watchful
waiting; watchful waiting is reserved for patients whose comorbidities contraindicate any subsequent active
treatment and do not require follow-up imaging, unless clinically indicated.
In the largest reported series of AS the growth of renal tumours was low and progression to
metastatic disease was reported in only a limited number of patients [320, 321].
A single-institutional comparative study evaluating patients aged > 75 years showed decreased
OS for those who underwent surveillance and nephrectomy relative to NSS for clinically T1 renal tumours.
However, at multivariate analysis, management type was not associated with OS after adjusting for age,
comorbidities, and other variables [317]. No statistically significant differences in OS and CSS were observed in
another study of RN vs. PN vs. AS for T1a renal masses with a follow-up of 34 months [322]. The prospective
non-randomised multi-institutional Delayed Intervention and Surveillance for Small Renal Masses (DISSRM)

32 RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021


study enrolled 497 patients with solid renal masses < 4 cm who selected either AS or primary active intervention.
Patients who selected AS were older, had worse ECOG scores, more comorbidities, smaller tumours, and
more often had multiple and bilateral lesions. In patients who elected AS in this study the overall median small
renal mass growth rate was 0.09 cm/year with a median follow-up of 1.83 years. The growth rate and variability
decreased with longer follow-up. No patients developed metastatic disease or died of RCC [323, 324].
Overall survival for primary intervention and AS was 98% and 96% at 2 years, and 92% and 75% at
5 years, respectively (p = 0.06). At 5 years, CSS was 99% and 100%, respectively (p = 0.3). Active surveillance
was not predictive of OS or CSS in regression modelling with relatively short follow-up [323]. Overall, both
short- and intermediate-term oncological outcomes indicate that in selected patients with advanced age
and/ or comorbidities, AS is appropriate for initially monitoring of small renal masses, followed, if required, by
treatment for progression [319-321, 325-328].
A multicentre study assessed QoL of patients undergoing immediate intervention vs. AS. Patients
undergoing immediate intervention had higher QoL scores at baseline, specifically for physical health. The
perceived benefit in physical health persisted for at least one year following intervention. Mental health, which
includes domains of depression and anxiety, was not adversely affected while on AS [329].

7.1.4.3 Role of renal tumour biopsy before active surveillance


Histological characterisation of small renal masses by renal tumour biopsy is useful to select tumours at
lower risk of progression based on grade and histotype, which can be safely managed with AS. Pathology
can also help to tailor surveillance imaging schedules. In the largest cohort of biopsy-proven, small, sporadic
RCCs followed with AS a significant difference in growth and progression among different RCC subtypes was
observed. Clear-cell RCC small renal masses grew faster than papillary type 1 small renal masses (0.25 and
0.02 cm/year on average, respectively, p = 0.0003) [330].

7.1.4.4 Tumour ablation


7.1.4.4.1 Role of renal mass biopsy
A RMB is required prior to tumour ablation (TA) (see Sections 5.3 Renal tumour biopsy and 5.4 Summary
of evidence and recommendations for the diagnostic assessment of RCC). Historically, up to 45% of
patients underwent TA of a benign or non-diagnostic mass [331, 332]. A RMB in a separate session reduces
over-treatment significantly, with 80% of patients with benign lesions opting not to proceed with TA [332].
Additionally, there is some evidence that the oncological outcome following TA differs according to RCC
subtype which should therefore be factored into the decision-making process. In a series of 229 patients with
cT1a tumours (mean size 2.5 cm) treated with RFA, the 5-year DFS rate was 90% for ccRCC and 100% for
pRCC (80 months: 100% vs. 87%, p = 0.04) [333]. In another series, the total TA effectiveness rate was 90.9%
for ccRCC and 100% for pRCC [334]. A study comparing RFA with surgery suggested worse outcomes of RFA
vs. PN in cT1b ccRCC, while no difference was seen in those with non-ccRCC [335]. Furthermore, patients with
high-grade RCC or metastasis may choose different treatments over TA. Finally, patients without biopsy or a
non-diagnostic biopsy are often assumed to have RCC and will undergo potentially unnecessary radiological
follow-up or further treatment.

7.1.4.4.2 Cryoablation
Cryoablation is performed using either a percutaneous- or a laparoscopic-assisted approach, with technical
success rates of > 95% [336]. In comparative studies, there was no significant difference in the overall
complication rates between laparoscopic- and percutaneous cryoablation [337-339]. One comparative study
reported similar OS, CSS, and RFS in 145 laparoscopic patients with a longer follow-up vs. 118 patients
treated percutaneously with a shorter follow-up [338]. A shorter average length of hospital stay was found
with the percutaneous technique [338-340]. A systematic review including 82 articles reported complication
rates ranging between 8 and 20% with most complications being minor [341]. Although a precise definition of
tumour recurrence is lacking, the authors reported a lower RFS as compared to that of PN.

Oncological outcomes after cryoablation have generally been favourable for cT1a tumours. In a recently
published series of 308 patients with cT1a and cT1b tumours undergoing percutaneous cryoablation, local
recurrence was seen in 7.7% of cT1a tumours vs. 34.5% of cT1b tumours. Disease-free survival for the entire
cohort was 92.5% at 1 year, 89.3% at 2 years, and 86.7% at 3 years. On multivariable regression, the risk of
disease progression increased by 32% with each 1 cm increase in tumour size (HR: 1.32, p < 0.001). Mean
decline in eGFR was 11.7 mL/min/1.73 m2 [342]. In another large series of 220 patients with biopsy proven
cT1 RCC, 5-year local RFS was 93.9%, while metastasis-free survival approached 94.4% [336].
For cT1b tumours, local tumour control rates drop significantly. One study showed local tumour
control in only 60.3% at 3 years [343]. In another series, the PFS rate was 66.7% at 12 months [344].

RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021 33


Furthermore, recent analyses demonstrated 5-year cancer-specific mortality rates of 7.6–9% [345, 346]. On
multivariable analysis, cryoablation of cT1b tumours was associated a 2.5-fold increased risk of death from
RCC compared with PN [345].
Recurrence after initial cryoablation is often managed with re-cryoablation, but only 45% of patients
remain disease-free at 2 years [347].

7.1.4.4.3 Radiofrequency ablation


Radiofrequency ablation is performed laparoscopically or percutaneously. Several studies compared patients
with cT1a tumours treated by laparoscopic or percutaneous RFA [348-351]. Complications occurred in up to
29% of patients but were mostly minor. Complication rates, recurrence rates and CSS were similar in patients
treated laparoscopically and percutaneously.
The initial technical success rate on early (i.e. 1 month) imaging after one session of RFA is 94%
for cT1a and 81% for cT1b tumours [352]. This is generally managed by re-RFA, approaching overall total
technical success rates > 95% with one or more sessions [353].
Long-term outcomes with over five years of follow-up following RFA have been reported. In recent
studies, the 5-year OS rate was 73–79% [352, 353], due to patient selection. Oncological outcomes for cT1a
tumours have been favourable. In a recent study, the 10-year disease-free survival rate was 82%, but there was
a significant drop to 68% for tumours > 3 cm [353]. In series focusing on clinical T1b tumours (4.1–7.0 cm), the
5-year DFS rate was 74.5% to 81% [352, 354]. Oncological outcomes appear to be worse than after surgery,
but comparative data are severely biased (see Section 7.1.4.3.4). In general, most disease recurrences occur
locally and recurrences beyond five years are rare [353, 354].

7.1.4.4.4 Tumour ablation versus surgery


The Guideline Panel performed a protocol-driven systematic review of comparative studies (including > 50
patients) of TA with PN for T1N0M0 renal masses [7]. Twenty-six non-randomised comparative studies
published between 2000 and 2019 were included, recruiting a total of 16,780 patients. Four studies compared
laparoscopic TA vs. laparoscopic/robotic PN; 16 studies compared laparoscopic or percutaneous TA vs.
open-, laparoscopic- or robotic PN; 2 studies compared different techniques of TA and 4 studies compared
TA vs. PN vs RN. In this systematic review, TA as treatment for T1 renal masses was found to be safe in
terms of complications and adverse events, but its long-term oncological effectiveness compared with PN
remained unclear. The primary reason for the persisting uncertainty was related to the nature of the available
data; most studies were retrospective observational studies with poorly matched controls, or single-arm case
series with short follow-up. Many studies were poorly described and lacked a clear comparator. There was
also considerable methodological heterogeneity. Another major limitation was the absence of clearly defined
primary outcome measures. Even when a clear endpoint such as OS was reported, data were difficult to
interpret because of the varying length and type of follow-up amongst studies. The Panel also appraised the
published systematic reviews based on the AMSTAR 2 tool which showed critically low or low ratings [7].

Tumour ablation has been demonstrated to be associated with good long-term survival in several single-arm
non-comparative studies [355, 356]. Due to the lack of controls, this apparent benefit is subject to significant
uncertainties. Whether such benefit is due to the favourable natural history of such tumours or due to the
therapeutic efficacy of TA, as compared to PN, remains unknown. In addition, there are data from comparative
studies suggesting TA may be associated with worse oncological outcomes in terms of local recurrence and
metastatic progression and cancer-specific mortality [237, 345, 346, 357, 358]. However, there appears to be
no clinically significant difference in 5-year cancer-specific mortality between TA and AS [313].

The Panel concluded that the current data are inadequate to reach conclusions regarding the clinical
effectiveness of TA as compared with PN. Given these uncertainties in the presence of only low-quality
evidence, TA can only be recommended to frail and/or comorbid patients with small renal masses.

7.1.4.4.5 Stereotactic ablative radiotherapy


Stereotactic ablative radiotherapy (SABR) has been emerging as a treatment option for medically inoperable
patients with localised cT1a and cT1b tumours. Patients usually receive 26 Gy in a single fraction, three
fractions of 14 Gy or five fractions of 6 Gy [359, 360]. In a systematic review or non-comparative single-arm
studies, the local control rate was 97.2% and the mean change in eGFR was 7.7 mL/min/1.73 m2. Grade 3 or
4 toxicities occurred in 1.5% of patients. However, viable tumour cells are often seen in post-SABR biopsies,
although their clinical significance remains unclear [360]. Although early results of SABR are encouraging, more
evidence from randomised trials is needed.

34 RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021


7.1.4.4.6 Other ablative techniques
Some studies have shown the feasibility of other ablative techniques, such as microwave ablation, high-
intensity focused US ablation and non-thermal irreversible electroporation. However, these techniques are still
considered experimental. The best evidence base for these techniques exists for percutaneous microwave
ablation. In a study of 185 patients with a median follow-up of 40 months, the 5-year local progression rate was
3.2%, while 4.3% developed distant metastases [361]. Results appear to be favourable for cT1b tumours as
well [362]. Overall, current data on cryoablation, RFA and microwave ablation of cT1a renal tumours indicate
short-term equivalence with regards to complications, oncological and renal functional outcomes [363].

7.1.4.4.7 Summary of evidence and recommendation for therapeutic approaches as alternative to surgery

Summary of evidence LE
Most population-based analyses show a significantly lower cancer-specific mortality for patients 3
treated with surgery compared to non-surgical management.
In AS cohorts, the growth of small renal masses is low in most cases and progression to metastatic 3
disease is rare (1–2%).
Low quality studies suggest high disease recurrence rates after RFA of tumours > 3 cm and after 3
cryoablation of tumours > 4 cm.
Low quality studies suggest a higher local recurrence rate for TA therapies compared to PN, but the 3
quality of data does not allow definitive conclusions.

Recommendation Strength rating


Offer active surveillance (AS) or thermal ablation (TA) to frail and/or comorbid patients with Weak
small renal masses.
Perform a percutaneous renal mass biopsy prior to, and not concomitantly with, TA. Strong
When TA or AS are offered, discuss with patients about the harms/benefits with regards to Strong
oncological outcomes and complications.
Do not routinely offer TA for tumours > 3 cm and cryoablation for tumours > 4 cm. Weak

7.2 Treatment of locally advanced RCC


7.2.1 Introduction
In addition to the summary of evidence and recommendations outlined in Section 7.1 for localised RCC, certain
therapeutic strategies arise in specific situations for locally advanced disease.

7.2.2 Role of lymph node invasion in locally advanced RCC


In locally advanced RCC, the role of LND is still controversial. The only available RCT demonstrated no
survival benefit for patients undergoing LND but this trial mainly included organ-confined disease cases [248].
In the setting of locally advanced disease, several papers addressed the topic with contradictory results, as
did several systematic reviews. Bhindi et al. could not confirm any survival benefit in patients at high risk of
progression treated with LND [364]. More recently, Luo et al. reported a systematic review and meta-analyses
showing a survival benefit in patients with locally advanced disease treated with LND [365]. More specifically,
thirteen studies on patients with LND and non-LND were identified and included in the analysis. In the
subgroup of locally advanced RCC (cT3-T4NxM0), LND showed a significantly better OS rate in patients who
had undergone LND compared to those without LND (HR: 0.73, 95% CI: 0.60–0.90, p = 0.003).

7.2.2.1 Management of clinically negative lymph nodes (cN-) in locally advanced RCC
In case of cN-, the probability of finding pathologically confirmed LN metastases ranges between 0 and 25%,
depending mainly on primary tumour size and the presence of distant metastases [366]. In case of clinically
negative LNs (cN-) at imaging, removal of LNs is justified only if visible or palpable during surgery [367], at
least for staging, prognosis and follow-up implications although a benefit in terms of cancer control has not yet
been demonstrated [250, 364]. Whether to extend the LND also to retroperitoneal areas without cN+ remains
controversial [249].

7.2.2.2 Management of clinically positive lymph nodes (cN+) in locally advanced RCC
In case of cN+, the probability to find pathologically confirmed LN metastases ranges between 10.3% (cT1
tumours) up to 54.5% in case of locally advanced disease. In cN+, removal of visible and palpable nodes during
lymphadenectomy is always justified [367], at least for staging, prognosis and follow-up implications, although a
benefit in terms of cancer control has not yet been demonstrated [250, 364].

RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021 35


7.2.3 Management of locally advanced unresectable RCC
In case of locally advanced unresectable RCC, a multidisciplinary evaluation, including urologists, medical
oncologists and radiation therapists is suggested to maximise cancer control, pain control and the best
supportive care. In patients with non-resectable disease, embolisation can control symptoms, including visible
haematuria or flank pain [264, 265, 368]. The use of systemic therapy to downsize tumours is experimental and
cannot be recommended outside clinical trials.

7.2.4 Management of RCC with venous tumour thrombus


Tumour thrombus formation in RCC patients is a significant adverse prognostic factor. Traditionally, patients
with venous tumour thrombus undergo surgery to remove the kidney and tumour thrombus. Aggressive surgical
resection is widely accepted as the default management option for patients with venous tumour thrombus
[369-377].

7.2.4.1 The evidence base for surgery in patients with venous tumour thrombus
Data whether patients with venous tumour thrombus should undergo surgery is derived from case series only.
In one of the largest published studies a higher level of thrombus was not associated with increased tumour
dissemination to LNs, perinephric fat or distant metastasis [374]. Therefore, all patients with non-metastatic
disease and venous tumour thrombus, and an acceptable PS, should be considered for surgical intervention,
irrespective of the extent of tumour thrombus at presentation. The surgical technique and approach for each
case should be selected based on the extent of tumour thrombus.

7.2.4.2 The evidence base for different surgical strategies


A systematic review was undertaken which included only comparative studies on the management of venous
tumour thrombus in non-metastatic RCC [378, 379]. Only 5 studies were eligible for final inclusion, with a high
risk of bias across all studies.
Minimal access techniques resulted in significantly shorter operating time compared with
traditional median sternotomy [380, 381]. No significant differences in oncological and process outcomes were
observed between cardiopulmonary bypass with deep hypothermic circulatory arrest or partial bypass under
normothermia or single caval clamp without circulatory support [382].
No surgical method was shown to be superior for the excision of venous tumour thrombus. The
surgical method selected depended on the level of tumour thrombus and the grade of occlusion of the IVC
[378, 380-382]. The relative benefits and harms of other strategies and approaches regarding access to the IVC
and the role of IVC filters and bypass procedures remain uncertain.

7.2.4.3 Summary of evidence and recommendations for the management of RCC with venous tumour
thrombus

Summary of evidence LE
In patients with locally advanced disease, the survival benefit of LN dissection is unproven but LN 3
dissection has significant staging, prognosis and follow-up implications.
Low quality data suggest that tumour thrombus excision in non-metastatic disease may be beneficial. 3

Recommendations Strength rating


In patients with clinically enlarged lymph nodes (LNs), perform LN dissection for staging, Weak
prognosis and follow-up implications.
Remove the renal tumour and thrombus in case of venous involvement in non-metastatic Strong
disease.
In case of metastatic disease, discuss surgery within the context of a multidisciplinary team. Weak

7.2.5 Neoadjuvant and adjuvant therapy


Neoadjuvant therapy is currently under investigation and available in clinical trials.
There is currently no evidence from a recent systematic review (including ten retrospective studies
and two RCTs) that adjuvant radiation therapy increases survival [383].
Similarly, there is currently no evidence from randomised phase III trials that medical adjuvant
therapy offers a survival benefit. The impact on OS of adjuvant tumour vaccination in selected patients
undergoing nephrectomy for T3 renal carcinomas remains unconfirmed [384-388] (LE: 1b). Results from
prior adjuvant trials studying interferon-alpha (IFN-α) and interleukin-2 (IL-2) did not show a survival benefit
[389]. A similar observation was made in an adjuvant trial of girentuximab, a monoclonal antibody against
carboanhydrase IX (CAIX) (ARISER Study) [390].

36 RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021


At present, there is no OS data supporting the use of adjuvant VEGFR or mTOR inhibitors. Thus far, several
RCTs comparing VEGFR-TKI vs. placebo have been published. One of the largest adjuvant trials compared
sunitinib vs. sorafenib vs. placebo (ASSURE). Its interim results published in 2015 demonstrated no significant
differences in DFS or OS between the experimental arms and placebo [391]. The study published its updated
analysis on a subset of high-risk patients in 2018, which demonstrated 5-year DFS rates of 47.7%, 49.9%,
and 50.0%, respectively for sunitinib, sorafenib, and placebo (HR: 0.94 for sunitinib vs. placebo; and HR: 0.90,
97.5% CI: 0.71–1.14 for sorafenib vs. placebo), and 5-year OS of 75.2%, 80.2%, and 76.5% (HR: 1.06, 97.5%
CI: 0.78–1.45, p = 0.66, for sunitinib vs. placebo; and HR: 0.80; 97.5% CI: 0.58–1.11, p = 0.12 for sorafenib vs.
placebo). The results indicated that adjuvant therapy with sunitinib or sorafenib should not be given [392].
The PROTECT study included 1,135 patients treated with pazopanib (n = 571) vs. placebo
(n = 564) in a 1:1 randomisation [393]. The primary endpoint was amended after 403 patients received a
starting dose of pazopanib 800 mg vs. placebo, to DFS with pazopanib 600 mg. The primary analysis results
of DFS in the intention to treat (ITT) pazopanib 600 mg arm were not significant (HR: 0.86, 95% CI: 0.7–1.06,
p = 0.16). Disease-free survival in the ITT pazopanib 800 mg population was improved (HR: 0.69, 95%
CI: 0.51–0.94, p = 0.02). No benefit in OS was seen in the ITT pazopanib 600 mg population (HR: 0.79 [0.57–1.09,
p = 0.16]). A subset analysis of these studies suggests that full-dose therapy is associated with improved DFS.
Furthermore, no strong association of DFS with OS has been established [394, 395].
The ATLAS study, a randomised, double-blind phase III trial including patients receiving (1:1) oral
twice-daily axitinib 5 mg or placebo for ≤ 3 years, for a minimum of one year unless patients experienced a
recurrence, had a second primary malignancy, significant toxicity, or withdrew consent. The primary endpoint
was DFS. A total of 724 patients (363 vs. 361, for axitinib vs. placebo) were randomised. The trial was stopped
due to futility at a preplanned interim analysis at 203 DFS events. There was no significant difference in DFS
per independent review committee (IRC) (HR: 0.870, 95% CI: 0.660–1.147, p=0.3211). In the highest-risk
subpopulation, a 36% and 27% reduction in risk of a DFS event (HR; 95% CI) was observed per investigator
(HR: 0.641, 95% CI: 0.468–0.879, p = 0.0051) and by IRC (HR: 0.735, 95% CI: 0.525–1.028, p = 0.0704),
respectively. Overall survival data were not mature. Similar adverse events (AEs; 99% vs. 92%) and serious AEs
(19% vs. 14%), but more grade 3/4 AEs (61% vs. 30%) were reported for axitinib vs. placebo [396].

In contrast, the S-TRAC study included 615 patients randomised to either sunitinib or placebo [397]. The
results showed a benefit of sunitinib over placebo for DFS (HR: 0.76, 95% CI: 0.59–0.98, p = 0.03). Grade 3/4
toxicity in the study was 60.5% for patients receiving sunitinib, which translated into significant differences in
QoL for loss of appetite and diarrhoea. The study published its updated results in 2018; the results for DFS had
not changed significantly (HR: 0.74, 95% CI: 0.55–0.99, p = 0.04) and median OS was not reached in either arm
(HR: 0.92, 95% CI: 0.66–1.28, p = 0.6).
To date, the results of two RCTs on the role of adjuvant sorafenib (SORCE) and everolimus
(EVEREST) in patients with RCC are still awaited. Their findings may provide additional insight into the role of
adjuvant targeted therapy in RCC.

A recent meta-analysis of phase III randomised clinical trials on adjuvant TKIs in ccRCC was published [398].
In the overall population, the pooled HR of OS and DFS was 0.89 (95% CI: 0.76–1.04) and 0.84 (95% CI: 0.76–
0.93), respectively. In the low- and high-risk populations, the pooled DFS HR was 0.98 (95% CI: 0.82–1.17)
and 0.85 (95% CI: 0.75–0.97), respectively. Adjuvant use of TKIs does not appear to provide a statistically
significant OS benefit. However, a benefit in DFS has been observed in overall and high-risk populations,
suggesting that better selection of patients might be important for the evaluation of adjuvant therapies in RCC,
although these results must be balanced against significant toxicity.

In summary, there is currently a lack of proven benefits of adjuvant therapy with VEGFR-TKIs for patients
with high-risk RCC after nephrectomy. The European Medicines Agency (EMA) has not approved sunitinib for
adjuvant treatment of high-risk RCC in adult patients after nephrectomy.

Immune checkpoint inhibitors, designed to restore and enhance immune activity against cancer cells, have
shown impressive efficacy in the metastatic setting. Several trials have tested these agents in metastatic
RCC, leading to a still-ongoing revolution in the treatment pathway. The inclusion of these drugs in
clinical practice has led to a third generation of adjuvant studies on immune checkpoint inhibitors. These
include the programmed death receptor-1 inhibitors nivolumab (PROSPER; NCT03055013), pembrolizumab
(KEYNOTE-564; NCT03142334), as well as the programmed death ligand-1 inhibitors atezolizumab
(IMmotion010; NCT03024996) and durvalumab (RAMPART [Renal Adjuvant MultiPle Arm Randomised Trial];
NCT03288532). Recruitment for most of these studies is still ongoing and results are awaited as of 2022.

RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021 37


7.2.5.1 Summary of evidence and recommendations for adjuvant therapy

Summary of evidence LE
Adjuvant therapy does not improve survival after nephrectomy. 1b
In one single RCT, in selected high-risk patients, adjuvant sunitinib improved disease-free survival 1b
(DFS) but not overall survival (OS).
Adjuvant sorafenib, pazopanib, everolimus, girentuximab or axitinib does not improve DFS or OS after 1b
nephrectomy.
Adjuvant RCTs are ongoing to evaluate the benefit of adjuvant immunotherapy after nephrectomy in 1b
high-risk patients.

Recommendations Strength rating


Do not offer adjuvant therapy with sorafenib, pazopanib, everolimus, girentuximab or Strong
axitinib.
Do not offer adjuvant sunitinib following surgically resected high-risk clear-cell renal cell Weak
carcinoma.

7.3 Advanced/metastatic RCC


7.3.1 Local therapy of advanced/metastatic RCC
7.3.1.1 Cytoreductive nephrectomy
Tumour resection is potentially curative only if all tumour deposits are excised. This includes patients with the
primary tumour in place and single- or oligometastatic resectable disease. For most patients with metastatic
disease, cytoreductive nephrectomy (CN) is palliative and systemic treatments are necessary. In a combined
analysis of two RCTs comparing CN+ IFN-based immunotherapy vs. IFN-based immunotherapy only, increased
long-term survival was found in patients treated with CN [399].
However, IFN-based immunotherapy is no longer relevant in contemporary clinical practice. In order
to investigate the role and sequence of CN in the era of targeted therapy, a structured literature assessment
was performed to identify relevant RCTs and systematic reviews published between July 1st - June 30th 2019.
Two RCTs [400, 401] and a narrative systematic review were identified [402]. The narrative systematic review
included both RCTs and 10 non-randomised studies. CARMENA, a phase III non-inferiority RCT investigating
immediate CN followed by sunitinib vs. sunitinib alone, showed that sunitinib alone was not inferior to CN
followed by sunitinib with regard to OS [403]. The trial included 450 patients with metastatic ccRCC of
intermediate- and MSKCC poor-risk of whom 226 were randomised to immediate CN followed by sunitinib
and 224 to sunitinib alone. Patients in both arms had a median of two metastatic sites. Patients in both arms
had a tumour burden of a median/mean of 140 mL of measurable disease by Response Evaluation Criteria In
Solid Tumours (RECIST) 1.1, of which 80 mL accounted for the primary tumour. The study did not reach the
full accrual of 576 patients and the Independent Data Monitoring Commission (IDMC) advised the trial steering
committee to close the study. In an ITT analysis after a median follow-up of 50.9 months, median OS with CN
was 13.9 months vs. 18.4 months with sunitinib alone (HR: 0.89, 95% CI: 0.71–1.10). This was found in both
risk groups. For MSKCC intermediate-risk patients (n = 256) median OS was 19.0 months with CN and 23.4
months with sunitinib alone (HR: 0.92, 95% CI: 0.60–1.24) and for MSKCC poor risk (n = 193) 10.2 months and
13.3 months, respectively (HR: 0.86, 95% CI: 0.62–1.17). Non-inferiority was also found in two per-protocol
analyses accounting for patients in the CN arm who either did not undergo surgery (n = 16) or did not receive
sunitinib (n = 40), and patients in the sunitinib-only arm who did not receive the study drug (n = 11). Median
PFS in the ITT population was 7.2 months with CN and 8.3 months with sunitinib alone (HR: 0.82, 95% CI:
0.67–1.00). The clinical benefit rate, defined as disease control beyond 12 weeks was 36.6% with CN and
47.9% with sunitinib alone (p = 0.022). Of note, 38 patients in the sunitinib-only arm required secondary CN
due to acute symptoms or for complete or near-complete response. The median time from randomisation to
secondary CN was 11.1 months.

The randomised EORTC SURTIME study revealed that the sequence of CN and sunitinib did not affect PFS
(HR: 0.88, 95% CI: 0.59–1.37, p = 0.569). The trial accrued poorly and therefore results are mainly exploratory.
However, in secondary endpoint analysis a strong OS benefit was observed in favour of the deferred CN
approach in the ITT population with a median OS of 32.4 (range 14.5–65.3) months in the deferred CN arm
vs. 15.0 (9.3–29.5) months in the immediate CN arm (HR: 0.57, 95% CI: 0.34–0.95, p = 0.032). The deferred
CN approach appears to select out patients with inherent resistance to systemic therapy [404]. This confirms
previous findings from single-arm phase II studies [404, 405]. Moreover, deferred CN and surgery appear safe
after sunitinib which supports the findings, with some caution, of the only available RCT.

38 RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021


In patients with poor PS or IMDC poor risk, small primaries and high metastatic volume and/or a
sarcomatoid tumour, CN is not recommended [406]. These data are confirmed by CARMENA [403] and upfront
pre-surgical VEGFR-targeted therapy followed by CN seems to be beneficial [407].

Meanwhile first-line therapy recommendations for patients with their primary tumour in place have changed
to immune checkpoint inhibitor combination therapy (see Section 7.4.2.4) with sunitinib and other VEGFR-
TKI monotherapies reserved for those who cannot tolerate immune checkpoint inhibitor (ICI) combination or
have no access to these drugs. High level evidence regarding CN is not available for ICI combinations but up
to 30% of patients with primary metastatic disease, treated with their tumour in place, were included in the
pivotal ICI combination trials (Table 7.1). The subgroup HRs, where available, suggest better outcomes for the
ICI combination compared to sunitinib monotherapy. In mRCC patients without a need for immediate drug
treatment, a recent systematic review evaluating effects of CN demonstrated an OS advantage of CN [402].
These data were supported by a nation-wide registry study showing that patients selected for primary CN had
a significant OS advantage across all age groups [408].

Table 7.1: Key trails on immune checkpoint inhibitor combinations for primary metastatic disease

Trial Drug Number and % of Number of patients Subgroup analyses


combination patients treated treated with the primary (HR with 95% CIs)
with primary tumour in place
tumour in place (ICI combination vs.
sunitinib)
ICI sunitinib PFS OS
combination
CheckMate 214 ipilimumab + 187/847 (22%) 84 103 NA 0.63
[409] nivolumab (0.42–0.94)
CheckMate 9ER cabozantinib + 196/651 (30.1%) 101 95 0.63 0.79
[410] nivolumab (0.43–0.92) (0.48–1.29)
Javelin 101 axitinib + 179/886 (20.2%) 90 89 0.75 NA
[411] avelumab (0.48–1.65)
KEYNOTE-426 axitinib + 143/861 (16.6%) 73 70 0.68 0.57
[412] pembrolizumab (0.45–1.03) (0.36–0.89)
CI = confidence interval; HR = hazard ratio; ICI = immune checkpoint inhibitor; NA = not available;
PFS = profession-free survival; OS = overall survival.

The results of CARMENA and SURTIME demonstrated that patients who require systemic therapy benefit from
immediate drug treatment. While randomised trials to investigate deferred vs. no cytoreductive nephrectomy
with ICI and ICI combinations are ongoing, the exploratory results from the ICI combination trials demonstrate
that the respective IO+IO or TKI+IO combinations have a superior effect on the primary tumour and metastatic
sites when compared to sunitinib alone (Table 7.1). In accordance with the CARMENA and SURTIME data
this suggests that mRCC patients and IMDC intermediate- and poor-risk groups with their primary tumour in
place should be treated with upfront IO-based combinations. In patients with a clinical response to IO-based
combinations, a subsequent CN may be considered.

7.3.1.1.1 Embolisation of the primary tumour


In patients unfit for surgery or with non-resectable disease, embolisation can control symptoms including
visible haematuria or flank pain [264, 265, 368] (see recommendations Section 7.1.2.2.4).

7.3.1.1.2 Summary of evidence and recommendations for local therapy of advanced/metastatic RCC

Summary of evidence LE
Deferred CN with pre-surgical sunitinib in intermediate-risk patients with cc-mRCC shows a survival 2b
benefit in secondary endpoint analyses and selects out patients with inherent resistance to systemic
therapy.
Sunitinib alone is non-inferior compared to immediate CN followed by sunitinib in patients with 1a
MSKCC intermediate and poor risk who require systemic therapy with VEGFR-TKI.
Cytoreductive nephrectomy in patients with simultaneous complete resection of a single metastasis or 3
oligometastases may improve survival and delay systemic therapy.

RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021 39


Patients with MSKCC or IMDC poor risk (≥ 4 risk factors) do not benefit from local therapy. 1a
Patients with their primary tumour in place treated with ICI-based combination therapy have better 2b
PFS and OS in exploratory subgroup analyses compared to treatment with sunitinib.

Recommendations Strength rating


Do not perform cytoreductive nephrectomy (CN) in MSKCC poor-risk patients. Strong
Do not perform immediate CN in intermediate-risk patients who have an asymptomatic Weak
synchronous primary tumour and require systemic therapy.
Start systemic therapy without CN in intermediate-risk patients who have an asymptomatic Weak
synchronous primary tumour and require systemic therapy.
Discuss delayed CN with patients who derive clinical benefit from systemic therapy. Weak
Perform immediate CN in patients with a good performance status who do not require Weak
systemic therapy.
Perform immediate CN in patients with oligometastases when complete local treatment of Weak
the metastases can be achieved.

7.3.2 Local therapy of metastases in metastatic RCC


A systematic review of the local treatment of metastases from RCC in any organ was undertaken [413].
Interventions included metastasectomy, various radiotherapy modalities, and no local treatment. The outcomes
assessed were OS, CSS and PFS, local symptom control and adverse events. A risk-of-bias assessment
was conducted [414]. Of the 2,235 studies identified only sixteen non-randomised comparative studies were
included.

Eight studies reported on local therapies of RCC-metastases in various organs [415-422]. This included
metastases to any single organ or multiple organs. Three studies reported on local therapies of RCC
metastases in bone, including the spine [423-425], two in the brain [426, 427] and one each in the liver [428]
lung [429] and pancreas [430]. Three studies were published as abstracts only [418, 420, 429]. Data were too
heterogeneous to meta-analyse. There was considerable variation in the type and distribution of systemic
therapies (cytokines and VEGF-inhibitors) and in reporting the results.

7.3.2.1 Complete versus no/incomplete metastasectomy


A systematic review, including only 8 studies, compared complete vs. no and/or incomplete metastasectomy of
RCC metastases in various organs [415-422]. In one study complete resection was achieved in only 45% of the
metastasectomy cohort, which was compared with no metastasectomy [422]. Non-surgical modalities were not
applied. Six studies [416-418, 420-422] reported a significantly longer median OS or CSS following complete
metastasectomy (the median value for OS or CSS was 40.75 months, range 23–122 months) compared with
incomplete and/or no metastasectomy (the median value for OS or CSS was 14.8 months, range 8.4–55.5
months). Of the two remaining studies, one [415] showed no significant difference in CSS between complete
and no metastasectomy, and one [419] reported a longer median OS for metastasectomy albeit no p-value was
provided.
Three studies reported on treatment of RCC metastases in the lung [429], liver [428], and pancreas
[430], respectively. The lung study reported a significantly higher median OS for metastasectomy vs. medical
therapy only for both targeted therapy and immunotherapy. Similarly, the liver and pancreas study reported a
significantly higher median OS and 5-year OS for metastasectomy vs. no metastasectomy.

7.3.2.2 Local therapies for RCC bone metastases


Of the three studies identified, one compared single-dose image-guided radiotherapy (IGRT) with
hypofractionated IGRT in patients with RCC bone metastases [425]. Single-dose IGRT (≥ 24 Gy) had a
significantly better 3-year actuarial local PFS rate, also shown by Cox regression analysis. Another study
compared metastasectomy/curettage and local stabilisation with no surgery of solitary RCC bone metastases
in various locations [423]. A significantly higher 5-year CSS rate was observed in the intervention group.
After adjusting for prior nephrectomy, gender and age, multi-variable analysis still favoured metastasectomy/
curettage and stabilisation. A third study compared the efficacy and durability of pain relief between single-
dose stereotactic body radiotherapy (SBRT) and conventional radiotherapy in patients with RCC bone
metastases to the spine [424]. Pain, ORR, time-to-pain relief and duration of pain relief were similar.

7.3.2.3 Local therapies for RCC brain metastases


Two studies on RCC brain metastases were included. A three-armed study compared stereotactic radiosurgery
(SRS) vs. whole brain radiotherapy (WBRT) vs. SRS and WBRT [426]. Each group was further subdivided into

40 RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021


recursive partitioning analysis (RPA) classes I to III (I favourable, II moderate and III poor patient status). Two-
year OS and intra-cerebral control were equivalent in patients treated with SRS alone and SRS plus WBRT.
Both treatments were superior to WBRT alone in the general study population and in the
RPA subgroup analyses. A comparison of SRS vs. SRS and WBRT in a subgroup analysis of RPA class I
showed significantly better 2-year OS and intra-cerebral control for SRS plus WBRT based on only three
participants. The other study compared fractionated stereotactic radiotherapy (FSRT) with metastasectomy and
conventional radiotherapy or conventional radiotherapy alone [427]. Several patients in all groups underwent
alternative surgical and non-surgical treatments after initial treatment. One-, two- and 3-year survival rates were
higher but not significantly so for FSRT as for metastasectomy and conventional radiotherapy, or conventional
radiotherapy alone. Fractionated stereotactic radiotherapy did not result in a significantly better 2-year local
control rate compared with metastasectomy plus conventional radiotherapy.

7.3.2.4 Embolisation of metastases


Embolisation prior to resection of hypervascular bone or spinal metastases can reduce intra-operative blood
loss [431]. In selected patients with painful bone or paravertebral metastases, embolisation can relieve
symptoms [432] (see recommendation Section 7.1.2.2.4).

7.3.2.5 Adjuvant treatment in cM0 patients after metastasectomy


Patients after metastasectomy and no evidence of disease (cM0) have a high risk of relapse. Recent attempts
to reduce RFS by offering adjuvant TKI treatment after metastasectomy did not demonstrate an improvement
in RFS. In a recent phase II trial 129 patients were randomised to either pazopanib 800 mg daily vs. placebo
for 52 weeks. The primary study endpoint of a 42% DFS improvement from 25% to 45% at three years was
not met. Hazard ratio for DFS in pazopanib vs. placebo-treated patients was 0.85 (0.55–1.31), p = 0.47 [433].
A second phase II trial randomised 69 ccRCC patients after metastasectomy and no evidence of disease to
either sorafenib (400 mg twice daily) or observation. The study was terminated early due to slow accrual and
the availability of new agents and multimodal treatment options, including surgery or a locoregional approach.
The primary endpoint of RFS was not reached with a RFS of 21 months in the sorafenib arms vs. 37 months in
the observation arm (p = 0.404) [434].

7.3.2.6 Summary of evidence and recommendations for local therapy of metastases in metastatic RCC

Summary of evidence LE
All studies included in the Panel systematic review were retrospective non-randomised comparative 3
studies, resulting in a high risk of bias associated with non-randomisation, attrition, and selective
reporting.
With the exception of brain and possibly bone metastases, metastasectomy remains by default the 3
only local treatment for most sites.
Retrospective comparative studies consistently point towards a benefit of complete metastasectomy 3
in mRCC patients in terms of overall survival, cancer-specific survival and delay of systemic therapy.
Radiotherapy to bone and brain metastases from RCC can induce significant relief from local 3
symptoms (e.g. pain).
Tyrosine kinase inhibitors treatment after metastasectomy in patients with no evidence of disease did 1b
not improve RFS when compared to placebo or observation.

Recommendations Strength rating


To control local symptoms, offer ablative therapy, including metastasectomy, to patients Weak
with metastatic disease and favourable disease factors and in whom complete resection is
achievable.
Offer stereotactic radiotherapy for clinically relevant bone or brain metastases for local Weak
control and symptom relief.
Do not offer tyrosine kinase inhibitor treatment to mRCC patients after metastasectomy and Strong
no evidence of disease.

7.4 Systemic therapy for advanced/metastatic RCC


7.4.1 Chemotherapy
Chemotherapy has proven to be generally ineffective in the treatment of RCC but can be offered in rare
patients, with the exception of collecting duct and medullary carcinoma [435].

RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021 41


7.4.1.1 Recommendation for systemic therapy in advanced/metastatic RCC

Recommendation Strength rating


Do not offer chemotherapy to patients with metastatic renal cell carcinoma. Strong

7.4.2 Immunotherapy
7.4.2.1 IFN-α monotherapy and combined with bevacizumab
All studies comparing targeted drugs to IFN-α monotherapy therapy showed superiority for sunitinib,
bevacizumab plus IFN-α, and temsirolimus [436-439]. Interferon-α has been superseded by targeted therapy in
cc-mRCC.

Table 7.2: The Metastatic Renal Cancer Database Consortium (IMDC) risk model [440]*

Risk factors** Cut-off point used


Karnofsky performance status < 80%
Time from diagnosis to treatment < 12 months
Haemoglobin < Lower limit of laboratory reference range
Corrected serum calcium > 10.0 mg/dL (2.4 mmol/L)
Absolute neutrophil count (neutrophilia) > upper limit of normal
Platelets (thrombocytosis) > upper limit of normal
*The MSKCC (Motzer) criteria are also widely used in this setting [225].
**Favourable (low) risk, no risk factors; intermediate risk, one or two risk factors; poor (high) risk, three to six risk
factors.

7.4.2.2 Interleukin-2
Interleukin-2 has been used to treat mRCC since 1985 with response rates ranging from 7–27% [439, 441, 442].
Complete and durable responses have been achieved with high-dose bolus IL-2, however, this can be achieved
at less toxicity with immune checkpoint inhibitor combination therapy and IL-2 is no longer widely used.

7.4.2.3 Immune checkpoint blockade


7.4.2.3.1 Immuno-oncology monotherapy
Immune checkpoint blockade with monoclonal antibodies targets and blocks the inhibitory T-cell receptor PD-1
or cytotoxic T-lymphocyte-associated antigen 4 (CTLA-4)-signalling to restore tumour-specific T-cell immunity
[443]. Immune checkpoint inhibitor monotherapy has been investigated as second- and third-line therapy. A
phase III trial of nivolumab vs. everolimus after one or two lines of VEGF-targeted therapy for mRCC with a
clear cell component (CheckMate 025, NCT01668784) reported a longer OS, better QoL and fewer grade 3 or
4 adverse events with nivolumab than with everolimus [444]. Nivolumab has superior OS to everolimus (HR:
0.73, 95% CI: 0.57–0.93, p < 0.002) in VEGF-refractory RCC with a median OS of 25 months for nivolumab
and 19.6 months for everolimus with a 5-year OS probability of 26% vs. 18% [445] (LE: 1b). Patients who had
failed multiple lines of VEGF-targeted therapy were included in this trial making the results broadly applicable.
The trial included 15% MSKCC poor-risk patients. There was no PFS advantage with nivolumab despite the OS
advantage. Progression-free survival does not appear to be a reliable surrogate of outcome for PD-1 therapy in
RCC. Currently PD-L1 biomarkers are not used to select patients for this therapy.

There are no RCTs supporting the use of single-agent immune checkpoint blockade in treatment-naive
patients. Randomised phase II data for atezolizumab vs. sunitinib showed a HR of 1.19 (95% CI: 0.82–1.71)
which did not justify further assessment of atezolizumab as single agent as first-line treatment option in this
group of patients, despite high complete response rates in the biomarker-positive population [446]. Single-arm
phase II data for pembrolizumab from the KEYNOTE-427 trial show high response rates of 38% (up to 50% in
PD-L1+ patients), but a PFS of 8.7 months (95% CI: 6.7–12.2) [447]. Based on these results and in the absence
of randomised phase III data, single-agent checkpoint inhibitor therapy is not recommended as an alternative in
a first-line therapy setting.

7.4.2.4 Immunotherapy/combination therapy


The phase III trial CheckMate 214 (NCT 02231749) showed a superiority of nivolumab and ipilimumab over
sunitinib. The primary endpoint population focused on the IMDC intermediate- and poor-risk population where
the combination demonstrated an OS benefit (HR: 0.63, 95% CI: 0.44–0.89) which led to regulatory approval
[409] and a paradigm shift in the treatment of mRCC [1]. Results from CheckMate 214 further established that

42 RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021


the combination of ipilimumab and nivolumab was associated with higher response rates (RR) (39% in the
ITT population), complete response rates (8% in the ITT population [central radiology review]) and duration
of response compared to sunitinib. Progression-free survival did not achieve the pre-defined endpoint. The
exploratory analysis of OS data in the PD-L1-positive population was 0.45 (95% CI: 0.29–0.41).
A recent update with 48-month data shows ongoing benefits for the immune combination with
independently assessed complete response rates of 10% and a HR for OS in the IMDC intermediate- and
poor-risk group of 0.65 (0.54–0.78). The 48-months OS propability was 50% for ipilimumab plus nivolumab
vs. 39% for sunitinib, respectively [448]. The IMDC good-risk group continues to perform better with sunitinib
although this appears less marked than in earlier analyses (HR for OS: 0.93 [95% CI: 0.62–1.40]) [448].

Nivolumab plus ipilimumab was associated with 15% grade 3-5 toxicity including 1.5% treatment-related
deaths. It should therefore be administered in centres with experience of immune combination therapy and
appropriate supportive care within the context of a multidisciplinary team (LE: 4). PD-L1 biomarker is currently
not used to select patients for therapy.
The frequency of steroid use has generated controversy and further analysis, as well as real world
data, are required. For these reasons the Panel continues to recommend ipilimumab and nivolumab in the
intermediate- and poor-risk population.

The KEYNOTE-426 trial (NCT02853331 reported results for the combination of axitinib plus pembrolizumab
vs. sunitinib in 861 treatment-naive cc-mRCC patients [449]. Overall survival and PFS assessed by central
independent review in the ITT population were the co-primary endpoints. Response rates and assessment in
the PD-L1-positive patient population were secondary endpoints. With a median follow-up of 12.8 months, at
first interim analysis both primary endpoints were reached, The median PFS in the pembrolizumab plus axitinib
arm was 15.1 months vs. 11.1 in the sunitinib arm (HR: 0.69, 95% CI: 0.57–0.84, p < 0.001). Median OS has
not been reached in either arm, but the risk of death was 47% lower in the axitinib plus pembrolizumab arm
when compared to the sunitinib arm (OS HR: 0.53, 95% CI: 0.38–0.74, p < 0.0001). Response rates were also
higher in the experimental arm (59.3% vs. 35.7%). Efficacy occurred irrespective of IMDC group and PD-L1
status. Treatment-related AEs (≥ grade 3) occurred in 63% of patients receiving axitinib and pembrolizumab vs.
58% of patients receiving sunitinib. Treatment-related deaths occurred in approximately 1% in both arms.

A recent update of KEYNOTE-426 with a minimum follow-up of 23.4 months (median 30.6 months)
demonstrated an ongoing OS benefit for axitinib plus pembrolizumab in the ITT population (HR: 0.68, 95% CI:
0.55–0.85, p < 0.001) and PFS benefit (HR: 0.71, 95% CI: 0.60–0.84, p < 0.0001 which was across all IMDC
subgoups for PFS, while OS was similar between axitinib plus pembrolizumab vs. sunitinib in the favourable
subgroup with an OS benefit in the IMDC intermediate- and poor-risk groups. The complete response rate by
independent review was 9% in the pembrolizumab plus axitinib arm and 3% in the sunitinib arm [450].

The phase III CheckMate 9ER trial randomised 651 patients to nivolumab plus cabozantinib (n = 323) or
vs. sunitinib (n = 328) in treatment-naive cc-mRCC patients. The primary endpoint of PFS assessed by
central independent review in the ITT population was siginifcantly prolonged for nivolumab plus cabozantinib
(16.6 months) vs. sunitinib (8.3 months, HR: 0.51, 95% CI: 0.41–0.64, p < 0.0001). The nivolumab/cabozantinib
combination also demonstrated a significant OS benefit in the secondary endpoint compared with sunitinib
(HR: 0.60, CI: 0.40–0.89, p = 0.0010) after a median follow-up of 18.1 months. The independantly assessed
ORR was 55.7% vs. 27.1% with a complete response rate of 8% for nivolumab plus cabozantinib vs. 4.6%
with sunitinib. The efficacy was observed independent of IMDC group and PD-L1 status. Treatment-related AEs
(≥ grade 3) occurred in 61% of patients receiving cabozantinib and nivolumab vs. 51% of patients receiving
sunitinib. Treatment-related deaths occurred in one patient in the nivolumab/cabozantinib arm and in two
patients in the sunitinib arm.

Recently, the randomised phase III trial CLEAR (Lenvatinib/Everolimus or Lenvatinib/Pembrolizumab Versus
Sunitinib Alone as Treatment of Advanced Renal Cell Carcinoma) was published [451]. CLEAR randomised a
total of 1,069 patients (in a 1:1:1 ratio) to lenvatinib plus pembrolizumab (n = 355) vs. lenvatinib plus everolimus
(n = 357) vs. sunitinib (n = 357). The trial reached its primary endpoint of independently assessed PFS at a
median of 23.9 vs. 9.2 months, for lenvatinib plus pembrolizumab vs. sunitinib, respectively (HR: 0.39, 95%
CI: 0.32–0.49, p < 0.001). Overall survival siginificantly improved with lenvatinib plus pembrolizumab vs.
sunitinib (HR: 0.66, 95% CI: 0.49–0.88, p = 0.005). Objective response for lenvatinib plus pembrolizumab
was 71% with 16% of the patients having a complete remission. Efficacy was observed across all IMDC
risk groups, independently of PD-L1 status. Treatment-related AEs of grade 3 and higher with lenvatinib
plus pembrolizumab were 72%. Treatment-related death occurred in four patients in the lenvatinib plus
pembrolizumab arm and in one patient in the sunitinib arm.

RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021 43


The JAVELIN trial investigated 886 patients in a phase III RCT of avelumab plus axitinib vs. sunitinib [411]. The
trial met one of its co-primary endpoints (PFS in the PD-L1-positive population at first interim analysis [median
follow up 11.5 months]). Hazard ratios for PFS and OS in the ITT population were 0.69 (95% CI: 0.56–0.84) and
0.78 (95% CI: 0.55–1.08), respectively. The same applies to the atezolizumab/bevacizumab combination which
also achieved a PFS advantage over sunitinib in the PD-L1-positive population at interim analysis and ITT
(HR: 0.74, 95% CI: 0.57–0.96), but has not yet shown a significant OS advantage (HR: 0.81, 95% CI: 0.63–1.03)
[452]. Final OS results are awaited and the combination cannot currently be recommended.

Table 7.3: First line immune checkpoint inhibitor combination trials for clear-cell RCC
Cross trial comparison is not recommended and should occur with caution

Study N Experimental Primary Risk PFS (mo) OS (mo)


arm endpoint groups Median (95% CI) Median (95% CI)
HR HR
KEYNOTE-426 861 Pembrolizumab PFS and IMDC (ITT) (ITT)
NCT02853331 200 mg. IV Q3W OS in the FAV 31% PEMBRO + AXI: PEMBRO + AXI: NR
Median plus axitinib 5 mg. ITT by IMD 56% 15.4 (12.7-18.9) SUN: 35.7 (33.3-NE)
follow-up PO BID BICR POOR 13% SUN: 11.1 (9.1-12.5)
30.6 months vs.
[449, 450] sunitinib 50 mg MSKCC HR: 0.71 HR: 0.68
PO QD 4/2 wk Not (95% CI: 0.60, 0.84) (95% CI: 0.55-0.85)
determined p < 0.0001 p = 0.0003
JAVELIN 101 886 Avelumab 10 mg/ PFS in the IMDC (PD-L1+) (PD-L1+)
NCT02684006 kg IV Q2W plus PD-L1+ FAV 22% AVE + AXI: AVE + AXI: NR
Median axitinib, 5 mg PO population IMD 62% 13.8 (10.1-20.7) SUN: 28.6 (27.4-NE)
follow-up BID and OS in POOR 16% SUN: 7.0 (5.7-9.6)
19 months vs. the ITT by HR: 0.83
[411, 453] sunitinib 50 mg BICR MSKCC HR: 0.62 (95% CI: 0.60-1.15)
PO QD 4/2 wk FAV 23% (95% CI: 0.49, 0.78) p = 0.1301
IMD 66% p < 0.0001
POOR 12%
Immotion 151 915 Atezolizumab PFS in the IMDC (PD-L1+) (ITT)
NCT02420821 1200 mg fixed PD-L1+ Not ATEZO + BEV: ATEZO + BEV:
Median dose IV plus population determined 11.2 (8.9-15.0) 33.6 (29.0-NE)
follow-up bevacizumab 15 and OS SUN: 7.7 (6.8-9.7) SUN: 34.9 (27.8-NE)
24 months mg/kg IV on days in the ITT MSKCC
[452] 1 and 22 of each by IR FAV 20% HR: 0.74 HR: 0.93
42-day cycle IMD 69% (95% CI: 0.57, 0.96) (95% CI: 0.76-1.14)
vs. POOR 12% p = 0.0217 p 0.4751
sunitinib 50 mg.
PO QD 4/2 wk
Checkmate 1096 Nivolumab 3 PFS and IMDC (IMDC IMD/poor) (IMDC IMD/poor)
214 mg/kg plus OS in the FAV 23% NIVO + IPI: NIVO + IPI:
NCT02231749 ipilimumab 1 mg/ IMDC IMD 61% 11.2 (8.4-16.1) 48.1 (35.6-NE)
Median kg IV Q3W for inter­ POOR 17% SUN: 8.3 (7.0-10.8) SUN:
follow-up 4 doses then mediate 26.6 (22.1-33.5)
48 months nivolumab 3 mg/ and poor MSKCC HR: 0.74
[409, 448] kg IV Q2W population Not (95% CI: 0.62, 0.88) HR: 0.65
vs. by BICR determined (0.54-0.78)
sunitinib 50 mg. p < 0.0001
PO QD 4/2 wk
CheckMate 651 Nivolumab 240 PFS in IMDC (ITT) (ITT)
9ER mg fixed dose IV the ITT by FAV 22% NIVO + CABO: NIVO + CABO:
NCT03141177 every 2 wk plus BICR IMD 58% 16.6 (12.5-24.9) NR (NE)
Median cabozantinib POOR 20% SUN: 8.3 (7.0-9.7) SUN: NR (22.6-NE)
follow-up 40 mg PO daily
18.1 months vs. MSKCC HR: 0.51 HR: 0.60
[410] sunitinib 50 mg Not (95% CI: 0.41-0.64) (98.9% CI: 0.40-0.89)
PO QD 4/2 wk determined p < 0.0001 p = 0.0010

44 RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021


CLEAR 712 Pembrolizumab PFS in IMDC (ITT) (ITT)
NCT02811861 200 mg IV Q3W the ITT by FAV 31% PEMBRO + LEN: PEMBRO + LEN:
Median plus lenvatinib BIRC IMD 59% 23.9 (20.8-27.7) NR (33.6-NE)
follow-up 20 mg PO QD POOR 9% SUN: 9.2 (6.0-11.0) SUN: NR (NE-NE)
26.6 months vs. NE 1%
[451] sunitinib 50 mg MSKCC HR: 0.39 HR: 0.66
PO QD 4/2 wk FAV 27% (95% CI: 0.32-0.49) (95% CI: 0.49-0.88)
IMD 64% p > 0.001 p = 0.005
POOR 9%
ATEZO = atezolizumab; AVE = avelumab; AXI = axitinib; BEV = bevacizumab; BICR = blinded independent
central review; BID = twice a day; CABO = cabozantinib; CI = confidence interval; FAV = favourable;
R = hazard ratio; IPI = ipilimumab; IMD = intermediate; IMDC = Metastatic Renal Cancer Database Consortium;
IR = investigator review; ITT = intention-to-treat; IV = intravenous; LEN = lenvatinib; mo = months;
MSKCC = Memorial Sloan Kettering Cancer Center; NE = non-estimable; NR = not reached; NIVO = nivolumab;
OS = overall survival; PEMBRO = pembrolizumab; PFS = profession-free survival; PO = by mouth; BID = twice
a day; QD = once a day; Q2W = every 2 weeks; Q3W = every 3 weeks; SUN = sunitinib; wk = weeks.

Patients who stop nivolumab plus ipilimumab because of toxicity require expert guidance and support from a
multidisciplinary team before re-challenge can occur (LE: 1). Patients who do not receive the full four doses of
ipilimumab due to toxicity should continue on single-agent nivolumab, where safe and feasible (LE: 4).
Treatment past progression with nivolumab plus ipilimumab can be justified but requires close
scrutiny and the support of an expert multidisciplinary team [454, 455] (LE: 1).
Patients who stop TKI and IO due to immune-related toxicity can receive single-agent TKI once
the adverse event has resolved (LE: 1). Adverse event management, including transaminitis and diarrhoea,
require particular attention as both agents may be causative. Expert advice should be sought on re-challenge
of immune checkpoint inhibitors after significant toxicity (LE: 4). Treatment past progression on axitinib plus
pembrolizumab or nivolumab plus cabozantinib requires careful consideration as it is biologically distinct from
treatment past progression on ipilimumab and nivolumab.

Generally, the Panel is of the opinion that nivolumab plus ipilimumab, pembrolizumab plus axitinib and
nivolumab plus cabozantinib should be administered in centres with experience of immune combination
therapy and appropriate supportive care within the context of a multidisciplinary team (LE: 4).

7.4.2.5 Summary of evidence and recommendations for immunotherapy in metastatic RCC

Summary of evidence LE
Interferon-α monotherapy is inferior to VEGF-targeted therapy or mTOR inhibition in mRCC. 1b
Nivolumab leads to superior OS compared to everolimus in patients failing one or two lines of VEGF- 1b
targeted therapy.
The combination of nivolumab and ipilimumab in treatment-naive patients with clear-cell-mRCC 1b
(cc-mRCC) of IMDC intermediate- and poor-risk demonstrated overall survival (OS) and objective
response rate (ORR) benefits compared to sunitinib.
The combination of pembrolizumab plus axitinib, lenvatinib plus pembrolizumab and nivolumab plus 1b
cabozantinib in treatment-naive patients with cc-mRCC across all IMDC risk group demonstrated
PFS, OS and ORR benefits compared to sunitinib.
Currently, PD-L1 expression is not used for patient selection. 2b
Axitinib, cabozantinib or lenvatinib can be continued if immune-related adverse events result 4
in cessation of axitinib plus pembrolizumab, cabozantinib plus nivolumab or lenvatinib plus
pembrolizumab. Re-challenge with immunotherapy requires expert support.
Patients who do not receive the full 4 doses of ipilimumab due to toxicity should continue on single- 4
agent nivolumab, where safe and feasible. Re-challenge with combination therapy requires expert
support.
Treatment past progression can be justified but requires close scrutiny and the support of an expert 1b
multidisciplinary team.
Nivolumab plus ipilimumab, pembrolizumab plus axitinib, nivolumab plus cabozantinib and lenvatinib 4
plus pembrolizumab should be administered in centres with experience of immune combination
therapy and appropriate supportive care within the context of a multidisciplinary team.

RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021 45


The combination of nivolumab plus ipilimumab in the IMDC intermediate- and poor-risk population of 2b
treatment-naive patients with cc-mRCC leads to superior survival compared to sunitinib while OS was
higher in IMDC good-risk patients with sunitinib.
Nivolumab plus ipilimumab was associated with 15% grade 3-5 toxicity and 1.5% treatment-related 1b
deaths.

Recommendations Strength rating


Offer pembrolizumab plus axitinib, lenvatinib plus pembrolizumab or nivolumab plus Strong
cabozantinib to treatment-naive patients in clear-cell metastatic renal cell carcinoma
(cc-mRCC).
Offer ipilimumab plus nivolumab to treatment-naive patients with IMDC intermediate- and Strong
poor-risk cc-mRCC.
Administer nivolumab plus ipilimumab, pembrolizumab plus axitinib, lenvatinib plus Weak
pembrolizumab and nivolumab and cabozantinib in centres with experience of immune
combination therapy and appropriate supportive care within the context of a
multidisciplinary team.
Patients who do not receive the full 4 doses of ipilimumab due to toxicity should continue Weak
on single-agent nivolumab, where safe and feasible.
Offer axitinib, cabozantinib or lenvatinib as subsequent treatment to patients who Weak
experience treatment-limiting immune-related adverse events after treatment with the
combination of axitinib plus pembrolizumab, cabozantinib plus nivolumab or lenvatinib plus
pembrolizumab.
Treatment past progression can be justified but requires close scrutiny and the support of Weak
an expert multidisciplinary team.
Do not re-challenge patients who stopped immune checkpoint inhibitors because of toxicity Strong
without expert guidance and support from a multidisciplinary team.
Offer sunitinib or pazopanib to treatment-naive patients with IMDC favourable-, intermediate-, Strong
and poor-risk cc-mRCC who cannot receive or tolerate immune checkpoint inhibition.
Offer cabozantinib to treatment-naive patients with IMDC intermediate- and poor-risk Stronga
cc-mRCC who cannot receive or tolerate immune checkpoint inhibition.
a While this is based on a randomised phase II trial, cabozantinib (weak) looks at least as good as sunitinib
in this population. This justified the same recommendation under exceptional circumstances.

7.4.3 Targeted therapies


In sporadic ccRCC, hypoxia-inducible factor (HIF) accumulation due to VHL-inactivation results in
overexpression of VEGF and platelet-derived growth factor (PDGF), which promote neo-angiogenesis [456-458].
This process substantially contributes to the development and progression of RCC. Several targeting drugs for
the treatment of mRCC are approved in both the USA and Europe.
Most published trials have selected for clear-cell carcinoma subtypes, thus no robust evidence-
based recommendations can be given for non-ccRCC subtypes.

In major trials leading to registration of the approved targeted agents, patients were stratified according to the
IMDC risk model (Table 7.2) [227].

Table 7.4: M
 edian OS and percentage of patients surviving two years treated in the era of targeted
therapy per IMDC risk group*,**

IMDC Model Patients** Median OS* 2-yr OS (95% CI)**


n % (months)
Favourable 157 18 43.2 75% (65–82%)
Intermediate 440 52 22.5 53% (46–59%)
Poor 252 30 7.8 7% (2–16%)
* Based on [227]; ** based on [440].
CI = confidence interval; IMDC = Metastatic Renal Cancer Database Consortium; n = number of patients;
OS = overall survival; yr = year.

46 RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021


7.4.3.1 Tyrosine kinase inhibitors
7.4.3.1.1 Sorafenib
Sorafenib is an oral multi-kinase inhibitor. A trial compared sorafenib and placebo after failure of prior systemic
immunotherapy or in patients unfit for immunotherapy. Sorafenib improved PFS (HR: 0.44, 95% CI: 0.35–0.55,
p < 0.01) [459]. Overall survival improved in patients initially assigned to placebo who were censored at
crossover [460].
In patients with previously untreated mRCC sorafenib was not superior to IFN-α (phase II study). A
number of studies have used sorafenib as the control arm in sunitinib-refractory disease vs. axitinib, dovitinib
or temsirolimus. None showed superior survival for the study drug compared to sorafenib.

7.4.3.1.2 Sunitinib
Sunitinib is an oral TKI inhibitor and has anti-tumour and anti-angiogenic activity. First-line monotherapy with
sunitinib demonstrated significantly longer PFS compared with IFN-α. Overall survival was greater in patients
treated with sunitinib (26.4 months) vs. IFN-α (21.8 months) despite crossover [461].
In the EFFECT trial, sunitinib 50 mg/day (4 weeks on/2 weeks off) was compared with continuous
uninterrupted sunitinib 37.5 mg/day in patients with cc-mRCC [462]. No significant differences in OS were
seen (23.1 vs. 23.5 months, p = 0.615). Toxicity was comparable in both arms. Because of the non-significant,
but numerically longer time to progression with the standard 50 mg dosage, the authors recommended using
this regimen. Alternate scheduling of sunitinib (2 weeks on/one week off) is being used to manage toxicity, but
robust data to support its use is lacking [463, 464].

7.4.3.1.3 Pazopanib
Pazopanib is an oral angiogenesis inhibitor. In a trial of pazopanib vs. placebo in treatment-naive mRCC patients
and cytokine-treated patients, a significant improvement in PFS and tumour response was observed [465].
A non-inferiority trial comparing pazopanib with sunitinib (COMPARZ) established pazopanib as
an alternative to sunitinib. It showed that pazopanib was not associated with significantly worse PFS or OS
compared to sunitinib. The two drugs had different toxicity profiles, and QoL was better with pazopanib [466].
In another patient-preference study (PISCES), patients preferred pazopanib to sunitinib (70% vs. 22%, p <
0.05) due to symptomatic toxicity [467]. Both studies were limited in that intermittent therapy (sunitinib) was
compared with continuous therapy (pazopanib).

7.4.3.1.4 Axitinib
Axitinib is an oral selective second-generation inhibitor of VEGFR-1, -2, and -3. Axitinib was first evaluated
as second-line treatment. In the AXIS trial, axitinib was compared to sorafenib in patients who had previously
failed cytokine treatment or targeted agents (mainly sunitinib) [468].
The overall median PFS was greater for axitinib than sorafenib. Axitinib was associated with
a greater PFS than sorafenib (4.8 vs. 3.4 months) after progression on sunitinib. Axitinib showed grade 3
diarrhoea in 11%, hypertension in 16%, and fatigue in 11% of patients. Final analysis of OS showed no
significant differences between axitinib or sorafenib [469, 470]. In a randomised phase III trial of axitinib vs.
sorafenib in first-line treatment-naive cc-mRCC, a significant difference in median PFS between the treatment
groups was not demonstrated, although the study was underpowered, raising the possibility of a type II error
[471]. As a result of this study, axitinib is not approved for first-line therapy.

7.4.3.1.5 Cabozantinib
Cabozantinib is an oral inhibitor of tyrosine kinase, including MET, VEGF and AXL. Cabozantinib was
investigated in a phase I study in patients resistant to VEGFR and mTOR inhibitors demonstrating objective
responses and disease control [197]. Based on these results an RCT investigated cabozantinib vs. everolimus in
patients with ccRCC failing one or more VEGF-targeted therapies (METEOR) [472, 473]. Cabozantinib delayed
PFS compared to everolimus in VEGF-targeted therapy refractory disease (HR: 0.58, 95% CI: 0.45–0.75) [472]
(LE: 1b). The median OS was 21.4 months (95% CI: 18.7 to not estimable) with cabozantinib and 16.5 months
(95% CI: 14.7–18.8) with everolimus in VEGF-resistant RCC. The HR for death was 0.66 (95% CI: 0.53–0.83,
p = 0.0003) [473]. Grade 3 or 4 adverse events were reported in 74% with cabozantinib and 65% with
everolimus. Adverse events were managed with dose reductions; doses were reduced in 60% of the patients
who received cabozantinib.
The Alliance A031203 CABOSUN randomised phase II trial comparing cabozatinib and sunitinib in
first-line in 157 intermediate- and poor-risk patients favoured cabozantinib for RR and PFS, but not OS [474,
475]. Cabozantinib significantly increased median PFS (8.2 vs. 5.6 months, adjusted HR: 0.66, 95% CI: 0.46 to
0.95; one-sided p = 0.012). Objective response rate was 46% (95% CI: 34–57) for cabozantinib vs. 18% (95%
CI: 10–28) for sunitinib. All-causality grade 3 or 4 adverse events were similar for cabozantinib and sunitinib. No
difference in OS was seen. Due to limitations of the statistical analyses within this trial the evidence is inferior
over existing choices.

RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021 47


7.4.3.1.6 Lenvatinib
Lenvatinib is an oral multi-target TKI of VEGFR1, VEGFR2, and VEGFR3, with inhibitory activity against
fibroblast growth factor receptors (FGFR1, FGFR2, FGFR3, and FGFR4), platelet growth factor receptor
(PDGFR-α), re-arranged during transfection (RET) and receptor for stem cell factor (KIT). It has recently been
investigated in a randomised phase II study in combination with everolimus vs. lenvatinib or everolimus alone
(see Section 7.4.6.1.1 for discussion of results) [476].

7.4.3.1.7 Tivozanib
Tivozanib is a potent and selective TKI of VEGFR1, VEGFR2, and VEGFR3 and was compared in two phase III
trials with sorafenib in patients with mRCC [477, 478]. Tivozanib was approved by the EMA in front-line mRCC.
While it was associated with a PFS advantage in both studies, no OS advantage was seen. In view of the
choice of sorafenib as the control arm in the front-line trial, the Panel considers there is too much uncertainty,
and too many attractive alternatives, to support its use in this front-line setting.

7.4.4 Monoclonal antibody against circulating VEGF


7.4.4.1 Bevacizumab monotherapy and bevacizumab plus IFN-α
Bevacizumab is a humanised monoclonal antibody. The double-blind AVOREN study compared bevacizumab
plus IFN-α with IFN-α monotherapy in mRCC. Overall response was higher in the bevacizumab plus IFN-α
group. Median PFS increased from 5.4 months with IFN-α to 10.2 months with bevacizumab plus IFN-α.
No benefit was seen in MSKCC poor-risk patients. Median OS in this trial, which allowed crossover after
progression, was not greater in the bevacizumab/IFN-α group (23.3 vs. 21.3 months) [479].
An open-label trial (CALGB 90206) of bevacizumab plus IFN-α vs. IFN-α showed a higher median
PFS for the combination group [480, 481]. Objective response rate was also higher in the combination group.
Overall toxicity was greater for bevacizumab plus IFN-α, with significantly more grade 3 hypertension, anorexia,
fatigue, and proteinuria. Bevacizumab, alone, or in combinations, is not widely recommended or used in mRCC
due to more attractive alternatives.

7.4.5 mTOR inhibitors


7.4.5.1 Temsirolimus
Temsirolimus is a specific inhibitor of mTOR [482]. Its use has been superseded as front-line treatment option.

7.4.5.2 Everolimus
Everolimus is an oral mTOR inhibitor, which is established in the treatment of VEGF-refractory disease. The
RECORD-1 study compared everolimus plus best supportive care (BSC) vs. placebo plus BSC in patients
with previously failed anti-VEGFR treatment (or previously intolerant of VEGF-targeted therapy) [483]. The data
showed a median PFS of 4 vs. 1.9 months for everolimus and placebo, respectively [483].

The Panel consider, even in the absence of conclusive data, that everolimus may present a therapeutic option
in patients who were intolerant to, or previously failed, immune- and VEGFR-targeted therapies (LE: 4). Recent
phase II data suggest adding lenvatinib is attractive.

7.4.6 Therapeutic strategies


7.4.6.1 Therapy for treatment-naïve patients with clear-cell metastatic RCC
The combination of pembrolizumab plus axitinib as well as nivolumab plus cabozantinib and lenvatinib plus
pembrolizumab is the standard of care in all IMDC-risk patients and ipilimumab plus nivolumab in IMDC
intermediate- and poor-risk patients (Figure 7.1). Therefore, the role of VEGFR-TKIs alone in front-line mRCC
has been superseded. Sunitinib, pazopanib, and cabozantinib (IMDC intermediate- and poor-risk disease),
remain alternative treatment options for patients who cannot receive or tolerate immune checkpoint inhibition in
this setting (Figure 7.1).

7.4.6.1.1 Sequencing systemic therapy in clear-cell metastatic RCC


The sequencing of targeted therapies is established in mRCC and maximises outcomes [444, 472, 476].
Pembrolizumab plus axitinib, nivolumab plus cabozantinib, lenvatinib plus pembrolizumab and nivolumab
plus ipilimumab are the new standard of care in front-line therapy. The impact of front-line immune checkpoint
inhibition on subsequent therapies is unclear. Randomised data on patients with disease refractory to either
nivolumab plus ipilimumab or TKI plus IO in a first-line setting are lacking, and available cohorts are limited
[484]. Prospective data on cabozantinib and axitinib are available for patients progressing on immunotherapy,
but these studies do not focus solely on the front-line setting, involve subset analyses, and are too small for
definitive conclusions [472, 485].

48 RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021


Retrospective data on VEGFR-TKI therapy after progression on front-line immune combinations exist but
have significant limitations. When considering this data in totality, there is some activity but it is still too early
to recommend one VEGFR-TKI over another after immunotherapy/immunotherapy or immunotherapy/VEGFR
combination (Figure 7.2). After the axitinib plus pembrolizumab combination, changing the VEGFR-TKI at
progression to cabozantinib or any other TKI not previously used is recommended.

The Panel do not support the use of mTOR inhibitors unless VEGF-targeted therapy is contraindicated as they
have been outperformed by other VEGF-targeted therapies in mRCC [486]. Drug choice in the third-line setting,
after immune checkpoint inhibitor combinations and subsequent VEGF-targeted therapy, is unknown. The
Panel recommends a subsequent agent which is approved in VEGF-refractory disease, with the exception of
re-challenge with immune checkpoint blockade. Cabozantinib is the only agent in VEGF-refractory disease with
RCT data showing a survival advantage and should be used preferentially [468]. Axitinib has positive PFS data
in VEGF-refractory disease. Both sorafenib and everolimus have been outperformed by other agents in VEGF-
refractory disease and are therefore less attractive [486]. The lenvatinib plus everolimus combination appears
superior to everolimus alone and has been granted EMA regulatory approval based on randomised phase II
data. This is an alternative despite the availability of phase II data only [476]. As shown in a study which also
included patients on immune checkpoint inhibitors tivozinib provides PFS superiority over sorafenib in VEGF-
refractory disease [487].

7.4.6.2 Non-clear-cell metastatic RCC


No phase III trials of patients with non-cc-mRCC have been reported. Expanded access programmes and
subset analyses from RCC studies suggest the outcome of these patients with targeted therapy is poorer than
for ccRCC. Targeted treatment in non-cc-mRCC has focused on temsirolimus, everolimus, sorafenib, sunitinib
and pembrolizumab [438, 488-490].
The most common non-clear-cell subtypes are papillary type I and non-type I papillary RCCs.
There are small single-arm trials for sunitinib and everolimus [490-493]. A trial of both types of pRCC treated
with everolimus (RAPTOR), showed a median PFS of 3.7 months per central review in the ITT population
with a median OS of 21.0 months [493]. In a non-randomised phase II trial, type 2 papillary RCC associated
to HLRCC, a familial cancer syndrome caused by germline mutations in the fumarate hydratase enzyme (FH)
gene, the combination of bevacizumab 10 mg/kg IV every 2 weeks and erlotinib 150 mg orally daily has been
evaluated [494]. The combination regimen reports interesting activity with an ORR of 64% (27/42; 95% CI:
49–77) in the HLRCC cohort, with a median PFS of 21.1 months (95% CI: 15.6–26.6). Grade ≥ 3 treatment-
related AEs occurred in 47% of patients, including hypertension (34%) and proteinuria (13%).

However, a randomised phase II trial of everolimus vs. sunitinib (ESPN) with crossover design in non-cc-mRCC
including 73 patients (27 with pRCC) was stopped after a futility analysis for PFS and OS [495]. The final results
showed a non-significant trend favouring sunitinib (6.1 vs. 4.1 months). Based on a systematic review including
subgroup analyses of the ESPN, RECORD-3 and another phase II trial (ASPEN), sunitinib and everolimus
remain options in this population, with a preference for sunitinib [8, 144, 496]. Patients with non-cc-mRCC
should be referred to a clinical trial, where appropriate. Efficacy for pembrolizumab (n = 165; response rates of
24%, PFS 4.1 months [95% CI: 2.8–5.6 months] 72% one-year OS) was noted but these results are based on
a single-arm phase II study [447]. Pembrolizumab can be conceded in this setting due to the high unmet need.

Subset analyses have shown impressive results for PD-L1 inhibitors combined with CTLA4 or VEGF-targeted
therapy in renal tumours with sarcomatoid features. Bevacizumab/atezolizumab, ipilimumab/nivolumab,
axitinib/pembrolizumab and avelumab/axitinib can all be recommended instead of VEFG-targeted therapy
alone. These options have impressive OS advantages over sunitinib and superseded VEGF-targeted therapy.
Collecting-duct cancers and renal medullary cancers are highly resistant to systemic therapy. Only
case reports have been published for a spectrum of treatment options so far and no clear recommendations
can be provided until data from international registries (RARECARE) or clinical trials become available.

RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021 49


Figure 7.1: U
 pdated EAU Guidelines recommendations for the first-line treatment of metastatic clear-cell
RCC

Alternative in patients who can


Standard of Care not receive or tolerate immune
checkpoint inhibitors

nivolumab/cabozantinib [1b]
sunitinib* [1b]
IMDC favourable risk pembrolizumab/axitinib [1b]
pazopanib* [1b]
pembrolizumab/lenvatinib [1b]

nivolumab/cabozantinib [1b]
cabozantinib* [2a]
IMDC intermediate and pembrolizumab/axitinib [1b]
sunitinib*[1b]
poor risk pembrolizumab/lenvatinib [1b]
pazopanib* [1b]
nivolumab/ipilimumab [1b]

IMDC = The International Metastatic Renal Cell Carcinoma Database Consortium


*pazopanib for intermediate-risk disease only.
[1b] = based on one randomised controlled phase III trial.
[2a] = based on a well-designed study without randomisation, or a subgroup analysis of a randomised
controlled trial.

Figure 7.2: EAU Guidelines recommendations for later-line therapy

Standard of care Alternative

Any VEGF-targeted therapy


that has not been used
Prior IO
previously in combination
with IO [4]

nivolumab [1b]
Prior TKI axitinib [2b]
cabozantinib [1b]

IO = immunotherapy; TKI = tyrosine kinase inhibitors; VEGF = vascular endothelial growth factor.
[1b] = based on one randomised controlled phase III trial.
[2b] = subgroup analysis of a randomised controlled phase III trial.
[4] = expert opinion.

7.4.7 Summary of evidence and recommendations for targeted therapy in metastatic RCC

Summary of evidence LE
Single-agent VEGF-targeted therapy has been superseded by immune checkpoint-based combination 1b
therapy.
Pazopanib is non-inferior to sunitinib in front-line mRCC. 1b
Cabozantinib in intermediate- and poor-risk treatment-naive clear-cell RCC leads to better response 2b
rates and PFS but not OS when compared to sunitinib.
Tivozanib has been EMA approved, but the evidence is still considered inferior over existing choices in 3
the front-line setting.
Single-agent VEGF-targeted therapies are preferentially recommended after front-line PD-L1-based 3
combinations. Re-challenge with treatments already used should be avoided.

50 RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021


Single-agent cabozantinib or nivolumab are superior to everolimus after one or more lines of VEGF- 1b
targeted therapy.
Everolimus prolongs PFS after VEGF-targeted therapy when compared to placebo. This is no longer 1b
widely recommended before third-line therapy.
Both mTOR inhibitors and VEGF-targeted therapies have limited activity in non-cc-mRCC. There is a 2a
non-significant trend for improved oncological outcomes for sunitinib over everolimus.
Lenvatinib in combination with everolimus improved PFS over everolimus alone in VEGF-refractory 2a
disease. Its role after immune checkpoint inhibitors is uncertain. There is a lack of robust data on this
combination making its recommendation challenging.

Recommendations Strength rating


Offer nivolumab or cabozantinib for immune checkpoint inhibitor-naive vascular endothelial Strong
growth factor receptor (VEGFR)-refractory clear-cell metastatic renal cell carcinoma
(cc-mRCC) after one or two lines of therapy.
Sequencing the agent not used as second-line therapy (nivolumab or cabozantinib) for Weak
third-line therapy is recommended.
Offer VEGF-tyrosine kinase inhibitors as second-line therapy to patients refractory to Weak
nivolumab plus ipilimumab or axitinib plus pembrolizumab or cabozantinib plus nivolumab
or lenvatinib plus pembrolizumab.
Offer cabozantinib after VEGF-targeted therapy in cc-mRCC. Strong
Sequence systemic therapy in treating mRCC. Strong

7.5 Locally recurrent RCC after treatment of localised disease


Locally recurrent disease can either affect the tumour-bearing kidney after PN or focal ablative therapy such
as RFA and cryotherapy. Local relapse may be due to the incomplete resection of the primary tumour (type
A), in a minority of the cases to the local spread of the tumour by microvascular embolisation (type B), or true
multifocality (type C) [497]. Most studies reporting on the oncological efficacy of surgery for recurrent disease
after removal of the kidney have not considered the traditional definition of local recurrence after RN, PN and
thermal ablation, which is: “tumour growth exclusively confined to the true renal fossa”. Instead, recurrences
within the renal vein, the ipsilateral adrenal gland or the regional LNs were included under this term. Isolated
tumour recurrence within the true renal fossa only is a rare event. In the existing literature the topic is poorly
investigated and available data are mainly related to positive surgical margins only [498, 499].

The prognosis of recurrent disease not due to multifocality (type A and B) is poor, despite salvage nephrectomy
[497]. Recurrent tumour growth in the regional LNs or ipsilateral adrenal gland may reflect metachronous
metastatic spread (see Section 7.3). After PN for pT1 disease, recurrences within the remaining kidney occur in
0.5–2% of patients [500, 501].

Following thermal ablation or cryotherapy generally intra-renal, but also peri-renal, recurrences have been
reported in up to 14% of cases [502]. Whereas repeat ablation is still recommended as the preferred
therapeutic option after treatment failure, the most effective salvage procedure as an alternative to complete
nephrectomy has not yet been defined.
Isolated local recurrence is associated with worse survival [503, 504]. Based on retrospective and
non-comparative data only, several approaches such as surgical excision, radiotherapy, systemic treatment
and observation have been suggested for the treatment of isolated local recurrence [505-507]. Among
these alternatives, surgical resection with negative margins remains the only therapeutic option shown to
be associated with improved survival [503] One of the largest series including 2,945 patients treated with
RN reported on 54 patients with recurrent disease localised in the renal fossa, the ipsilateral adrenal gland
or the regional LNs as sole metastatic sites [505]. Another recent series identified 33 patients with isolated
local recurrences and 30 local recurrences with synchronous metastases within a cohort of 2,502 surgically
treated patients, confirming the efficacy of locally directed treatment vs. conservative approaches (observation,
systemic therapy) [508]. In a series of 1,955 patients with clinical T1 RCCs treated with PN, 95 patients (4.9%)
had a pT3a upstaging, indicating a high risk for local and intra-renal recurrence and reduced survival [506].
Open surgery has been successfully reported in studies [509, 510]. However, minimaly invasive
approaches, including standard and hand-assisted laparoscopic- and robotic approaches for the resection
of isolated RCC recurrences have been occasionally reported. Ablative therapies including cryoablation,
radiofrequency and microwave ablation, may also have a role in managing recurrent RCC patients, but further
validation will be needed [511].

RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021 51


In summary, the limited available evidence suggests that in selected patients surgical removal of locally
recurrent disease can induce durable tumour control, although with expected high risk of complications.
Johnson et al. published on 51 planned repeat PNs in 47 patients with locally recurrent disease, reporting
a total of 40 peri-operative complications, with temporary urinary extravasation being the most prevalent
[512]. Since local recurrences develop early, with a median time interval of 10–20 months after treatment of
the primary tumour [513], a guideline-adapted follow-up scheme for early detection is recommended (see
Chapter 8 - Follow-up) even though benefit in terms of cancer control has not yet been demonstrated [514].

Adverse prognostic parameters are a short time interval since treatment of the primary tumour (< 3–12 months)
[515], sarcomatoid differentiation of the recurrent lesion and incomplete surgical resection [505]. In case
complete surgical removal is unlikely to be performed or when significant comorbidities are present (especially
when combined with poor prognostic tumour features), palliative therapeutic approaches including radiation
therapy aimed at symptom control and prevention of local complications should be considered (see Sections
7.3 and 7.4).

7.5.1 Summary of evidence and recommendation on locally recurrent RCC after treatment of
localised disease

Summary of evidence LE
Isolated recurrence is a rare entity (< 2%). 3
In the absence of adverse prognostic factors such as sarcomatoid features or median time interval of 3
< 12 months since treatment of the primary tumour, treatment of local recurrences can induce durable
local control.
The most optimal modality of local treatment for locally recurrent RCC is still under debate. 3

Recommendation Strength rating


Offer local treatment of locally recurrent disease when technically possible and significant Weak
comorbidities are absent.

8. FOLLOW-UP IN RCC
8.1 Introduction
Surveillance after treatment for RCC allows the urologist to monitor or identify:
• post-operative complications;
• renal function;
• local recurrence;
• recurrence in the contralateral kidney;
• distant metastases;
• cardiovascular events.

There is no consensus on follow-up strategies after RCC treatment, with limited evidence suggesting that more
frequent post-operative imaging intervals do not provide any improvement for early detection of recurrence that
would lead to improved survival [514]. As such, intensive radiological surveillance may not be necessary for all
patients. Follow-up is also important to assess functional outcomes and to limit long-term sequelae such as
renal function impairment, end-stage renal disease and cardiovascular events [516].
Currently, the key question is whether any recurrence detection during follow-up and subsequent
treatment will lead to any meaningful change in survival outcome for these patients.

In contrast to high-grade and/or locally advanced disease, the outcome after surgery for T1a low-grade
tumours is almost always excellent. It is therefore reasonable to stratify follow-up, taking into account the risk
of each different RCC to develop a local or distant recurrence. Although there is no randomised evidence, large
studies have examined prognostic factors with long follow-up [168, 517, 518] (LE: 4). One study has shown a
survival benefit in patients who were followed within a structured surveillance protocol vs. patients who were
not [519]; patients undergoing follow-up seem to have a longer OS when compared to patients not undergoing
routine follow-up [519].

52 RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021


Furthermore, an individualised and risk-based approach to RCC follow-up has recently been
proposed. The authors used competing risk models, incorporating patient age, pathologic stage, relapse
location and comorbidities, to calculate when the risk of non-RCC death exceeds the risk of RCC recurrence
[520]. For patients with low-stage disease but with a Charlson comorbidity index ≥ 2, the risk of non-RCC
death exceeded that of abdominal recurrence risk already one month after surgery, regardless of patient age.
The RECUR consortium, initiated by this Panel, collects similar data with the aim to provide comparators for
guideline recommendations. Recently published RECUR data support a risk-based approach; more specifically
a competing-risk analysis showed that for low-risk patients, the risk of non-RCC related death exceeded the
risk of RCC recurrence shortly after the intial surgery. For intermediate-risk patients, the corresponding time
point was reached around four to five years after surgery. In high-risk patients, the risk of RCC recurrence
continously exceeded the risk of non-RCC related death [10]. In the near future, genetic profiling may refine
the existing prognostic scores and external validation in datasets from adjuvant trials have been promising in
improving stratification of patient’s risk of recurrence [10, 521].

Recurrence after PN is rare, but early diagnosis is relevant, as the most effective treatment is surgery [509,
522]. Recurrence in the contralateral kidney is rare (1–2%) and can occur late (median 5–6 years) [523]
(LE: 3). Follow-up can identify local recurrences or metastases at an early stage. In metastatic disease,
extended tumour growth can limit the opportunity for surgical resection, which is considered the standard
therapy in cases of resectable and preferably solitary lesions. In addition, early diagnosis of tumour recurrence
may enhance the efficacy of systemic treatment if the tumour burden is low.

8.2 Which imaging investigations for which patients, and when?


• The sensitivity of chest radiography and US for detection of small RCC metastases is poor. The sensitivity
of chest radiography is significantly lower than CT-scans, as proven in comparative studies including
histological evaluation [524-526]. Therefore, follow-up for recurrence detection with chest radiography
and US are less sensitive [527].
• Positron-emission tomography and PET-CT as well as bone scintigraphy should not be used routinely in
RCC follow-up, due to their limited specificity and sensitivity [6, 119].
• Surveillance should also include evaluation of renal function and cardiovascular risk factors [516].
• Outside the scope of regular follow-up imaging of the chest and abdomen, targeted imaging should be
considered in patients with organ-specific symptoms, e.g. CT or MRI imaging of the brain in patients
experiencing neurological symptoms [528].

Controversy exists on the optimal duration of follow-up. Some authors argue that follow-up with imaging is
not cost-effective after five years; however, late metastases are more likely to be solitary and justify more
aggressive therapy with curative intent. In addition, patients with tumours that develop in the contralateral
kidney can be treated with NSS if the tumours are detected early. Several authors have designed scoring
systems and nomograms to quantify the likelihood of patients to develop tumour recurrences, metastases, and
subsequent death [215, 217, 529, 530]. These models, of which the most utilised are summarised in Chapter 6
- Prognosis, have been compared and validated [531] (LE: 2). Using prognostic variables, several stage-based
follow-up regimens have been proposed, although, none propose follow-up strategies after ablative therapies
[532, 533]. A post-operative nomogram is available to estimate the likelihood of freedom from recurrence at five
years [212]. Recently, a pre-operative prognostic model based on age, symptoms and TNM staging has been
published and validated [534] (LE: 3).
A follow-up algorithm for monitoring patients after treatment for RCC is needed, recognising not
only the patient’s risk of recurrence profile, but also the efficacy of the treatment given (Table 8.1). These
prognostic systems can be used to adapt the follow-up schedule according to predicted risk of recurrence.
Ancillary to the above, life-expectancy calculations based on comorbidity and age at diagnosis may be useful
in counselling patients on duration of follow-up [535].

RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021 53


Table 8.1: Proposed follow-up schedule following treatment for localised RCC, taking into account
patient risk of recurrence profile and treatment efficacy (based on expert opinion [LE: 4])

Risk profile (*) Oncological follow-up after date of surgery


3 mo 6 mo 12 mo 18 mo 24 mo 30 mo 36 mo > 3 yr (**) (***) > 5 yr (**) (***)
Low risk of recurrence - CT - CT - CT - CT once every -
two yrs
For ccRCC:
Leibovich Score 0-2

For non-ccRCC:
pT1a-T1b pNx-0 M0 and
histological grade 1 or 2.
Intermediate risk of - CT CT - CT - CT CT once yr CT once
recurrence every
two yrs
For ccRCC:
Leibovich Score 3-5

For non-ccRCC:
pT1b pNx-0 and/or
histological grade 3 or 4.
High risk of recurrence CT CT CT CT CT - CT CT once yr CT once
every
For ccRCC: two yrs
Leibovich Score ≥ 6

For non-ccRCC:
pT2-pT4 with any
histological grade
or
pT any, pN1 cM0 with
any histological grade
ccRCC = clear cell renal cell carcinoma, CT = computed tomography, mo = months,
non-ccRCC = non clear cell renal cell carcinoma; yr = years.

The table above provides recommendations on follow-up strategies for low, intermediate and high risk of
recurrence in patients curatively treated for localised RCC either with NSS or RN. Computed tomography in
the table refers to imaging of both chest and abdomen. Alternatively, MRI of the abdomen can be performed
instead of a CT-scan.
* R
 isk of recurrence profiles should be based on validated prognostic models. The EAU RCC Guidelines
Panel recommends the 2003 Leibovich model for ccRCC [215]. However, other validated models can
be used by physicians based on their own national/regional recommendations. In a similar fashion, for
curatively treated localised non-ccRCC, the Panel recommends the use of the University of California Los
Angeles integrated staging system (UISS) to determine risk of recurrence [216].
** for all risk of recurrence profiles, functional follow-up, mainly monitoring renal and cardiovascular function,
may continue according to specific clinical needs irrespective of the length of the oncological follow-up.
*** F or low-risk profiles at > 3 years and intermediate-risk at > 5 years of follow-up respectively, consider
counselling patients about terminating oncological follow-up imaging based on assessment of
comorbidities, age, life expectancy and/or patient wishes.

54 RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021


8.3 Summary of evidence and recommendations for surveillance following RN or PN or
ablative therapies in RCC

Summary of evidence LE
Functional follow-up after curative treatment for RCC is useful to prevent renal and cardiovascular 4
deterioration.
Oncological follow-up can detect local recurrence or metastatic disease while the patient may still be 4
surgically curable.
After NSS, there is an increased risk of recurrence for larger (> 7 cm) tumours, or when there is a 3
positive surgical margin.
Patients undergoing follow-up have a better overall survival than patients not undergoing surveillance. 3
Prognostic models provide stratification of RCC risk of recurrence based on TNM and histological 3
features.
In competing-risk models, risk of non-RCC-related death exceeds that of RCC recurrence or related 3
death in low-risk patients.
Life expectancy estimation is feasible and may support counselling of patients on duration of follow-up. 4

Recommendations Strength rating


Base follow-up after treatment of localised RCC on the risk of recurrence. Strong
Perform functional follow-up (renal function assessment and prevention of cardiovascular Weak
events) both in nephron-sparing (NSS) and radical nephrectomy patients.
Intensify follow-up in patients after nephron-sparing surgery for tumours > 7 cm or in Weak
patients with a positive surgical margin.
Consider curtailing follow-up when the risk of dying from other causes is double that of Weak
recurrence risk.
Base risk of recurrence stratification on validated subtype-specific models such as the Weak
Leibovich Score for ccRCC or the University of California Los Angeles integrated staging
system (UISS) for non-ccRCC.

8.4 Research priorities


There is a clear need for future research to determine whether follow-up can optimise patient survival. Data
evaluating at which time point follow-up has the highest chance to detect recurrence will be most valuable for
clinical practice.
Novel prognostic markers at surgery should be investigated to determine the risk of relapse over time.

9. REFERENCES
1. Ljungberg, B., et al. European Association of Urology Guidelines on Renal Cell Carcinoma: The 2019
Update. Eur Urol, 2019. 75: 799.
https://pubmed.ncbi.nlm.nih.gov/30803729
2. Guyatt, G.H., et al. GRADE: an emerging consensus on rating quality of evidence and strength of
recommendations. BMJ, 2008. 336: 924.
https://pubmed.ncbi.nlm.nih.gov/18436948
3. Phillips, B., et al. Oxford Centre for Evidence-based Medicine Levels of Evidence. Updated by
Jeremy Howick March 2009.
https://www.cebm.net/2009/06/oxford-centre-evidence-based-medicine-levels-evidence-march-2009/
4. Guyatt, G.H., et al. Going from evidence to recommendations. BMJ, 2008. 336: 1049.
https://pubmed.ncbi.nlm.nih.gov/18467413
5. Fernández-Pello, S., et al. Management of Sporadic Renal Angiomyolipomas: A Systematic Review
of Available Evidence to Guide Recommendations from the European Association of Urology Renal
Cell Carcinoma Guidelines Panel. Eur Urol Oncol, 2020. 3: 57.
https://pubmed.ncbi.nlm.nih.gov/31171501
6. Vogel, C., et al. Imaging in Suspected Renal-Cell Carcinoma: Systematic Review. Clin Genitourin
Cancer, 2019. 17: e345.
https://pubmed.ncbi.nlm.nih.gov/30528378

RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021 55


7. Abu-Ghanem, Y., et al. Limitations of Available Studies Prevent Reliable Comparison Between
Tumour Ablation and Partial Nephrectomy for Patients with Localised Renal Masses: A Systematic
Review from the European Association of Urology Renal Cell Cancer Guideline Panel. Eur Urol
Oncol, 2020. 3: 433.
https://pubmed.ncbi.nlm.nih.gov/32245655
8. Fernández-Pello, S., et al. A Systematic Review and Meta-analysis Comparing the Effectiveness and
Adverse Effects of Different Systemic Treatments for Non-clear Cell Renal Cell Carcinoma. Eur Urol,
2017. 71: 426.
https://pubmed.ncbi.nlm.nih.gov/27939075
9. Hofmann, F., et al. Targeted therapy for metastatic renal cell carcinoma. Cochrane Database Syst
Rev, 2020. CD012796.
https://pubmed.ncbi.nlm.nih.gov/33058158
10. Dabestani, S., et al. Long-term Outcomes of Follow-up for Initially Localised Clear Cell Renal Cell
Carcinoma: RECUR Database Analysis. Eur Urol Focus, 2019. 5: 857.
https://pubmed.ncbi.nlm.nih.gov/29525381
11. Dabestani, S., et al. Intensive Imaging-based Follow-up of Surgically Treated Localised Renal
Cell Carcinoma Does Not Improve Post-recurrence Survival: Results from a European Multicentre
Database (RECUR). Eur Urol, 2019. 75: 261.
https://pubmed.ncbi.nlm.nih.gov/30318330
12. Ferlay, J., et al. Cancer incidence and mortality patterns in Europe: Estimates for 40 countries and
25 major cancers in 2018. Eur J Cancer, 2018. 103: 356.
https://pubmed.ncbi.nlm.nih.gov/23485231
13. Capitanio, U., et al. Epidemiology of Renal Cell Carcinoma. Eur Urol, 2019. 75: 74.
https://pubmed.ncbi.nlm.nih.gov/30243799
14. Levi, F., et al. The changing pattern of kidney cancer incidence and mortality in Europe. BJU Int,
2008. 101: 949.
https://pubmed.ncbi.nlm.nih.gov/18241251
15. Moch, H., et al. The 2016 WHO Classification of Tumours of the Urinary System and Male Genital
Organs-Part A: Renal, Penile, and Testicular Tumours. Eur Urol, 2016. 70: 93.
https://pubmed.ncbi.nlm.nih.gov/26935559
16. Tahbaz, R., et al. Prevention of kidney cancer incidence and recurrence: lifestyle, medication and
nutrition. Curr Opin Urol, 2018. 28: 62.
https://pubmed.ncbi.nlm.nih.gov/29059103
17. Al-Bayati, O., et al. Systematic review of modifiable risk factors for kidney cancer. Urol Oncol, 2019.
37: 359.
https://pubmed.ncbi.nlm.nih.gov/30685335
18. Moch H, et al. WHO Classification of Tumours of the Urinary System and Male Genital Organs, ed.
WHO. 2016, IARC, Lyon.
https://publications.iarc.fr/Book-And-Report-Series/Who-Classification-Of-Tumours/WHO-
Classification-Of-Tumours-Of-The-Urinary-System-And-Male-Genital-Organs-2016
19. Klatte, T., et al. Prognostic factors and prognostic models for renal cell carcinoma: a literature
review. World J Urol, 2018. 36: 1943.
https://pubmed.ncbi.nlm.nih.gov/29713755
20. Keegan, K.A., et al. Histopathology of surgically treated renal cell carcinoma: survival differences by
subtype and stage. J Urol, 2012. 188: 391.
https://pubmed.ncbi.nlm.nih.gov/22698625
21. Linehan, W.M., et al. Comprehensive Molecular Characterization of Papillary Renal-Cell Carcinoma.
N Engl J Med, 2016. 374: 135.
https://pubmed.ncbi.nlm.nih.gov/26536169
22. Hora, M. Re: Philip S. Macklin, Mark E. Sullivan, Charles R. Tapping, et al. Tumour Seeding in the
Tract of Percutaneous Renal Tumour Biopsy: A Report on Seven Cases from a UK Tertiary Referral
Centre. Eur Urol 2019;75:861-7. Eur Urol, 2019. 76: e96.
https://pubmed.ncbi.nlm.nih.gov/31255420
23. Ledezma, R.A., et al. Clinically localized type 1 and 2 papillary renal cell carcinomas have similar
survival outcomes following surgery. World J Urol, 2016. 34: 687.
https://pubmed.ncbi.nlm.nih.gov/26407582
24. Volpe, A., et al. Chromophobe renal cell carcinoma (RCC): oncological outcomes and prognostic
factors in a large multicentre series. BJU Int, 2012. 110: 76.
https://pubmed.ncbi.nlm.nih.gov/22044519

56 RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021


25. Amin, M.B., et al. Collecting duct carcinoma versus renal medullary carcinoma: an appeal for
nosologic and biological clarity. Am J Surg Pathol, 2014. 38: 871.
https://pubmed.ncbi.nlm.nih.gov/24805860
26. Shah, A.Y., et al. Management and outcomes of patients with renal medullary carcinoma: a
multicentre collaborative study. BJU Int, 2017. 120: 782.
https://pubmed.ncbi.nlm.nih.gov/27860149
27. Iacovelli, R., et al. Clinical outcome and prognostic factors in renal medullary carcinoma: A pooled
analysis from 18 years of medical literature. Can Urol Assoc J, 2015. 9: E172.
https://pubmed.ncbi.nlm.nih.gov/26085875
28. Alvarez, O., et al. Renal medullary carcinoma and sickle cell trait: A systematic review. Pediatr Blood
Cancer, 2015. 62: 1694.
https://pubmed.ncbi.nlm.nih.gov/26053587
29. Msaouel, P., et al. Updated Recommendations on the Diagnosis, Management, and Clinical Trial
Eligibility Criteria for Patients With Renal Medullary Carcinoma. Clin Genitourin Cancer, 2019. 17: 1.
https://pubmed.ncbi.nlm.nih.gov/30287223
30. Beckermann, K.E., et al. Clinical and immunologic correlates of response to PD-1 blockade in a
patient with metastatic renal medullary carcinoma. J Immunother Cancer, 2017. 5: 1.
https://pubmed.ncbi.nlm.nih.gov/28105368
31. Sodji, Q., et al. Predictive role of PD-L1 expression in the response of renal Medullary carcinoma to
PD-1 inhibition. J Immunother Cancer, 2017. 5: 62.
https://pubmed.ncbi.nlm.nih.gov/28807004
32. Beckermann, K.E., et al. Renal Medullary Carcinoma: Establishing Standards in Practice. J Oncol
Pract, 2017. 13: 414.
https://pubmed.ncbi.nlm.nih.gov/28697319
33. Rathmell, W.K., et al. High-dose-intensity MVAC for Advanced Renal Medullary Carcinoma: Report
of Three Cases and Literature Review. Urology, 2008. 72: 659.
https://pubmed.ncbi.nlm.nih.gov/18649931
34. Breda, A., et al. Clinical and pathological outcomes of renal cell carcinoma (RCC) in native kidneys
of patients with end-stage renal disease: a long-term comparative retrospective study with RCC
diagnosed in the general population. World J Urol, 2015. 33: 1.
https://pubmed.ncbi.nlm.nih.gov/24504760
35. Breda, A., et al. Erratum to: Clinical and pathological outcomes of renal cell carcinoma (RCC) in
native kidneys of patients with end-stage renal disease: a long-term comparative retrospective
study with RCC diagnosed in the general population. World J Urol, 2015. 33: 9.
https://pubmed.ncbi.nlm.nih.gov/24577798
36. Tsuzuki, T., et al. Renal tumors in end-stage renal disease: A comprehensive review. Int J Urol, 2018.
25: 780.
https://pubmed.ncbi.nlm.nih.gov/30066367
37. Eble J.N., et al. Pathology and genetics of tumours of the urinary system and male genital organs.
World Health Organization Classification of Tumours. In: Pathology and genetics of tumours of the
urinary systemand male genital organs. World Health Organization Classification of Tumours., Eble
JN, Epstein JI, et al Editors. 2004, IARC: Lyon.
38. Shuch, B., et al. Defining early-onset kidney cancer: implications for germline and somatic mutation
testing and clinical management. J Clin Oncol, 2014. 32: 431.
https://pubmed.ncbi.nlm.nih.gov/24378414
39. Srigley, J.R., et al. The International Society of Urological Pathology (ISUP) Vancouver Classification
of Renal Neoplasia. Am J Surg Pathol, 2013. 37: 1469.
https://pubmed.ncbi.nlm.nih.gov/24025519
40. Pignot, G., et al. Survival analysis of 130 patients with papillary renal cell carcinoma: prognostic
utility of type 1 and type 2 subclassification. Urology, 2007. 69: 230.
https://pubmed.ncbi.nlm.nih.gov/17275070
41. Przybycin, C.G., et al. Hereditary syndromes with associated renal neoplasia: a practical guide to
histologic recognition in renal tumor resection specimens. Adv Anat Pathol, 2013. 20: 245.
https://pubmed.ncbi.nlm.nih.gov/23752087
42. Shuch, B., et al. The surgical approach to multifocal renal cancers: hereditary syndromes, ipsilateral
multifocality, and bilateral tumors. Urol Clin North Am, 2012. 39: 133.
https://pubmed.ncbi.nlm.nih.gov/22487757
43. Bratslavsky, G., et al. Salvage partial nephrectomy for hereditary renal cancer: feasibility and
outcomes. J Urol, 2008. 179: 67.
https://pubmed.ncbi.nlm.nih.gov/17997447

RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021 57


44. Grubb, R.L., 3rd, et al. Hereditary leiomyomatosis and renal cell cancer: a syndrome associated with
an aggressive form of inherited renal cancer. J Urol, 2007. 177: 2074.
https://pubmed.ncbi.nlm.nih.gov/17509289
45. Nielsen, S.M., et al. Von Hippel-Lindau Disease: Genetics and Role of Genetic Counseling in a
Multiple Neoplasia Syndrome. J Clin Oncol, 2016. 34: 2172.
https://pubmed.ncbi.nlm.nih.gov/27114602
46. Kauffman, E.C., et al. Molecular genetics and cellular features of TFE3 and TFEB fusion kidney
cancers. Nat Rev Urol, 2014. 11: 465.
https://pubmed.ncbi.nlm.nih.gov/25048860
47. Jonasch, E., et al. Phase II study of the oral HIF-2α inhibitor MK-6482 for Von Hippel-Lindau
disease–associated renal cell carcinoma. J Clin Oncol, 2020. 38: 5003.
https://ascopubs.org/doi/abs/10.1200/JCO.2020.38.15_suppl.5003
48. Bhatt, J.R., et al. Natural History of Renal Angiomyolipoma (AML): Most Patients with Large AMLs
>4cm Can Be Offered Active Surveillance as an Initial Management Strategy. Eur Urol, 2016. 70: 85.
https://pubmed.ncbi.nlm.nih.gov/26873836
49. Fittschen, A., et al. Prevalence of sporadic renal angiomyolipoma: a retrospective analysis of 61,389
in- and out-patients. Abdom Imaging, 2014. 39: 1009.
https://pubmed.ncbi.nlm.nih.gov/24705668
50. Nese, N., et al. Pure epithelioid PEComas (so-called epithelioid angiomyolipoma) of the kidney:
A clinicopathologic study of 41 cases: detailed assessment of morphology and risk stratification.
Am J Surg Pathol, 2011. 35: 161.
https://pubmed.ncbi.nlm.nih.gov/21263237
51. Tsai, H.Y., et al. Clinicopathologic analysis of renal epithelioid angiomyolipoma: Consecutively
excised 23 cases. Kaohsiung J Med Sci, 2019. 35: 33.
https://pubmed.ncbi.nlm.nih.gov/30844148
52. Ramon, J., et al. Renal angiomyolipoma: long-term results following selective arterial embolization.
Eur Urol, 2009. 55: 1155.
https://pubmed.ncbi.nlm.nih.gov/18440125
53. Nelson, C.P., et al. Contemporary diagnosis and management of renal angiomyolipoma. J Urol,
2002. 168: 1315.
https://pubmed.ncbi.nlm.nih.gov/12352384
54. Bhatt, N.R., et al. Dilemmas in diagnosis and natural history of renal oncocytoma and implications
for management. Can Urol Assoc J, 2015. 9: E709.
https://pubmed.ncbi.nlm.nih.gov/26664505
55. Bissler, J.J., et al. Everolimus for renal angiomyolipoma in patients with tuberous sclerosis complex
or sporadic lymphangioleiomyomatosis: extension of a randomized controlled trial. Nephrol Dial
Transplant, 2016. 31: 111.
https://pubmed.ncbi.nlm.nih.gov/23312829
56. Bissler, J.J., et al. Everolimus long-term use in patients with tuberous sclerosis complex: Four-year
update of the EXIST-2 study. PLoS One, 2017. 12: e0180939.
https://pubmed.ncbi.nlm.nih.gov/28792952
57. Geynisman, D.M., et al. Sporadic Angiomyolipomas Growth Kinetics While on Everolimus: Results of
a Phase II Trial. J Urol, 2020. 204: 531.
https://pubmed.ncbi.nlm.nih.gov/32250730
58. Patel, H.D., et al. Surgical histopathology for suspected oncocytoma on renal mass biopsy: a
systematic review and meta-analysis. BJU Int, 2017. 119: 661.
https://pubmed.ncbi.nlm.nih.gov/28058773
59. Liu, S., et al. Active surveillance is suitable for intermediate term follow-up of renal oncocytoma
diagnosed by percutaneous core biopsy. BJU Int, 2016. 118 Suppl 3: 30.
https://pubmed.ncbi.nlm.nih.gov/27457972
60. Kawaguchi, S., et al. Most renal oncocytomas appear to grow: observations of tumor kinetics with
active surveillance. J Urol, 2011. 186: 1218.
https://pubmed.ncbi.nlm.nih.gov/21849182
61. Richard, P.O., et al. Active Surveillance for Renal Neoplasms with Oncocytic Features is Safe. J Urol,
2016. 195: 581.
https://pubmed.ncbi.nlm.nih.gov/26388501
62. Abdessater, M., et al. Renal Oncocytoma: An Algorithm for Diagnosis and Management. Urology,
2020. 143: 173.
https://pubmed.ncbi.nlm.nih.gov/32512107

58 RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021


63. Roubaud, G., et al. Combination of gemcitabine and doxorubicin in rapidly progressive metastatic
renal cell carcinoma and/or sarcomatoid renal cell carcinoma. Oncology, 2011. 80: 214.
https://pubmed.ncbi.nlm.nih.gov/21720184
64. Abern, M.R., et al. Characteristics and outcomes of tumors arising from the distal nephron. Urology,
2012. 80: 140.
https://pubmed.ncbi.nlm.nih.gov/22626576
65. Husillos, A., et al. [Collecting duct renal cell carcinoma]. Actas Urol Esp, 2011. 35: 368.
https://pubmed.ncbi.nlm.nih.gov/21450372
66. Hora, M., et al. MiT translocation renal cell carcinomas: two subgroups of tumours with
translocations involving 6p21 [t (6; 11)] and Xp11.2 [t (X;1 or X or 17)]. Springerplus, 2014. 3: 245.
https://pubmed.ncbi.nlm.nih.gov/24877033
67. Forde, C., et al. Hereditary Leiomyomatosis and Renal Cell Cancer: Clinical, Molecular, and
Screening Features in a Cohort of 185 Affected Individuals. Eur Urol Oncol, 2020. 3: 764.
https://pubmed.ncbi.nlm.nih.gov/31831373
68. Schoots, I.G., et al. Bosniak Classification for Complex Renal Cysts Reevaluated: A Systematic
Review. J Urol, 2017. 198: 12.
https://pubmed.ncbi.nlm.nih.gov/28286071
69. Defortescu, G., et al. Diagnostic performance of contrast-enhanced ultrasonography and magnetic
resonance imaging for the assessment of complex renal cysts: A prospective study. Int J Urol, 2017.
24: 184.
https://pubmed.ncbi.nlm.nih.gov/28147450
70. Silverman, S.G., et al. Bosniak Classification of Cystic Renal Masses, Version 2019: An Update
Proposal and Needs Assessment. Radiology, 2019. 292: 475.
https://pubmed.ncbi.nlm.nih.gov/31210616
71. Donin, N.M., et al. Clinicopathologic outcomes of cystic renal cell carcinoma. Clin Genitourin
Cancer, 2015. 13: 67.
https://pubmed.ncbi.nlm.nih.gov/25088469
72. Park, J.J., et al. Postoperative Outcome of Cystic Renal Cell Carcinoma Defined on Preoperative
Imaging: A Retrospective Study. J Urol, 2017. 197: 991.
https://pubmed.ncbi.nlm.nih.gov/27765694
73. Chandrasekar, T., et al. Natural History of Complex Renal Cysts: Clinical Evidence Supporting Active
Surveillance. J Urol, 2018. 199: 633.
https://pubmed.ncbi.nlm.nih.gov/28941915
74. Nouhaud, F.X., et al. Contemporary assessment of the correlation between Bosniak classification
and histological characteristics of surgically removed atypical renal cysts (UroCCR-12 study). World
J Urol, 2018. 36: 1643.
https://pubmed.ncbi.nlm.nih.gov/29730837
75. Sobin LH., G.M., Wittekind C. (eds). TNM classification of malignant tumors, ed. U.I.U.A. Cancer.
Vol. 7th edn. 2009.
https://www.wiley.com/en-us/
TNM+Classification+of+Malignant+Tumours%2C+7th+Edition-p-9781444358964
76. Gospodarowicz, M.K., et al. The process for continuous improvement of the TNM classification.
Cancer, 2004. 100: 1.
https://pubmed.ncbi.nlm.nih.gov/14692017
77. Kim, S.P., et al. Independent validation of the 2010 American Joint Committee on Cancer TNM
classification for renal cell carcinoma: results from a large, single institution cohort. J Urol, 2011.
185: 2035.
https://pubmed.ncbi.nlm.nih.gov/21496854
78. Novara, G., et al. Validation of the 2009 TNM version in a large multi-institutional cohort of patients
treated for renal cell carcinoma: are further improvements needed? Eur Urol, 2010. 58: 588.
https://pubmed.ncbi.nlm.nih.gov/20674150
79. Waalkes, S., et al. Is there a need to further subclassify pT2 renal cell cancers as implemented by
the revised 7th TNM version? Eur Urol, 2011. 59: 258.
https://pubmed.ncbi.nlm.nih.gov/21030143
80. Bertini, R., et al. Renal sinus fat invasion in pT3a clear cell renal cell carcinoma affects outcomes of
patients without nodal involvement or distant metastases. J Urol, 2009. 181: 2027.
https://pubmed.ncbi.nlm.nih.gov/19286201
81. Poon, S.A., et al. Invasion of renal sinus fat is not an independent predictor of survival in pT3a renal
cell carcinoma. BJU Int, 2009. 103: 1622.
https://pubmed.ncbi.nlm.nih.gov/19154464

RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021 59


82. Bedke, J., et al. Perinephric and renal sinus fat infiltration in pT3a renal cell carcinoma: possible
prognostic differences. BJU Int, 2009. 103: 1349.
https://pubmed.ncbi.nlm.nih.gov/19076147
83. Heidenreich, A., et al. Preoperative imaging in renal cell cancer. World J Urol, 2004. 22: 307.
https://pubmed.ncbi.nlm.nih.gov/15290202
84. Sheth, S., et al. Current concepts in the diagnosis and management of renal cell carcinoma: role of
multidetector ct and three-dimensional CT. Radiographics, 2001. 21 Spec No: S237.
https://pubmed.ncbi.nlm.nih.gov/11598260
85. Amin, M.B., et al. The Eighth Edition AJCC Cancer Staging Manual: Continuing to build a bridge
from a population-based to a more “personalized” approach to cancer staging. CA Cancer J Clin,
2017. 67: 93.
https://pubmed.ncbi.nlm.nih.gov/28094848
86. Klatte, T., et al. A Literature Review of Renal Surgical Anatomy and Surgical Strategies for Partial
Nephrectomy. Eur Urol, 2015. 68: 980.
https://pubmed.ncbi.nlm.nih.gov/25911061
87. Spaliviero, M., et al. An Arterial Based Complexity (ABC) Scoring System to Assess the Morbidity
Profile of Partial Nephrectomy. Eur Urol, 2016. 69: 72.
https://pubmed.ncbi.nlm.nih.gov/26298208
88. Hakky, T.S., et al. Zonal NePhRO scoring system: a superior renal tumor complexity classification
model. Clin Genitourin Cancer, 2014. 12: e13.
https://pubmed.ncbi.nlm.nih.gov/24120084
89. Jayson, M., et al. Increased incidence of serendipitously discovered renal cell carcinoma. Urology,
1998. 51: 203.
https://pubmed.ncbi.nlm.nih.gov/9495698
90. Patard, J.J., et al. Correlation between symptom graduation, tumor characteristics and survival in
renal cell carcinoma. Eur Urol, 2003. 44: 226.
https://pubmed.ncbi.nlm.nih.gov/12875943
91. Lee, C.T., et al. Mode of presentation of renal cell carcinoma provides prognostic information. Urol
Oncol, 2002. 7: 135.
https://pubmed.ncbi.nlm.nih.gov/12474528
92. Sacco, E., et al. Paraneoplastic syndromes in patients with urological malignancies. Urol Int, 2009.
83: 1.
https://pubmed.ncbi.nlm.nih.gov/19641351
93. Kim, H.L., et al. Paraneoplastic signs and symptoms of renal cell carcinoma: implications for
prognosis. J Urol, 2003. 170: 1742.
https://pubmed.ncbi.nlm.nih.gov/14532767
94. Magera, J.S., Jr., et al. Association of abnormal preoperative laboratory values with survival after
radical nephrectomy for clinically confined clear cell renal cell carcinoma. Urology, 2008. 71: 278.
https://pubmed.ncbi.nlm.nih.gov/18308103
95. Uzzo, R.G., et al. Nephron sparing surgery for renal tumors: indications, techniques and outcomes.
J Urol, 2001. 166: 6.
https://pubmed.ncbi.nlm.nih.gov/11435813
96. Huang, W.C., et al. Chronic kidney disease after nephrectomy in patients with renal cortical tumours:
a retrospective cohort study. Lancet Oncol, 2006. 7: 735.
https://pubmed.ncbi.nlm.nih.gov/16945768
97. Israel, G.M., et al. How I do it: evaluating renal masses. Radiology, 2005. 236: 441.
https://pubmed.ncbi.nlm.nih.gov/16040900
98. Israel, G.M., et al. Pitfalls in renal mass evaluation and how to avoid them. Radiographics, 2008.
28: 1325.
https://pubmed.ncbi.nlm.nih.gov/18794310
99. Choudhary, S., et al. Renal oncocytoma: CT features cannot reliably distinguish oncocytoma from
other renal neoplasms. Clin Radiol, 2009. 64: 517.
https://pubmed.ncbi.nlm.nih.gov/19348848
100. Rosenkrantz, A.B., et al. MRI features of renal oncocytoma and chromophobe renal cell carcinoma.
AJR Am J Roentgenol, 2010. 195: W421.
https://pubmed.ncbi.nlm.nih.gov/21098174
101. Hindman, N., et al. Angiomyolipoma with minimal fat: can it be differentiated from clear cell renal cell
carcinoma by using standard MR techniques? Radiology, 2012. 265: 468.
https://pubmed.ncbi.nlm.nih.gov/23012463

60 RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021


102. Pedrosa, I., et al. MR imaging of renal masses: correlation with findings at surgery and pathologic
analysis. Radiographics, 2008. 28: 985.
https://pubmed.ncbi.nlm.nih.gov/18635625
103. Yamashita, Y. et al. The therapeutic value of lymph node dissection for renal cell carcinoma.
Nishinihon J Urol, 1989: 777. [No abstract available].
104. Gong, I.H., et al. Relationship among total kidney volume, renal function and age. J Urol, 2012.
187: 344.
https://pubmed.ncbi.nlm.nih.gov/22099987
105. Ferda, J., et al. Assessment of the kidney tumor vascular supply by two-phase MDCT-angiography.
Eur J Radiol, 2007. 62: 295.
https://pubmed.ncbi.nlm.nih.gov/17324548
106. Shao, P., et al. Precise segmental renal artery clamping under the guidance of dual-source
computed tomography angiography during laparoscopic partial nephrectomy. Eur Urol, 2012. 62:
1001.
https://pubmed.ncbi.nlm.nih.gov/22695243
107. Fan, L., et al. Diagnostic efficacy of contrast-enhanced ultrasonography in solid renal parenchymal
lesions with maximum diameters of 5 cm. J Ultrasound Med, 2008. 27: 875.
https://pubmed.ncbi.nlm.nih.gov/18499847
108. Correas, J.M., et al. [Guidelines for contrast enhanced ultrasound (CEUS)--update 2008]. J Radiol,
2009. 90: 123.
https://pubmed.ncbi.nlm.nih.gov/19212280
109. Mitterberger, M., et al. Contrast-enhanced ultrasound for diagnosis of prostate cancer and kidney
lesions. Eur J Radiol, 2007. 64: 231.
https://pubmed.ncbi.nlm.nih.gov/17881175
110. Janus, C.L., et al. Comparison of MRI and CT for study of renal and perirenal masses. Crit Rev
Diagn Imaging, 1991. 32: 69.
https://pubmed.ncbi.nlm.nih.gov/1863349
111. Mueller-Lisse, U.G., et al. Imaging of advanced renal cell carcinoma. World J Urol, 2010. 28: 253.
https://pubmed.ncbi.nlm.nih.gov/20458484
112. Kabala, J.E., et al. Magnetic resonance imaging in the staging of renal cell carcinoma. Br J Radiol,
1991. 64: 683.
https://pubmed.ncbi.nlm.nih.gov/1884119
113. Hallscheidt, P.J., et al. Preoperative staging of renal cell carcinoma with inferior vena cava thrombus
using multidetector CT and MRI: prospective study with histopathological correlation. J Comput
Assist Tomogr, 2005. 29: 64.
https://pubmed.ncbi.nlm.nih.gov/15665685
114. Putra, L.G., et al. Improved assessment of renal lesions in pregnancy with magnetic resonance
imaging. Urology, 2009. 74: 535.
https://pubmed.ncbi.nlm.nih.gov/19604560
115. Giannarini, G., et al. Potential and limitations of diffusion-weighted magnetic resonance imaging in
kidney, prostate, and bladder cancer including pelvic lymph node staging: a critical analysis of the
literature. Eur Urol, 2012. 61: 326.
https://pubmed.ncbi.nlm.nih.gov/22000497
116. Johnson, B.A., et al. Diagnostic performance of prospectively assigned clear cell Likelihood scores
(ccLS) in small renal masses at multiparametric magnetic resonance imaging. Urol Oncol, 2019.
37: 941.
https://pubmed.ncbi.nlm.nih.gov/31540830
117. Steinberg, R.L., et al. Prospective performance of clear cell likelihood scores (ccLS) in renal masses
evaluated with multiparametric magnetic resonance imaging. Eur Radiol, 2021. 31: 314.
https://pubmed.ncbi.nlm.nih.gov/32770377
118. Capogrosso, P., et al. Follow-up After Treatment for Renal Cell Carcinoma: The Evidence Beyond the
Guidelines. Eur Urol Focus, 2016. 1: 272.
https://pubmed.ncbi.nlm.nih.gov/28723399
119. Park, J.W., et al. Significance of 18F-fluorodeoxyglucose positron-emission tomography/computed
tomography for the postoperative surveillance of advanced renal cell carcinoma. BJU Int, 2009.
103: 615.
https://pubmed.ncbi.nlm.nih.gov/19007371
120. Bechtold, R.E., et al. Imaging approach to staging of renal cell carcinoma. Urol Clin North Am, 1997.
24: 507.
https://pubmed.ncbi.nlm.nih.gov/9275976

RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021 61


121. Miles, K.A., et al. CT staging of renal carcinoma: a prospective comparison of three dynamic
computed tomography techniques. Eur J Radiol, 1991. 13: 37.
https://pubmed.ncbi.nlm.nih.gov/1889427
122. Lim, D.J., et al. Computerized tomography in the preoperative staging for pulmonary metastases in
patients with renal cell carcinoma. J Urol, 1993. 150: 1112.
https://pubmed.ncbi.nlm.nih.gov/8371366
123. Larcher, A., et al. When to perform preoperative chest computed tomography for renal cancer
staging. BJU Int, 2017. 120: 490.
https://pubmed.ncbi.nlm.nih.gov/27684653
124. Voss, J., et al. Chest computed tomography for staging renal tumours: validation and simplification
of a risk prediction model from a large contemporary retrospective cohort. BJU Int, 2020. 125: 561.
https://pubmed.ncbi.nlm.nih.gov/31955483
125. Marshall, M.E., et al. Low incidence of asymptomatic brain metastases in patients with renal cell
carcinoma. Urology, 1990. 36: 300.
https://pubmed.ncbi.nlm.nih.gov/2219605
126. Koga, S., et al. The diagnostic value of bone scan in patients with renal cell carcinoma. J Urol, 2001.
166: 2126.
https://pubmed.ncbi.nlm.nih.gov/11696720
127. Henriksson, C., et al. Skeletal metastases in 102 patients evaluated before surgery for renal cell
carcinoma. Scand J Urol Nephrol, 1992. 26: 363.
https://pubmed.ncbi.nlm.nih.gov/1292074
128. Seaman, E., et al. Association of radionuclide bone scan and serum alkaline phosphatase in
patients with metastatic renal cell carcinoma. Urology, 1996. 48: 692.
https://pubmed.ncbi.nlm.nih.gov/8911510
129. Beuselinck, B., et al. Whole-body diffusion-weighted magnetic resonance imaging for the detection
of bone metastases and their prognostic impact in metastatic renal cell carcinoma patients treated
with angiogenesis inhibitors. Acta Oncol, 2020. 59: 818.
https://pubmed.ncbi.nlm.nih.gov/32297532
130. Warren, K.S., et al. The Bosniak classification of renal cystic masses. BJU Int, 2005. 95: 939.
https://pubmed.ncbi.nlm.nih.gov/15839908
131. Bosniak, M.A. The use of the Bosniak classification system for renal cysts and cystic tumors. J Urol,
1997. 157: 1852.
https://pubmed.ncbi.nlm.nih.gov/9112545
132. Richard, P.O., et al. Renal Tumor Biopsy for Small Renal Masses: A Single-center 13-year
Experience. Eur Urol, 2015. 68: 1007.
https://pubmed.ncbi.nlm.nih.gov/25900781
133. Shannon, B.A., et al. The value of preoperative needle core biopsy for diagnosing benign lesions
among small, incidentally detected renal masses. J Urol, 2008. 180: 1257.
https://pubmed.ncbi.nlm.nih.gov/18707712
134. Maturen, K.E., et al. Renal mass core biopsy: accuracy and impact on clinical management. AJR Am
J Roentgenol, 2007. 188: 563.
https://pubmed.ncbi.nlm.nih.gov/17242269
135. Volpe, A., et al. Contemporary results of percutaneous biopsy of 100 small renal masses: a single
center experience. J Urol, 2008. 180: 2333.
https://pubmed.ncbi.nlm.nih.gov/18930274
136. Veltri, A., et al. Diagnostic accuracy and clinical impact of imaging-guided needle biopsy of renal
masses. Retrospective analysis on 150 cases. Eur Radiol, 2011. 21: 393.
https://pubmed.ncbi.nlm.nih.gov/20809129
137. Abel, E.J., et al. Percutaneous biopsy of primary tumor in metastatic renal cell carcinoma to predict
high risk pathological features: comparison with nephrectomy assessment. J Urol, 2010. 184: 1877.
https://pubmed.ncbi.nlm.nih.gov/20850148
138. Richard, P.O., et al. Is Routine Renal Tumor Biopsy Associated with Lower Rates of Benign
Histology following Nephrectomy for Small Renal Masses? J Urol, 2018. 200: 731.
https://pubmed.ncbi.nlm.nih.gov/29653161
139. Marconi, L., et al. Systematic Review and Meta-analysis of Diagnostic Accuracy of Percutaneous
Renal Tumour Biopsy. Eur Urol, 2016. 69: 660.
https://pubmed.ncbi.nlm.nih.gov/26323946
140. Leveridge, M.J., et al. Outcomes of small renal mass needle core biopsy, nondiagnostic
percutaneous biopsy, and the role of repeat biopsy. Eur Urol, 2011. 60: 578.
https://pubmed.ncbi.nlm.nih.gov/21704449

62 RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021


141. Breda, A., et al. Comparison of accuracy of 14-, 18- and 20-G needles in ex-vivo renal mass biopsy:
a prospective, blinded study. BJU Int, 2010. 105: 940.
https://pubmed.ncbi.nlm.nih.gov/19888984
142. Cate, F., et al. Core Needle Biopsy and Fine Needle Aspiration Alone or in Combination: Diagnostic
Accuracy and Impact on Management of Renal Masses. J Urol, 2017. 197: 1396.
https://pubmed.ncbi.nlm.nih.gov/28093293
143. Yang, C.S., et al. Percutaneous biopsy of the renal mass: FNA or core needle biopsy? Cancer
Cytopathol, 2017. 125: 407.
https://pubmed.ncbi.nlm.nih.gov/28334518
144. Motzer, R.J., et al. Phase II randomized trial comparing sequential first-line everolimus and second-
line sunitinib versus first-line sunitinib and second-line everolimus in patients with metastatic renal
cell carcinoma. J Clin Oncol, 2014. 32: 2765.
https://pubmed.ncbi.nlm.nih.gov/25049330
145. Wood, B.J., et al. Imaging guided biopsy of renal masses: indications, accuracy and impact on
clinical management. J Urol, 1999. 161: 1470.
https://pubmed.ncbi.nlm.nih.gov/10210375
146. Somani, B.K., et al. Image-guided biopsy-diagnosed renal cell carcinoma: critical appraisal of
technique and long-term follow-up. Eur Urol, 2007. 51: 1289.
https://pubmed.ncbi.nlm.nih.gov/17081679
147. Vasudevan, A., et al. Incidental renal tumours: the frequency of benign lesions and the role of
preoperative core biopsy. BJU Int, 2006. 97: 946.
https://pubmed.ncbi.nlm.nih.gov/16643475
148. Neuzillet, Y., et al. Accuracy and clinical role of fine needle percutaneous biopsy with computerized
tomography guidance of small (less than 4.0 cm) renal masses. J Urol, 2004. 171: 1802.
https://pubmed.ncbi.nlm.nih.gov/15076280
149. Schmidbauer, J., et al. Diagnostic accuracy of computed tomography-guided percutaneous biopsy
of renal masses. Eur Urol, 2008. 53: 1003.
https://pubmed.ncbi.nlm.nih.gov/18061339
150. Wunderlich, H., et al. The accuracy of 250 fine needle biopsies of renal tumors. J Urol, 2005. 174: 44.
https://pubmed.ncbi.nlm.nih.gov/15947574
151. Abel, E.J., et al. Multi-Quadrant Biopsy Technique Improves Diagnostic Ability in Large
Heterogeneous Renal Masses. J Urol, 2015. 194: 886.
https://pubmed.ncbi.nlm.nih.gov/25837535
152. Macklin, P.S., et al. Tumour Seeding in the Tract of Percutaneous Renal Tumour Biopsy: A Report on
Seven Cases from a UK Tertiary Referral Centre. Eur Urol, 2019. 75: 861.
https://pubmed.ncbi.nlm.nih.gov/30591353
153. Cooper, S., et al. Diagnostic Yield and Complication Rate in Percutaneous Needle Biopsy of Renal
Hilar Masses With Comparison With Renal Cortical Mass Biopsies in a Cohort of 195 Patients. AJR
Am J Roentgenol, 2019. 212: 570.
https://pubmed.ncbi.nlm.nih.gov/30645159
154. Amin, M.B., et al., AJCC Cancer Staging Manual. 8th ed. 2017.
https://www.springer.com/gp/book/9783319406176#aboutBook
155. Bierley, J.D., et al., UICC TNM classification of malignant tumours. 2017, Chichester, UK.
https://www.uicc.org/resources/tnm
156. Sun, M., et al. Prognostic factors and predictive models in renal cell carcinoma: a contemporary
review. Eur Urol, 2011. 60: 644.
https://pubmed.ncbi.nlm.nih.gov/21741163
157. Zhang, L., et al. Tumor necrosis as a prognostic variable for the clinical outcome in patients with
renal cell carcinoma: a systematic review and meta-analysis. BMC Cancer, 2018. 18: 870.
https://pubmed.ncbi.nlm.nih.gov/30176824
158. Fuhrman, S.A., et al. Prognostic significance of morphologic parameters in renal cell carcinoma. Am
J Surg Pathol, 1982. 6: 655.
https://pubmed.ncbi.nlm.nih.gov/7180965
159. Delahunt, B., et al. The International Society of Urological Pathology (ISUP) grading system for renal
cell carcinoma and other prognostic parameters. Am J Surg Pathol, 2013. 37: 1490.
https://pubmed.ncbi.nlm.nih.gov/24025520
160. Paner, G.P., et al. Updates in the Eighth Edition of the Tumor-Node-Metastasis Staging Classification
for Urologic Cancers. Eur Urol, 2018. 73: 560.
https://pubmed.ncbi.nlm.nih.gov/29325693

RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021 63


161. Dagher, J., et al. Clear cell renal cell carcinoma: validation of World Health Organization/International
Society of Urological Pathology grading. Histopathology, 2017. 71: 918.
https://pubmed.ncbi.nlm.nih.gov/28718911
162. Leibovich, B.C., et al. Histological subtype is an independent predictor of outcome for patients with
renal cell carcinoma. J Urol, 2010. 183: 1309.
https://pubmed.ncbi.nlm.nih.gov/20171681
163. Adibi, M., et al. Percentage of sarcomatoid component as a prognostic indicator for survival in renal
cell carcinoma with sarcomatoid dedifferentiation. Urol Oncol, 2015. 33: 427.e17.
https://pubmed.ncbi.nlm.nih.gov/26004164
164. Kim, T., et al. Using percentage of sarcomatoid differentiation as a prognostic factor in renal cell
carcinoma. Clin Genitourin Cancer, 2015. 13: 225.
https://pubmed.ncbi.nlm.nih.gov/25544725
165. Ohashi, R., et al. Multi-institutional re-evaluation of prognostic factors in chromophobe renal cell
carcinoma: proposal of a novel two-tiered grading scheme. Virchows Arch, 2020. 476: 409.
https://pubmed.ncbi.nlm.nih.gov/31760491
166. Cheville, J.C., et al. Comparisons of outcome and prognostic features among histologic subtypes of
renal cell carcinoma. Am J Surg Pathol, 2003. 27: 612.
https://pubmed.ncbi.nlm.nih.gov/12717246
167. Patard, J.J., et al. Prognostic value of histologic subtypes in renal cell carcinoma: a multicenter
experience. J Clin Oncol, 2005. 23: 2763.
https://pubmed.ncbi.nlm.nih.gov/15837991
168. Capitanio, U., et al. A critical assessment of the prognostic value of clear cell, papillary and
chromophobe histological subtypes in renal cell carcinoma: a population-based study. BJU Int,
2009. 103: 1496.
https://pubmed.ncbi.nlm.nih.gov/19076149
169. Wagener, N., et al. Outcome of papillary versus clear cell renal cell carcinoma varies significantly in
non-metastatic disease. PLoS One, 2017. 12: e0184173.
https://pubmed.ncbi.nlm.nih.gov/28934212
170. Wong, E.C.L., et al. Morphologic subtyping as a prognostic predictor for survival in papillary renal
cell carcinoma: Type 1 vs. type 2. Urol Oncol: Sem Orig Invest, 2019. 37: 721.
https://pubmed.ncbi.nlm.nih.gov/31176614
171. Klatte, T., et al. The VENUSS prognostic model to predict disease recurrence following surgery
for non-metastatic papillary renal cell carcinoma: Development and evaluation using the ASSURE
prospective clinical trial cohort. BMC Med, 2019. 17: 182.
https://pubmed.ncbi.nlm.nih.gov/31578141
172. Deng, J., et al. A comparison of the prognosis of papillary and clear cell renal cell carcinoma:
Evidence from a meta-analysis. Medicine (Baltimore), 2019. 98: e16309.
https://pubmed.ncbi.nlm.nih.gov/31277173
173. Klatte, T., et al. Renal cell carcinoma associated with transcription factor E3 expression and Xp11.2
translocation: incidence, characteristics, and prognosis. Am J Clin Pathol, 2012. 137: 761.
https://pubmed.ncbi.nlm.nih.gov/22523215
174. Linehan, W.M., et al. Genetic basis of cancer of the kidney: disease-specific approaches to therapy.
Clin Cancer Res, 2004. 10: 6282S.
https://pubmed.ncbi.nlm.nih.gov/15448018
175. Yang, X.J., et al. A molecular classification of papillary renal cell carcinoma. Cancer Res, 2005.
65: 5628.
https://pubmed.ncbi.nlm.nih.gov/15994935
176. Furge, K.A., et al. Identification of deregulated oncogenic pathways in renal cell carcinoma: an
integrated oncogenomic approach based on gene expression profiling. Oncogene, 2007. 26: 1346.
https://pubmed.ncbi.nlm.nih.gov/17322920
177. Boissier, R., et al. Long-term oncological outcomes of cystic renal cell carcinoma according to the
Bosniak classification. Int Urol Nephrol, 2019. 51: 951.
https://pubmed.ncbi.nlm.nih.gov/30977021
178. Wahlgren, T., et al. Treatment and overall survival in renal cell carcinoma: a Swedish population-
based study (2000-2008). Br J Cancer, 2013. 108: 1541.
https://pubmed.ncbi.nlm.nih.gov/23531701
179. Li, P., et al. Survival among patients with advanced renal cell carcinoma in the pretargeted versus
targeted therapy eras. Cancer Med, 2016. 5: 169.
https://pubmed.ncbi.nlm.nih.gov/26645975

64 RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021


180. Golijanin, B., et al. The natural history of renal cell carcinoma with isolated lymph node metastases
following surgical resection from 2006 to 2013. Urol Oncol, 2019. 37: 932.
https://pubmed.ncbi.nlm.nih.gov/31570248
181. Lee, Z., et al. Local Recurrence Following Resection of Intermediate-High Risk Non-metastatic
Renal Cell Carcinoma: An Anatomic Classification and Analysis of the ASSURE (ECOG-ACRIN
E2805) Adjuvant Trial. J Urol, 2019: 101097.
https://www.cochranelibrary.com/es/central/doi/10.1002/central/CN-01997604/full
182. Bensalah, K., et al. Prognostic value of thrombocytosis in renal cell carcinoma. J Urol, 2006. 175: 859.
https://pubmed.ncbi.nlm.nih.gov/16469566
183. Kim, H.L., et al. Cachexia-like symptoms predict a worse prognosis in localized t1 renal cell
carcinoma. J Urol, 2004. 171: 1810.
https://pubmed.ncbi.nlm.nih.gov/15076282
184. Patard, J.J., et al. Multi-institutional validation of a symptom based classification for renal cell
carcinoma. J Urol, 2004. 172: 858.
https://pubmed.ncbi.nlm.nih.gov/15310983
185. Cho, D.S., et al. Prognostic significance of modified Glasgow Prognostic Score in patients with non-
metastatic clear cell renal cell carcinoma. Scand J Urol, 2016. 50: 186.
https://pubmed.ncbi.nlm.nih.gov/26878156
186. Shao, Y., et al. Prognostic value of pretreatment neutrophil-to-lymphocyte ratio in renal cell
carcinoma: a systematic review and meta-analysis. BMC Urol, 2020. 20: 90.
https://pubmed.ncbi.nlm.nih.gov/32631294
187. Albiges, L., et al. Body Mass Index and Metastatic Renal Cell Carcinoma: Clinical and Biological
Correlations. J Clin Oncol, 2016. 34: 3655.
https://pubmed.ncbi.nlm.nih.gov/27601543
188. Donin, N.M., et al. Body Mass Index and Survival in a Prospective Randomized Trial of Localized
High-Risk Renal Cell Carcinoma. Cancer Epidemiol Biomarkers Prev, 2016. 25: 1326.
https://pubmed.ncbi.nlm.nih.gov/27418270
189. Choi, Y., et al. Body mass index and survival in patients with renal cell carcinoma: a clinical-based
cohort and meta-analysis. Int J Cancer, 2013. 132: 625.
https://pubmed.ncbi.nlm.nih.gov/22610826
190. Bagheri, M., et al. Renal cell carcinoma survival and body mass index: a dose-response meta-
analysis reveals another potential paradox within a paradox. Int J Obes (Lond), 2016. 40: 1817.
https://pubmed.ncbi.nlm.nih.gov/27686524
191. Hu, X., et al. Sarcopenia predicts prognosis of patients with renal cell carcinoma: A systematic
review and meta-analysis. Int Braz J Urol, 2020. 46: 705.
https://pubmed.ncbi.nlm.nih.gov/32213202
192. Dai, J., et al. The prognostic value of body fat components in metastasis renal cell carcinoma
patients treated with TKIs. Cancer Manag Res, 2020. 12: 891.
https://pubmed.ncbi.nlm.nih.gov/32104071
193. A Phase 3, Randomized, Open-Label Study of Nivolumab Combined With Ipilimumab Versus
Sunitinib Monotherapy in Subjects With Previously Untreated, Advanced or Metastatic Renal Cell
Carcinoma. 2015. NCT02231749. [Accessed March 2021]
https://clinicaltrials.gov/ct2/show/NCT02231749
194. Sim, S.H., et al. Prognostic utility of pre-operative circulating osteopontin, carbonic anhydrase IX
and CRP in renal cell carcinoma. Br J Cancer, 2012. 107: 1131.
https://pubmed.ncbi.nlm.nih.gov/22918393
195. Sabatino, M., et al. Serum vascular endothelial growth factor and fibronectin predict clinical
response to high-dose interleukin-2 therapy. J Clin Oncol, 2009. 27: 2645.
https://pubmed.ncbi.nlm.nih.gov/19364969
196. Li, G., et al. Serum carbonic anhydrase 9 level is associated with postoperative recurrence of
conventional renal cell cancer. J Urol, 2008. 180: 510.
https://pubmed.ncbi.nlm.nih.gov/18550116
197. Choueiri, T.K., et al. A phase I study of cabozantinib (XL184) in patients with renal cell cancer. Ann
Oncol, 2014. 25: 1603.
https://pubmed.ncbi.nlm.nih.gov/24827131
198. Lu, Y., et al. The prevalence and prognostic and clinicopathological value of PD-L1 and PD-L2 in
renal cell carcinoma patients: A systematic review and meta-analysis involving 3,389 patients. Transl
Androl Urol, 2020. 9: 367.
https://pubmed.ncbi.nlm.nih.gov/32420142

RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021 65


199. Raimondi, A., et al. Predictive Biomarkers of Response to Immunotherapy in Metastatic Renal Cell
Cancer. Front Oncol, 2020. 10: 1644.
https://pubmed.ncbi.nlm.nih.gov/32903369
200. Motzer, R.J., et al. Avelumab plus axitinib versus sunitinib in advanced renal cell carcinoma:
biomarker analysis of the phase 3 JAVELIN Renal 101 trial. Nat Med, 2020. 26: 1733.
https://pubmed.ncbi.nlm.nih.gov/32895571
201. Rini, B.I., et al. Molecular correlates differentiate response to atezolizumab+ bevacizumab vs
sunitinib: results from a phase III study (IMmotion151) in untreated metastatic renal cell carcinoma.
Ann Oncol, 2018. 29: LBA31.
https://www.annalsofoncology.org/article/S0923-7534(19)50428-8/fulltext
202. Motzer, R.J., et al. Biomarker analyses from the phase III CheckMate 214 trial of nivolumab plus
ipilimumab (N+I) or sunitinib (S) in advanced renal cell carcinoma (aRCC). J Clin Oncol, 2020. 38: 5009.
https://ascopubs.org/doi/abs/10.1200/JCO.2020.38.15_suppl.5009
203. Scelo, G., et al. KIM-1 as a Blood-Based Marker for Early Detection of Kidney Cancer: A
Prospective Nested Case-Control Study. Clin Cancer Res, 2018. 24: 5594.
https://pubmed.ncbi.nlm.nih.gov/30037816
204. Zhang, K.J., et al. Diagnostic role of kidney injury molecule-1 in renal cell carcinoma. Int Urol
Nephrol, 2019. 51: 1893.
https://link.springer.com/article/10.1007/s11255-019-02231-0
205. Minardi, D., et al. Loss of nuclear BAP1 protein expression is a marker of poor prognosis in patients
with clear cell renal cell carcinoma. Urol Oncol, 2016. 34: 338 e11.
https://pubmed.ncbi.nlm.nih.gov/27085487
206. Kapur, P., et al. Effects on survival of BAP1 and PBRM1 mutations in sporadic clear-cell renal-cell
carcinoma: a retrospective analysis with independent validation. Lancet Oncol, 2013. 14: 159.
https://pubmed.ncbi.nlm.nih.gov/23333114
207. Joseph, R.W., et al. Clear Cell Renal Cell Carcinoma Subtypes Identified by BAP1 and PBRM1
Expression. J Urol, 2016. 195: 180.
https://pubmed.ncbi.nlm.nih.gov/26300218
208. Klatte, T., et al. Cytogenetic profile predicts prognosis of patients with clear cell renal cell carcinoma.
J Clin Oncol, 2009. 27: 746.
https://pubmed.ncbi.nlm.nih.gov/19124809
209. Turajlic, S., et al. Tracking Cancer Evolution Reveals Constrained Routes to Metastases: TRACERx
Renal. Cell, 2018. 173: 581.
https://pubmed.ncbi.nlm.nih.gov/29656895
210. Kroeger, N., et al. Deletions of chromosomes 3p and 14q molecularly subclassify clear cell renal cell
carcinoma. Cancer, 2013. 119: 1547.
https://pubmed.ncbi.nlm.nih.gov/23335244
211. Rini, B., et al. A 16-gene assay to predict recurrence after surgery in localised renal cell carcinoma:
development and validation studies. Lancet Oncol, 2015. 16: 676.
https://pubmed.ncbi.nlm.nih.gov/25979595
212. Sorbellini, M., et al. A postoperative prognostic nomogram predicting recurrence for patients with
conventional clear cell renal cell carcinoma. J Urol, 2005. 173: 48.
https://pubmed.ncbi.nlm.nih.gov/15592023
213. Zisman, A., et al. Improved prognostication of renal cell carcinoma using an integrated staging
system. J Clin Oncol, 2001. 19: 1649.
https://pubmed.ncbi.nlm.nih.gov/11250993
214. Frank, I., et al. An outcome prediction model for patients with clear cell renal cell carcinoma treated
with radical nephrectomy based on tumor stage, size, grade and necrosis: the SSIGN score. J Urol,
2002. 168: 2395.
https://pubmed.ncbi.nlm.nih.gov/12441925
215. Leibovich, B.C., et al. Prediction of progression after radical nephrectomy for patients with clear cell
renal cell carcinoma: a stratification tool for prospective clinical trials. Cancer, 2003. 97: 1663.
https://pubmed.ncbi.nlm.nih.gov/12655523
216. Patard, J.J., et al. Use of the University of California Los Angeles integrated staging system to
predict survival in renal cell carcinoma: an international multicenter study. J Clin Oncol, 2004.
22: 3316.
https://pubmed.ncbi.nlm.nih.gov/15310775
217. Karakiewicz, P.I., et al. Multi-institutional validation of a new renal cancer-specific survival
nomogram. J Clin Oncol, 2007. 25: 1316.
https://pubmed.ncbi.nlm.nih.gov/17416852

66 RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021


218. Zigeuner, R., et al. External validation of the Mayo Clinic stage, size, grade, and necrosis (SSIGN)
score for clear-cell renal cell carcinoma in a single European centre applying routine pathology. Eur
Urol, 2010. 57: 102.
https://pubmed.ncbi.nlm.nih.gov/19062157
219. Okita, K., et al. Impact of Disagreement Between Two Risk Group Models on Prognosis in Patients
With Metastatic Renal-Cell Carcinoma. Clin Genitourin Cancer, 2019. 17: e440.
https://pubmed.ncbi.nlm.nih.gov/30772204
220. Massari, F., et al. Addition of Primary Metastatic Site on Bone, Brain, and Liver to IMDC Criteria in
Patients With Metastatic Renal Cell Carcinoma: A Validation Study. Clin Genitourin Cancer, 2021.
19: 32.
https://pubmed.ncbi.nlm.nih.gov/32694008
221. Martini, D.J., et al. Novel Risk Scoring System for Patients with Metastatic Renal Cell Carcinoma
Treated with Immune Checkpoint Inhibitors. Oncologist, 2020. 25: e484.
https://pubmed.ncbi.nlm.nih.gov/32162798
222. Zisman, A., et al. Risk group assessment and clinical outcome algorithm to predict the natural
history of patients with surgically resected renal cell carcinoma. J Clin Oncol, 2002. 20: 4559.
https://pubmed.ncbi.nlm.nih.gov/12454113
223. Leibovich, B.C., et al. Predicting Oncologic Outcomes in Renal Cell Carcinoma After Surgery. Eur
Urol, 2018. 73: 772.
https://pubmed.ncbi.nlm.nih.gov/29398265
224. Buti, S., et al. Validation of a new prognostic model to easily predict outcome in renal cell
carcinoma: the GRANT score applied to the ASSURE trial population. Ann Oncol, 2017. 28: 2747.
https://pubmed.ncbi.nlm.nih.gov/28945839
225. Motzer, R.J., et al. Interferon-alfa as a comparative treatment for clinical trials of new therapies
against advanced renal cell carcinoma. J Clin Oncol, 2002. 20: 289.
https://pubmed.ncbi.nlm.nih.gov/11773181
226. Karnofsky, D., et al. The use of the nitrogen mustards in the palliative treatment of carcinoma. With
particular reference to bronchogenic carcinoma. Cancer 1948. 1: 634.
https://acsjournals.onlinelibrary.wiley.com/doi/abs/10.1002/1097-0142%28194811%291%3A4%3C
634%3A%3AAID-CNCR2820010410%3E3.0.CO%3B2-L
227. Heng, D.Y., et al. External validation and comparison with other models of the International
Metastatic Renal-Cell Carcinoma Database Consortium prognostic model: a population-based
study. Lancet Oncol, 2013. 14: 141.
https://pubmed.ncbi.nlm.nih.gov/23312463
228. MacLennan, S., et al. Systematic review of perioperative and quality-of-life outcomes following
surgical management of localised renal cancer. Eur Urol, 2012. 62: 1097.
https://pubmed.ncbi.nlm.nih.gov/22841673
229. Kunath, F., et al. Partial nephrectomy versus radical nephrectomy for clinical localised renal masses.
Cochrane Database Syst Rev, 2017. 5: CD012045.
https://pubmed.ncbi.nlm.nih.gov/28485814
230. Van Poppel, H., et al. A prospective, randomised EORTC intergroup phase 3 study comparing the
oncologic outcome of elective nephron-sparing surgery and radical nephrectomy for low-stage renal
cell carcinoma. Eur Urol, 2011. 59: 543.
https://pubmed.ncbi.nlm.nih.gov/21186077
231. Thompson, R.H., et al. Radical nephrectomy for pT1a renal masses may be associated with
decreased overall survival compared with partial nephrectomy. J Urol, 2008. 179: 468.
https://pubmed.ncbi.nlm.nih.gov/18076931
232. Huang, W.C., et al. Partial nephrectomy versus radical nephrectomy in patients with small renal
tumors--is there a difference in mortality and cardiovascular outcomes? J Urol, 2009. 181: 55.
https://pubmed.ncbi.nlm.nih.gov/19012918
233. Miller, D.C., et al. Renal and cardiovascular morbidity after partial or radical nephrectomy. Cancer,
2008. 112: 511.
https://pubmed.ncbi.nlm.nih.gov/18072263
234. Capitanio, U., et al. Nephron-sparing techniques independently decrease the risk of cardiovascular
events relative to radical nephrectomy in patients with a T1a-T1b renal mass and normal
preoperative renal function. Eur Urol, 2015. 67: 683.
https://pubmed.ncbi.nlm.nih.gov/25282367
235. Scosyrev, E., et al. Renal function after nephron-sparing surgery versus radical nephrectomy: results
from EORTC randomized trial 30904. Eur Urol, 2014. 65: 372.
https://pubmed.ncbi.nlm.nih.gov/23850254

RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021 67


236. Kates, M., et al. Increased risk of overall and cardiovascular mortality after radical nephrectomy for
renal cell carcinoma 2 cm or less. J Urol, 2011. 186: 1247.
https://pubmed.ncbi.nlm.nih.gov/21849201
237. Thompson, R.H., et al. Comparison of partial nephrectomy and percutaneous ablation for cT1 renal
masses. Eur Urol, 2015. 67: 252.
https://pubmed.ncbi.nlm.nih.gov/25108580
238. Sun, M., et al. Management of localized kidney cancer: calculating cancer-specific mortality and
competing risks of death for surgery and nonsurgical management. Eur Urol, 2014. 65: 235.
https://pubmed.ncbi.nlm.nih.gov/23567066
239. Sun, M., et al. Comparison of partial vs radical nephrectomy with regard to other-cause mortality
in T1 renal cell carcinoma among patients aged >/=75 years with multiple comorbidities. BJU Int,
2013. 111: 67.
https://pubmed.ncbi.nlm.nih.gov/22612472
240. Shuch, B., et al. Overall survival advantage with partial nephrectomy: a bias of observational data?
Cancer, 2013. 119: 2981.
https://pubmed.ncbi.nlm.nih.gov/23674264
241. Lane, B.R., et al. Survival and Functional Stability in Chronic Kidney Disease Due to Surgical
Removal of Nephrons: Importance of the New Baseline Glomerular Filtration Rate. Eur Urol, 2015.
68: 996.
https://pubmed.ncbi.nlm.nih.gov/26012710
242. Van Poppel, H., et al. A prospective randomized EORTC intergroup phase 3 study comparing the
complications of elective nephron-sparing surgery and radical nephrectomy for low-stage renal cell
carcinoma. Eur Urol, 2007. 51: 1606.
https://pubmed.ncbi.nlm.nih.gov/17140723
243. Poulakis, V., et al. Quality of life after surgery for localized renal cell carcinoma: comparison between
radical nephrectomy and nephron-sparing surgery. Urology, 2003. 62: 814.
https://pubmed.ncbi.nlm.nih.gov/14624900
244. Mir, M.C., et al. Partial Nephrectomy Versus Radical Nephrectomy for Clinical T1b and T2 Renal
Tumors: A Systematic Review and Meta-analysis of Comparative Studies. Eur Urol, 2017. 71: 606.
https://pubmed.ncbi.nlm.nih.gov/27614693
245. Janssen, M.W.W., et al. Survival outcomes in patients with large (>/=7cm) clear cell renal
cell carcinomas treated with nephron-sparing surgery versus radical nephrectomy: Results of a
multicenter cohort with long-term follow-up. PLoS One, 2018. 13: e0196427.
https://pubmed.ncbi.nlm.nih.gov/29723225
246. Lane, B.R., et al. Management of the adrenal gland during partial nephrectomy. J Urol, 2009.
181: 2430.
https://pubmed.ncbi.nlm.nih.gov/19371896
247. Bekema, H.J., et al. Systematic review of adrenalectomy and lymph node dissection in locally
advanced renal cell carcinoma. Eur Urol, 2013. 64: 799.
https://pubmed.ncbi.nlm.nih.gov/23643550
248. Blom, J.H., et al. Radical nephrectomy with and without lymph-node dissection: final results of
European Organization for Research and Treatment of Cancer (EORTC) randomized phase 3 trial
30881. Eur Urol, 2009. 55: 28.
https://pubmed.ncbi.nlm.nih.gov/18848382
249. Capitanio, U., et al. Lymph node dissection in renal cell carcinoma. Eur Urol, 2011. 60: 1212.
https://pubmed.ncbi.nlm.nih.gov/21940096
250. Gershman, B., et al. Radical Nephrectomy with or without Lymph Node Dissection for High Risk
Nonmetastatic Renal Cell Carcinoma: A Multi-Institutional Analysis. J Urol, 2018. 199: 1143.
https://pubmed.ncbi.nlm.nih.gov/29225056
251. Kim S., et al. The relationship of lymph node dissection with recurrence and survival for patients
treated with nephrectomy for high-risk renal cell carcinoma. J Urol, 2012. 187: e233.
https://www.auajournals.org/doi/full/10.1016/j.juro.2012.02.649
252. Dimashkieh, H.H., et al. Extranodal extension in regional lymph nodes is associated with outcome in
patients with renal cell carcinoma. J Urol, 2006. 176: 1978.
https://pubmed.ncbi.nlm.nih.gov/17070225
253. Terrone, C., et al. Reassessing the current TNM lymph node staging for renal cell carcinoma. Eur
Urol, 2006. 49: 324.
https://pubmed.ncbi.nlm.nih.gov/16386352

68 RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021


254. Whitson, J.M., et al. Lymphadenectomy improves survival of patients with renal cell carcinoma and
nodal metastases. J Urol, 2011. 185: 1615.
https://pubmed.ncbi.nlm.nih.gov/21419453
255. Capitanio, U., et al. Extent of lymph node dissection at nephrectomy affects cancer-specific survival
and metastatic progression in specific sub-categories of patients with renal cell carcinoma (RCC).
BJU Int, 2014. 114: 210.
https://pubmed.ncbi.nlm.nih.gov/24854206
256. Gershman, B., et al. Perioperative Morbidity of Lymph Node Dissection for Renal Cell Carcinoma:
A Propensity Score-based Analysis. Eur Urol, 2018. 73: 469.
https://pubmed.ncbi.nlm.nih.gov/29132713
257. Herrlinger, A., et al. What are the benefits of extended dissection of the regional renal lymph nodes
in the therapy of renal cell carcinoma. J Urol, 1991. 146: 1224.
https://pubmed.ncbi.nlm.nih.gov/1942267
258. Chapin, B.F., et al. The role of lymph node dissection in renal cell carcinoma. Int J Clin Oncol, 2011.
16: 186.
https://pubmed.ncbi.nlm.nih.gov/21523561
259. Kwon, T., et al. Reassessment of renal cell carcinoma lymph node staging: analysis of patterns of
progression. Urology, 2011. 77: 373.
https://pubmed.ncbi.nlm.nih.gov/20817274
260. Bex, A., et al. Intraoperative sentinel node identification and sampling in clinically node-negative
renal cell carcinoma: initial experience in 20 patients. World J Urol, 2011. 29: 793.
https://pubmed.ncbi.nlm.nih.gov/21107845
261. Sherif, A.M., et al. Sentinel node detection in renal cell carcinoma. A feasibility study for detection of
tumour-draining lymph nodes. BJU Int, 2012. 109: 1134.
https://pubmed.ncbi.nlm.nih.gov/21883833
262. May, M., et al. Pre-operative renal arterial embolisation does not provide survival benefit in patients
with radical nephrectomy for renal cell carcinoma. Br J Radiol, 2009. 82: 724.
https://pubmed.ncbi.nlm.nih.gov/19255117
263. Subramanian, V.S., et al. Utility of preoperative renal artery embolization for management of renal
tumors with inferior vena caval thrombi. Urology, 2009. 74: 154.
https://pubmed.ncbi.nlm.nih.gov/19428069
264. Maxwell, N.J., et al. Renal artery embolisation in the palliative treatment of renal carcinoma.
Br J Radiol, 2007. 80: 96.
https://pubmed.ncbi.nlm.nih.gov/17495058
265. Lamb, G.W., et al. Management of renal masses in patients medically unsuitable for nephrectomy--
natural history, complications, and outcome. Urology, 2004. 64: 909.
https://pubmed.ncbi.nlm.nih.gov/15533476
266. Brewer, K., et al. Perioperative and renal function outcomes of minimally invasive partial
nephrectomy for T1b and T2a kidney tumors. J Endourol, 2012. 26: 244.
https://pubmed.ncbi.nlm.nih.gov/22192099
267. Sprenkle, P.C., et al. Comparison of open and minimally invasive partial nephrectomy for renal
tumors 4-7 centimeters. Eur Urol, 2012. 61: 593.
https://pubmed.ncbi.nlm.nih.gov/22154728
268. Peng B., et al. Retroperitoneal laparoscopic nephrectomy and open nephrectomy for radical
treatment of renal cell carcinoma: A comparison of clinical outcomes. Acad J Second Military Med
Univ, 2006: 1167.
https://www.researchgate.net/publication/283136329
269. Steinberg, A.P., et al. Laparoscopic radical nephrectomy for large (greater than 7 cm, T2) renal
tumors. J Urol, 2004. 172: 2172.
https://pubmed.ncbi.nlm.nih.gov/15538225
270. Gratzke, C., et al. Quality of life and perioperative outcomes after retroperitoneoscopic radical
nephrectomy (RN), open RN and nephron-sparing surgery in patients with renal cell carcinoma. BJU
Int, 2009. 104: 470.
https://pubmed.ncbi.nlm.nih.gov/19239445
271. Hemal, A.K., et al. Laparoscopic versus open radical nephrectomy for large renal tumors: a long-
term prospective comparison. J Urol, 2007. 177: 862.
https://pubmed.ncbi.nlm.nih.gov/17296361
272. Laird, A., et al. Matched pair analysis of laparoscopic versus open radical nephrectomy for the
treatment of T3 renal cell carcinoma. World J Urol, 2015. 33: 25.
https://pubmed.ncbi.nlm.nih.gov/24647880

RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021 69


273. Patel, P., et al. A Multicentered, Propensity Matched Analysis Comparing Laparoscopic and Open
Surgery for pT3a Renal Cell Carcinoma. J Endourol, 2017. 31: 645.
https://pubmed.ncbi.nlm.nih.gov/28381117
274. Desai, M.M., et al. Prospective randomized comparison of transperitoneal versus retroperitoneal
laparoscopic radical nephrectomy. J Urol, 2005. 173: 38.
https://pubmed.ncbi.nlm.nih.gov/15592021
275. Nambirajan, T., et al. Prospective, randomized controlled study: transperitoneal laparoscopic versus
retroperitoneoscopic radical nephrectomy. Urology, 2004. 64: 919.
https://pubmed.ncbi.nlm.nih.gov/15533478
276. Nadler, R.B., et al. A prospective study of laparoscopic radical nephrectomy for T1 tumors--is
transperitoneal, retroperitoneal or hand assisted the best approach? J Urol, 2006. 175: 1230.
https://pubmed.ncbi.nlm.nih.gov/16515966
277. Gabr, A.H., et al. Approach and specimen handling do not influence oncological perioperative and
long-term outcomes after laparoscopic radical nephrectomy. J Urol, 2009. 182: 874.
https://pubmed.ncbi.nlm.nih.gov/19616234
278. Jeong, I.G., et al. Association of Robotic-Assisted vs Laparoscopic Radical Nephrectomy With
Perioperative Outcomes and Health Care Costs, 2003 to 2015. JAMA, 2017. 318: 1561.
https://pubmed.ncbi.nlm.nih.gov/29067427
279. Asimakopoulos, A.D., et al. Robotic radical nephrectomy for renal cell carcinoma: a systematic
review. BMC Urol, 2014. 14: 75.
https://pubmed.ncbi.nlm.nih.gov/25234265
280. Soga, N., et al. Comparison of radical nephrectomy techniques in one center: minimal incision
portless endoscopic surgery versus laparoscopic surgery. Int J Urol, 2008. 15: 1018.
https://pubmed.ncbi.nlm.nih.gov/19138194
281. Park Y., et al. Laparoendoscopic single-site radical nephrectomy for localized renal cell carcinoma:
comparison with conventional laparoscopic surgery. J Endourol 2009. 23: A19.
https://pubmed.ncbi.nlm.nih.gov/20370595
282. Gill, I.S., et al. Comparison of 1,800 laparoscopic and open partial nephrectomies for single renal
tumors. J Urol, 2007. 178: 41.
https://pubmed.ncbi.nlm.nih.gov/17574056
283. Lane, B.R., et al. 7-year oncological outcomes after laparoscopic and open partial nephrectomy.
J Urol, 2010. 183: 473.
https://pubmed.ncbi.nlm.nih.gov/20006866
284. Gong, E.M., et al. Comparison of laparoscopic and open partial nephrectomy in clinical T1a renal
tumors. J Endourol, 2008. 22: 953.
https://pubmed.ncbi.nlm.nih.gov/18363510
285. Marszalek, M., et al. Laparoscopic and open partial nephrectomy: a matched-pair comparison of
200 patients. Eur Urol, 2009. 55: 1171.
https://pubmed.ncbi.nlm.nih.gov/19232819
286. Kaneko, G., et al. The benefit of laparoscopic partial nephrectomy in high body mass index patients.
Jpn J Clin Oncol, 2012. 42: 619.
https://pubmed.ncbi.nlm.nih.gov/22561514
287. Muramaki, M., et al. Prognostic Factors Influencing Postoperative Development of Chronic Kidney
Disease in Patients with Small Renal Tumors who Underwent Partial Nephrectomy. Curr Urol, 2013.
6: 129.
https://pubmed.ncbi.nlm.nih.gov/24917730
288. Tugcu, V., et al. Transperitoneal versus retroperitoneal laparoscopic partial nephrectomy: initial
experience. Arch Ital Urol Androl, 2011. 83: 175.
https://pubmed.ncbi.nlm.nih.gov/22670314
289. Minervini, A., et al. Simple enucleation is equivalent to traditional partial nephrectomy for renal cell
carcinoma: results of a nonrandomized, retrospective, comparative study. J Urol, 2011. 185: 1604.
https://pubmed.ncbi.nlm.nih.gov/21861225
290. Bazzi, W.M., et al. Comparison of laparoendoscopic single-site and multiport laparoscopic radical
and partial nephrectomy: a prospective, nonrandomized study. Urology, 2012. 80: 1039.
https://pubmed.ncbi.nlm.nih.gov/22990064
291. Masson-Lecomte, A., et al. A prospective comparison of the pathologic and surgical outcomes
obtained after elective treatment of renal cell carcinoma by open or robot-assisted partial
nephrectomy. Urol Oncol, 2013. 31: 924.
https://pubmed.ncbi.nlm.nih.gov/21906969

70 RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021


292. Peyronnet, B., et al. Comparison of 1800 Robotic and Open Partial Nephrectomies for Renal
Tumors. Ann Surg Oncol, 2016. 23: 4277.
https://pubmed.ncbi.nlm.nih.gov/27411552
293. Nisen, H., et al. Hand-assisted laparoscopic versus open partial nephrectomy in patients with T1 renal
tumor: Comparative perioperative, functional and oncological outcome. Scand J Urol, 2015: 49: 446.
https://pubmed.ncbi.nlm.nih.gov/26317448
294. Chang, K.D., et al. Functional and oncological outcomes of open, laparoscopic and robot-assisted
partial nephrectomy: a multicentre comparative matched-pair analyses with a median of 5 years’
follow-up. BJU Int, 2018. 122: 618.
https://pubmed.ncbi.nlm.nih.gov/29645344
295. Alimi, Q., et al. Comparison of Short-Term Functional, Oncological, and Perioperative Outcomes
Between Laparoscopic and Robotic Partial Nephrectomy Beyond the Learning Curve.
J Laparoendosc Adv Surg Tech A, 2018. 28: 1047.
https://pubmed.ncbi.nlm.nih.gov/29664692
296. Choi, J.E., et al. Comparison of perioperative outcomes between robotic and laparoscopic partial
nephrectomy: a systematic review and meta-analysis. Eur Urol, 2015. 67: 891.
https://pubmed.ncbi.nlm.nih.gov/25572825
297. Arora, S., et al. What is the hospital volume threshold to optimize inpatient complication rate after
partial nephrectomy? Urol Oncol, 2018. 36: 339.e17.
https://pubmed.ncbi.nlm.nih.gov/29773492
298. Xia, L., et al. Hospital volume and outcomes of robot-assisted partial nephrectomy. BJU Int, 2018.
121: 900.
https://pubmed.ncbi.nlm.nih.gov/29232025
299. Peyronnet, B., et al. Impact of hospital volume and surgeon volume on robot-assisted partial
nephrectomy outcomes: a multicentre study. BJU Int, 2018. 121: 916.
https://pubmed.ncbi.nlm.nih.gov/29504226
300. Schiavina, R., et al. Predicting positive surgical margins in partial nephrectomy: A prospective
multicentre observational study (the RECORd 2 project). Eur J Surg Oncol, 2020. 46: 1353.
https://pubmed.ncbi.nlm.nih.gov/32007380
301. Tabayoyong, W., et al. Variation in Surgical Margin Status by Surgical Approach among Patients
Undergoing Partial Nephrectomy for Small Renal Masses. J Urol, 2015. 194: 1548.
https://pubmed.ncbi.nlm.nih.gov/26094808
302. Porpiglia, F., et al. Partial Nephrectomy in Clinical T1b Renal Tumors: Multicenter Comparative Study
of Open, Laparoscopic and Robot-assisted Approach (the RECORd Project). Urology, 2016. 89: 45.
https://pubmed.ncbi.nlm.nih.gov/26743388
303. Steinestel, J., et al. Positive surgical margins in nephron-sparing surgery: risk factors and
therapeutic consequences. World J Surg Oncol, 2014. 12: 252.
https://pubmed.ncbi.nlm.nih.gov/25103683
304. Wood, E.L., et al. Local Tumor Bed Recurrence Following Partial Nephrectomy in Patients with Small
Renal Masses. J Urol, 2018. 199: 393.
https://pubmed.ncbi.nlm.nih.gov/28941919
305. Bensalah, K., et al. Positive surgical margin appears to have negligible impact on survival of renal
cell carcinomas treated by nephron-sparing surgery. Eur Urol, 2010. 57: 466.
https://pubmed.ncbi.nlm.nih.gov/19359089
306. Lopez-Costea, M.A., et al. Oncological outcomes and prognostic factors after nephron-sparing
surgery in renal cell carcinoma. Int Urol Nephrol, 2016. 48: 681.
https://pubmed.ncbi.nlm.nih.gov/26861062
307. Shah, P.H., et al. Positive Surgical Margins Increase Risk of Recurrence after Partial Nephrectomy
for High Risk Renal Tumors. J Urol, 2016. 196: 327.
https://pubmed.ncbi.nlm.nih.gov/26907508
308. Tellini, R., et al. Positive Surgical Margins Predict Progression-free Survival After Nephron-sparing
Surgery for Renal Cell Carcinoma: Results From a Single Center Cohort of 459 Cases With a
Minimum Follow-up of 5 Years. Clin Genitourin Cancer, 2019. 17: e26.
https://pubmed.ncbi.nlm.nih.gov/30266249
309. Sundaram, V., et al. Positive margin during partial nephrectomy: does cancer remain in the renal
remnant? Urology, 2011. 77: 1400.
https://pubmed.ncbi.nlm.nih.gov/21411126
310. Kim, S.P., et al. Treatment of Patients with Positive Margins after Partial Nephrectomy. J Urol, 2016.
196: 301.
https://pubmed.ncbi.nlm.nih.gov/27188474

RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021 71


311. Antic, T., et al. Partial nephrectomy for renal tumors: lack of correlation between margin status and
local recurrence. Am J Clin Pathol, 2015. 143: 645.
https://pubmed.ncbi.nlm.nih.gov/25873497
312. Zini, L., et al. A population-based comparison of survival after nephrectomy vs nonsurgical
management for small renal masses. BJU Int, 2009. 103: 899.
https://pubmed.ncbi.nlm.nih.gov/19154499
313. Xing, M., et al. Comparative Effectiveness of Thermal Ablation, Surgical Resection, and Active
Surveillance for T1a Renal Cell Carcinoma: A Surveillance, Epidemiology, and End Results (SEER)-
Medicare-linked Population Study. Radiology, 2018. 288: 81.
https://pubmed.ncbi.nlm.nih.gov/29737950
314. Sun, M., et al. 1634 Management of localized kidney cancer: calculating cancer-specific mortality and
competing-risks of death tradeoffs between surgery and acive surveillance. J Urol, 2013. 189: e672.
https://www.sciencedirect.com/science/article/pii/S0022534713033764
315. Huang W.C., et al. Surveillance for the management of small renal masses: outcomes in a
population-based cohort. J Urol, 2013: e483.
https://ascopubs.org/doi/abs/10.1200/jco.2013.31.6_suppl.343
316. Hyams E.S., et al. Partial nephrectomy vs. Non-surgical management for small renal massess: a
population-based comparison of disease-specific and overall survival. J Urol, 2012. 187: E678.
https://www.jurology.com/article/S0022-5347(12)01914-3/abstract
317. Lane, B.R., et al. Active treatment of localized renal tumors may not impact overall survival in
patients aged 75 years or older. Cancer, 2010. 116: 3119.
https://pubmed.ncbi.nlm.nih.gov/20564627
318. Hollingsworth, J.M., et al. Five-year survival after surgical treatment for kidney cancer: a population-
based competing risk analysis. Cancer, 2007. 109: 1763.
https://pubmed.ncbi.nlm.nih.gov/17351954
319. Volpe, A., et al. The natural history of incidentally detected small renal masses. Cancer, 2004. 100: 738.
https://pubmed.ncbi.nlm.nih.gov/14770429
320. Jewett, M.A., et al. Active surveillance of small renal masses: progression patterns of early stage
kidney cancer. Eur Urol, 2011. 60: 39.
https://pubmed.ncbi.nlm.nih.gov/21477920
321. Smaldone, M.C., et al. Small renal masses progressing to metastases under active surveillance: a
systematic review and pooled analysis. Cancer, 2012. 118: 997.
https://pubmed.ncbi.nlm.nih.gov/21766302
322. Patel, N., et al. Active surveillance of small renal masses offers short-term oncological efficacy
equivalent to radical and partial nephrectomy. BJU Int, 2012. 110: 1270.
https://pubmed.ncbi.nlm.nih.gov/22564495
323. Pierorazio, P.M., et al. Five-year analysis of a multi-institutional prospective clinical trial of delayed
intervention and surveillance for small renal masses: the DISSRM registry. Eur Urol, 2015. 68: 408.
https://pubmed.ncbi.nlm.nih.gov/25698065
324. Uzosike, A.C., et al. Growth Kinetics of Small Renal Masses on Active Surveillance: Variability and
Results from the DISSRM Registry. J Urol, 2018. 199: 641.
https://pubmed.ncbi.nlm.nih.gov/28951284
325. Abou Youssif, T., et al. Active surveillance for selected patients with renal masses: updated results
with long-term follow-up. Cancer, 2007. 110: 1010.
https://pubmed.ncbi.nlm.nih.gov/17628489
326. Abouassaly, R., et al. Active surveillance of renal masses in elderly patients. J Urol, 2008. 180: 505.
https://pubmed.ncbi.nlm.nih.gov/18550113
327. Crispen, P.L., et al. Natural history, growth kinetics, and outcomes of untreated clinically localized
renal tumors under active surveillance. Cancer, 2009. 115: 2844.
https://www.ncbi.nlm.nih.gov/pmc/articles/PMC2860784/
328. Rosales, J.C., et al. Active surveillance for renal cortical neoplasms. J Urol, 2010. 183: 1698.
https://pubmed.ncbi.nlm.nih.gov/20299038
329. Pierorazio P., et al. Quality of life on active surveillance for small masses versus immediate
intervention: interim analysis of the DISSRM (delayed intervention and surveillance for small renal
masses) registry. J Urol, 2013. 189: e259.
https://www.jurology.com/article/S0022-5347(13)00461-8/fulltext
330. Finelli, A., et al. Small Renal Mass Surveillance: Histology-specific Growth Rates in a Biopsy-
characterized Cohort. Eur Urol, 2020. 78: 460.
https://pubmed.ncbi.nlm.nih.gov/32680677

72 RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021


331. Atwell, T.D., et al. Percutaneous ablation of renal masses measuring 3.0 cm and smaller:
comparative local control and complications after radiofrequency ablation and cryoablation. AJR
Am J Roentgenol, 2013. 200: 461.
https://pubmed.ncbi.nlm.nih.gov/23345372
332. Widdershoven, C.V., et al. Renal biopsies performed before versus during ablation of T1 renal
tumors: implications for prevention of overtreatment and follow-up. Abdom Radiol (NY), 2021.
46: 373.
https://pubmed.ncbi.nlm.nih.gov/32564209
333. Lay, A.H., et al. Oncologic Efficacy of Radio Frequency Ablation for Small Renal Masses: Clear Cell
vs Papillary Subtype. J Urol, 2015. 194: 653.
https://pubmed.ncbi.nlm.nih.gov/25846416
334. McClure, T., et al. Efficacy of percutaneous radiofrequency ablation may vary with clear cell renal
cell cancer histologic subtype. Abdom Radiol (NY), 2018. 43: 1472.
https://pubmed.ncbi.nlm.nih.gov/28936542
335. Liu, N., et al. Percutaneous radiofrequency ablation for renal cell carcinoma vs. partial nephrectomy:
Comparison of long-term oncologic outcomes in both clear cell and non-clear cell of the most
common subtype. Urol Oncol, 2017. 35: 530.e1.
https://pubmed.ncbi.nlm.nih.gov/28408296
336. Breen, D.J., et al. Image-guided Cryoablation for Sporadic Renal Cell Carcinoma: Three- and 5-year
Outcomes in 220 Patients with Biopsy-proven Renal Cell Carcinoma. Radiology, 2018. 289: 554.
https://pubmed.ncbi.nlm.nih.gov/30084744
337. Sisul, D.M., et al. RENAL nephrometry score is associated with complications after renal
cryoablation: a multicenter analysis. Urology, 2013. 81: 775.
https://pubmed.ncbi.nlm.nih.gov/23434099
338. Kim E.H., et al. Outcomes of laparoscopic and percutaneous cryoablation for renal masses. J Urol,
2013. 189: e492. [No abstract available].
339. Goyal, J., et al. Single-center comparative oncologic outcomes of surgical and percutaneous
cryoablation for treatment of renal tumors. J Endourol, 2012. 26: 1413.
https://pubmed.ncbi.nlm.nih.gov/22642574
340. Jiang, K., et al. Laparoscopic cryoablation vs. percutaneous cryoablation for treatment of small renal
masses: a systematic review and meta-analysis. Oncotarget, 2017. 8: 27635.
https://pubmed.ncbi.nlm.nih.gov/28199973
341. Zargar, H., et al. Cryoablation for Small Renal Masses: Selection Criteria, Complications, and
Functional and Oncologic Results. Eur Urol, 2016. 69: 116.
https://pubmed.ncbi.nlm.nih.gov/25819723
342. Pickersgill, N.A., et al. Ten-Year Experience with Percutaneous Cryoablation of Renal Tumors: Tumor
Size Predicts Disease Progression. J Endourol, 2020. 34: 1211.
https://pubmed.ncbi.nlm.nih.gov/32292059
343. Hebbadj, S., et al. Safety Considerations and Local Tumor Control Following Percutaneous Image-
Guided Cryoablation of T1b Renal Tumors. Cardiovasc Intervent Radiol, 2018. 41: 449.
https://pubmed.ncbi.nlm.nih.gov/29075877
344. Grange, R., et al. Computed tomography-guided percutaneous cryoablation of T1b renal tumors:
safety, functional and oncological outcomes. Int J Hyperthermia, 2019. 36: 1065.
https://pubmed.ncbi.nlm.nih.gov/31648584
345. Pecoraro, A., et al. Cryoablation Predisposes to Higher Cancer Specific Mortality Relative to Partial
Nephrectomy in Patients with Nonmetastatic pT1b Kidney Cancer. J Urol, 2019. 202: 1120.
https://pubmed.ncbi.nlm.nih.gov/31347950
346. Andrews, J.R., et al. Oncologic Outcomes Following Partial Nephrectomy and Percutaneous
Ablation for cT1 Renal Masses. Eur Urol, 2019. 76: 244.
https://pubmed.ncbi.nlm.nih.gov/31060824
347. Sundelin, M.O., et al. Repeated Cryoablation as Treatment Modality after Failure of Primary Renal
Cryoablation: A European Registry for Renal Cryoablation Multinational Analysis. J Endourol, 2019.
33: 909.
https://pubmed.ncbi.nlm.nih.gov/31507206
348. Lian, H., et al. Single-center comparison of complications in laparoscopic and percutaneous
radiofrequency ablation with ultrasound guidance for renal tumors. Urology, 2012. 80: 119.
https://pubmed.ncbi.nlm.nih.gov/22633890
349. Young, E.E., et al. Comparison of safety, renal function outcomes and efficacy of laparoscopic and
percutaneous radio frequency ablation of renal masses. J Urol, 2012. 187: 1177.
https://pubmed.ncbi.nlm.nih.gov/22357170

RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021 73


350. Kim, S.D., et al. Radiofrequency ablation of renal tumors: four-year follow-up results in 47 patients.
Korean J Radiol, 2012. 13: 625.
https://pubmed.ncbi.nlm.nih.gov/22977331
351. Trudeau, V., et al. Comparison of Postoperative Complications and Mortality Between Laparoscopic
and Percutaneous Local Tumor Ablation for T1a Renal Cell Carcinoma: A Population-based Study.
Urology, 2016. 89: 63.
https://pubmed.ncbi.nlm.nih.gov/26514977
352. Psutka, S.P., et al. Long-term oncologic outcomes after radiofrequency ablation for T1 renal cell
carcinoma. Eur Urol, 2013. 63: 486.
https://pubmed.ncbi.nlm.nih.gov/22959191
353. Johnson, B.A., et al. Ten-Year Outcomes of Renal Tumor Radio Frequency Ablation. J Urol, 2019.
201: 251.
https://pubmed.ncbi.nlm.nih.gov/30634350
354. Chang, X., et al. Radio frequency ablation versus partial nephrectomy for clinical T1b renal cell
carcinoma: long-term clinical and oncologic outcomes. J Urol, 2015. 193: 430.
https://pubmed.ncbi.nlm.nih.gov/25106899
355. Guazzoni, G., et al. Oncologic results of laparoscopic renal cryoablation for clinical T1a tumors:
8 years of experience in a single institution. Urology, 2010. 76: 624.
https://pubmed.ncbi.nlm.nih.gov/20579705
356. Larcher, A., et al. Long-term oncologic outcomes of laparoscopic renal cryoablation as primary
treatment for small renal masses. Urol Oncol, 2015. 33: 22.e1.
https://pubmed.ncbi.nlm.nih.gov/25301741
357. Haber, G.P., et al. Tumour in solitary kidney: laparoscopic partial nephrectomy vs laparoscopic
cryoablation. BJU Int, 2012. 109: 118.
https://pubmed.ncbi.nlm.nih.gov/21895929
358. Turna, B., et al. Minimally invasive nephron sparing management for renal tumors in solitary kidneys.
J Urol, 2009. 182: 2150.
https://pubmed.ncbi.nlm.nih.gov/19758655
359. Siva, S., et al. Stereotactic ablative body radiotherapy for inoperable primary kidney cancer: a
prospective clinical trial. BJU Int, 2017. 120: 623.
https://pubmed.ncbi.nlm.nih.gov/28188682
360. Correa, R.J.M., et al. The Emerging Role of Stereotactic Ablative Radiotherapy for Primary Renal
Cell Carcinoma: A Systematic Review and Meta-Analysis. Eur Urol Focus, 2019. 5: 958.
https://pubmed.ncbi.nlm.nih.gov/31248849
361. Yu, J., et al. Percutaneous Microwave Ablation versus Laparoscopic Partial Nephrectomy for cT1a
Renal Cell Carcinoma: A Propensity-matched Cohort Study of 1955 Patients. Radiology, 2020.
294: 698.
https://pubmed.ncbi.nlm.nih.gov/31961239
362. Shapiro, D.D., et al. Comparing Outcomes for Patients with Clinical T1b Renal Cell Carcinoma
Treated With Either Percutaneous Microwave Ablation or Surgery. Urology, 2020. 135: 88.
https://pubmed.ncbi.nlm.nih.gov/31585198
363. Zhou, W., et al. Radiofrequency Ablation, Cryoablation, and Microwave Ablation for T1a Renal
Cell Carcinoma: A Comparative Evaluation of Therapeutic and Renal Function Outcomes. J Vasc
Intervent Radiol, 2019. 30: 1035.
https://pubmed.ncbi.nlm.nih.gov/30956075
364. Bhindi, B., et al. The role of lymph node dissection in the management of renal cell carcinoma: a
systematic review and meta-analysis. BJU Int, 2018. 121: 684.
https://pubmed.ncbi.nlm.nih.gov/29319926
365. Luo, X., et al. Influence of lymph node dissection in patients undergoing radical nephrectomy
for non-metastatic renal cell carcinoma: a systematic review and meta-analysis. Eur Rev Med
Pharmacol Sci, 2019. 23: 6079.
https://pubmed.ncbi.nlm.nih.gov/31364109
366. Capitanio, U., et al. When to perform lymph node dissection in patients with renal cell carcinoma:
a novel approach to the preoperative assessment of risk of lymph node invasion at surgery and of
lymph node progression during follow-up. BJU Int, 2013. 112: E59.
https://pubmed.ncbi.nlm.nih.gov/23795799
367. Tsui, K.H., et al. Prognostic indicators for renal cell carcinoma: a multivariate analysis of 643 patients
using the revised 1997 TNM staging criteria. J Urol, 2000. 163: 1090.
https://pubmed.ncbi.nlm.nih.gov/10737472

74 RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021


368. Hallscheidt, P., et al. [Preoperative and palliative embolization of renal cell carcinomas: follow-up of
49 patients]. Rofo, 2006. 178: 391.
https://pubmed.ncbi.nlm.nih.gov/16612730
369. Nesbitt, J.C., et al. Surgical management of renal cell carcinoma with inferior vena cava tumor
thrombus. Ann Thorac Surg, 1997. 63: 1592.
https://pubmed.ncbi.nlm.nih.gov/9205155
370. Hatcher, P.A., et al. Surgical management and prognosis of renal cell carcinoma invading the vena
cava. J Urol, 1991. 145: 20.
https://pubmed.ncbi.nlm.nih.gov/1984092
371. Neves, R.J., et al. Surgical treatment of renal cancer with vena cava extension. Br J Urol, 1987.
59: 390.
https://pubmed.ncbi.nlm.nih.gov/3594097
372. Haferkamp, A., et al. Renal cell carcinoma with tumor thrombus extension into the vena cava:
prospective long-term followup. J Urol, 2007. 177: 1703.
https://pubmed.ncbi.nlm.nih.gov/17437789
373. Kirkali, Z., et al. A critical analysis of surgery for kidney cancer with vena cava invasion. Eur Urol,
2007. 52: 658.
https://pubmed.ncbi.nlm.nih.gov/17548146
374. Moinzadeh, A., et al. Prognostic significance of tumor thrombus level in patients with renal cell
carcinoma and venous tumor thrombus extension. Is all T3b the same? J Urol, 2004. 171: 598.
https://pubmed.ncbi.nlm.nih.gov/14713768
375. Kaplan, S., et al. Surgical management of renal cell carcinoma with inferior vena cava tumor
thrombus. Am J Surg, 2002. 183: 292.
https://pubmed.ncbi.nlm.nih.gov/11943130
376. Bissada, N.K., et al. Long-term experience with management of renal cell carcinoma involving the
inferior vena cava. Urology, 2003. 61: 89.
https://pubmed.ncbi.nlm.nih.gov/12559273
377. Skinner, D.G., et al. Vena caval involvement by renal cell carcinoma. Surgical resection provides
meaningful long-term survival. Ann Surg, 1989. 210: 387.
https://pubmed.ncbi.nlm.nih.gov/2774709
378. Lardas, M., et al. Systematic Review of Surgical Management of Nonmetastatic Renal Cell
Carcinoma with Vena Caval Thrombus. Eur Urol, 2016. 70: 265.
https://pubmed.ncbi.nlm.nih.gov/26707869
379. Ljungberg B., et al. Systematic Review Methodology for the European Association of Urology
Guidelines for Renal Cell Carcinoma (2014 update).
https://uroweb.org/wp-content/uploads/Systematic_methodology_RCC_2014_update.pdf
380. Wotkowicz, C., et al. Management of renal cell carcinoma with vena cava and atrial thrombus:
minimal access vs median sternotomy with circulatory arrest. BJU Int, 2006. 98: 289.
https://pubmed.ncbi.nlm.nih.gov/16879667
381. Faust W., et al. Minimal access versus median sternotomy for cardiopulmonary bypass in the
management of renal cell carcinoma with vena caval and atrial involvement. J Urol, 2013. 189
(Suppl.): e255.
https://www.researchgate.net/publication/274614629
382. Orihashi, K., et al. Deep hypothermic circulatory arrest for resection of renal tumor in the inferior
vena cava: beneficial or deleterious? Circ J, 2008. 72: 1175.
https://pubmed.ncbi.nlm.nih.gov/18577831
383. Rodríguez-Fernández, I.A., et al. Adjuvant Radiation Therapy After Radical Nephrectomy in Patients
with Localized Renal Cell Carcinoma: A Systematic Review and Meta-analysis. Eur Urol Oncol,
2019. 2: 448.
https://pubmed.ncbi.nlm.nih.gov/31277782
384. Galligioni, E., et al. Adjuvant immunotherapy treatment of renal carcinoma patients with autologous
tumor cells and bacillus Calmette-Guerin: five-year results of a prospective randomized study.
Cancer, 1996. 77: 2560.
https://pubmed.ncbi.nlm.nih.gov/8640706
385. Figlin, R.A., et al. Multicenter, randomized, phase III trial of CD8(+) tumor-infiltrating lymphocytes in
combination with recombinant interleukin-2 in metastatic renal cell carcinoma. J Clin Oncol, 1999.
17: 2521.
https://pubmed.ncbi.nlm.nih.gov/10561318

RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021 75


386. Clark, J.I., et al. Adjuvant high-dose bolus interleukin-2 for patients with high-risk renal cell
carcinoma: a cytokine working group randomized trial. J Clin Oncol, 2003. 21: 3133.
https://pubmed.ncbi.nlm.nih.gov/12810695
387. Atzpodien, J., et al. Adjuvant treatment with interleukin-2- and interferon-alpha2a-based
chemoimmunotherapy in renal cell carcinoma post tumour nephrectomy: results of a prospectively
randomised trial of the German Cooperative Renal Carcinoma Chemoimmunotherapy Group
(DGCIN). Br J Cancer, 2005. 92: 843.
https://pubmed.ncbi.nlm.nih.gov/15756254
388. Jocham, D., et al. Adjuvant autologous renal tumour cell vaccine and risk of tumour progression in
patients with renal-cell carcinoma after radical nephrectomy: phase III, randomised controlled trial.
Lancet, 2004. 363: 594.
https://pubmed.ncbi.nlm.nih.gov/14987883
389. Janowitz, T., et al. Adjuvant therapy in renal cell carcinoma-past, present, and future. Semin Oncol,
2013. 40: 482.
https://pubmed.ncbi.nlm.nih.gov/23972712
390. Chamie, K., et al. Adjuvant Weekly Girentuximab Following Nephrectomy for High-Risk Renal Cell
Carcinoma: The ARISER Randomized Clinical Trial. JAMA Oncol, 2017. 3: 913.
https://pubmed.ncbi.nlm.nih.gov/25823535
391. Haas, N.B., et al. Adjuvant Treatment for High-Risk Clear Cell Renal Cancer: Updated Results of a
High-Risk Subset of the ASSURE Randomized Trial. JAMA Oncol, 2017. 3: 1249.
https://pubmed.ncbi.nlm.nih.gov/28278333
392. Haas, N.B., et al. Initial results from ASSURE (E2805): Adjuvant sorafenib or sunitinib for unfavorable
renal carcinoma, an ECOG-ACRIN-led, NCTN phase III trial. ASCO Meeting Abstracts, 2015. 33: 403.
https://ascopubs.org/doi/abs/10.1200/jco.2015.33.7_suppl.403
393. Motzer, R.J., et al. Randomized Phase III Trial of Adjuvant Pazopanib Versus Placebo After
Nephrectomy in Patients With Localized or Locally Advanced Renal Cell Carcinoma. J Clin Oncol,
2017. 35: 3916.
https://pubmed.ncbi.nlm.nih.gov/28902533
394. Harshman, L.C., et al. Meta-analysis of disease free survival (DFS) as a surrogate for overall survival
(OS) in localized renal cell carcinoma (RCC). J Clin Oncol, 2017. 35: 459.
https://pubmed.ncbi.nlm.nih.gov/29266178
395. Lenis, A.T., et al. Adjuvant Therapy for High Risk Localized Kidney Cancer: Emerging Evidence and
Future Clinical Trials. J Urol, 2018. 199: 43.
https://pubmed.ncbi.nlm.nih.gov/28479237
396. Gross-Goupil, M., et al. Axitinib versus placebo as an adjuvant treatment of renal cell carcinoma:
results from the phase III, randomized ATLAS trial. Ann Oncol, 2018. 29: 2371.
https://pubmed.ncbi.nlm.nih.gov/30346481
397. Motzer, R.J., et al. Adjuvant Sunitinib for High-risk Renal Cell Carcinoma After Nephrectomy:
Subgroup Analyses and Updated Overall Survival Results. Eur Urol, 2018. 73: 62.
https://pubmed.ncbi.nlm.nih.gov/28967554
398. Massari, F., et al. Adjuvant Tyrosine Kinase Inhibitors in Treatment of Renal Cell Carcinoma: A Meta-
Analysis of Available Clinical Trials. Clin Genitourin Cancer, 2019. 17: e339.
https://pubmed.ncbi.nlm.nih.gov/30704796
399. Flanigan, R.C., et al. Cytoreductive nephrectomy in patients with metastatic renal cancer: a
combined analysis. J Urol, 2004. 171: 1071.
https://pubmed.ncbi.nlm.nih.gov/14767273
400. Clinical Trial to Assess the Importance of Nephrectomy (CARMENA). 2009. 2019 p. NCT00930033.
https://pubmed.ncbi.nlm.nih.gov/https://clinicaltrials.gov/ct2/show/NCT00930033
401. Immediate Surgery or Surgery After Sunitinib Malate in Treating Patients With Metastatic Kidney
Cancer (SURTIME). 2019. [Accessed March 2021]
https://clinicaltrials.gov/ct2/show/results/NCT01099423
402. Bhindi, B., et al. Systematic Review of the Role of Cytoreductive Nephrectomy in the Targeted
Therapy Era and Beyond: An Individualized Approach to Metastatic Renal Cell Carcinoma. Eur Urol,
2019. 75: 111.
https://pubmed.ncbi.nlm.nih.gov/30467042
403. Mejean, A., et al. Sunitinib Alone or after Nephrectomy in Metastatic Renal-Cell Carcinoma. N Engl
J Med, 2018. 379: 417.
https://www.nejm.org/doi/full/10.1056/NEJMoa1803675

76 RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021


404. Bex, A., et al. Comparison of Immediate vs Deferred Cytoreductive Nephrectomy in Patients With
Synchronous Metastatic Renal Cell Carcinoma Receiving Sunitinib: The SURTIME Randomized
Clinical Trial. JAMA Oncol, 2019. 5: 164.
https://pubmed.ncbi.nlm.nih.gov/30543350
405. Powles, T., et al. The outcome of patients treated with sunitinib prior to planned nephrectomy in
metastatic clear cell renal cancer. Eur Urol, 2011. 60: 448.
https://pubmed.ncbi.nlm.nih.gov/21612860
406. Heng, D.Y., et al. Cytoreductive nephrectomy in patients with synchronous metastases from
renal cell carcinoma: results from the International Metastatic Renal Cell Carcinoma Database
Consortium. Eur Urol, 2014. 66: 704.
https://pubmed.ncbi.nlm.nih.gov/24931622
407. de Bruijn, R., et al. Deferred Cytoreductive Nephrectomy Following Presurgical Vascular Endothelial
Growth Factor Receptor-targeted Therapy in Patients with Primary Metastatic Clear Cell Renal Cell
Carcinoma: A Pooled Analysis of Prospective Trial Data. Eur Urol Oncol, 2020. 3: 168.
https://pubmed.ncbi.nlm.nih.gov/31956080
408. Ljungberg, B., et al. Survival advantage of upfront cytoreductive nephrectomy in patients with
primary metastatic renal cell carcinoma compared with systemic and palliative treatments in a real-
world setting. Scand J Urol, 2020. 54: 487.
https://pubmed.ncbi.nlm.nih.gov/32897123
409. Motzer, R.J., et al. Nivolumab plus Ipilimumab versus Sunitinib in Advanced Renal-Cell Carcinoma.
N Engl J Med, 2018. 378: 1277.
https://pubmed.ncbi.nlm.nih.gov/29562145
410. Choueiri, T.K., et al. Nivolumab plus Cabozantinib versus Sunitinib for Advanced Renal-Cell
Carcinoma. N Engl J Med. 2021. 384: 829.
https://pubmed.ncbi.nlm.nih.gov/33657295
411. Motzer, R.J., et al. Avelumab plus Axitinib versus Sunitinib for Advanced Renal-Cell Carcinoma.
N Engl J Med, 2019. 380: 1103.
https://pubmed.ncbi.nlm.nih.gov/30779531
412. Soulières, D., et al. Pembrolizumab Plus Axitinib Versus Sunitinib as First-Line Therapy for
Advanced Renal Cell Carcinoma (RCC): Subgroup Analysis From KEYNOTE-426 by Prior
Nephrectomy 19th annual meeting of the International Kidney Cancer Symposium, 2020. A Virtual
Experience. [No abstract available].
413. Dabestani, S., et al. Local treatments for metastases of renal cell carcinoma: a systematic review.
Lancet Oncol, 2014. 15: e549.
https://pubmed.ncbi.nlm.nih.gov/25439697
414. Dabestani, S., et al. EAU Renal Cell Carcinoma Guideline Panel. Systematic review methodology for
the EAU RCC Guideline 2013.
https://uroweb.org/wp-content/uploads/Systematic_methodology_RCC_2014_update.pdf
415. Brinkmann, O.A., et al. The Role of Residual Tumor Resection in Patients with Metastatic Renal Cell
Carcinoma and Partial Remission following Immunochemotherapy. Eur Urol Suppl, 2007. 6: 641.
https://www.eusupplements.europeanurology.com/article/S1569-9056(07)00097-8/pdf
416. Alt, A.L., et al. Survival after complete surgical resection of multiple metastases from renal cell
carcinoma. Cancer, 2011. 117: 2873.
https://pubmed.ncbi.nlm.nih.gov/21692048
417. Kwak, C., et al. Metastasectomy without systemic therapy in metastatic renal cell carcinoma:
comparison with conservative treatment. Urol Int, 2007. 79: 145.
https://pubmed.ncbi.nlm.nih.gov/17851285
418. Petralia, G., et al. 450 Complete metastasectomy is an independent predictor of cancer-specific
survival in patients with clinically metastatic renal cell carcinoma. Eur Urol Suppl, 2010. 9: 162.
https://www.eusupplements.europeanurology.com/article/S1569-9056(10)60446-0/abstract
419. Russo, P., et al. Cytoreductive nephrectomy and nephrectomy/complete metastasectomy for
metastatic renal cancer. Sci World J, 2007. 7: 768.
https://pubmed.ncbi.nlm.nih.gov/17619759
420. Staehler, M.D., et al. Metastasectomy significantly prolongs survival in patients with metastatic renal
cancer. Eur Urol Suppl, 2009: 181: 498.
https://www.jurology.com/article/S0022-5347(09)61409-9/pdf
421. Eggener, S.E., et al. Risk score and metastasectomy independently impact prognosis of patients
with recurrent renal cell carcinoma. J Urol, 2008. 180: 873.
https://pubmed.ncbi.nlm.nih.gov/18635225

RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021 77


422. Lee, S.E., et al. Metastatectomy prior to immunochemotherapy for metastatic renal cell carcinoma.
Urol Int, 2006. 76: 256.
https://pubmed.ncbi.nlm.nih.gov/16601390
423. Fuchs, B., et al. Solitary bony metastasis from renal cell carcinoma: significance of surgical
treatment. Clin Orthop Relat Res, 2005: 187.
https://pubmed.ncbi.nlm.nih.gov/15685074
424. Hunter, G.K., et al. The efficacy of external beam radiotherapy and stereotactic body radiotherapy
for painful spinal metastases from renal cell carcinoma. Pract Radiat Oncol, 2012. 2: e95.
https://pubmed.ncbi.nlm.nih.gov/24674192
425. Zelefsky, M.J., et al. Tumor control outcomes after hypofractionated and single-dose stereotactic
image-guided intensity-modulated radiotherapy for extracranial metastases from renal cell
carcinoma. Int J Radiat Oncol Biol Phys, 2012. 82: 1744.
https://pubmed.ncbi.nlm.nih.gov/21596489
426. Fokas, E., et al. Radiotherapy for brain metastases from renal cell cancer: should whole-brain
radiotherapy be added to stereotactic radiosurgery?: analysis of 88 patients. Strahlenther Onkol,
2010. 186: 210.
https://pubmed.ncbi.nlm.nih.gov/20165820
427. Ikushima, H., et al. Fractionated stereotactic radiotherapy of brain metastases from renal cell
carcinoma. Int J Radiat Oncol Biol Phys, 2000. 48: 1389.
https://pubmed.ncbi.nlm.nih.gov/11121638
428. Staehler, M.D., et al. Liver resection for metastatic disease prolongs survival in renal cell carcinoma:
12-year results from a retrospective comparative analysis. World J Urol, 2010. 28: 543.
https://pubmed.ncbi.nlm.nih.gov/20440505
429. Amiraliev, A. Treatment strategy in patients with pulmonary metastases of renal cell cancer. Int
Cardiovasc Thor Surg, 2012. S20.
https://www.researchgate.net/publication/284295837
430. Zerbi, A., et al. Pancreatic metastasis from renal cell carcinoma: which patients benefit from surgical
resection? Ann Surg Oncol, 2008. 15: 1161.
https://pubmed.ncbi.nlm.nih.gov/18196343
431. Kickuth, R., et al. Interventional management of hypervascular osseous metastasis: role of
embolotherapy before orthopedic tumor resection and bone stabilization. AJR Am J Roentgenol,
2008. 191: W240.
https://pubmed.ncbi.nlm.nih.gov/19020210
432. Forauer, A.R., et al. Selective palliative transcatheter embolization of bony metastases from renal
cell carcinoma. Acta Oncol, 2007. 46: 1012.
https://pubmed.ncbi.nlm.nih.gov/17851849
433. Appleman, L.J., et al. Randomized, double-blind phase III study of pazopanib versus placebo
in patients with metastatic renal cell carcinoma who have no evidence of disease following
metastasectomy: A trial of the ECOG-ACRIN cancer research group (E2810). J Clin Oncol, 2019.
37: 4502.
https://ascopubs.org/doi/10.1200/JCO.2019.37.15_suppl.4502
434. Procopio, G., et al. Sorafenib Versus Observation Following Radical Metastasectomy for Clear-cell
Renal Cell Carcinoma: Results from the Phase 2 Randomized Open-label RESORT Study. Eur Urol
Oncol, 2019. 2: 699.
https://pubmed.ncbi.nlm.nih.gov/31542243
435. Amato, R.J. Chemotherapy for renal cell carcinoma. Semin Oncol, 2000. 27: 177.
https://pubmed.ncbi.nlm.nih.gov/10768596
436. Negrier, S., et al. Medroxyprogesterone, interferon alfa-2a, interleukin 2, or combination of both
cytokines in patients with metastatic renal carcinoma of intermediate prognosis: results of a
randomized controlled trial. Cancer, 2007. 110: 2468.
https://pubmed.ncbi.nlm.nih.gov/17932908
437. Motzer, R.J., et al. Sunitinib versus interferon alfa in metastatic renal-cell carcinoma. N Engl J Med,
2007. 356: 115.
https://pubmed.ncbi.nlm.nih.gov/17215529
438. Hudes, G., et al. Temsirolimus, interferon alfa, or both for advanced renal-cell carcinoma. N Engl
J Med, 2007. 356: 2271.
https://pubmed.ncbi.nlm.nih.gov/17538086

78 RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021


439. Rosenberg, S.A., et al. Prospective randomized trial of high-dose interleukin-2 alone or in
conjunction with lymphokine-activated killer cells for the treatment of patients with advanced
cancer. J Natl Cancer Inst, 1993. 85: 622.
https://pubmed.ncbi.nlm.nih.gov/8468720
440. Heng, D.Y., et al. Prognostic factors for overall survival in patients with metastatic renal cell
carcinoma treated with vascular endothelial growth factor-targeted agents: results from a large,
multicenter study. J Clin Oncol, 2009. 27: 5794.
https://pubmed.ncbi.nlm.nih.gov/19826129
441. Fyfe, G., et al. Results of treatment of 255 patients with metastatic renal cell carcinoma who
received high-dose recombinant interleukin-2 therapy. J Clin Oncol, 1995. 13: 688.
https://pubmed.ncbi.nlm.nih.gov/7884429
442. McDermott, D.F., et al. Randomized phase III trial of high-dose interleukin-2 versus subcutaneous
interleukin-2 and interferon in patients with metastatic renal cell carcinoma. J Clin Oncol, 2005. 23: 133.
https://pubmed.ncbi.nlm.nih.gov/15625368
443. Ribas, A. Tumor immunotherapy directed at PD-1. N Engl J Med, 2012. 366: 2517.
https://pubmed.ncbi.nlm.nih.gov/22658126
444. Motzer, R.J., et al. Nivolumab versus Everolimus in Advanced Renal-Cell Carcinoma. N Engl J Med,
2015. 373: 1803.
https://pubmed.ncbi.nlm.nih.gov/26406148
445. Motzer, R.J., et al. Nivolumab versus everolimus in patients with advanced renal cell carcinoma:
Updated results with long-term follow-up of the randomized, open-label, phase 3 CheckMate 025
trial. Cancer, 2020. 126: 4156.
https://pubmed.ncbi.nlm.nih.gov/32673417
446. McDermott, D.F., et al. Clinical activity and molecular correlates of response to atezolizumab alone or
in combination with bevacizumab versus sunitinib in renal cell carcinoma. Nat Med, 2018. 24: 749.
https://pubmed.ncbi.nlm.nih.gov/29867230
447. McDermott, D.F., et al. Pembrolizumab monotherapy as first-line therapy in advanced clear cell renal
cell carcinoma (accRCC): Results from cohort A of KEYNOTE-427. J Clin Oncol, 2018. 36.
https://ascopubs.org/doi/abs/10.1200/JCO.2018.36.15_suppl.4500
448. Albiges, L., et al. 711P Nivolumab + ipilimumab (N+I) vs sunitinib (S) for first-line treatment of
advanced renal cell carcinoma (aRCC) in CheckMate 214: 4-year follow-up and subgroup analysis
of patients (pts) without nephrectomy. Ann Oncol, 2020. 31: S559.
https://www.annalsofoncology.org/article/S0923-7534(20)40779-3/fulltext
449. Rini, B.I., et al. Pembrolizumab plus Axitinib versus Sunitinib for Advanced Renal-Cell Carcinoma.
N Engl J Med, 2019. 380: 1116.
https://pubmed.ncbi.nlm.nih.gov/30779529
450. Powles, T., et al. Pembrolizumab plus axitinib versus sunitinib monotherapy as first-line treatment of
advanced renal cell carcinoma (KEYNOTE-426): extended follow-up from a randomised, open-label,
phase 3 trial. Lancet Oncol, 2020. 21: 1563.
https://pubmed.ncbi.nlm.nih.gov/33284113
451. Motzer, R., et al. Lenvatinib plus Pembrolizumab or Everolimus for Advanced Renal Cell Carcinoma.
N Engl J Med, 2021.
https://pubmed.ncbi.nlm.nih.gov/33616314/
452. Rini, B.I., et al. Atezolizumab plus bevacizumab versus sunitinib in patients with previously untreated
metastatic renal cell carcinoma (IMmotion151): a multicentre, open-label, phase 3, randomised
controlled trial. Lancet, 2019. 393: 2404.
https://pubmed.ncbi.nlm.nih.gov/31079938
453. Choueiri, T.K., et al. Updated efficacy results from the JAVELIN Renal 101 trial: first-line avelumab
plus axitinib versus sunitinib in patients with advanced renal cell carcinoma. Ann Oncol, 2020.
31: 1030.
https://pubmed.ncbi.nlm.nih.gov/32339648
454. Tannir, N.M., et al. Thirty-month follow-up of the phase III CheckMate 214 trial of first-line nivolumab
+ ipilimumab (N+I) or sunitinib (S) in patients (pts) with advanced renal cell carcinoma (aRCC). J Clin
Oncol, 2019. 37: 547.
https://ascopubs.org/doi/10.1200/JCO.2019.37.7_suppl.547
455. Motzer R.J., et al. Nivolumab + Ipilimumab (N+I) vs Sunitinib (S) for treatment-naïve advanced or
metastatic renal cell carcinoma (aRCC): results from CheckMate 214, including overall survival by
subgroups J Immunother Cancer, 2017. Late breaking abstracts, 32nd Annual Meeting and Pre-
comference Programs of the Society for Immunotherapy of Cancer: 038.

RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021 79


456. Patel, P.H., et al. Targeting von Hippel-Lindau pathway in renal cell carcinoma. Clin Cancer Res,
2006. 12: 7215.
https://pubmed.ncbi.nlm.nih.gov/17189392
457. Yang, J.C., et al. A randomized trial of bevacizumab, an anti-vascular endothelial growth factor
antibody, for metastatic renal cancer. N Engl J Med, 2003. 349: 427.
https://pubmed.ncbi.nlm.nih.gov/12890841
458. Patard, J.J., et al. Understanding the importance of smart drugs in renal cell carcinoma. Eur Urol,
2006. 49: 633.
https://pubmed.ncbi.nlm.nih.gov/16481093
459. Escudier, B., et al. Sorafenib in advanced clear-cell renal-cell carcinoma. N Engl J Med, 2007. 356: 125.
https://pubmed.ncbi.nlm.nih.gov/17215530
460. Bellmunt, J., et al. The medical treatment of metastatic renal cell cancer in the elderly: position
paper of a SIOG Taskforce. Crit Rev Oncol Hematol, 2009. 69: 64.
https://pubmed.ncbi.nlm.nih.gov/18774306
461. Motzer, R.J., et al. Overall survival and updated results for sunitinib compared with interferon alfa in
patients with metastatic renal cell carcinoma. J Clin Oncol, 2009. 27: 3584.
https://pubmed.ncbi.nlm.nih.gov/19487381
462. Motzer, R.J., et al. Randomized phase II trial of sunitinib on an intermittent versus continuous dosing
schedule as first-line therapy for advanced renal cell carcinoma. J Clin Oncol, 2012. 30: 1371.
https://pubmed.ncbi.nlm.nih.gov/22430274
463. Bracarda, S., et al. Sunitinib administered on 2/1 schedule in patients with metastatic renal cell
carcinoma: the RAINBOW analysis. Ann Oncol, 2016. 27: 366.
https://pubmed.ncbi.nlm.nih.gov/26685011
464. Jonasch, E., et al. A randomized phase 2 study of MK-2206 versus everolimus in refractory renal cell
carcinoma. Ann Oncol, 2017. 28: 804.
https://pubmed.ncbi.nlm.nih.gov/28049139
465. Sternberg, C.N., et al. Pazopanib in locally advanced or metastatic renal cell carcinoma: results of a
randomized phase III trial. J Clin Oncol, 2010. 28: 1061.
https://pubmed.ncbi.nlm.nih.gov/20100962
466. Motzer, R.J., et al. Pazopanib versus sunitinib in metastatic renal-cell carcinoma. N Engl J Med,
2013. 369: 722.
https://pubmed.ncbi.nlm.nih.gov/23964934
467. Escudier, B., et al. Randomized, controlled, double-blind, cross-over trial assessing treatment
preference for pazopanib versus sunitinib in patients with metastatic renal cell carcinoma: PISCES
Study. J Clin Oncol, 2014. 32: 1412.
https://pubmed.ncbi.nlm.nih.gov/24687826
468. Rini, B.I., et al. Comparative effectiveness of axitinib versus sorafenib in advanced renal cell
carcinoma (AXIS): a randomised phase 3 trial. Lancet, 2011. 378: 1931.
https://pubmed.ncbi.nlm.nih.gov/22056247
469. Dror Michaelson M., et al. Phase III AXIS trial of axitinib versus sorafenib in metastatic renal cell
carcinoma: Updated results among cytokine-treated patients. J Clin Oncol 2012. J Clin Oncol 30:
15 suppl; abstr 4546.
https://ascopubs.org/doi/abs/10.1200/jco.2012.30.15_suppl.4546
470. Motzer, R.J., et al. Axitinib versus sorafenib as second-line treatment for advanced renal cell
carcinoma: overall survival analysis and updated results from a randomised phase 3 trial. Lancet
Oncol, 2013. 14: 552.
https://pubmed.ncbi.nlm.nih.gov/23598172
471. Hutson, T.E., et al. Axitinib versus sorafenib as first-line therapy in patients with metastatic renal-cell
carcinoma: a randomised open-label phase 3 trial. Lancet Oncol, 2013. 14: 1287.
https://pubmed.ncbi.nlm.nih.gov/24206640
472. Choueiri, T.K., et al. Cabozantinib versus Everolimus in Advanced Renal-Cell Carcinoma. N Engl
J Med, 2015. 373: 1814.
https://pubmed.ncbi.nlm.nih.gov/26406150
473. Choueiri, T.K., et al. Cabozantinib versus everolimus in advanced renal cell carcinoma (METEOR):
final results from a randomised, open-label, phase 3 trial. Lancet Oncol, 2016. 17: 917.
https://pubmed.ncbi.nlm.nih.gov/27279544
474. Choueiri, T.K., et al. Cabozantinib Versus Sunitinib As Initial Targeted Therapy for Patients With
Metastatic Renal Cell Carcinoma of Poor or Intermediate Risk: The Alliance A031203 CABOSUN
Trial. J Clin Oncol, 2017. 35: 591.
https://pubmed.ncbi.nlm.nih.gov/28199818

80 RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021


475. Choueiri, T.K., et al. Cabozantinib versus sunitinib as initial therapy for metastatic renal cell
carcinoma of intermediate or poor risk (Alliance A031203 CABOSUN randomised trial): Progression-
free survival by independent review and overall survival update. Eur J Cancer, 2018. 94: 115.
https://pubmed.ncbi.nlm.nih.gov/29550566
476. Motzer, R.J., et al. Lenvatinib, everolimus, and the combination in patients with metastatic renal cell
carcinoma: a randomised, phase 2, open-label, multicentre trial. Lancet Oncol, 2015. 16: 1473.
https://pubmed.ncbi.nlm.nih.gov/26482279
477. Motzer, R.J., et al. Tivozanib versus sorafenib as initial targeted therapy for patients with metastatic
renal cell carcinoma: results from a phase III trial. J Clin Oncol, 2013. 31: 3791.
https://pubmed.ncbi.nlm.nih.gov/24019545
478. Molina, A.M., et al. Efficacy of tivozanib treatment after sorafenib in patients with advanced renal
cell carcinoma: crossover of a phase 3 study. Eur J Cancer, 2018. 94: 87.
https://pubmed.ncbi.nlm.nih.gov/29547835
479. Escudier B., et al. Phase III trial of bevacizumab plus interferon alfa-2a in patients with metastatic
renal cell carcinoma (AVOREN): final analysis of overall survival. J Clin Oncol, 2010. 28: 2144.
https://pubmed.ncbi.nlm.nih.gov/16860997
480. Rini, B.I., et al. Bevacizumab plus interferon alfa compared with interferon alfa monotherapy in
patients with metastatic renal cell carcinoma: CALGB 90206. J Clin Oncol, 2008. 26: 5422.
https://pubmed.ncbi.nlm.nih.gov/18936475
481. Rini, B.I., et al. Phase III trial of bevacizumab plus interferon alfa versus interferon alfa monotherapy
in patients with metastatic renal cell carcinoma: final results of CALGB 90206. J Clin Oncol, 2010.
28: 2137.
https://pubmed.ncbi.nlm.nih.gov/20368558
482. Larkin, J.M., et al. Kinase inhibitors in the treatment of renal cell carcinoma. Crit Rev Oncol Hematol,
2006. 60: 216.
https://www.sciencedirect.com/science/article/pii/S104084280600117X
483. Motzer, R.J., et al. Efficacy of everolimus in advanced renal cell carcinoma: a double-blind,
randomised, placebo-controlled phase III trial. Lancet, 2008. 372: 449.
https://pubmed.ncbi.nlm.nih.gov/18653228
484. Auvray, M., et al. Second-line targeted therapies after nivolumab-ipilimumab failure in metastatic
renal cell carcinoma. Eur J Cancer, 2019. 108: 33.
https://pubmed.ncbi.nlm.nih.gov/30616146
485. Ornstein, M.C., et al. Prospective phase II multi-center study of individualized axitinib (Axi) titration
for metastatic renal cell carcinoma (mRCC) after treatment with PD-1 / PD-L1 inhibitors. J Clin
Oncol, 2018. 36.
https://ascopubs.org/doi/abs/10.1200/JCO.2018.36.15_suppl.4517
486. Coppin, C., et al. Targeted therapy for advanced renal cell cancer (RCC): a Cochrane systematic
review of published randomised trials. BJU Int, 2011. 108: 1556.
https://pubmed.ncbi.nlm.nih.gov/21952069
487. Rini, B.I., et al. Tivozanib versus sorafenib in patients with advanced renal cell carcinoma (TIVO-3): a
phase 3, multicentre, randomised, controlled, open-label study. Lancet Oncol, 2020. 21: 95.
https://pubmed.ncbi.nlm.nih.gov/31810797
488. Gore, M.E., et al. Safety and efficacy of sunitinib for metastatic renal-cell carcinoma: an expanded-
access trial. Lancet Oncol, 2009. 10: 757.
https://pubmed.ncbi.nlm.nih.gov/19615940
489. Sánchez P., et al. Non-clear cell advanced kidney cancer: is there a gold standard? Anticancer
Drugs 2011. 22 S9.
https://pubmed.ncbi.nlm.nih.gov/21173605
490. Koh, Y., et al. Phase II trial of everolimus for the treatment of nonclear-cell renal cell carcinoma. Ann
Oncol, 2013. 24: 1026.
https://pubmed.ncbi.nlm.nih.gov/23180114
491. Tannir, N.M., et al. A phase 2 trial of sunitinib in patients with advanced non-clear cell renal cell
carcinoma. Eur Urol, 2012. 62: 1013.
https://pubmed.ncbi.nlm.nih.gov/22771265
492. Ravaud A, et al. First-line sunitinib in type I and II papillary renal cell carcinoma (PRCC): SUPAP,
a phase II study of the French Genito-Urinary Group (GETUG) and the Group of Early Phase trials
(GEP) J. Clin Oncol, 2009. Vol 27, No 15S: 5146.
https://ascopubs.org/doi/abs/10.1200/jco.2009.27.15_suppl.5146

RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021 81


493. Escudier, B., et al. Open-label phase 2 trial of first-line everolimus monotherapy in patients with
papillary metastatic renal cell carcinoma: RAPTOR final analysis. Eur J Cancer, 2016. 69: 226.
https://pubmed.ncbi.nlm.nih.gov/27680407
494. Srinivasan, R., et al. Results from a phase II study of bevacizumab and erlotinib in subjects with
advanced hereditary leiomyomatosis and renal cell cancer (HLRCC) or sporadic papillary renal cell
cancer. J Clin Oncol, 2020. 38: 15 Suppl. 5004.
https://ascopubs.org/doi/abs/10.1200/JCO.2020.38.15_suppl.5004
495. Tannir, N.M., et al. Everolimus Versus Sunitinib Prospective Evaluation in Metastatic Non-Clear Cell
Renal Cell Carcinoma (ESPN): A Randomized Multicenter Phase 2 Trial. Eur Urol, 2016. 69: 866.
https://pubmed.ncbi.nlm.nih.gov/26626617
496. Armstrong, A.J., et al. Everolimus versus sunitinib for patients with metastatic non-clear cell renal
cell carcinoma (ASPEN): a multicentre, open-label, randomised phase 2 trial. Lancet Oncol, 2016.
17: 378.
https://ascopubs.org/doi/abs/10.1200/jco.2015.33.15_suppl.4507
497. Antonelli, A., et al. Features of Ipsilateral Renal Recurrences After Partial Nephrectomy: A Proposal
of a Pathogenetic Classification. Clin Genitourin Cancer, 2017. 15: 540.
https://pubmed.ncbi.nlm.nih.gov/28533051
498. Petros, F.G., et al. Oncologic outcomes of patients with positive surgical margin after partial
nephrectomy: a 25-year single institution experience. World J Urol, 2018. 36: 1093.
https://pubmed.ncbi.nlm.nih.gov/29488096
499. Bansal, R.K., et al. Positive surgical margins during partial nephrectomy for renal cell carcinoma:
Results from Canadian Kidney Cancer information system (CKCis) collaborative. Can Urol Assoc J,
2017. 11: 182.
https://pubmed.ncbi.nlm.nih.gov/28652876
500. Bertolo, R., et al. Low Rate of Cancer Events After Partial Nephrectomy for Renal Cell Carcinoma:
Clinicopathologic Analysis of 1994 Cases with Emphasis on Definition of “Recurrence”. Clin
Genitourin Cancer, 2019. 17: 209.
https://pubmed.ncbi.nlm.nih.gov/31000486
501. Kreshover, J.E., et al. Renal cell recurrence for T1 tumors after laparoscopic partial nephrectomy.
J Endourol, 2013. 27: 1468.
https://pubmed.ncbi.nlm.nih.gov/24074156
502. Wah, T.M., et al. Radiofrequency ablation (RFA) of renal cell carcinoma (RCC): experience in 200
tumours. BJU Int, 2014. 113: 416.
https://pubmed.ncbi.nlm.nih.gov/24053769
503. Itano, N.B., et al. Outcome of isolated renal cell carcinoma fossa recurrence after nephrectomy.
J Urol, 2000. 164: 322.
https://pubmed.ncbi.nlm.nih.gov/10893575
504. Lee, Z., et al. Local Recurrence Following Resection of Intermediate-High Risk Nonmetastatic Renal
Cell Carcinoma: An Anatomical Classification and Analysis of the ASSURE (ECOG-ACRIN E2805)
Adjuvant Trial. J Urol, 2020. 203: 684.
https://pubmed.ncbi.nlm.nih.gov/31596672
505. Margulis, V., et al. Predictors of oncological outcome after resection of locally recurrent renal cell
carcinoma. J Urol, 2009. 181: 2044.
https://pubmed.ncbi.nlm.nih.gov/19286220
506. Russell, C.M., et al. Multi-institutional Survival Analysis of Incidental Pathologic T3a Upstaging in
Clinical T1 Renal Cell Carcinoma Following Partial Nephrectomy. Urology, 2018. 117: 95.
https://pubmed.ncbi.nlm.nih.gov/29678662
507. Srivastava, A., et al. Incidence of T3a up-staging and survival after partial nephrectomy: Size-
stratified rates and implications for prognosis. Urol Oncol, 2018. 36: 12.e7.
https://pubmed.ncbi.nlm.nih.gov/28970053
508. Psutka, S.P., et al. Renal fossa recurrence after nephrectomy for renal cell carcinoma: prognostic
features and oncological outcomes. BJU Int, 2017. 119: 116.
https://pubmed.ncbi.nlm.nih.gov/27489013
509. Sandhu, S.S., et al. Surgical excision of isolated renal-bed recurrence after radical nephrectomy for
renal cell carcinoma. BJU Int, 2005. 95: 522.
https://pubmed.ncbi.nlm.nih.gov/15705072
510. Master, V.A., et al. Management of isolated renal fossa recurrence following radical nephrectomy.
J Urol, 2005. 174: 473.
https://pubmed.ncbi.nlm.nih.gov/16006867

82 RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021


511. Ierardi, A.M., et al. Percutaneous microwave ablation therapy of renal cancer local relapse after
radical nephrectomy: a feasibility and efficacy study. Med Oncol, 2020. 37: 27.
https://pubmed.ncbi.nlm.nih.gov/32166412
512. Johnson, A., et al. Feasibility and outcomes of repeat partial nephrectomy. J Urol, 2008. 180: 89.
https://pubmed.ncbi.nlm.nih.gov/18485404
513. Mouracade, P., et al. Imaging strategy and outcome following partial nephrectomy. Urol Oncol,
2017. 35: 660.e1.
https://pubmed.ncbi.nlm.nih.gov/28863862
514. Dabestani, S., et al. Increased use of cross-sectional imaging for follow-up does not improve post-
recurrence survival of surgically treated initially localized R.C.C.: results from a European multicenter
database (R.E.C.U.R.). Scand J Urol, 2019. 53: 14.
https://pubmed.ncbi.nlm.nih.gov/30907214
515. Rieken, M., et al. Predictors of Cancer-specific Survival After Disease Recurrence in Patients With
Renal Cell Carcinoma: The Effect of Time to Recurrence. Clin Genitourin Cancer, 2018. 16: e903.
https://pubmed.ncbi.nlm.nih.gov/29653814
516. Capitanio, U., et al. Hypertension and Cardiovascular Morbidity Following Surgery for Kidney
Cancer. Eur Urol Oncol, 2020. 3: 209.
https://pubmed.ncbi.nlm.nih.gov/31411993
517. Lam, J.S., et al. Renal cell carcinoma 2005: new frontiers in staging, prognostication and targeted
molecular therapy. J Urol, 2005. 173: 1853.
https://pubmed.ncbi.nlm.nih.gov/15879764
518. Scoll, B.J., et al. Age, tumor size and relative survival of patients with localized renal cell carcinoma:
a surveillance, epidemiology and end results analysis. J Urol, 2009. 181: 506.
https://pubmed.ncbi.nlm.nih.gov/19084868
519. Beisland, C., et al. A prospective risk-stratified follow-up programme for radically treated renal cell
carcinoma patients: evaluation after eight years of clinical use. World J Urol, 2016. 34: 1087.
https://pubmed.ncbi.nlm.nih.gov/26922650
520. Stewart-Merrill, S.B., et al. Oncologic Surveillance After Surgical Resection for Renal Cell
Carcinoma: A Novel Risk-Based Approach. J Clin Oncol, 2015. 33: 4151.
https://pubmed.ncbi.nlm.nih.gov/26351352
521. Rini, B.I., et al. Validation of the 16-Gene Recurrence Score in Patients with Locoregional, High-Risk
Renal Cell Carcinoma from a Phase III Trial of Adjuvant Sunitinib. Clin Cancer Res, 2018. 24: 4407.
https://pubmed.ncbi.nlm.nih.gov/29773662
522. Bruno, J.J., 2nd, et al. Renal cell carcinoma local recurrences: impact of surgical treatment and
concomitant metastasis on survival. BJU Int, 2006. 97: 933.
https://pubmed.ncbi.nlm.nih.gov/16643473
523. Bani-Hani, A.H., et al. Associations with contralateral recurrence following nephrectomy for renal cell
carcinoma using a cohort of 2,352 patients. J Urol, 2005. 173: 391.
https://pubmed.ncbi.nlm.nih.gov/15643178
524. Schaner, E.G., et al. Comparison of computed and conventional whole lung tomography in detecting
pulmonary nodules: a prospective radiologic-pathologic study. Am J Roentgenol, 1978. 131: 51.
https://pubmed.ncbi.nlm.nih.gov/97985
525. Patel, T. Lung Metastases Imaging. 2017.
https://emedicine.medscape.com/article/358090-overview
526. Chang, A.E., et al. Evaluation of computed tomography in the detection of pulmonary metastases: a
prospective study. Cancer, 1979. 43: 913.
https://pubmed.ncbi.nlm.nih.gov/284842
527. Doornweerd, B.H., et al. Chest X-ray in the follow-up of renal cell carcinoma. World J Urol, 2014.
32: 1015.
https://pubmed.ncbi.nlm.nih.gov/24096433
528. Sountoulides, P., et al. Atypical presentations and rare metastatic sites of renal cell carcinoma: a
review of case reports. J Med Case Rep, 2011. 5: 429.
https://pubmed.ncbi.nlm.nih.gov/21888643
529. Kattan, M.W., et al. A postoperative prognostic nomogram for renal cell carcinoma. J Urol, 2001.
166: 63.
https://pubmed.ncbi.nlm.nih.gov/11435824
530. Lam, J.S., et al. Postoperative surveillance protocol for patients with localized and locally advanced
renal cell carcinoma based on a validated prognostic nomogram and risk group stratification
system. J Urol, 2005. 174: 466.
https://pubmed.ncbi.nlm.nih.gov/16006866

RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021 83


531. Cindolo, L., et al. Comparison of predictive accuracy of four prognostic models for nonmetastatic
renal cell carcinoma after nephrectomy: a multicenter European study. Cancer, 2005. 104: 1362.
https://pubmed.ncbi.nlm.nih.gov/16116599
532. Skolarikos, A., et al. A review on follow-up strategies for renal cell carcinoma after nephrectomy. Eur
Urol, 2007. 51: 1490.
https://pubmed.ncbi.nlm.nih.gov/17229521
533. Chin, A.I., et al. Surveillance strategies for renal cell carcinoma patients following nephrectomy. Rev
Urol, 2006. 8: 1.
https://pubmed.ncbi.nlm.nih.gov/16985554
534. Karakiewicz, P.I., et al. A preoperative prognostic model for patients treated with nephrectomy for
renal cell carcinoma. Eur Urol, 2009. 55: 287.
https://pubmed.ncbi.nlm.nih.gov/18715700
535. Cho, H., et al. Comorbidity-adjusted life expectancy: a new tool to inform recommendations for
optimal screening strategies. Ann Intern Med, 2013. 159: 667.
https://pubmed.ncbi.nlm.nih.gov/24247672

10. CONFLICT OF INTEREST


All members of the Renal Cell Cancer Guidelines Panel have provided disclosure statements of all relationships
that they have that might be perceived as a potential source of a conflict of interest. This information is
publically accessible through the European Association of Urology website: https://uroweb.org/guideline/
renalcell-carcinoma/?type=panel/.
This guidelines document was developed with the financial support of the European Association of
Urology. No external sources of funding and support have been involved. The EAU is a non-profit organisation
and funding is limited to administrative assistance and travel and meeting expenses. No honoraria or other
reimbursements have been provided.

11. CITATION INFORMATION


The format in which to cite the EAU Guidelines will vary depending on the style guide of the journal in which the
citation appears. Accordingly, the number of authors or whether, for instance, to include the publisher, location,
or an ISBN number may vary.

The compilation of the complete Guidelines should be referenced as:


EAU Guidelines. Edn. presented at the EAU Annual Congress Milan 2021. ISBN 978-94-92671-13-4.

If a publisher and/or location is required, include:


EAU Guidelines Office, Arnhem, The Netherlands. http://uroweb.org/guidelines/compilations-of-all-guidelines/

References to individual guidelines should be structured in the following way:


Contributors’ names. Title of resource. Publication type. ISBN. Publisher and publisher location, year.

84 RENAL CELL CARCINOMA - LIMITED UPDATE MARCH 2021


EAU Guidelines on
Sexual and
Reproductive Health
A. Salonia (Chair), C. Bettocchi, J. Carvalho, G. Corona,
T.H. Jones, A. Kadioglu,
ˇ J.I. Martinez-Salamanca,
S. Minhas (Vice-chair), E.C. Serefoglu,
ˇ P. Verze

Guidelines Associates: L. Boeri, P. Capogrosso,


A. Cocci, K. Dimitropoulos, M. Gül,
G. Hatzichristodoulou, A. Kalkanli, V. Modgil,
U. Milenkovic, G. Russo, T. Tharakan

© European Association of Urology 2021


TABLE OF CONTENTS PAGE
1. INTRODUCTION 10
1.1 Aims and Objectives 10
1.2 Panel composition 10
1.3 Available Publications 10
1.4 Publication History 10
1.5 Changes in the Guideline for 2021 10

2. METHODOLOGY 10
2.1 Methods 10
2.2 Review 11
2.3 Future goals 11

3. MALE HYPOGONADISM 11
3.1 Epidemiology and prevalence of male hypogonadism 11
3.1.1 Body Composition and Metabolic Profile 11
3.1.2 Metabolic Syndrome/Type 2 Diabetes 12
3.2 Physiology of testosterone production 12
3.2.1 Circulation and transport of testosterone 13
3.2.2 Androgen receptor 14
3.3 Role of testosterone in male sexual and reproductive health 14
3.3.1 Sexual development and maturation 14
3.3.2 Sexual function 14
3.4 Classification and causes of male hypogonadism 15
3.5 Late-onset hypogonadism 17
3.5.1 Diagnostic evaluation 17
3.5.2 History taking 20
3.5.3 Physical examination 20
3.5.4 Summary of evidence and recommendations for the diagnostic evaluation of
LOH 20
3.5.5 Recommendations for screening men with LOH 21
3.6 Treatment of LOH 21
3.6.1 Indications and contraindications for treatment of LOH 21
3.6.2 Testosterone therapy outcomes 22
3.6.2.1 Sexual dysfunction 22
3.6.2.2 Body composition and metabolic profile 22
3.6.2.3 Mood and cognition 22
3.6.2.4 Bone 23
3.6.2.5 Vitality and physical strength 23
3.6.2.6 Summary of evidence and recommendations for testosterone
therapy outcome 23
3.6.3 Choice of treatment 24
3.6.3.1 Lifestyle factors 24
3.6.3.2 Medical preparations 24
3.6.3.2.1 Oral formulations 24
3.6.3.2.2 Parenteral formulations 25
3.6.3.2.3 Transdermal testosterone preparations 25
3.6.3.2.4 Transmucosal formulations 25
3.6.3.2.4.1 Transbuccal Testosterone preparations 25
3.6.3.2.4.2 Transnasal testosterone preparations 25
3.6.3.2.5 Subdermal depots 25
3.6.3.2.6 Anti-oestrogens 25
3.6.3.2.7 Gonadotropins 26
3.6.3.3 Summary of evidence and recommendations for choice of treatment
for LOH 28
3.7 Safety and follow-up in hypogonadism management 28
3.7.1 Hypogonadism and fertility issues 28
3.7.2 Male breast cancer 28
3.7.3 Lower urinary tract symptoms/benign prostatic hyperplasia 28

2 SEXUAL AND REPRODUCTIVE HEALTH - MARCH 2021


3.7.4 Prostate cancer (PCa) 29
3.7.5 Cardiovascular Disease 29
3.7.5.1 Cardiac Failure 30
3.7.6 Erythrocytosis 30
3.7.7 Obstructive Sleep Apnoea 31
3.7.8 Follow up 31
3.7.9 Summary of evidence and recommendations on risk factors in testosterone
treatment 32

4. EPIDEMIOLOGY AND PREVALENCE OF SEXUAL DYSFUNCTION AND DISORDERS


OF MALE REPRODUCTIVE HEALTH 33
4.1 Erectile dysfunction 33
4.2 Premature ejaculation 33
4.3 Other ejaculatory disorders 34
4.3.1 Delayed ejaculation 34
4.3.2 Anejaculation and Anorgasmia 34
4.3.3 Retrograde ejaculation 34
4.3.4 Painful ejaculation 34
4.3.5 Haemospermia 34
4.4 Low sexual desire 35

5. MANAGEMENT OF ERECTILE DYSFUNCTION 43


5.1 Definition and classification 43
5.2 Risk factors 43
5.3 Pathophysiology 44
5.3.1 Pelvic surgery and prostate cancer treatment 46
5.3.2 Summary of evidence on the epidemiology/aetiology/pathophysiology of ED 47
5.4 Diagnostic evaluation (basic work-up) 47
5.4.1 Medical and sexual history 47
5.4.2 Physical examination 48
5.4.3 Laboratory testing 48
5.4.4 Cardiovascular system and sexual activity: the patient at risk 49
5.4.4.1 Low-risk category 51
5.4.4.2 Intermediate- or indeterminate-risk category 51
5.4.4.3 High-risk category 51
5.5 Diagnostic Evaluation (advanced work-up) 51
5.5.1 Nocturnal penile tumescence and rigidity test 51
5.5.2 Intracavernous injection test 51
5.5.3 Dynamic duplex ultrasound of the penis 51
5.5.4 Arteriography and dynamic infusion cavernosometry or cavernosography 51
5.5.5 Psychiatric and psychosocial assessment 52
5.5.6 Recommendations for diagnostic evaluation of ED 53
5.6 Treatment of erectile dysfunction 53
5.6.1 Patient education - consultation and referrals 53
5.6.2 Treatment options 53
5.6.2.1 Oral pharmacotherapy 55
5.6.2.2 Topical/Intraurethral alprostadil 60
5.6.2.3 Shockwave therapy 61
5.6.2.4 Psychosexual counselling and therapy 61
5.6.2.5 Hormonal treatment 61
5.6.2.6 Vacuum erection devises 61
5.6.2.7 Intracavernous injections therapy 61
5.6.2.7.1 Alprostadil 62
5.6.2.8 Combination therapy 62
5.6.2.8.1 Erectile dysfunction after radical prostatectomy 63
5.6.2.9 Vascular surgery 64
5.6.2.9.1 Surgery for post-traumatic arteriogenic ED 64
5.6.2.9.2 Venous ligation surgery 64
5.6.2.9.3 Penile prostheses 65
5.6.2.9.4 Penile prostheses implantation: complications 65

SEXUAL AND REPRODUCTIVE HEALTH - MARCH 2021 3


5.6.2.9.4.1 Conclusions about penile prostheses
implantation 66
5.6.3 Recommendations for treatment of ED 66
5.6.4 Follow-up 67

6. DISORDERS OF EJACULATION 67
6.1 Introduction 67
6.2 Premature ejaculation 67
6.2.1 Epidemiology 67
6.2.2 Pathophysiology and risk factors 67
6.2.3 Impact of PE on quality of life 68
6.2.4 Classification 68
6.2.5 Diagnostic evaluation 69
6.2.5.1 Intravaginal ejaculatory latency time (IELT) 69
6.2.5.2 Premature ejaculation assessment questionnaires 69
6.2.5.3 Physical examination and investigations 70
6.2.5.4 Recommendations for the diagnostic evaluation of PE 70
6.2.6 Disease management 70
6.2.6.1 Psychological aspects and intervention 71
6.2.6.1.1 Recommendation for the assessment and treatment
(psychosexual approach) of PE 72
6.2.6.2 Pharmacotherapy 72
6.2.6.2.1 Dapoxetine 72
6.2.6.2.2 Off-label use of antidepressants: selective serotonin
reuptake inhibitors and clomipramine 73
6.2.6.2.3 Topical anaesthetic agents 74
6.2.6.2.3.1 Lidocaine/prilocaine cream 74
6.2.6.2.3.2 Lidocaine/prilocaine spray 74
6.2.6.2.4 Tramadol 74
6.2.6.2.5 Phosphodiesterase type 5 inhibitors 75
6.2.6.2.6 Other drugs 75
6.2.7 Summary of evidence on the epidemiology/aetiology/pathophysiology of PE 76
6.2.8 Recommendations for the treatment of PE 76
6.3 Delayed Ejaculation 76
6.3.1 Definition and classification 76
6.3.2 Pathophysiology and risk factors 76
6.3.3 Investigation and treatment 77
6.3.3.1 Psychological aspects and intervention 77
6.3.3.2 Pharmacotherapy 78
6.4 Anejaculation 78
6.4.1 Definition and classification 78
6.4.2 Pathophysiology and risk factors 78
6.4.3 Investigation and treatment 78
6.5 Painful Ejaculation 78
6.5.1 Definition and classification 78
6.5.2 Pathophysiology and risk factors 78
6.5.3 Investigation and treatment 79
6.5.3.1 Surgical intervention 79
6.6 Retrograde ejaculation 79
6.6.1 Definition and classification 79
6.6.2 Pathophysiology and risk factors 79
6.6.3 Disease management 80
6.6.3.1 Pharmacological 80
6.6.3.2 Management of infertility 80
6.7 Anorgasmia 80
6.7.1 Definition and classification 80
6.7.2 Pathophysiology and risk factors 81
6.7.3 Disease management 81
6.7.3.1 Psychological/behavioural strategies 81
6.7.3.2 Pharmacotherapy 81

4 SEXUAL AND REPRODUCTIVE HEALTH - MARCH 2021


6.7.3.3 Management of infertility 81
6.8 Haemospermia 81
6.8.1 Definition and classification 81
6.8.2 Pathophysiology and risk factors 81
6.8.3 Investigations 82
6.8.4 Disease management 83
6.9 Recommendations for the management of recurrent haemospermia 84

7. LOW SEXUAL DESIRE AND MALE HYPOACTIVE SEXUAL DESIRE DISORDER 85


7.1 Definition, classification and epidemiology 85
7.2 Pathophysiology and risk factors 85
7.2.1 Psychological aspects 85
7.2.2 Biological aspects 85
7.2.3 Risk factors 86
7.3 Diagnostic work-up 86
7.3.1 Assessment questionnaires 86
7.3.2 Physical examination and investigations 86
7.4 Disease management 86
7.4.1 Psychological intervention 86
7.4.2 Pharmacotherapy 87
7.5 Recommendations for the treatment of low sexual desire 87

8. PENILE CURVATURE 88
8.1 Congenital penile curvature 88
8.1.1 Epidemiology/aetiology/pathophysiology 88
8.1.2 Diagnostic evaluation 88
8.1.3 Disease management 88
8.1.4 Summary of evidence for congenital penile curvature 88
8.1.5 Recommendation for the treatment congenital penile curvature 88
8.2 Peyronie’s Disease 88
8.2.1 Epidemiology/aetiology/pathophysiology 88
8.2.1.1 Epidemiology 88
8.2.1.2 Aetiology 88
8.2.1.3 Risk factors 90
8.2.1.4 Pathophysiology 90
8.2.1.5 Summary of evidence on epidemiology/aetiology/pathophysiology
of Peyronie’s disease 91
8.2.2 Diagnostic evaluation 91
8.2.2.1 Summary of evidence for diagnosis of Peyronie’s disease 92
8.2.2.2 Recommendations for diagnosis of Peyronie’s disease 92
8.2.3 Disease management 92
8.2.3.1 Conservative treatment 92
8.2.3.1.1 Oral treatment 93
8.2.3.1.2 Intralesional treatment 93
8.2.3.1.3 Topical treatments 95
8.2.3.1.4 Multimodal treatment 98
8.2.3.1.5 Summary of evidence for conservative treatment of
Peyronie’s disease 98
8.2.3.1.6 Recommendations for non-operative treatment of
Peyronie’s disease 99
8.2.3.2 Surgical treatment 99
8.2.3.2.1 Tunical shortening procedures 100
8.2.3.2.2 Tunical lengthening procedures 101
8.2.3.2.3 Penile prosthesis 103
8.2.3.2.4 Summary of evidence for non-operative treatment of
Peyronie’s disease 103
8.2.3.2.5 Recommendations for surgical treatment of penile
curvature 104
8.2.3.3 Treatment algorithm 104

SEXUAL AND REPRODUCTIVE HEALTH - MARCH 2021 5


9. PRIAPISM 106
9.1 Ischaemic (Low-Flow or Veno-Occlusive) Priapism 106
9.1.1 Epidemiology, aetiology, pathophysiology and Diagnosis 106
9.1.1.1 Summary of evidence on the epidemiology, aetiology and
pathophysiology of ischaemic priapism 108
9.1.2 Diagnostic evaluation 109
9.1.2.1 History 109
9.1.2.2 Physical examination 109
9.1.2.3 Laboratory testing 109
9.1.2.4 Penile imaging 110
9.1.2.5 Recommendations for the diagnosis of ischaemic priapism 110
9.1.3 Disease management 111
9.1.3.1 Medical Management 111
9.1.3.1.1 First-line treatments 112
9.1.3.1.2 Penile anaesthesia/analgesia 112
9.1.3.1.3 Aspiration ± irrigation with 0.9% w/v saline solution 112
9.1.3.1.4 Aspiration ± irrigation with 0.9% w/v saline solution in
combination with intracavernous injection of
pharmacological agents. 113
9.1.3.2 Surgical management 115
9.1.3.2.1 Second-line treatments 115
9.1.3.2.1.1 Penile shunt surgery 115
9.1.4 Summary of evidence for treatment of ischaemic priapism 119
9.1.5 Recommendations for the treatment of ischaemic priapism 119
9.2 Priapism in Special Situations 120
9.2.1 Stuttering (recurrent or intermittent) priapism 120
9.2.1.1 Epidemiology/aetiology/pathophysiology 120
9.2.1.1.1 Summary of evidence on the epidemiology, aetiology
and pathophysiology of stuttering priapism 120
9.2.1.2 Classification 121
9.2.1.3 Diagnostic evaluation 121
9.2.1.3.1 History 121
9.2.1.3.2 Physical examination 121
9.2.1.3.3 Laboratory testing 121
9.2.1.3.4 Penile imaging 121
9.2.1.4 Disease management 121
9.2.1.4.1 α-Adrenergic agonists 121
9.2.1.4.2 Hormonal manipulations of circulating testosterone 121
9.2.1.4.3 Digoxin 122
9.2.1.4.4 Terbutaline 122
9.2.1.4.5 Gabapentin 122
9.2.1.4.6 Baclofen 122
9.2.1.4.7 Hydroxyurea 122
9.2.1.4.8 Phosphodiesterase type 5 inhibitors 122
9.2.1.4.9 Intracavernosal injections 123
9.2.1.4.10 Penile prosthesis 123
9.2.1.5 Summary of evidence for treatment of stuttering priapism 123
9.2.1.6 Recommendations for treatment of stuttering priapism 123
9.2.1.7 Follow-up 123
9.2.2 Priapism in children 123
9.3 Non-ischaemic (high-flow or arterial) priapism 124
9.3.1 Epidemiology/aetiology/pathophysiology 124
9.3.1.1 Summary of evidence on the epidemiology, aetiology and
pathophysiology of arterial priapism 124
9.3.2 Classification 124
9.3.3 Diagnostic evaluation 125
9.3.3.1 History 125
9.3.3.2 Physical examination 125
9.3.3.3 Laboratory testing 125
9.3.3.4 Penile imaging 125

6 SEXUAL AND REPRODUCTIVE HEALTH - MARCH 2021


9.3.3.5 Recommendations for the diagnosis of non-ischaemic priapism 125
9.3.4 Disease management 125
9.3.4.1 Conservative management 126
9.3.4.2 Selective arterial embolisation 126
9.3.4.3 Surgical management 126
9.3.4.4 Summary of evidence for the treatment of arterial priapism 126
9.3.4.5 Recommendations for the treatment of arterial priapism 127
9.3.4.6 High-flow priapism in children 127
9.3.4.7 Follow-up 127
9.4 Controversies and future areas of focus in the management of priapism 127

10. MALE INFERTILITY 128


10.1 Definition and classification 128
10.2 Epidemiology/aetiology/pathophysiology/risk factors 128
10.2.1 Introduction 128
10.2.2 Recommendations on epidemiology and aetiology 129
10.3 Diagnostic work-up 129
10.3.1 Medical/reproductive history and physical examination 129
10.3.1.1 Medical and reproductive history 129
10.3.1.2 Physical examination 130
10.3.2 Semen analysis 130
10.3.3 Measurement of sperm DNA Fragmentation Index (DFI) 131
10.3.4 Hormonal determinations 131
10.3.5 Genetic testing 132
10.3.5.1 Chromosomal abnormalities 132
10.3.5.1.1 Sex chromosome abnormalities (Klinefelter syndrome
and variants [47,XXY; 46,XY/47, XX mosaicism]) 132
10.3.5.1.2 Autosomal abnormalities 133
10.3.5.2 Cystic fibrosis gene mutations 133
10.3.5.2.1 Unilateral or bilateral absence/abnormality of the vas
and renal anomalies 134
10.3.5.3 Y microdeletions - partial and complete 134
10.3.5.3.1 Clinical implications of Y microdeletions 134
10.3.5.3.1.1 Testing for Y microdeletions 134
10.3.5.3.1.2 Genetic counselling for AZF deletions 135
10.3.5.3.1.3 Y-chromosome: ‘gr/gr’ deletion 135
10.3.5.3.1.4 Autosomal defects with severe
phenotypic abnormalities and infertility 135
10.3.5.4 Sperm chromosomal abnormalities 135
10.3.5.5 Measurement of Oxidative Stress 135
10.3.5.6 Outcomes from assisted reproductive technology and long-term
health implications to the male and offspring 136
10.3.6 Imaging in the infertile men 136
10.3.6.1 Scrotal US 136
10.3.6.1.1 Testicular neoplasms 136
10.3.6.1.2 Varicocele 137
10.3.6.1.3 Other 137
10.3.6.2 Transrectal US 137
10.3.7 Recommendations for the diagnostic work-up of male infertility 138
10.4 Special Conditions and Relevant Clinical Entities 139
10.4.1 Cryptorchidism 139
10.4.1.1 Classification 139
10.4.1.1.1 Aetiology and pathophysiology 139
10.4.1.1.2 Pathophysiological effects in maldescended testes 139
10.4.1.1.2.1 Degeneration of germ cells 139
10.4.1.1.2.2 Relationship with fertility 139
10.4.1.1.2.3 Germ cell tumours 139
10.4.1.2 Disease management 140
10.4.1.2.1 Hormonal treatment 140
10.4.1.2.2 Surgical treatment 140

SEXUAL AND REPRODUCTIVE HEALTH - MARCH 2021 7


10.4.1.3 Summary of evidence recommendations for cryptorchidism 140
10.4.2 Germ cell malignancy and male infertility 140
10.4.2.1 Testicular germ cell cancer and reproductive function 141
10.4.2.2 Testicular microcalcification (TM) 142
10.4.2.3 Recommendations for germ cell malignancy and testicular
microcalcification 143
10.4.3 Varicocele 143
10.4.3.1 Classification 143
10.4.3.2 Diagnostic evaluation 143
10.4.3.3 Basic considerations 143
10.4.3.3.1 Varicocele and fertility 143
10.4.3.3.2 Varicocelectomy 144
10.4.3.3.3 Prophylactic varicocelectomy 144
10.4.3.3.4 Varicocelectomy for assisted reproductive technology
and raised DNA fragmentation 145
10.4.3.4 Disease management 145
10.4.3.5 Summary of evidence and recommendations for varicocele 147
10.4.4 Male accessory gland infections and infertility 147
10.4.4.1 Introduction 147
10.4.4.2 Diagnostic evaluation 147
10.4.4.2.1 Semen analysis 147
10.4.4.2.2 Microbiological findings 147
10.4.4.2.3 White blood cells 148
10.4.4.2.4 Sperm quality 148
10.4.4.2.5 Seminal plasma alterations 148
10.4.4.2.6 Glandular secretory dysfunction 148
10.4.4.2.7 Reactive oxygen species 148
10.4.4.2.8 Disease management 149
10.4.4.3 Epididymitis 149
10.4.4.3.1 Diagnostic evaluation 149
10.4.4.3.1.1 Ejaculate analysis 149
10.4.4.3.1.2 Disease management 149
10.4.4.4 Summary of evidence and recommendation for male accessory
gland infections 149
10.5 Non-Invasive Male Infertility Management 150
10.5.1 Idiopathic male infertility and oligo-astheno-terato-zoospermia 150
10.5.2 Empirical treatments 150
10.5.2.1 Life-style 150
10.5.2.1.1 Weight loss 150
10.5.2.1.2 Physical activity 150
10.5.2.1.3 Smoking 150
10.5.2.1.4 Alcohol consumption 150
10.5.2.2 Antioxidant treatment 150
10.5.2.3 Selective oestrogen receptor modulators 151
10.5.2.4 Aromatase inhibitors 151
10.5.3 Hormonal therapy 151
10.5.3.1 Gonadotrophins 151
10.5.3.2 Secondary hypogonadism 152
10.5.3.2.1 Pre-Pubertal-Onset 152
10.5.3.2.2 Post-Pubertal Onset Secondary 152
10.5.3.3 Primary Hypogonadism 153
10.5.3.4 Idiopathic Male Factor Infertility 153
10.5.3.5 Anabolic Steroid Abuse 153
10.5.3.6 Recommendations for treatment of male infertility with hormonal
therapy 153
10.6 Invasive Male Infertility Management 153
10.6.1 Obstructive azoospermia 153
10.6.1.1 Classification of obstructive azoospermia 154
10.6.1.1.1 Intratesticular obstruction 154
10.6.1.1.2 Epididymal obstruction 154

8 SEXUAL AND REPRODUCTIVE HEALTH - MARCH 2021


10.6.1.1.3 Vas deferens obstruction 154
10.6.1.1.4 Ejaculatory duct obstruction 154
10.6.1.1.4.1 Functional obstruction of the distal
seminal ducts 154
10.6.1.2 Diagnostic evaluation 155
10.6.1.2.1 Clinical history 155
10.6.1.2.2 Clinical examination 155
10.6.1.2.3 Semen analysis 155
10.6.1.2.4 Hormone levels 155
10.6.1.2.5 Genetic Testing 155
10.6.1.2.6 Testicular biopsy 155
10.6.1.3 Disease management 155
10.6.1.3.1 Intratesticular obstruction 155
10.6.1.3.2 Epididymal obstruction 155
10.6.1.3.3 Vas deferens obstruction after vasectomy 156
10.6.1.3.4 Vas deferens obstruction at the inguinal level 156
10.6.1.3.5 Ejaculatory duct obstruction 156
10.6.1.4 Summary of evidence and recommendations for obstructive
azoospermia 156
10.6.2 Non-obstructive azoospermia 157
10.6.2.1 Investigation of non-obstructive azoospermia 157
10.6.2.2 Surgery for non-obstructive azoospermia 157
10.6.2.3 3 Indications and techniques of sperm retrieval 157
10.6.2.4 Recommendations for Non-Obstructive Azoospermia 160
10.7 Assisted Reproductive Technologies 161
10.7.1 Types of assisted reproductive technology 161
10.7.1.1 Intra-uterine insemination (IUI) 161
10.7.1.2 In vitro fertilisation (IVF) 161
10.7.1.3 Intracytoplasmic sperm injection 162
10.7.1.4 Testicular sperm in men with raised DNA fragmentation in
ejaculated sperm 163
10.7.1.5 Intra-cytoplasmic morphologically selected sperm injection 163
10.7.1.6 Physiological ICSI (PICSI) technique: a selection based on
membrane maturity of sperm 164
10.7.1.7 Magnetic-activated cell sorting (MACS) 164
10.7.2 Safety 164

11. LATE EFFECTS, SURVIVORSHIP AND MEN’S HEALTH 165

12. REFERENCES 166

13. CONFLICT OF INTEREST 269

14. CITATION INFORMATION 270

APPENDICES 271

SEXUAL AND REPRODUCTIVE HEALTH - MARCH 2021 9


1. INTRODUCTION
1.1 Aims and Objectives
The European Association of Urology (EAU) Sexual and Reproductive Health Guidelines aim to provide a
comprehensive overview of the medical aspects relating to sexual and reproductive health in adult men. These
Guidelines cover the former EAU Guidelines on Male Sexual Dysfunction, Male Infertility and Male Hypogonadism.
It must be emphasised that guidelines present the best evidence available to the experts. However
following guideline recommendations will not necessarily result in the best outcome. Guidelines can never
replace clinical expertise when making treatment decisions for individual patients, but rather help to focus
decisions - while taking personal values and preferences/individual circumstances of patients into account.
Guidelines are not mandates and do not purport to be a legal standard of care.

1.2 Panel composition


The EAU Sexual and Reproductive Health Guidelines Panel consists of an international multi-disciplinary group
of urologists, endocrinologists and a psychologist. All experts involved in the production of this document have
submitted potential conflict of interest statements which can be viewed on the EAU website:
http://www.uroweb.org/guideline/sexualandreproductivehealth/.

1.3 Available Publications


Alongside the full text version, a quick reference document (Pocket Guidelines) is available in print and as an
app for iOS and android devices. These are abridged versions that may require consultation together with the
full text version. All documents can be viewed through the EAU website: http://www.uroweb.org/guideline/
sexualandreproductivehealth/.

1.4 Publication History


This document is a further update of the 2020 Guidelines which already included a comprehensive update
of the 2018 versions of Male Sexual Dysfunction, Male Infertility and Male Hypogonadism guidelines, along
with several new topics. Additional sections will be added in the coming years to address male contraception,
vasectomy, and penile cosmetic surgery, which have not been previously addressed.

1.5 Changes in the Guideline for 2021


The results of ongoing and new systematic reviews have been included in the 2021 update of the Sexual and
Reproductive Health Guidelines. Systematic reviews undertaken in 2020 were:
• What is the effectiveness of non-surgical therapies in the treatment of ischaemic priapism in patients with
sickle cell disease?
• What is the effectiveness of non-surgical therapies in the management of priapism in patients without
sickle cell disease?
• What is the effectiveness of surgical therapies in the treatment of priapism?

2. METHODOLOGY
2.1 Methods
For the 2021 Sexual and Reproductive Health Guidelines, further new evidence has been identified, collated
and appraised through a structured assessment of the literature.

For each recommendation within the Guidelines there is an accompanying online strength rating form; the
basis of which is a modified Grading of Recommendations Assessment, Development and Evaluation (GRADE)
methodology [1, 2]. Each strength rating form addresses several key elements namely:
1. the overall quality of the evidence which exists for the recommendation, references used in
this text are graded according to a classification system modified from the Oxford Centre for
Evidence-Based Medicine Levels of Evidence [3];
2. the magnitude of the effect (individual or combined effects);
3. the certainty of the results (precision, consistency, heterogeneity and other statistical or
study-related factors);
4. the balance between desirable and undesirable outcomes;
5. the impact of patient values and preferences on the intervention;
6. the certainty of those patient values and preferences.

10 SEXUAL AND REPRODUCTIVE HEALTH - MARCH 2021


These key elements are the basis that panels use to define the strength rating of each recommendation.
The strength of each recommendation is represented by the term ‘strong’ or ‘weak’ [4]. The strength of each
recommendation is determined by the balance between desirable and undesirable consequences of alternative
management strategies, the quality of the evidence (including certainty of estimates), and nature and variability
of patient values and preferences. Additional information can be found in the general Methodology section of
this print, and online at the European Association of Urology (EAU) website: http://www.uroweb.org/guideline/.
A list of associations endorsing the EAU Guidelines can also be viewed online at this address.

2.2 Review
The existing sections of the Sexual and Reproductive Health Guidelines were peer reviewed prior to publication
in 2020. The new section for priapism was reviewed prior to publication in 2021.

2.3 Future goals


The results of ongoing and new systematic reviews will be included in the 2022 update of the Sexual and
Reproductive Health Guidelines. Systematic reviews planned for 2021 are:
• Penile cosmetic surgery;
• Vasectomy.

3. MALE HYPOGONADISM
3.1 Epidemiology and prevalence of male hypogonadism
Male hypogonadism is associated with decreased testicular function, with decreased production of androgens
and/or impaired sperm production [5]. This is caused by poor testicular function or as a result of inadequate
stimulation of the testes by the hypothalamic-pituitary axis. Several congenital or acquired disorders causing
impaired action of androgens are also described [5]. Hypogonadism may adversely affect multiple organ
functions and quality of life (QoL) [6]. Late-onset hypogonadism (LOH) is a clinical condition in ageing
men, which, by definition, must comprise both persistent specific symptoms and biochemical evidence
of testosterone deficiency [5, 7]. Late-onset hypogonadism is frequently diagnosed in the absence of an
identifiable classical cause of hypogonadism, which becomes more prevalent with age, usually occurring, but
not exclusively, in men aged > 40 years.
Male hypogonadism has also been called Testosterone Deficiency. The Panel has agreed to use the
term Male Hypogonadism, which may better reflect and explain the underlying pathophysiology. Likewise, the
Panel has further agreed to continue with the term testosterone therapy. The present Guidelines specifically
address the management of adult male hypogonadism also called LOH. Some insights related to congenital or
pre-pubertal hypogonadism are also provided and summarised.
The prevalence of hypogonadism increases with age and the major causes are central obesity,
co-morbidity (e.g., diabetes) and overall poor health [8]. In healthy ageing men, there is only a small gradual
decline in testosterone; up to the age of 80 years, age accounts for a low percentage of hypogonadism [8].
In men aged 40-79 years, the incidence of symptomatic hypogonadism varies between 2.1-5.7% [9-11]. The
incidence of hypogonadism has been reported to be 12.3 and 11.7 cases per 1,000 people per year [9, 12].
There is a high prevalence of hypogonadism within specific populations, including patients with
type 2 diabetes (T2DM), metabolic syndrome (MetS), obesity, cardiovascular disease (CVD), chronic obstructive
pulmonary disease, renal disease and cancer [11]. Low testosterone levels are common in men with T2DM [13]
and a high prevalence of hypogonadism (42%) has been reported in T2DM patients [14].
Klinefelter syndrome, a trisomy associated with a 47,XXY karyotype, is the most prevalent genetic
cause of primary hypogonadism (hypergonadotropic hypogonadism), with a global prevalence of 1/500-1,000 live
male births [15-17]. However, < 50% of individuals with Klinefelter syndrome are diagnosed in their lifetime [18].

3.1.1 Body Composition and Metabolic Profile


Low testosterone levels are common in men with obesity. Male hypogonadism is associated with a greater
percentage of fat mass and a lesser lean mass compared to men with adequate testosterone levels [19]. There
is much evidence that low testosterone level is strongly associated with increased visceral adiposity, but it also
leads to lipid deposition in the liver and muscle and is associated with atherosclerosis [19]. In vitro studies have
suggested that hypogonadism impairs glucose and triglyceride uptake into subcutaneous fat depots [19]. This
enhances the uptake of glucose and triglycerides into ectopic fat depots.

SEXUAL AND REPRODUCTIVE HEALTH - MARCH 2021 11


Testosterone therapy has been associated with a reduced percentage of body fat and increase of lean body
mass [20]. Data from a registry study have suggested that over a period of 8 years, testosterone therapy
with long-acting intramuscular testosterone undecanoate was associated with a substantial but gradual loss
of weight along with a reduction in waist circumference [21]. Testosterone also reduces liver fat content and
muscle fat storage [19].

3.1.2 Metabolic Syndrome/Type 2 Diabetes


Metabolic Syndrome (MetS) is characterised by several specific components, including increased waist
circumference, dyslipidaemia, hypertension, and impaired glucose tolerance. Hypogonadism is associated
with central obesity, hyperglycaemia, insulin resistance and dyslipidaemia [low high-density lipoprotein (HDL)
cholesterol, raised total and low-density lipoprotein (LDL) cholesterol and triglycerides], hypertension and
predisposition to T2DM, which are all components of MetS [22].
Several randomised controlled trials (RCTs) have shown that testosterone therapy might improve
insulin resistance and hyperglycaemia and lower cholesterol and LDL-cholesterol [23-27]. Testosterone
therapy in hypogonadal T2DM improved glycaemic control in some RCTs and registry trials; however, there
is no conclusive evidence from RCTs and meta-analyses [24, 28, 29]. A recent large placebo-controlled RCT,
including 1,007 patients with impaired glucose tolerance or newly-diagnosed T2DM and total testosterone
< 14 nmol/L, showed that testosterone therapy for 2 years reduced the proportion of patients with T2DM
regardless of a lifestyle programme [30]. Similarly, a previously published registry study reported that
testosterone therapy was associated in time with remission of T2DM [28]. HDL-cholesterol may decrease,
remain unchanged or increase with testosterone therapy. Testosterone therapy in men with MetS and low
testosterone has been shown to reduce mortality compared to that in untreated men [31, 32], although no
conclusive evidence is available.
Erectile dysfunction (ED) is common in men with MetS and T2DM (up to 70% of patients).
The causes of ED are multi-factorial and 30% of men with ED have co-existing testosterone-deficiency
hypogonadism. Some evidence has suggested that for patients with T2DM this is only found in men with
clearly reduced testosterone levels (< 8 nmol/L or 2.31 ng/mL) [33]. From a pathophysiological point of view,
it has been reported that this is because ED is predominantly caused by vascular and neuropathic disease,
and therefore not likely in men who do not have established vascular disease. Therefore, men presenting
with ED should be screened for MetS. Likewise, patients with ED and diabetes may be offered testosterone
measurement.
Placebo-controlled RCTs of testosterone therapy in T2DM have demonstrated improved sexual
desire and satisfaction, but not erectile function [24, 33]. The presence of multiple comorbidity in this group
of patients may confound the response to testosterone therapy alone. In a long-term registry study in men
with T2DM, parenteral testosterone undecanoate therapy led to one third of patients entering remission from
diabetes during 11 years’ follow-up [34]. A large 2-year RCT of testosterone undecanoate vs. placebo showed
that testosterone therapy significantly decreased progression of 999 men with low testosterone (< 14 nmol/L)
from pre-diabetes to overt T2DM [30].

3.2 Physiology of testosterone production


The pituitary gland regulates testicular activity through secretion of luteinising hormone (LH), which regulates
testosterone production in Leydig cells and follicle-stimulating hormone (FSH), which mainly controls sperm
production in seminiferous tubules [35, 36]. The production and secretion of gonadotropins is stimulated
by hypothalamic gonadotropin releasing hormone (GnRH) and inhibited by negative feedback mediated
by the central action of sex steroids and inhibin B (Figure 1) [35, 36]. Gonadotropin releasing hormone is
secreted in a pulsatile manner and negatively controlled by the activity of hypothalamic neurons, including
corticotrophin-releasing hormone (CRH) and β endorphin neurons [35, 36]. Conversely, kisspeptin-1 (Kiss-1)
neurons, neurokinin-B and tachykinin-3 are involved in GnRH stimulation. Leptin is involved in activation of
Kiss-1 signalling [37]. About 25 mg of testosterone is present in the normal testes, and, on average, 5-10 mg of
testosterone are secreted daily [35, 36]. The testes also produce lesser amounts of other androgens, such as
androstenedione and dihydrotestosterone (DHT). A small amount of extra-gonadal testosterone is derived from
circulating weak adrenal androgen precursor dehydroepiandrosterone (DHEA), although its specific contribution
to daily testosterone production is limited in men [38, 39]. In physiological terms, DHT formation accounts for
6-8% of testosterone metabolism, and the ratio of plasma testosterone/DHT is approximately 1:20 [35, 36].
Finally, testosterone and its precursor, Δ4 androstenedione, can be aromatised through P450 aromatase to
other bioactive metabolites, such as oestrone (E1) and 17-β-oestradiol (E2), with a daily production of ~45 μg
[35, 36]. Leydig cells, can also directly produce and release into the bloodstream small amounts of oestrogens,
with a daily production rate of 5-10 μg (up to 20% of circulating oestrogens) [40].

12 SEXUAL AND REPRODUCTIVE HEALTH - MARCH 2021


Figure 1: Physiology of testosterone production

HYPOTHALAMUS

Kisspeptin GnRH CRH


Neurokinin B β endorphin
Tachykinin-3
+ - -
+

PITUITARY

LH FSH

- -

T, E2 Inhibin
LH DHT B FSH

TESTES
+ +
Leydig Seroli
cells cells

GnRH = gonadotropin releasing hormone; LH = luteinising hormone; FSH = follicle-stimulating hormone;


T = testosterone; E2 = 7-β-estradiol; DHT = dehydroepiandrosterone; CRH = corticotrophin releasing hormone.

3.2.1 Circulation and transport of testosterone


In healthy men, 60-70% of circulating testosterone is bound to the high-affinity sex-hormone-binding globulin
(SHBG), a protein produced by the liver, which prevents its bound testosterone sub-fraction from biological
action. The remaining circulating testosterone binds to lower affinity, high-capacity binding proteins, (albumin,
α-1 acid glycoprotein and corticosteroid-binding protein), and only 1-2% of testosterone remains non-
protein bound [41]. There is a general agreement that testosterone bound to lower-affinity proteins can easily
dissociate in the capillary bed of many organs, accounting for so-called ‘bioavailable’ testosterone [41]. It is
important to recognise that several clinical conditions and ageing itself can modify SHBG levels, thus altering
circulating total testosterone levels (Table 1). Therefore, if not recognised, these factors could lead to an
incorrect estimation of male androgen status. Therefore, when indicated (Table 1), SHBG should be tested and
free testosterone calculated.

SEXUAL AND REPRODUCTIVE HEALTH - MARCH 2021 13


Table 1: Main factors associated with an increase or reduction of SHBG circulating levels

SHBG increase • Drugs: anticonvulsants, oestrogens, thyroid hormone


• Hyperthyroidism
• Hepatic disease
• Ageing
• Smoking
• AIDS/HIV
SHBG decrease • Drugs: growth hormone (GH), glucocorticoids, testosterone, anabolic
androgenic steroids
• Hypothyroidism
• Obesity
• Acromegaly [42]
• Cushing’s disease
• Insulin resistance (MetS/T2DM)
• Non-alcoholic fatty liver disease (NAFLD),
• Nephrotic syndrome

3.2.2 Androgen receptor


Testosterone and DHT exert their biological action through activation of a specific nuclear receptor. The
androgen receptor (AR) gene is localised on the X chromosome (Xq11–12), encoded in eight exons [43]. Exon
1 includes two polymorphic trinucleotide repeat segments encoding polyglutamine (CAG) and polyglycine
(GGN) tracts in the N-terminal transactivation domain of its protein. Activity of the AR is inversely associated
with the length of the CAG repeat chains [43]. However, the specific role of AR CAG repeat number in relation to
hypogonadal symptoms or to clinical management of testosterone deficiency remains unclear [44, 45]. A RCT
has shown that a higher CAG repeat number is positively associated with a change in fasting insulin, triglyceride
and diastolic blood pressure, demonstrating the more sensitive the receptor, the greater the benefit [46].

3.3 Role of testosterone in male sexual and reproductive health


3.3.1 Sexual development and maturation
Testosterone production in the foetal testis starts between the eighth and ninth week of gestation after the
expression of the SRY gene, which regulates organisation of the undifferentiated gonadal ridge into the
testis [47]. During the first trimester, the testes drive the virilisation of internal and external genitalia through
placental human chorionic gonadotropin (hCG)-stimulated androgen secretion by Leydig cells. During foetal
life, testosterone mainly controls the differentiation of internal genitalia and testicular descent (regression
of gubernaculum testis), whereas DHT is mainly involved in the development of the external male genitalia
[48]. During puberty, reactivation of the hypothalamus–pituitary-gonadal (HPG) axis allows the development
of secondary sexual characteristics, spermatogenesis maturation and, along with the contribution of other
hormonal axes, completion of the adolescent growth spurt [5, 49]. Clinical models of aromatase deficiency and
oestrogen receptor insensitivity have demonstrated that testosterone conversion to oestradiol is essential for
epiphyseal closure and growth arrest [50].

3.3.2 Sexual function


Testosterone is involved in the regulation of all steps of the male sexual response. Sexual thoughts and
motivations are universally accepted as the most testosterone-dependent aspects of male sexual behaviour
[20]. The European Male Aging Study (EMAS), a population-based survey including 3,369 subjects aged
40-79 years from eight European countries, showed that sexual symptoms, particularly impairment of sexual
desire, ED and decreased frequency of morning erections, were the most specific symptoms associated
with age-depended decline of testosterone [10]. Similar findings were reported in patients consulting for
sexual dysfunctions [51]. Accordingly, several brain areas, including the amygdala, medial preoptic area,
paraventricular nucleus of the hypothalamus, and peri-aqueductal grey matter express androgen receptors
[51, 52]. Experimental and clinical studies have both documented that testosterone plays a crucial role in
regulating penile function. In particular, testosterone controls the structural integrity necessary for penile
erection, as well as several enzymatic activities within the corpus cavernosum, including a positive action on
nitric oxide (NO) formation and a negative influence on the activity of the Ras homolog gene family member
A/Rho-associated kinase (RhoA/ROCK) pathways [51, 53]. Testosterone is also involved in penile adrenergic
response and cavernous smooth muscle cell turnover [51, 53]. Although some authors have suggested a
positive role for testosterone in regulating penile phosphodiesterase 5 (PDE5) expression and activity, others
have shown an inhibitory role of oestrogens on this pathway [51, 54].

14 SEXUAL AND REPRODUCTIVE HEALTH - MARCH 2021


More limited evidence has indicated a possible role of testosterone in regulating ejaculation, acting
either at the central or peripheral level. Androgen receptors are expressed in several central spinal and super-
spinal areas involved in the control of the ejaculatory reflex [55]. Additionally, the male genital tract expresses
NO-PDE5 and RhoA/ROCK pathways, which are modulated by testosterone [55].

3.4 Classification and causes of male hypogonadism


Male hypogonadism can be classified according to the origin of the underlying problem into primary, if a
consequence of testicular dysfunction, or secondary, if due to a pituitary or hypothalamic dysfunction (Table 2).
Primary hypogonadism is also called hypergonadotropic hypogonadism, since the pituitary
tries to compensate for testicular dysfunction by increasing central stimulation. Conversely, in secondary
hypogonadism the testes are inadequately stimulated by gonadotropins, usually with inappropriately normal
or reduced gonadotropin levels [5, 36]. A compensated or subclinical form of hypogonadism, characterised
by normal testosterone serum levels and elevated LH production, has also been reported [56]; the clinical
significance of the latter condition is unclear [56-58]. Finally, hypogonadism can also result from several
conditions leading to reduced sensitivity/insensitivity to testosterone and its metabolites [5, 36] (Table 2). This
classification, based on the aetiology of hypogonadism, allows clinicians to adequately select appropriate
treatment. In patients with secondary hypogonadism, both fertility and testosterone normalisation can be
theoretically achieved with adequate treatment whereas in primary hypogonadism only testosterone therapy
can be considered, which impairs fertility due to suppression of the HGP axis [5, 36] (Table 2). However, it
should also be recognised that symptoms and signs of hypogonadism can be similarly independent of the site
of origin of the disease. Conversely, the age of onset of hypogonadism can influence the clinical phenotype
[37]. Accordingly, when the problem starts early, such as during foetal life, clinical phenotype can span from
an almost complete female phenotype (e.g., complete androgen insensitivity or enzymatic defects blocking
androgen synthesis) to various defects in virilisation. In the case of a pre- or peri-pubertal appearance of
hypogonadism due to a milder central (isolated hypogonadotropic hypogonadism) or a peripheral defect (such
as in Klinefelter’s syndrome), there might be delayed puberty with an overall eunochoid phenotype. Finally,
when hypogonadism develops after puberty and especially with ageing (i.e., LOH; see below), symptoms can
be mild, and often confused the with ageing process per se [5, 37].
In 2017, Grossmann and Matsumoto suggested a new classification of adult male hypogonadism,
distinguishing functional versus organic hypogonadism [59]. Accordingly, organic hypogonadism is
characterised by any proven pathology affecting the HPG axis and should be treated with conventional
medication (i.e., gonadotropins or testosterone therapy). Conversely, functional hypogonadism is based on
the absence of any recognised organic alterations in the HPG axis and should be treated first by resolving or
improving the associated comorbidity. These Guidelines refer to the validated international classification of
adult male hypogonadism.

Table 2: Classification of male hypogonadism

PRIMARY HYPOGONADISM (hypergonadotropic hypogonadism)


Congenital or developmental disorders
Common causes Uncommon causes
Klinefelter syndrome - Rare chromosomal abnormalities
- XX male syndrome
- 47 XYY syndrome
- 48 XXYY syndrome
- 21 Trisomy (Down syndrome)
- Noonan syndrome
- Autosomal translocations1
- Defects of testosterone biosynthesis
- CAH (testicular adrenal rest tumours)
- Disorders of sex development (gonadal dysgenesis)
- LHR gene mutations
- Myotonic dystrophy (including type I and II)
- Uncorrected cryptorchidism (including INSL3 and LGR8 mutations)
- Bilateral congenital anorchia
- Sickle cell disease
- Adreno-leukodystrophy

SEXUAL AND REPRODUCTIVE HEALTH - MARCH 2021 15


Acquired disorders
Drug-induced Localised problems
- Chemotherapy agents - Bilateral surgical castration or trauma
Alkylating agents - Testicular irradiation
Methotrexate - Orchitis (including mumps orchitis)
- Testosterone synthesis inhibitors - Autoimmune testicular failure
• Ketoconazole - Testicular Torsion
• Aminoglutethimide - Alcohol/Cirrhosis
• Mitotane - Environmental Toxins
• Metyrapon
Systemic diseases/conditions with hypothalamus/pituitary impact
- Chronic systemic diseases* - Malignancies
- Chronic organ failure* - Lymphoma
- Glucocorticoid excess (Cushing - Testis cancer
syndrome)* - Spinal cord injury
- Aging* - Vasculitis
- HIV - Infiltrative diseases (amyloidosis; leukaemia)
SECONDARY HYPOGONADISM (hypogonadotropic hypogonadism)
Congenital or developmental disorders
Common causes Uncommon causes
- Haemochromatosis* - Combined hormone pituitary deficiency
- Idiopathic hypogonadotropic hypogonadism
- (IHH) with variants:
- Normosmic IHH
- Kallmann syndrome
- Isolated LH β gene mutations
- Prader-Willi Syndrome
Acquired disorders
Drug-induced Localised problems
- Oestrogens - Traumatic brain injury
- Testosterone or androgenic - Pituitary neoplasm (micro/macro-adenomas)
anabolic steroids - Hypothalamus tumours
- Progestogens (including - Pituitary stalk diseases
cyproterone acetate) - Iatrogenic
- Hyperprolactinaemia-induced - Surgical hypophisectomy
drugs - Pituitary or cranial irradiation
- Opiates - Inflammatory and infectious diseases
- GnRH agonist or antagonist - Lymphocytic hypophysitis
- Glucocorticoids - Pituitary infections
- Granulomatous lesions
- Sarcoidosis
- Wegener’s granulomatosis
- Other granulomatosis
- Encephalitis
- Langerhans’ histiocytosis
- Hyperprolactinaemia as a consequence of localised problems
(hypothalamus-pituitary mass)
Systemic diseases/conditions impacting the hypothalamus/pituitary
- Chronic systemic diseases* - Spinal cord injury
- Metabolic diseases - Transfusion-related iron overload (β-thalassemia)
- HIV infection
- Chronic organ failure
- Chronic Inflammatory Arthritis
- Glucocorticoid excess (Cushing
syndrome)*
- Eating disorders*
- Endurance exercise
- Acute and critical illness
- Ageing*

16 SEXUAL AND REPRODUCTIVE HEALTH - MARCH 2021


ANDROGEN RESISTANCE/DECREASED TESTOSTERONE BIOACTIVITY
Congenital or developmental disorders
- Aromatase deficiency
- Kennedy diseases (spinal and bulbar muscular atrophy) and other extensions of CAG repeats
- Partial or complete androgen insensitivity
- 5α reductase type II (5αR) deficiency
Acquired disorders
Drug-induced Localised problems
- Drug-induced AR blockage - Coeliac disease
- Steroidal antiandrogen
- Cyproterone acetate
- Spironolactone
- Non-steroidal antiandrogen
- Flutamide
- Bicalutamide
- Nilutamide
-D  rug-induced 5α reductase (5αR)
activity blockade
- Finasteride
- Dutasteride
- Drug-induced ER blockade
- Clomiphene
- Tamoxifen
- Raloxifene
-D  rug-induced aromatase activity
blockade
- Letrozole
- Anastrazole
- Exemestane
- Increased SHBG
* Conditions acting and central and peripheral levels resulting in either primary and secondary hypogonadism.
1 Different autosomal translocations can cause rare cases of hypogonadism and infertility.

3.5 Late-onset hypogonadism


Testosterone production declines with ageing. The EMAS study reported a 0.4% per annum (log hormone-
age) decrease in total testosterone and a 1.3% per annum decline in free testosterone (fT) [8]. Late onset
hypogonadism is the term frequently used to describe this phenomenon and the detection of hypogonadism in
adulthood, in particular. Evidence has documented that several associated diseases and chronic co-morbidity
can interfere with the HPG axis leading to development of primary hypogonadism or, more frequently,
secondary hypogonadism in adulthood, thus significantly influencing the physiological age-dependent decline
of testosterone. By combining the data from three different waves of the Massachusetts Male Aging Study
(MMAS), a population-based, observational study including 1,709 men aged 40–70 years, showed that
associated comorbidity and obesity significantly decreased, whereas smoking tended to increase total, free
and bio-available testosterone concentrations [60]. Similarly, data derived from the EMAS study confirm
these findings [8, 57]. Based upon these data and other evidence, the concept of functional and organic
hypogonadism has been recently introduced [59]. The diagnosis of functional hypogonadism is based on
the exclusion of a classical (organic) aetiology. The main causes of functional hypogonadism are obesity,
co-morbidity and ageing with the first two accounting for most cases. Inflammatory cytokines released in
chronic inflammation, and adipocytokines and oestradiol in obesity, can suppress the HPG axis. The role of
ageing up to age 80 years seems relatively small [59]. Considering that suppression of HPG axis activity is
functional, and potentially reversible by empiric measures, such as weight loss, the need for testosterone
therapy has been questioned [59].

3.5.1 Diagnostic evaluation


The phenotype of the hypogonadal patient appears independent of the aetiology causing the problem, but
is more often affected by the age of onset of hypogonadism. When androgen deficiency is complete and
develops during foetal life, symptoms can be dramatic, spanning from an almost complete female phenotype
(complete androgen insensitivity or enzymatic defects blocking androgen synthesis) to various defects in
virilisation and ambiguous genitalia (micropenis, hypospadias and cryptorchidism) [5, 36]. Delay in puberty

SEXUAL AND REPRODUCTIVE HEALTH - MARCH 2021 17


with an overall eunochoidal phenotype (scant body hair, high-pitched voice and small testes, penis and
prostate) is typical of defects manifesting over the pre- or peri-pubertal period due to milder central (isolated
HH) or peripheral defects (such as in Klinefelter syndrome) [5, 36]. When hypogonadism occurs in adulthood,
especially functional hypogonadism, symptoms can often be mild, difficult to recognise and frequently
confused with the ageing process [5, 36] or with chronic comorbidity. Several non-specific clinical features,
such as fatigue, weakness, and decreased energy, as well as sexual impairment may be clinical manifestations.
The EMAS study showed that a triad of sexual symptoms, including low libido, reduced spontaneous erections
and ED, are typically associated with a decrease in serum testosterone levels [10]. Conversely, psychological
and physical symptoms were less informative [10].

The mainstay of LOH diagnosis includes signs and symptoms consistent with hypogonadism, coupled with
biochemical evidence of low morning serum total testosterone levels on two or more occasions, measured
with a reliable assay. Testosterone levels show a circadian variation, which persist in ageing men [61, 62].
Likewise, testosterone levels are potentially influenced by food intake [63]; hence, serum total testosterone
should be measured in fasting conditions and in the morning (between 07.00 and 11.00 hours). A confirmatory
measurement should always be undertaken in the case of a primary pathological value, and certainly before
starting any testosterone therapy.

Liquid Chromatography-Tandem Mass Spectrometry (LC-MS/MS) represents the gold standard and most
accurate method for sex steroid evaluation; however, standardised automated platform immuno-assays for
total testosterone assessment demonstrate a good correlation with LC-MS/MS [64]. Conversely, available
immuno-assays are not able to provide an accurate estimation of fT; therefore, direct fT evaluation with these
methods is not recommended and should be avoided [41]. Liquid chromatography-tandem mass spectrometry
remains the standard method for fT determination. Alternatively, fT can be derived from specific mathematical
calculations taking into account serum SHBG and albumin levels [65] (http://www.issam.ch/freetesto.htm).

Data from available meta-analyses have documented that testosterone therapy is ineffective when baseline
levels are > 12 nmol/L (3.5 ng/mL). Positive outcomes are documented when testosterone levels are
< 12 nmol/L, being higher in symptomatic patients with more severe forms of hypogonadism (< 8 nmol/L).
Hence, 12 nmol/L should be considered as a possible threshold for starting testosterone therapy in the
presence of hypogonadal symptoms [21, 66]. As reported above, clinical conditions that may interfere with
SHBG levels, evaluation of fT should be considered to better estimate actual androgen levels (Figure 2).
Unfortunately, despite its potential clinical value [67], no validated thresholds for fT are available from clinical
studies and this represents an area of uncertainty; however, some data indicate that fT levels < 225 pmol/L (6.5
ng/dL) are associated with hypogonadal symptoms [10, 51, 68, 69].

The determination of LH must be performed along with prolactin (PRL) when pathological total testosterone
levels are detected, in order to correctly define the underlying conditions and exclude possible organic causes
(Figure 2). Due to its negative influence on libido, PRL can also be considered as first-line screening in patients
with reduced sexual desire. In addition, pituitary magnetic resonance imaging (MRI) scanning, as well as other
pituitary hormone evaluations, is required in the presence of specific symptoms such as visual disturbances,
headache [70, 71] or when hyperprolactinemia is confirmed. Limited evidence suggests performing pituitary
MRI also in the case of severe hypogonadism (< 6 nmol/L, 1.75 ng/mL) with inadequate gonadotropin levels
(Figure 2) [70, 71].

18 SEXUAL AND REPRODUCTIVE HEALTH - MARCH 2021


Figure 2: Diagnostic evaluation of Late-Onset Hypogonadism

Check symptoms and signs


suggesve for hypogonadism

Check for drugs and substances that can interfere with T producon/acon
Check for concomitant metabolic diseases: obesity/metabolic syndrome/diabetes
Check for potenal testosterone therapy contraindicaons

Measure fasng and morning (7-11 am) total T


(consider PRL measurement if low desire or other suggesve symptoms are present)
(consider SHBG and free-T calculaon when indicated)
(consider LH when T deficiency pathophysiology must be invesgated)

TT < 12 nM TT > 12 nM/


TT > 12 nM
hypogonadism reduced cFT
hypogonadism
possible hypogonadism
unlikely
possible

Repeat TT TT < 12 nM
measurements (reduced cFT)
along with LH and LH reduced/
PRL +/-SHBG cFT inappropriate
normal

TT < 12 nM Secondary
(reduced cFT) hypogonadism
and LH elevated

Check for drugs and substances that can interfere


Primary with T producon/acon.
hypogonadism Check for concomitant metabolic disease: obesity/
metabolic syndrome/diabetes

TT < 8 nM TT < 6nM/ Perform


elevated PRL pituitary
Headache/visual MRI
disturbances

Possible
specific
therapy
Invesgate if drugs or substances that may interfere with
hypothalamic-pituitary axis can be eliminated. Ferlity
Suggest modifying potenal interfering condions obesity /underweight desired
or other metabolic disturbances

Gonadotropin
Rule out testosterone therapy possible contraindicaons therapy

Testosterone therapy
trial

TT = total testosterone; cFT = calculated free testosterone; PRL = prolactin; SHBG = sex hormone-binding
globulin; LH = luteinising hormone; MRI = Magnetic resonance imaging.

SEXUAL AND REPRODUCTIVE HEALTH - MARCH 2021 19


3.5.2 History taking
Specific symptoms associated with LOH are shown in Table 3. History of surgical intervention for
cryptorchidism or hypospadias must be taken into account as possible signs of congenital defects. Likewise,
chronic and systemic comorbidity must be comprehensively investigated in every patient. Use of drugs that
potentially interfere with the HPG axis should be ruled out (Table 2). Acute diseases are associated with
development of functional hypogonadism and determination of serum total testosterone levels should be
avoided in these conditions. However, recent data derived from SARS-CoV-2 infected patients showing worse
outcomes in hypogonadal subjects suggest that the role of testosterone in the case of acute illness should
be clarified [72, 73]. Several self-reported questionnaires or structural interviews have been developed for
screening of hypogonadism. Although these case-history tools have demonstrated clinical utility in supporting
the biochemical diagnosis of hypogonadism, or in the assessment of testosterone therapy outcomes, their
specificity remains poor and they should not be used for a systematic screening of hypogonadal men [74].

Table 3: Specific symptoms associated with LOH

Sexual symptoms Physical symptoms Psychological symptoms


More specific - Reduced libido - Decreased vigorous activity - Low mood/mood deflection
- Erectile dysfunction - Difficulty walking >1 km - Decreased motivation
- Decreased spontaneous/ - Decreased bending - Fatigue
morning erections
Less specific - Reduced frequency of - Hot flushes - Concentration or
sexual intercourse - Decreased energy mnemonic difficulties
- Reduced frequency of -D ecreased physical - Sleep disturbances
masturbation strength/function/activity
- Delayed ejaculation

3.5.3 Physical examination


Since obesity is frequently associated with hypogonadism (mostly functional), the determination of body
mass index (BMI) and the measurement of waist circumference are strongly recommended in all individuals.
Testicular and penile size, as well the presence of sexual secondary characteristics can provide useful
information regarding overall androgen status. In addition, upper segment/lower segment ratio (n.v. > 0.92)
and arm-span to height ratio (n.v. < 1.0) can be useful to identify a eunochoid body shape, especially in
subjects with pre-pubertal hypogonadism or delayed puberty. Finally, digital rectal examination (DRE) should be
performed in all subjects to exclude prostate abnormalities before testosterone therapy (any type) or to support
suspicion of hypogonadism [75].

3.5.4 Summary of evidence and recommendations for the diagnostic evaluation of LOH

Summary of evidence
Sexual symptoms are the most specific symptoms associated with LOH.
Diagnosis of LOH should be based on specific signs and symptoms of androgen deficiency, together with
consistently low serum testosterone levels.
Functional hypogonadism is a consequence of comorbidity/concomitant drugs, which can impair
testosterone production in adulthood. The diagnosis of functional hypogonadism is a diagnosis of exclusion,
after ruling out organic causes of hypogonadism.

20 SEXUAL AND REPRODUCTIVE HEALTH - MARCH 2021


Recommendations Strength rating
Check for concomitant diseases, drugs and substances that can interfere with testosterone Strong
production/action.
Total testosterone must be measured in the morning (07.00 and 11.00 hours) and in the Strong
fasting state, with a reliable method.
Repeat total testosterone on at least two separate occasions when < 12 nmol/L and before Strong
starting testosterone therapy.
12 nmol/L total testosterone (3.5 ng/mL) represents a reliable threshold to diagnose late Strong
onset hypogonadism (LOH).
Consider sex hormone-binding globulin and free-testosterone calculation when indicated. Strong
Calculated free-testosterone < 225 pmol/L has been suggested as a possible cut-off to Weak
diagnose LOH.
Analyse luteinising hormone (LH) and follicle-stimulating hormone (FSH) serum levels to Strong
differentiate between primary and secondary hypogonadism.
Consider prolactin (PRL) measurement if low sexual desire (or other suggestive signs/ Strong
symptoms) and low or low-normal testosterone is present.
Perform pituitary magnetic resonance imaging (MRI) in secondary hypogonadism, with Strong
elevated PRL or specific symptoms of a pituitary mass and/or presence of other anterior
pituitary hormone deficiencies.
Perform pituitary MRI in secondary severe hypogonadism (total testosterone < 6 nmol/L). Weak

3.5.5 Recommendations for screening men with LOH

Recommendations Strength rating


Screen for late onset hypogonadism (LOH) (including in T2DM) only in symptomatic men. Strong
Do not use structured interviews and self-reported questionnaires for systematic screening Strong
for LOH as they have low specificity.

3.6 Treatment of LOH


3.6.1 Indications and contraindications for treatment of LOH
Patients with symptomatic hypogonadism (total testosterone < 12 nmol/L) without specific contraindications
are suitable candidates to receive testosterone therapy (Table 4).

Absolute contraindications are untreated breast cancer and prostate cancer (PCa). Acute cardiovascular events
as well as uncontrolled or poorly controlled congestive heart failure and severe lower urinary tract symptoms
(LUTS) [International Prostate Symptom Score (IPSS) score > 19] represent other contraindications, as there
is insufficient information on the long-term effects of testosterone therapy in these patients [66]. A positive
family history for venous thromboembolism requires further analysis to exclude a condition of undiagnosed
thrombophilia-hypofibrinolysis [76]. These patients need to be carefully counselled prior to testosterone
therapy initiation. A haematocrit (HCT) > 54% should require testosterone therapy withdrawal, reduction in
dose, change of formulation and venesection depending on the clinical situation to avoid any potential cardio-
vascular complications. Lower baseline HTC (48-50%) should be carefully evaluated before testosterone
therapy initiation, to avoid pathological increases during treatment, especially in high-risk men such as
those with chronic obstructive pulmonary disease (COPD) or Obstructive Sleep Apnoea Syndrome (OSAS).
Accordingly, the Framingham Heart Study showed that HCT > 48% represented a condition associated with
increased risk of coronary artery disease (CAD) and mortality and was associated with cardiovascular disorders
[77]. Finally, testosterone therapy suppresses gonadotropin and endogenous testosterone secretion as well
as spermatogenesis. Hence, testosterone therapy is contraindicated in individuals who desire fertility [78].
Secondary hypogonadism is characterised by low or inappropriately normal gonadotropin levels; therefore, the
rationale is to substitute the gonadotropin deficiency with FSH and LH analogues, if fertility is desired [79].

Table 4: Main contraindications of testosterone therapy

SEXUAL AND REPRODUCTIVE HEALTH - MARCH 2021 21


Absolute contraindications Locally advanced or metastatic prostate cancer (PCa)
Male breast cancer
Men with an active desire to have children
Haematocrit ≥ 54%
Uncontrolled or poorly controlled congestive heart failure
Relative contraindication IPSS score > 19
Baseline haematocrit 48-50%
Familial history of venous thromboembolism

3.6.2 Testosterone therapy outcomes


3.6.2.1 Sexual dysfunction
Sexual concerns are the main symptoms of hypogonadal patient [5, 10, 80, 81]. A consistent body of
evidence shows that testosterone therapy in hypogonadal men (total testosterone < 12 nmol/L) may have
a beneficial effect on several aspects of sexual life; in contrast, there is no evidence of benefits in using
testosterone therapy for treating sexual dysfunction in eugonadal men [53, 66, 82, 83]. The beneficial effect
on sexual function seems to be more related to testosterone level normalisation than the specific testosterone
formulations used [83, 84].

A recent meta-analysis of only placebo-controlled RCTs using the International Index of Erectile Function
(IIEF) [85] as a possible tool for outcome evaluation, showed that testosterone therapy significantly improves
erectile function (as measured by IIEF-Erectile Function domain score) and that patients with more severe
hypogonadism (i.e., total testosterone < 8 nmol/L) are more likely to achieve better improvement than patients
with milder hypogonadism (i.e., total testosterone < 12 nmol/L). Similar results were observed for sexual desire;
however, the presence of metabolic comorbidity (such as diabetes and obesity) decreased the magnitude of
these improvements. In particular, testosterone therapy alone resulted in a clinically effective outcome only
in patients with milder ED [66]. Other sexual function parameters, such as intercourse, orgasm and overall
satisfaction, were all improved compared with placebo [66]. Men with comorbidity such as diabetes usually
show modest improvements in terms of sexual function after testosterone therapy and may potentially require
concomitant phosphodiesterase type 5 inhibitors (PDE5Is) to improve effectiveness [5, 83]. However, the
specific beneficial effect derived from the combined use of testosterone therapy and PDE5Is is not completely
clear [53]. Similarly, information related to the combined use of testosterone therapy with other ED drug
therapies is lacking [5, 83].

The Sexual Function Trial of the Testosterone Trials (TTrials) (one of the largest placebo-controlled trials on
testosterone therapy) documented consistent improvements in 10 of 12 measures of sexual activities in older
(≥ 65 years) hypogonadal men, particularly in frequency of intercourse, masturbation and nocturnal erections
(as measured by PDQ-Q4) [86]. The magnitude in improvement was shown to be proportional to the increase
in serum total testosterone, fT and oestradiol levels, it was not possible to demonstrate a threshold level
[87]. A study of 220 men with MetS with or without T2DM also found that sexual function improved in men
who reported sexual problems with improvement in IIEF scores with specific increases in libido and sexual
satisfaction [24].

3.6.2.2 Body composition and metabolic profile


Late onset hypogonadism is associated with a greater percentage fat mass and a lesser lean mass compared
to testosterone-replete men [88]. The major effect of low testosterone is to increase visceral adiposity but also
leads to deposition of lipids in the liver and muscle and is associated with atherosclerosis [19]. Some published
data have suggested that testosterone therapy reduces percentage body fat and increases lean mass [89].
Testosterone therapy has also been found to decrease waist circumference, body weight and BMI, with these
effects more predominant after 12 months of treatment [89-91]. However, it should be recognised that these
results are mainly derived from registry and observational trials, which have important limitations due to the
risk of selection bias for the non-random assignment of testosterone exposure. Accordingly, data derived from
RCTs showed only an improvement of fat mass and lean mass of the same amount without any modifications
in body weight [21].

3.6.2.3 Mood and cognition


Several observational studies have documented a relationship between depressive symptoms, reduced QoL
and hypogonadism [92, 93]. However, the specific relationship between hypogonadism and the incidence of
depression is still unclear [93]. Only a few placebo-controlled RCTs have investigated the role of testosterone

22 SEXUAL AND REPRODUCTIVE HEALTH - MARCH 2021


therapy in improving depressive symptoms. Data derived from TTrials showed that testosterone therapy
improved mood, and depressive symptoms as continuous measures using several instruments [86]. However,
the final effect was small in magnitude. In line with these data, the largest meta-analysis of available studies,
including 1,890 hypogonadal men (baseline total testosterone < 12 nmol/L or fT < 225 pmol/L) men from
27 RCTs, documented that the positive effect of testosterone therapy was particularly evident in patients
with milder symptoms [94]. The BLAST study of testosterone therapy in T2DM reported that those men with
depression were less likely to respond with regard to symptoms of sexual dysfunction compared to men
without depression [29].
Robust data on the effect of testosterone therapy on QoL are limited. Although recent meta-
analyses suggest a significant effect of testosterone therapy over placebo, the magnitude is low and the
heterogeneity high, therefore reducing the scientific value of the effect [84, 95].
The role of testosterone therapy in patients with cognitive impairment is even more uncertain. The
TTrials evaluated the effect of testosterone therapy in 493 individuals with age-associated memory impairment
in order to assess possible improvement of several aspects of cognitive function. However, the final results
failed to demonstrate any beneficial effect of testosterone therapy in improving cognitive function [86].

3.6.2.4 Bone
Evidence suggests that bone mineralisation requires circulating sex steroids within the normal range [96].
The possible association between mild hypogonadism and osteopenia/osteoporosis is weak, whereas severe
hypogonadism (total testosterone < 3.5 nM) is frequently associated with bone loss and osteoporosis,
independent of patient age [96]. Two independent meta-analyses showed a positive effect of testosterone
therapy on bone mineral density (BMD), with the highest effect at the lumber level [97, 98]. Similarly, data
derived from TTrials confirmed that testosterone therapy increased BMD in hypogonadal ageing men,
particularly at the lumbar level [86]. However, available data are insufficient to determine the effect of
testosterone therapy alone on the risk of fractures [96]. The use of testosterone therapy as an adjunct to
anti-resorptive treatment in hypogonadal patients at high risk of fractures has not been established. Therefore,
anti-resorptive therapy must be the first-choice treatment in hypogonadal men at high risk for bone fractures.
The combination of anti-resorptive treatment and testosterone therapy should be offered only in conjunction
with hypogonadism-related symptoms.

3.6.2.5 Vitality and physical strength


The role of testosterone in stimulating muscle growth and strength is well established. Accordingly, androgenic-
anabolic steroids (AAS) have been used as performance-enhancing agents to increase physical performance
in competitive sport [99]. In this regard, testosterone therapy in hypogonadal men has been shown to increase
muscle mass and reduce fat mass, with limited effects on final weight [21]. Despite this evidence, the role of
testosterone therapy in older men with mobility limitations remains unclear. The National Health and Nutrition
Examination Survey 1999-2004 [100] was unable to detect any association between overall circulating
testosterone levels and the amount of physical activity. However, among non-obese men, those in the highest
physical activity tertile were significantly less likely to have low or low-normal testosterone than those in the
lowest tertile. Data from TTrials indicated that testosterone therapy did not substantially increase the fraction
of men whose 6-minute walking distance increased > 50 m or the absolute increase in the distance walked
by those enrolled in the physical function trial [86]. However, when the whole population of the TTrials was
considered, a significant, although modest, positive effect on these two parameters was reported [86]. Similar
data were derived from the Vitality Trial [86].

3.6.2.6 Summary of evidence and recommendations for testosterone therapy outcome

Summary of evidence
Testosterone therapy can improve milder forms of ED and libido in hypogonadal men.
Testosterone therapy can improve other sexual symptoms, including intercourse frequency, orgasm and
overall satisfaction.
Testosterone therapy can similarly increase lean mass, reduce fat mass, and improves insulin resistance.
Testosterone therapy can improve weight, waist circumference and lipid profile, but findings are not unique.
Testosterone therapy can improve milder depressive symptoms in hypogonadal men.
Testosterone therapy can improve bone mineral density, but information related to fracture risk is lacking.

SEXUAL AND REPRODUCTIVE HEALTH - MARCH 2021 23


Recommendations Strength rating
The use of testosterone therapy in eugonadal men is not indicated. Strong
Use testosterone therapy as first-line treatment in patients with symptomatic hypogonadism Strong
and mild erectile dysfunction (ED).
Use combination of phosphodiesterase type 5 inhibitors and testosterone therapy in more Weak
severe forms of ED as it may result in better outcomes.
Use conventional medical therapies for severe depressive symptoms and osteoporosis. Strong
Do not use testosterone therapy to improve body composition, reduce weight and benefit Weak
cardio-metabolic profile.
Do not use testosterone therapy to improve cognition vitality and physical strength in ageing Strong
men.

3.6.3 Choice of treatment


3.6.3.1 Lifestyle factors
As reported above, functional hypogonadism is frequently associated with obesity and metabolic disorders
[101]. Therefore, weight loss and lifestyle changes should be the first approach for all overweight and obese
men with hypogonadism. A previous meta-analysis documented that a low-calorie diet is able to revert
obesity-associated secondary hypogonadism by increasing total testosterone and fT, reducing oestrogens and
restoring normal gonadotropin circulating levels [102]. This was confirmed in a recent updated meta-analysis
showing that the increase in testosterone is significantly associated with weight reduction [103]. Similar results
can be obtained through physical activity, which is associated with the duration of scheduled exercise and
weight loss obtained [103]. However, it should be recognised that the increase in testosterone levels observed
after a low- calorie diet and physical activity is small (1-2 nmol) [102, 103]. It should also be recognised that
60-86% of weight lost is regained after 3 years and 75-121% after 5 years [104]. A greater testosterone
increase can be achieved through bariatric surgery, which results in an average increase of about 10 nmol/L
depending on the degree of weight loss [103]. Lifestyle changes represent an essential part of the management
of obesity; however, some evidence suggests that when compared to lifestyle modifications alone, testosterone
therapy-treated obese men benefit most from relief of their symptoms associated with testosterone deficiency,
whereas those not treated did not benefit [79]. There is limited evidence to suggest that combination of life-
style interventions and testosterone therapy in symptomatic hypogonadal men might result in better outcomes
[88]. There is a large, on-going placebo-controlled RCT that aims to determine whether lifestyle intervention
alone or in combination with testosterone therapy reduces T2DM incidence and improves glucose tolerance
[105].

3.6.3.2 Medical preparations


Several testosterone formulations are available (Table 5). Direct comparisons among different testosterone
products are still lacking. Candidates for testosterone therapy should be adequately informed about the
possible risks and benefits of all available testosterone preparations. The final choice should be based on the
clinical situation, testosterone formulation availability, and patient needs and expectations [106].

3.6.3.2.1 Oral formulations


The esterification of testosterone with a long-chain fatty acid (testosterone undecanoate; TU) enables
testosterone to be absorbed by the intestine through the lymphatic system, by-passing liver metabolism. This
formulation has been available in oleic acid since the 1970s, and it has been recently reformulated in a mixture
of castor oil and propylene glycol laureate (TU caps), to allow the drug to be maintained at room temperature
without degradation [107]. The main limitation is related to the poor bioavailability, which is strongly dependent
on dietary fat content [107]. Recently, the US Food and Drug Administration (FDA) approved a new formulation
of oral TU incorporating a liquid-filled hard capsule drug delivery system and a higher amount (225 mg)
of the compound, which improves oral availability (https://www.fda.gov/media/110187/download). In an
open label study of approximately 4 months’ duration (NCT02722278), 145 (87%) of 166 hypogonadal men
enrolled who received the TU caps formulation had mean total testosterone concentration within the normal
eugonadal range at the end of treatment (https://www.fda.gov/media/110187/download). However, the TU
caps compound is not available in Europe.
Mesterolone is a 5α-DHT derivate available for oral administration. Along with DHT, mesterolone
cannot be converted to oestrogens and can only be used for a limited period and specific indications, such as
the presence of painful gynaecomastia. However, the lack of a full spectrum of testosterone bioactivity strongly
limits its long-term use [107].

24 SEXUAL AND REPRODUCTIVE HEALTH - MARCH 2021


3.6.3.2.2 Parenteral formulations
Injectable testosterone preparations can be classified according to their half-lives (Table 5). Testosterone
propionate is a short-term ester formulation requiring multiple fractionated doses (usually 50 mg, every 2-3
days), thus representing a major limitation for its use [107]. Cypionate and enanthate-T esters are short-term
formulations, requiring administration every 2-4 weeks. A formulation containing mixed testosterone esters (TU,
isocaproate, phenyl propionate, propionate - Sustanon®) which allows some benefit of a smoother release of
testosterone into the circulation is available in some countries. The use of these older formulations is associated
with wide fluctuations in plasma testosterone concentrations and is often reported as unpleasant by patients
and potentially resulting in adverse effects, such as polycythaemia [107, 108]. A longer-lasting TU injectable
formulation is widely available [107]; which has a good safety/benefit profile allowing the maintenance of normal
stable testosterone levels at a dose of 1,000 mg initially every 12 weeks, following a 6-week loading dose, but
can be adjusted to a frequency of 10-14 weeks dependent on the trough (pre-injection level) after 3-5 injections
to maintain levels in the therapeutic range (usually > 12 and < 18 nmol/L) [107, 109].

3.6.3.2.3 Transdermal testosterone preparations


Among the available transdermal formulations, testosterone gels represent the most frequently used
preparations. The gel is quickly absorbed by the stratum corneum, creating a reservoir within the subcutaneous
tissues from where testosterone is continuously delivered for 24 hours, after a single daily application. These
formulations have been shown to normalise serum testosterone levels with an excellent safety profile [107].
The introduction of specific devices and skin enhancers has resulted in better skin penetration of the drugs,
thus reducing potential adverse effects. Local skin adverse effects are limited when compared to those with
traditional testosterone patches, but they potentially allow transference of testosterone during close contact
with the skin surface. The risk can be reduced by wearing clothing or by applying the gel on skin surfaces
not usually touched (e.g., the inner thigh surface) [107]. To reduce the total amount of gel applied and
residual quantities remaining on the skin, new formulations of testosterone gel have been introduced with a
testosterone concentration of 1.62-2% [107]. Another transdermal testosterone formulation includes a topical,
alcohol-based testosterone (2%) solution, which must be applied to the underarm once daily, using a metered
dose applicator [107]. This testosterone formulation is not available in Europe. Testosterone levels should be
monitored to optimise the testosterone dose. Blood collection is best taken at 2-4 hours after gel application
to use the peak level of testosterone absorbed as a reference for adequate therapeutic levels. Levels of
testosterone after application can vary and a repeat measurement may be indicated especially as sometimes,
inadvertently, the skin over the vene-puncture site can be contaminated by the gel, leading to falsely elevated
results.
In some European countries, DHT is available as a hydroalcoholic 2.5% gel. It is rapidly absorbed,
reaching a steady state in 2-3 days. Similar to that reported for mesterolone, DHT is not aromatised but can be
useful for treating particular conditions, such as gynaecomastia and microphallus [107].

3.6.3.2.4 Transmucosal formulations


3.6.3.2.4.1 Transbuccal Testosterone preparations
A testosterone buccal system is still available in several countries. It consists on a sustained-release muco-
adhesive buccal-testosterone-tablet requiring twice-daily application to the upper gums. The tablet does not
dissolve completely in the mouth and must be removed after 12 hours. This formulation has been proven to
restore testosterone levels within the physiological range with minimal or transient local problems, including
gum oedema, blistering and gingivitis [107].

3.6.3.2.4.2 Transnasal testosterone preparations


A gel for intranasal administration is available in some countries, including the USA and Canada. It requires
administration two or three times daily using a specific metered-dose pump. The application is rapid, non-
invasive, convenient, and avoids secondary transference observed with other topical products [107].

3.6.3.2.5 Subdermal depots


The implantation of testosterone pellets, available in the USA, UK and Australia, represents the longest
available testosterone formulation lasting from 4-7 months. However, the procedure is invasive and may be
unattractive to patients [107].

3.6.3.2.6 Anti-oestrogens
Anti-oestrogens, including selective oestrogen receptor (ER) modulators (SERMs) and aromatase inhibitors (AI)
have been suggested as off-label treatments to restore testosterone levels and fertility in men with functional
secondary hypogonadism or idiopathic infertility. They work by preventing down-regulation of the HPG axis
by oestrogens and, for this reason are particularly useful in men with obesity and metabolic disorders [103]. In

SEXUAL AND REPRODUCTIVE HEALTH - MARCH 2021 25


the latter case, the hypothesis is that the excess of adipose tissue leads to increased aromatase activity and
oestrogens levels resulting in impairment of the HPG [101]. Due to their putative mechanism of action, they
require an intact HPG axis and cannot work in primary hypogonadism or secondary hypogonadism due to
organic damage of the HPG axis. Both types of SERMs, which bind ERs with an agonist or antagonist effect
depending upon the target tissue, and AIs, which prevent androgens from being converted into oestrogens by
aromatase, have been used in clinical practice [107]. The evidence published so far is poor; all these products
are off-label treatments and SERMs, due to their agonistic effect on venous vessels, could predispose men to
the development of venous thromboembolism [107]. In this context patients should be warned of the potential
increased risk of venous thromboembolism, although data are lacking. Long-term use of these agents can lead
to reduced bone density and development of osteoporosis, potentially increasing fracture risk.

3.6.3.2.7 Gonadotropins
Considering the aforementioned limitations regarding the use of anti-oestrogens, gonadotropin therapy
should be considered the standard in men with secondary hypogonadism who desire paternity (Table 5)
[107]. The treatment is based on the use of human chorionic gonadotropin (hCG), purified from the urine of
pregnant women. The most expensive recombinant hCG (rhCG) and LH (rLH) formulations do not offer clinical
advantages [107]. According to a meta-analysis of the available evidence, hCG should be administered with
FSH as combined therapy results in better outcomes. Similar to recombinant hCG, recombinant FSH (rFSH)
does not seem to offer any advantages compared to urinary-derived preparations [110]. More details on the
use of gonadotropins are provided in Section 10.

Table 5: Available preparations for hypogonadism treatment

Formulation Chemical structure t 1/2 Standard Advantages Disadvantages


dosage
GONADOTROPINS
Human chorionic gonadotrophin (HCG)
Extractive HCG purified from NA 1,000-2,000 Low cost Multiple weekly
the urine of pregnant IU administration
women 3 times/week
Recombinant Human recombinant NA No data in NA
HCG men
Luteotropic hormone (LH)
Recombinant Human recombinant NA No data in NA
LH men
Follicle-stimulating hormone (FSH)
Extractive FSH purified from NA 75-150 IU Low cost Multiple weekly
urine of pregnant 3 times/week administration
women
Recombinant Human recombinant NA 75-150 IU Multiple weekly
FSH 3 times/week administration
TESTOSTERONE PREPARATIONS
Oral
Testosterone 17-α-hydroxylester 4 120-240 mg - Reduction of liver - Unpredictable
undecanoate hours 2-3 times involvement absorption
daily - Oral convenience depending on
- Modifiable dosage dietary fat content
- Must be taken with
meals
Mesterolone 1α-methyl-4, 12 50-100 mg - Oral convenience - Not aromatisable
5α-dihydro- hours 2-3 times - Modifiable dosage
testosterone daily - Useful in
gynaecomastia

26 SEXUAL AND REPRODUCTIVE HEALTH - MARCH 2021


Parental
Testosterone 17-α-hydroxylester 4-5 250 mg every - Low cost - Fluctuations
enanthate days 2-3 weeks -S hort-acting in circulating
preparation allowing testosterone levels
drug withdrawal - Multiple injections
in case of adverse - Relative risk of
effects polycythemia
Testosterone 17-α-hydroxylester 8 days 200 mg every - Low cost - Fluctuations
cypionathe 2-3 weeks - Short-acting in circulating
preparation allowing testosterone levels
drug withdrawal - Multiple injections
in case of adverse - Relative risk of
effects polycythemia
Testosterone 17-α-hydroxylester 20 100 mg every - Low cost - Fluctuations
propionate hours 2 days - Very short-acting in circulating
preparation allowing testosterone levels
drug withdrawal - Multiple injections
in case of adverse - Relative risk of
effects polycythemia
Testosterone 4-androsten-3-one- 4-5 250 mg every - Low cost - Fluctuations
ester mixture 17 beta-hydroxy- days 3 weeks - Short-acting in circulating
Propionate androst-4-en-3-one preparation allowing testosterone levels
(30mg) drug withdrawal - Multiple injections
Phenylpropionate in case of adverse - Relative risk of
(60 mg) effects polycythemia
Isocaproate
(60 mg)
Decanoate
(100 mg)
Testosterone 17-α-hydroxylester 34 1,000 mg - Steady-state - Pain at injection
undecanoate in days every 10-14 testosterone level site
castor oil weeks without fluctuation - Long-acting
*750 mg - Long-lasting preparation not
every 10 - Less frequent allowing rapid drug
weeks administration withdrawal in case
of adverse effects
Surgical implants Native testosterone -- 4-6 200 mg - Long duration and - Placement is
implants constant serum invasive
lasting up to testosterone level - Risk of extrusion
6 months and site infections
TRANSDERMAL
Testosterone Native testosterone 10 50-100 mg/ Steady-state - Skin irritation
patches hours day testosterone level - Daily
without fluctuation administration
Testosterone gel Native testosterone 6 50-100 mg/ Steady-state - Possible transfer
1-2% hours day testosterone level during intimate
without fluctuation contact
- Daily
administration
Underarm Native testosterone NA 60-120 mg/ Steady-state - Possible transfer
testosterone day testosterone level during intimate
(testosterone without fluctuation contact
solution 2%) - Daily
administration
Dihydro- Native dihydro- NA 34-70 mg/day - Steady-state - Possible transfer
testosterone gel testosterone testosterone level during intimate
2.5% without fluctuation contact
- Useful in - Daily
gynaecomastia administration
- Not aromatisable

SEXUAL AND REPRODUCTIVE HEALTH - MARCH 2021 27


TRANSMUCOSAL
Testosterone Native testosterone 12 60 mg Steady-state - Possible oral
buccal system hours 3 times daily testosterone level irritation
without fluctuation - Twice-daily dosing
- Unpleasant taste
Testosterone Native testosterone 6 33 mg Steady-state - Nasal irritation
nasal hours 3 times daily testosterone level - Multiple daily
without fluctuation administration
NA = not applicable.

3.6.3.3 Summary of evidence and recommendations for choice of treatment for LOH

Summary of evidence
Weight loss obtained through a low-calorie diet and regular physical activity result in a small improvement in
testosterone levels.
Testosterone gels and long-acting injectable TU represent T preparations with the best safety profile.
Gonadotropins treatment can be used to restore fertility in men with secondary hypogonadism.

Recommendations Strength rating


Treat, when indicated, organic causes of hypogonadism (e.g., pituitary masses, Strong
hyperprolactinemia, etc).
Improve lifestyle and reduce weight (e.g., obesity); withdraw, when possible, concomitant Weak
drugs that can impair testosterone production; treat co-morbidity before starting
testosterone therapy.
Fully inform patients about expected benefits and adverse effects of any treatment option. Strong
Select the testosterone preparation in a joint decision process, only with fully informed
patients.
The aim of testosterone therapy is to restore serum testosterone concentration to the Weak
average normal range for young men.
Use testosterone gels rather than long-acting depot administration when starting initial Weak
treatment, so that therapy can be adjusted or stopped in the case of treatment-related
adverse effects.

3.7 Safety and follow-up in hypogonadism management


3.7.1 Hypogonadism and fertility issues
The aim of pharmacological management of hypogonadism is to increase testosterone levels. The first choice
is to administer exogenous testosterone. However, while exogenous testosterone has a beneficial effect on the
clinical symptoms of hypogonadism, it also inhibits gonadotropin secretion by the pituitary gland, resulting in
impaired spermatogenesis and sperm cell maturation [111]. Therefore testosterone therapy is contraindicated
in hypogonadal men seeking fertility treatment [78]. When secondary hypogonadism is present, gonadotropin
therapy can maintain normal testosterone levels and restore sperm production [5].

3.7.2 Male breast cancer


In vitro and in vivo studies have clearly documented that breast cancer growth is significantly influenced by
testosterone and/or by its conversion to E2 through different mechanisms and pathways [112]. Accordingly,
the use of SERMs still represents an important therapeutic option in the management of this cancer [112].
No information is available on the role of testosterone therapy in patients successfully treated for male breast
cancer; therefore, treated and active male breast cancer should be recognised as absolute contraindications
for testosterone therapy.

3.7.3 Lower urinary tract symptoms/benign prostatic hyperplasia


Based on the assumption that prostate growth is dependent on the presence of androgens, historically
testosterone therapy historically has raised some concerns regarding the possibility of aggravating LUTS
in patients affected by benign prostatic hyperplasia (BPH) associated with prostate enlargement [75, 113].
However, pre-clinical and clinical data have indicated that low rather than high androgen levels may decrease
bladder capacity, alter tissue histology and decrease the ratio of smooth muscle to connective tissue, thus
impairing urinary dynamics [75, 113].

28 SEXUAL AND REPRODUCTIVE HEALTH - MARCH 2021


A trial of 60 patients undergoing testosterone therapy for 6 months showed no significant differences
in post-voidal residual urine and prostate volume, while storage symptoms as measured by IPSS significantly
improved, despite an increase in prostate-specific antigen (PSA) level. A larger pre-treatment prostate volume was
a predictive factor of improvement in LUTS [114]. A long-term study of 428 men undergoing testosterone therapy
for 8 years demonstrated significant improvements in IPSS, no changes max flow rate (Qmax) and residual urine
volume, but also a significant increase in prostate volume [115]. Similar data from the Registry of Hypogonadism
in Men (RHYME), including 999 patients with a follow-up of 3 years, did not document any difference in PSA
levels or total IPSS in men undergoing testosterone therapy, compared to untreated patients [116]. Similar results
were reported in an Italian registry (SIAMO-NOI), collecting data from 432 hypogonadal men from 15 centres
[117]. Meta-analyses have not found significant changes in LUTS between patients treated with testosterone
or placebo [118-124]. According to the most recent literature, there are no grounds to discourage testosterone
therapy in hypogonadal patients with BPH/LUTS and there is evidence of limited benefit from androgen
administration. The only concern is related to patients with severe LUTS (IPSS > 19), as they are usually excluded
from RCTs, therefore limiting the long-term safety data of testosterone therapy in this specific setting [75].

3.7.4 Prostate cancer (PCa)


A considerable number of observational studies have failed to demonstrate any association between circulating
higher testosterone levels and PCa [125]. In contrast, studies investigating the relationship between low
levels of testosterone and risk of PCa have found that men with very low levels of fT have a reduced risk of
developing low-to-intermediate-grade PCa, but have a non-significantly increased chance of developing high-
grade PCa [125]. This peculiar pattern was also reported in trials such as the Health Professionals Follow-up
Study, the Prostate Cancer Prevention Trial (PCPT) and the Reduction by Dutasteride of Prostate Cancer
Events (REDUCE), with varying magnitudes of significance [126].
The most recent meta-analysis, including 27 placebo-controlled, RCTs, found no evidence of
increased PSA levels following testosterone therapy for one year. When considering 11 studies reporting on
the occurrence of PCa, the meta-analysis found no evidence of increased risk of PCa. However, a 1-year
follow-up may be considered too short to draw conclusions on PCa occurrence. Furthermore, the analysis was
restricted to studies with > 1-year follow-up, but no significant changes in PSA levels nor increased risk of PCa
were found [119]. After 5-years’ median follow-up in three independent registry studies with > 1,000 patients
undergoing testosterone therapy, PCa occurrence remained at all times below the reported incidence rate in
the general population [127]. Similar results were reported by a more recent large observational study including
10,311 men treated with testosterone therapy and 28,029 controls with a median follow-up of 5.3 years [128].
The same study, also showed that the risk of PCa was decreased for men in the highest tertile of testosterone
therapy cumulative dose exposure as compared with controls [128].

With regards to PCa survivors, safety in terms of the risk of recurrence and progression has not yet been
established. Limited data are available in the literature, with most case series not providing sufficient data to
draw definitive conclusions (e.g., insufficient follow-up, small samples, lack of control arms, heterogeneity in
study population and treatment regimen, etc.) [129]. More recently, a meta-analysis derived from 13 studies
including 608 patients, of whom 109 had a history of high-risk PCa, with follow-up of 1-189.3 months [130],
suggested that testosterone therapy did not increase the risk of biochemical recurrence, but the available
evidence is poor, limiting data interpretation [130]. Similar considerations can be derived from another, larger
meta-analysis of 21 studies [131]. It is important to recognise that most of the studies analysed included low-
risk patients with Gleason score < 8 [130].
In conclusion, recent literature does not support an increased risk of PCa in hypogonadal men
undergoing testosterone therapy. Although it is necessary to avoid testosterone administration in men with
advanced PCa, insufficient long-term prospective data on the safety of androgen administration in PCa
survivors [131], without recurrence should prompt caution in choosing to treat symptomatic hypogonadal
men in this setting. Specifically, patients should be fully counselled that the long-term effects of testosterone
therapy in this setting are still unknown and requires further investigation. If an occult PCa is not detected
before initiation of testosterone therapy, treatment may unmask the cancer detected by an early rise in PSA
over 6-9 months of therapy. Due to the lack of strong evidence-based data on safety, the possible use of
testosterone therapy in symptomatic hypogonadal men previously treated for PCa should be adequately
discussed with the patients and limited to low-risk individuals.

3.7.5 Cardiovascular Disease


Evidence suggests that hypogonadal men have an increased risk of CVD [132, 133]. Whether or not LOH is
a cause or a consequence of atherosclerosis has not been clearly determined. Late-onset hypogonadism is
associated with CV risk factors, including central obesity, insulin resistance and hyperglycaemia, dyslipidaemia
(elevated total cholesterol, LDL-cholesterol, triglycerides and low HDL-cholesterol), pro-thrombotic tendency

SEXUAL AND REPRODUCTIVE HEALTH - MARCH 2021 29


and chronic inflammatory state [133]. Atherosclerosis is a chronic inflammatory disease, that releases pro-
inflammatory cytokines into the circulation, which are known to suppress testosterone release from the HPG
axis. Evidence from RCTs of testosterone therapy in men with MetS and/or T2DM demonstrates some benefit
in CV risk, including reduced central adiposity, insulin resistance, total cholesterol and LDL-cholesterol and
suppression of circulating cytokines [14, 23-25, 29, 133]. However, due to the equivocal nature of these
studies, testosterone therapy cannot be recommended for indications outside the specific symptoms.

Published data show that LOH is associated with an increase in all-cause and CVD-related mortality [12, 134-
137]. These studies are supported by a meta-analysis that concluded that hypogonadism is a risk factor for
cardiovascular morbidity [123] and mortality [138]. Importantly, men with low testosterone when compared
to eugonadal men with angiographically proven coronary disease have twice the risk of earlier death [133].
Longitudinal population studies have reported that men with testosterone in the upper quartile of the normal
range have a reduced number of CV events compared to men with testosterone in the lower three quartiles
[134]. Androgen deprivation therapy for PCa is linked to an increased risk of CVD and sudden death [139].
Conversely, two long-term epidemiological studies have reported reduced CV events in men with high normal
serum testosterone levels [140, 141]. Erectile dysfunction is independently associated with CVD and may be
the first clinical presentation in men with atherosclerosis.

The knowledge that men with hypogonadism and/or ED may have underlying CVD should prompt individual
assessment of their CV risk profile. Individual risk factors (e.g., lifestyle, diet, exercise, smoking, hypertension,
diabetes and dyslipidaemia) should be assessed and treated in men with pre-existing CVD and in patients
receiving androgen deprivation therapy. Cardiovascular risk reduction can be managed by primary care
clinicians, but patients should be appropriately counselled by clinicians active in prescribing testosterone
therapy [80]. If appropriate, they could be referred to cardiologists for risk stratification and treatment of
co-morbidity.

No RCTs have provided a clear answer on whether testosterone therapy affects CV outcomes. The TTrial
(n=790) in older men [142], the TIMES2 (n=220) [24] and the BLAST studies in men with MetS and T2DM
and the pre-frail and frail study in elderly men - all of 1-year duration - did not reveal any increase in Major
Adverse Cardiovascular Events (MACE) [24, 27, 142, 143]. In this context, MACE is defined as the composite
of CV death, non-fatal acute myocardial infarction, acute coronary syndromes, stroke and cardiac failure.
Randomised controlled trials between 3 and 12 months in men with known heart disease treated with
testosterone have not found an increase in MACE, but have reported improvement in cardiac ischaemia,
angina and functional exercise capacity [144-146]. The European Medicines Agency has stated that ‘The
Co-ordination Group for Mutual recognition and Decentralisation Procedures-Human (CMDh), a regulatory
body representing EU Member States, has agreed by consensus that there is no consistent evidence of
an increased risk of heart problems with testosterone in men who lack the hormone (a condition known as
hypogonadism). However, the product information is to be updated in line with the most current available
evidence on safety, and with warnings that the lack of testosterone should be confirmed by signs and
symptoms and laboratory tests before treating men with these drugs [147].

As a whole, as for MACE, current available data from interventional studies suggest that there is no increased
risk with up to 3 years of testosterone therapy [148-151]. The weight of the currently published evidence has
reported that testosterone therapy in men with diagnosed hypogonadism has neutral or beneficial actions on
MACE in patients with normalised testosterone levels. The findings could be considered sufficiently reliable
for a 3-year course of testosterone therapy, after which no available study can exclude further or long-term CV
events [152, 153].

3.7.5.1 Cardiac Failure


Testosterone treatment is contraindicated in men with severe chronic cardiac failure because fluid retention
may lead to exacerbation of the condition. Some studies including one of 12 months’ duration have shown
that men with moderate chronic cardiac failure may benefit from low doses of testosterone, which achieve
mid-normal range testosterone levels [145, 154, 155]. If a decision is made to treat hypogonadism in men with
chronic cardiac failure, it is essential that the patient is followed up carefully with clinical assessment and both
testosterone and haematocrit measurements on a regular basis. An interesting observation is that untreated
hypogonadism increased the re-admission and mortality rate in men with heart failure [156].

3.7.6 Erythrocytosis
An elevated haematocrit is the most common adverse effect of testosterone therapy. Stimulation of
erythropoiesis is a normal biological action that enhances delivery of oxygen to testosterone-sensitive tissues

30 SEXUAL AND REPRODUCTIVE HEALTH - MARCH 2021


(e.g., striated, smooth and cardiac muscle). Any elevation above the normal range for haematocrit usually
becomes evident between 3 and 12 months after testosterone therapy initiation. However, polycythaemia
can also occur after any subsequent increase in testosterone dose, switching from topical to parenteral
administration and, development of co-morbidity, which can be linked to an increase in haematocrit (e.g.,
respiratory or haematological diseases).
There is no evidence that an increase of haematocrit up to and including 54% causes any adverse
effects. If the haematocrit exceeds 54% there is a testosterone independent, but weak associated rise in CV
events and mortality [77, 157-159]. Any relationship is complex as these studies were based on patients with
any cause of secondary polycythaemia, which included smoking and respiratory diseases. There have been no
specific studies in men with only testosterone-induced erythrocytosis.

Three large studies have not shown any evidence that testosterone therapy is associated with an increased risk
of venous thromboembolism [160, 161]. However, one study showed that an increased risk peaked at 6 months
after initiation of testosterone therapy, then declined over the subsequent period [162]. No study reported
whether there was monitoring of haematocrit, testosterone and/or E2 levels. High endogenous testosterone
or E2 levels are not associated with a greater risk of venous thromboembolism [163]. In one study venous
thromboembolism was reported in 42 cases and 40 of these had diagnosis of an underlying thrombophilia
(including factor V Leiden deficiency, prothrombin mutations and homocysteinuria) [164]. In a RCT of
testosterone therapy in men with chronic stable angina there were no adverse effects on coagulation, by
assessment of tissue plasminogen activator or plasminogen activator inhibitor-1 enzyme activity or fibrinogen
levels [165]. A meta-analysis of RCTs of testosterone therapy reported that venous thromboembolism was
frequently related to underlying undiagnosed thrombophilia-hypofibrinolysis disorders [76].

With testosterone therapy an elevated haematocrit is more likely to occur if the baseline level is toward the
upper limit of normal prior to initiation. Added risks for raised haematocrit on testosterone therapy include
smoking or respiratory conditions at baseline. Higher haematocrit is more common with parenteral rather
than topical formulations. In men with pre-existing CVD extra caution is advised with a definitive diagnosis of
hypogonadism before initiating testosterone therapy and monitoring of testosterone as well as haematocrit
during treatment.
Elevated haematocrit in the absence of co-morbidity or acute CV or venous thromboembolism
can be managed by a reduction in testosterone dose, change in formulation or if the elevated haematocrit is
very high by venesection (500 mL), even repeated if necessary, with usually no need to stop the testosterone
therapy.

3.7.7 Obstructive Sleep Apnoea


There is also no evidence that testosterone therapy can result in onset or worsening of sleep apnoea.
Combined therapy with Continuous Positive Airway Pressure (CPAP) and testosterone gel was more effective
than CPAP alone in the treatment of obstructive sleep apnoea [166]. In one RCT, testosterone therapy in men
with severe sleep apnoea reported a reduction in oxygen saturation index and nocturnal hypoxaemia after 7
weeks of therapy compared to placebo, but this change was not evident after 18 weeks’ treatment and there
was no association with baseline testosterone levels [167].

3.7.8 Follow up
Testosterone therapy alleviates symptoms and signs of hypogonadism in men in a specific time-dependent
manner. The TTrials clearly showed that testosterone therapy improved sexual symptoms as early as 3 months
after initiation [86]. Similar results have been derived from meta-analyses [53, 76]. Hence, the first evaluation
should be planned after 3 months of treatment. Further evaluation may be scheduled at 6 months or 12
months, according to patient characteristics, as well as results of biochemical testing (see below). Table 6
summarises the clinical and biochemical parameters that should be monitored during testosterone therapy.

Trials were designed to maintain the serum testosterone concentration within the normal range for young men
(280–873 ng/dL or 9.6-30 nmol/L) [86]. This approach resulted in a good benefit/risk ratio. A similar approach
could be considered during follow-up. The correct timing for evaluation of testosterone levels varies according
to the type of preparation used (Table 5). Testosterone is involved in the regulation of erythropoiesis [108]
and prostate growth [75], hence evaluation of PSA and haematocrit should be mandatory before and during
testosterone therapy. However, it is important to recognise that the risk of PCa in men aged < 40 years is low.
Similarly, the mortality risk for PCa in men aged > 70 years is not been considered high enough to warrant
monitoring in the general population [168]. Hence, any screening for PCa through determination of PSA and
DRE in men aged < 40 or > 70 years during testosterone therapy should be discussed with the patients.
Baseline and, at least, annually glyco-metabolic profile evaluation may be a reasonable

SEXUAL AND REPRODUCTIVE HEALTH - MARCH 2021 31


consideration, particularly in the management of functional hypogonadism. Testosterone therapy may be
beneficial for hypogonadal men with low or moderate fracture risk [96]; therefore, dual energy X-ray
absorptiometry (DEXA) bone scan may also be considered at baseline and 18-24 months following
testosterone therapy, particularly in patients with more severe hypogonadism [96].
Digital rectal examination may detect prostate abnormalities that can be present even in men with
normal PSA values. Hence, DRE is mandatory in all men at baseline and during testosterone therapy.
The decision to stop testosterone therapy or to perform prostate biopsy due to PSA increase or
prostate abnormalities should be based on local PCa guidelines. There is a large consensus that any increase
of haematocrit > 54% during testosterone therapy requires therapy withdrawal and phlebotomy to avoid
potential adverse effects including venous-thromboembolism and CVD, especially in high-risk individuals. In
patients with lower risk of relevant clinical sequelae, the situation can be alternatively managed by reducing
testosterone dose and switching formulation along with venesection. A positive family history of venous-
thromboembolism should be carefully investigated and the patient counselled with regard to testosterone
therapy to avoid/prevent thrombophilia-hypofibrinolysis [76]. Finally, caution should be exercised in men with
pre-existing CVD or at higher risk of CVD.

Table 6: Clinical and biochemical parameters to be checked during testosterone therapy

Parameters Year 1 of treatment After year 1 of treatment


Baseline 3 months 6/12 months Annually 18-24 months
Clinical
Symptoms X X X X
Body Mass Index
Waist circumference X X X X
Digital rectal examination X X X X
Blood pressure X X X X
Biochemistry
PSA (ng/mL) X X X X
Haematocrit (%) X X X X
Testosterone X X X X
Lipid and glycaemic profile X X X
Instrumental
DEXA X X

3.7.9 Summary of evidence and recommendations on risk factors in testosterone treatment

Summary of evidence
Testosterone therapy is contraindicated in men with secondary hypogonadism who desire fertility.
Testosterone therapy is contraindicated in men with active prostate cancer or breast cancer.
Testosterone therapy does not increase the risk of prostate cancer, but long-term prospective follow-up data
are required to validate this statement.
The effect of testosterone therapy in men with severe lower-urinary tract symptoms is limited, as these
patients are usually excluded from RCTs.
There is no substantive evidence that testosterone therapy, when replaced to normal levels, results in the
development of major adverse cardiovascular events.
There is no evidence of a relationship between testosterone therapy and mild, moderate or CPAP-treated
severe sleep apnoea.

Recommendations Strength rating


Fully counsel symptomatic hypogonadal men who have been surgically treated for localised Weak
prostate cancer (PCa) and who are currently without evidence of active disease considering
testosterone therapy, emphasising the benefits and lack of sufficient safety data on long-
term follow-up.
Restrict treatment to patients with a low risk for recurrent PCa (i.e., pre-operative PSA Weak
< 10 ng/mL; Gleason score < 7 (International Society for Urological Pathology grade 1);
cT1-2a)* and treatment should start after at least 1 year follow-up with PSA level < 0.01
ng/mL.
Safety data on the use of testosterone therapy in men treated for breast cancer are unknown. Strong

32 SEXUAL AND REPRODUCTIVE HEALTH - MARCH 2021


Assess for cardiovascular risk factors before commencing testosterone therapy. Strong
Assess men with known cardiovascular disease (CVD) for cardiovascular symptoms before Strong
testosterone therapy and with close clinical assessment and evaluation during treatment.
Treat men with hypogonadism and pre-existing CVD, venous-thromboembolism or chronic Weak
cardiac failure, who require testosterone therapy with caution, by careful clinical monitoring
and regular measurement of haematocrit (not exceeding 54%) and testosterone levels.
Exclude a family history of venous-thromboembolism before starting testosterone therapy. Strong
Monitor testosterone, haematocrit at 3, 6 and 12 months after testosterone therapy Strong
initiation, and thereafter annually. A haematocrit > 54% should require testosterone therapy
withdrawal and phlebotomy. Re-introduce testosterone therapy at a lower dose once the
haematocrit has normalised and consider switching to topical testosterone preparations.
*As for EAU risk groups for biochemical recurrence of localised or locally advanced prostate cancer (see EAU
guidelines 2021 on prostate cancer)

4. EPIDEMIOLOGY AND PREVALENCE OF


SEXUAL DYSFUNCTION AND DISORDERS OF
MALE REPRODUCTIVE HEALTH
4.1 Erectile dysfunction
Epidemiological data have shown a high prevalence and incidence of ED worldwide [169]. Among others,
the Massachusetts Male Aging Study (MMAS) [170] reported an overall prevalence of 52% ED in non-
institutionalised men aged 40-70 years in the Boston area; specific prevalence for minimal, moderate, and
complete ED was 17.2%, 25.2%, and 9.6%, respectively. In the Cologne study of men aged 30-80 years, the
prevalence of ED was 19.2%, with a steep age-related increase from 2.3% to 53.4% [171]. The incidence rate
of ED (new cases per 1,000 men annually) was 26 in the long-term data from the MMAS study [172] and 19.2
(mean follow-up of 4.2 years) in a Dutch study [173]. In a cross-sectional real-life study among men seeking
first medical help for new-onset ED, one in four patients was younger than 40 years, with almost 50% of the
young men complaining of severe ED [174]. Differences among these studies can be explained by differences
in methodology, ages, and socio-economic and cultural status of the populations studied. The prevalence rates
of ED studies are reported in Table 7.

4.2 Premature ejaculation


As evidenced by the highly discrepant prevalence rates reported in Table 8 [175], the method of recruitment for
study participation, method of data collection and operational criteria can all greatly affect reported prevalence
rates of premature ejaculation (PE). The major problem in assessing the prevalence of PE was the lack of
a universally recognised definition at the time the surveys were conducted [176]. Vague definitions without
specific operational criteria, different manners of sampling, and non-standardised data acquisition have led to
heterogeneity in estimated prevalence [176-180]. The highest prevalence rate of 31% (men aged 18-59 years)
was found by the National Health and Social Life Survey (NHSLS), which determines adult sexual behaviour in
the USA [181]. Prevalence rates were 30% (18-29 years), 32% (30-39 years), 28% (40-49 years) and 55% (50-59
years). It is, however, unlikely that the PE prevalence is as high as 20-30% based on the relatively low number of
men who seek medical help for PE. These high prevalence rates may be a result of the dichotomous scale (yes/
no) in a single question asking if ejaculation occurred too early, as the prevalence rates in European studies have
been significantly lower [182]. Two separate observational, cross-sectional surveys from different continents found
that overall prevalence of PE was 19.8 and 25.80%, respectively [183, 184]. Further stratifying these complaints
into the classifications defined by Waldinger et al. [185], rates of lifelong PE were 2.3 and 3.18%, acquired PE 3.9
and 4.48%, variable PE 8.5 and 11.38% and subjective PE 5.1 and 6.4% [183, 184]. Both studies showed that
men with acquired PE were more likely to seek treatment compared to men with lifelong PE. Treatment-seeking
behaviour may have contributed to errors in the previously reported rates of PE, as it is possible that men with
lifelong PE came to terms with their problem and did not seek treatment. The additional psychological burden
of a new change in ejaculatory latency in acquired PE may have prompted more frequent treatment seeking
[186]. Thus, it is likely that there is disparity between the incidence of the various PE sub-types in the general
community and in men actively seeking treatment for PE [187, 188]. This disparity could be a further barrier to
understanding the true incidence of each sub-type of PE. An approximately 5% prevalence of acquired PE and
lifelong PE in the general population is consistent with epidemiological data indicating that around 5% of the
population have an ejaculation latency of < 2 minutes [189].

SEXUAL AND REPRODUCTIVE HEALTH - MARCH 2021 33


4.3 Other ejaculatory disorders
4.3.1 Delayed ejaculation
Due to its rarity and uncertain definitions, the epidemiology of delayed ejaculation (DE) is not clear [190]. However,
several well-designed epidemiological studies have revealed that its prevalence is around 3% among sexually
active men [181, 191]. According to data from the NHSLS, 7.78% of a national probability sample of 1,246 men
aged 18-59 years reported inability achieving climax or ejaculation [181]. In a similar stratified national probability
sample survey completed over 6 months among 11,161 men and women aged 16-44 years in Britain, 0.7%
of men reported inability to reach orgasm [192]. In an international survey of sexual problems among 13,618
men aged 40–80 years from 29 countries, 1.1-2.8% of men reported that they frequently experience inability to
reach orgasm [193]. Another study conducted in the United States, in a national probability sample of 1,455 men
aged 57-85 years, 20% of men reported inability to climax and 73% reported that they were bothered by this
problem. [194]. Considering the findings of these epidemiological studies and their clinical experiences, some
urologists and sex therapists have postulated that the prevalence of DE may be higher among older men [195-
197]. Similar to the general population, the prevalence of men with DE is low among patients who seek treatment
for their sexual problems. An Indian study that evaluated the data on 1,000 consecutive patients with sexual
disorders who attended a psychosexual clinic demonstrated that the prevalence of DE was 0.6% and it was more
frequent in elderly people with diabetes [198]. Nazareth et al. [199] evaluated the prevalence of International
Classification of Diseases 10th edition (ICD-10) diagnosed sexual dysfunctions among 447 men attending 13
general practices in London, UK and found that 2.5% of the men reported inhibited orgasm during intercourse.
Similar to PE, there are distinctions among lifelong, acquired and situational DE [200]. Although the evidence is
limited, the prevalence of lifelong and acquired DE is estimated at 1 and 4%, respectively [201].

4.3.2 Anejaculation and Anorgasmia


Establishing the exact prevalence of anejaculation and anorgasmia is difficult since many men cannot
distinguish between ejaculation and orgasm. The rarity of these clinical conditions further hampers the
attempts to conduct epidemiological studies. In a report from the USA, 8% of men reported unsuccessfully
achieving orgasm during the past year [181].
According to Kinsey et al. [202], 0.14% of the general population have anejaculation. The most
common causes of anejaculation were spinal cord injury, diabetes mellitus and multiple sclerosis. Especially
in most cases of spinal cord injury, medical assistance is the only way to ejaculate. While masturbation leads
to the lowest rates of ejaculation, higher response rates can be obtained with penile vibratory stimulation or
acetylcholine esterase inhibitors followed by masturbation in patients with spinal cord injury [203].

4.3.3 Retrograde ejaculation


Similar to anejaculation, it is difficult to estimate the true incidence of retrograde ejaculation (RE). Although
RE is generally reported in 0.3-2% of patients attending fertility clinics [204], diabetes may increase these
rates by leading to autonomic neuropathy. Autonomic neuropathy results in ED and ejaculatory dysfunctions
ranging from DE to RE and anejaculation, depending on the degree of sympathetic autonomic neuropathy
involved [205]. In 54 diabetic patients with sexual dysfunction, RE was observed with a 6% incidence [206]. In
a controlled trial, RE was observed in 34.6% of diabetic men [207]. A more recent trial reported the rate of RE
among 57 type-1-diabetes mellitus patients (aged 18-50 years) was at least 8.8% [208]. Retrograde ejaculation
was also reported in studies of patients who had undergone transurethral resection of prostate (TURP) or
open prostatectomy due to disrupted bladder neck integrity. A study of the effect of prostatectomy on QoL in
5,276 men after TURP, found that 68% reported post-surgical RE [209]. However, with the development of less
invasive techniques, the incidence of RE decreases following the surgical treatment of LUTS [210-214].

4.3.4 Painful ejaculation


Painful ejaculation is a common but poorly understood clinical phenomenon, which is associated with
sexual dysfunction. Several studies demonstrated its prevalence to range between 1 and 10% in the general
population [215-217]; however, it may increase to 30-75% among men with chronic prostatitis/chronic pelvic
pain syndrome (CP/CPPS) [218-222]. It should be noted that the design of most of these studies was not
scientifically sound and the condition was probably under-reported due to the lack of an evidence-based
definition and well-defined prognostic criteria.

4.3.5 Haemospermia
The exact incidence and prevalence of haemospermia are difficult to elucidate due to a number of factors
including its covert presentation, usually self-limiting nature and patient embarrassment. The symptom
represents 1-1.5% of all urological referrals and occurs in all age groups, with a mean age of 37 years [223,
224]. In a PCa screening study of 26,126 men, aged ≥ 50 years or older than 40 with a history of PCa or of
black ethnicity, haemospermia was found in 0.5% on entry to the trial [225].

34 SEXUAL AND REPRODUCTIVE HEALTH - MARCH 2021


4.4 Low sexual desire
The global prevalence of low sexual desire in men is 3-28% [193, 226, 227]. Low solitary and dyadic sexual
desires, have been reported in 68% and 14% of men, respectively [228]. Also, low sexual desire has been
observed as a common complaint in gay men, with a prevalence of 19-57% [229, 230]. Despite its relationship
with age, low sexual desire has been reported among young men (18-29 years), with prevalence of 6-19% [181,
231, 232].

Table 7: Prevalence rates of erectile dysfunction [169]

Date Authors Population Response Age Measurement Principal findings Correlates


rate (years) technique
1993 Solstad 439 men; 81% 51 Interview Overall, 4% of men Not reported
et al. random and self- had ED as assessed
[233, 234] population administered by questionnaire,
sample questionnaire interviews identified a
(Denmark) higher frequency of
ED (40%)
1994 Feldman 1,290 men; 40% 40-70 Self- Overall, 52% of men Age
et al. [170] random administered had ED
*MMAS population questionnaire 17.2% of men had
sample minimal ED
(United States) 25.2% of men had
moderate ED
9.6% of men had
complete ED
1995 Panser et al. 2,155 men; 55% 40-79 Self- 1% ED in men aged
[235] random administered 40-49 years
population questionnaire 6% ED in men aged
sample 50-59 years
(United States) 22% ED in men aged
60-69 years
44% ED in men aged
70-79 years
1996 Helgason 319 men; 73% 50-80 Self- 3% ED in men aged Age, Prostate cancer,
et al. [236] random administered 50–59 years Diabetes, Myocardial
population questionnaire 24% ED in men aged infarction, Diuretic
sample 60–69 years use, Warfarin use, H2
(Sweden) 49% ED in men aged receptor blocker use
70–80 years
1996 MacFarlane 1,734 men; 86% 50-80 Self- 20% ED in men aged Age
et al. [237] random administered 50–59 years
population questionnaire 33% ED in men aged
sample 60–69 years
(France) 38% ED in men aged
70–80 years
1996 Fugl-Meyer 1,288 men; 52% 18-74 Structured Overall, 5% of men Age
[226] random interview had ED
population 3% ED in men aged
sample men 18–24 years
(Sweden) 2% ED in men aged
25–34 years
2% ED in men aged
35–49 years
7% ED in men aged
50–65 years
24% ED in men aged
66–74 years

SEXUAL AND REPRODUCTIVE HEALTH - MARCH 2021 35


1999 Laumann 1,244 men; 70% 18-59 Structured Overall, 10% of men Age, Race, Emotional
et al. [181] random interview had ED (moderate stress, Urinary
*NHSLS population plus severe) symptoms, Poor
sample 7% ED in men aged health, Low income
(United States) 18–29 years
9% ED in men aged
30–39 years
11% ED in men aged
40–49 years
18% ED in men aged
50–59 years
1999 Pinnock 427 men; 69.8% > 40 Self- 6% ED in men aged Age, Hyper-
et al. [238] random administered 40–49 years cholesterolemia,
population questionnaire 12% ED in men aged
sample 50–59 years
(Australia) 41% ED in men aged
60–69 years
63% ED in men aged
70–79 years
81% ED in men aged
80+ years
2000 Braun 8,000 men 56% 30-80 Self- Prevalence of ED Age, Hypertension,
et al. [171] administered was 19.2% Diabetes, Pelvic
(COLOGNE questionnaire surgery, LUTS
Study) by mail
(Cologne ED
Questionnaire)
2001 Moreira 1,170 men; 91% >18 Self- Overall, 14.7% Age, Education,
et al. [239] attending administered of men had ED Racial origin,
public places questionnaire (moderate plus Diabetes,
(heavy bias severe); Hypertension,
toward 9.4% ED in men Depression
younger men) aged 18–39 years
(Brazil) 15.5% ED in men
aged 40–49 years
22.1% ED in men
aged 50–59 years
37% ED in men aged
60–69 years
39.6% ED in men
aged >70 years
2001 Meuleman 1,233 men; 70% 40-79 Self- Overall, 13% of men Age
et al. [240] random administered had ED
population questionnaire 6% ED in men aged
sample (the 40–49 years
Netherlands) 9% ED in men aged
50–59 years
22% ED in men aged
60–69 years
38% ED in men aged
70–79 years
2001 Blanker 1,688 men; 50% 50-75 Self- 3% ED in men aged Age, Smoking,
et al. random administered 50–54 years Obesity, LUTS,
[216, 241] population questionnaire 5% ED in men aged COPD, Treatment for
sample (the 55–59 years CV disease
Netherlands) 11% ED in men aged
60–64 years
19% ED in men aged
65–69 years
26% ED in men aged
70–78 years

36 SEXUAL AND REPRODUCTIVE HEALTH - MARCH 2021


2001 Martin- 2,476 men; 75% 25-70 Self- Overall, 12.1% of Age, Hypertension,
Morales random administered men had ED (single Diabetes, Cardiac
et al. [242] population questionnaire question) and 18.9% disease, Pulmonary
sample (Spain) and single for questionnaire disease, Circulatory
question According to single disease, Rheumatic
question: disease, High
3.9% ED in men cholesterol, Prostatic
aged 25–39 years disease, Allergy,
6.3% ED in men Medication “for
aged 40–49 years nerves”, Sleeping
15.9% ED in men tablets, Heavy
aged 50–59 years smoking, Alcohol
32.2% ED in men abuse
aged 60–70 years
IIEF identified milder
ED, and single
question identified
more moderate and
severe ED
2002 Moreira 602 men; 92% 40-70 Interview Overall, 14.4% Age, Marital
et al. [243] random of men had ED status, Diabetes,
population (moderate or severe) Depression, IPSS,
sample (Brazil) 9.9% ED in men Decreased physical
aged 40–49 years activity
11.8% ED in men
aged 50–59 years
31.7% ED in men
aged 60–69 years
2002 Moreira 342 men; 47.6% 40-70 Self- Overall, 12.0% Age, Diabetes,
et al. [244] random administered of men had ED Hypertension, Heavy
population questionnaire (moderate or severe) smoking
sample (Brazil) 3.5% ED in men
aged 40–49 years
16.7% ED in men
aged 50–59 years
39.6% ED in men
aged 60–69 years
2002 Morillo et al. 1,963 men; 82% > 40 Standardised Overall, 19.8% Age, Diabetes,
[245] random questionnaire of men had ED Hypertension, BPH
population (moderate or severe)
sample
(Columbia,
Venezuela and
Ecuador)
2003 Richters 8,517 men; 69.4% 16-59 Computer- Overall, 9.5% of men Age
et al. [246] random assisted had ED
population telephone 4.3% ED in men
sample interview aged 16–19 years
(Australia) 4.5% ED in men
aged 20–29 years
5.1% ED in men
aged 30–39 years
12.5% ED in men
aged 40–49 years
19.2% ED in men
aged 50–59 years

SEXUAL AND REPRODUCTIVE HEALTH - MARCH 2021 37


2003 Rosen et al. 12,815 men; 36.8% 50-80 Self- According to DAN- Age, LUTS, Diabetes,
[247] random administered PSS: Overall, 48.9% Hypertension,
population questionnaire of men had ED Cardiac disease,
sample (IIEF and DAN- 30.8% ED in men Hyperlipidemia,
(multinational: PSS) aged 50-59 years Tobacco use
United States, 55.1% ED in men
United aged 60-69 years
Kingdom, 76% ED in men aged
France, 70-80 years
Germany, the
Netherlands,
Italy, Spain)
2004 Rosen et al. 27,839 men US: 45%; 20-75 Random digit Overall prevalence Age, Hypertension,
[248] random UK: 48%; dialing and of ED in the MALES Heart trouble
*MALES population Germany: interviewed sample was 16% or angina, High
sample 45%; via computer- cholesterol, Diabetes,
(multinational: France: assisted Depression or anxiety
United 48%; telephone
States, United Italy: 53%; interviewing. A
Kingdom, Spain: standardised
Germany, 50%; questionnaire
France, Italy, Mexico:
Spain, Mexico, 55% and
and Brazil) Brazil:
51%.
2004 Shiri et al. 2,198 men; 70% 50,60, Self- 48% of men had Age, Diabetes,
[249] stratified and 70 administered minimal ED Hypertension,
birth cohort at first questionnaire 15.2% of men had Heart disease,
(Finland) survey at two separate moderate ED Cerebrovascular
55, 65, time points, 13.2% of men had disease,Smoking
and 5 years apart complete ED
75 at
second
survey
2005 Laumann 13,750 men; 19% 40-80 Telephone Overall: Age
et al. [193] random survey (random In Northern Europe,
*GSSAB population dialed digit) 13.3% had ED
sample (world) In Southern Europe,
12.9% had ED
In non-European
West,
20.6% had ED
In Central/South
America,
13.7% had ED
In Middle East,
14.1% had ED
In East Asia,
13.3% had ED
In Southeast Asia,
28.1% had ED
2005 Moreira 750 men; 23% 40-80 Telephone Overall, 12.7% had Age
et al. [250] random survey (random ED
population digit dialing)
sample (Spain)
2005 Moreira 750 men; 17.4% 40-80 Telephone Overall, 7.9% had ED Age
et al. [251] random survey (random
population digit dialing)
sample
(Germany)

38 SEXUAL AND REPRODUCTIVE HEALTH - MARCH 2021


2005 Moreira 471 men; 18% 40-80 Telephone Overall, 13.1% of Age, Depression
Junior et al. random survey (random men had ED
[252] population digit dialing)
sample (Brazil)
2006 Brock et al. 500 men; 9.7% 40-80 Telephone Overall, 16% of men Age, Depression,
[253] random survey (random had ED Diabetes
population digit dialing)
sample
(Canada)
2007 De Almeida 2,000 men; Not >20 Standardised Overall, 1.7% of men Age
Claro et al. random reported interview with had ED
[254] population self-reported 0.2% ED in men
study (Brazil) questionnaire aged 20-30 years
(IIEF) 0.22% ED in men
aged 31-40 years
1.0% ED in men
aged 41-50 years
2.8% ED in men
aged 51-60 years
7.0% ED in men
aged > 61 years
2007 Ahn et al. 1,570 men; Not 40-79 Self- Overall, 13.4% had Age, Single
[255] geographically reported administered self-reported ED status, Low
stratified questionnaire ED prevalence as income, Diabetes,
random (IIEF-5) defined by IIEF-5 Hypertension,
population score less than 17 Hyperlipidemia,
study was 32.4% Heart disease,
According to single Musculoskeletal
question: disorders, Alcohol,
4.2% ED in men Depression, Coffee
aged 40-49 years intake
13.0% ED in men
aged 50-59 years
30.1% ED in men
aged 60-69 years
41.1% ED in men
aged 70-79 years
2008 Moreira 750 men; 16.9% 40-80 Telephone Overall, 32% of men Age
et al. [256] random survey (random had ED
population digit dialing)
sample
(Australia)
2008 Chew et al. 1,580 men; 37.3% >20 Postal 15.7% ED in men Age, Marital status
[257] random survey Self- aged 20-29 years
population administered 8.7% ED in men
sample questionnaire aged 30-39 years
(Australia) (IIEF-5) 12.9% ED in men
aged 40-49 years
31.6% ED in men
aged 50-59 years
52.4% ED in men
aged 60-69 years
69.4% ED in men
aged 70-79 years
68.2% ED in men
aged > 80 years

SEXUAL AND REPRODUCTIVE HEALTH - MARCH 2021 39


2008 Teles et al. 3,067 men; 81.3% 40-69 Self- Overall, 48.1% of Age, Diabetes,
[258] random administered men had ED Cardiac insufficiency,
population questionnaire, 29% ED in men aged Psychiatric illness
sample including IIEF 40-49 years
(Portugal) 50% ED in men aged
50-59 years
74% ED in men aged
60-69 years
2008 Moreira 750 men; 17% 40-80 Telephone Overall, 17.8% of Age
et al. [259] random survey (random men had ED
population digit dialing)
sample (United
Kingdom)
2009 Laumann 742 men; 9% 40-80 Telephone Overall, 22.5% of Age, Depression
et al. [260] random survey (random men had ED
population digit dialing)
sample (United
States)
2009 Buvat et al. 750 men; 23.8% 40-80 Telephone Overall, 15% of men Age
[261] random survey (random had ED
population digit dialing)
sample
(France)
2010 Corona 3,369 men; 40% 40-80 Self- Overall, 30% of men Age, Depression,
et al. [262] random administered had ED LUTS, Cardiovascular
population questionnaire 6% ED in men aged disease, Diabetes,
study (Europe: 40-49 years Obesity
Italy, Belgium, 19% ED in men aged
United 50-59 years
Kingdom, 38% ED in men aged
Spain, Poland, 60-69 years
Hungary, 64% ED in men 70
Estonia) and over
2016 Oyelade 241 men; 99% 30-80 Self- General prevalence Age, Hypertension,
et al. [263] random administered of ED was 58.9% Use of anti-
sampling questionnaire hypertensive drugs,
cross- (IIEF-5) Diabetes mellitus,
sectional Heart disease
population
based survey
(Nigeria)
2017 Cayan et al. 2,760 men; Non- > 40 Self- Prevalence of ED Age, Diabetes,
[264] random reported administered was calculated as Hypertension,
population questionnaire 33% among all men Atherosclerosis,
study (Turkey) (IIEF-5) aged ≥ 40 years. Dyslipidaemia, LUTS,
ED prevalence Educational status,
rates were Monthly income
17% for 40-49
years,
35.5% for 50-59
years,
68.8% for 60-69
years, and
82.9% for ≥ 70 years
2017 Quilter et al. Randomly 30% 40-70 Self-reported Prevalence of ED was Age, Anxiety or
[265] selected erectile 42% (22% mild, 10% depression
age-stratified function (IIEF-5) mild to moderate,
population- and a single- 6% moderate, and
based sample question self- 4% severe)
of 2,000 men assessment
(New Zealand) tool.

40 SEXUAL AND REPRODUCTIVE HEALTH - MARCH 2021


2020 Calzo et al. 2,660 sexually 18-31 Self- Prevalence of mild Demographic (age;
[266] active men administered ED was 11.3% and marital status)
(USA) questionnaire moderate-to-severe Metabolic (body
(IIEF-5) ED was 2.9% mass index; waist
circumference;
history of diabetes,
hypertension,
hyperch
olesterolaemia)
Mental health
(depression, anxiety,
antidepressant,
tranquiliser use)
Four baseline studies estimating the prevalence of Erectile Dysfunction:
MMAS = the Massachusetts Male Aging Study; NHSLS = the National Health of Social Life Survey;
MALES = the multi-national men’s attitudes to life events and sexuality; GSSAB = Global Study of Sexual
Attitudes and Behaviours.
BPH = Benign Prostate Hyperplasia; COPD = Chronic Obstructive Pulmonary Disease; ED = Erectile
Dysfunction; IIEF = International Index of Erectile Function; IPSS = International Prostate Symptom Score;
LUTS = Lower urinary tract symptoms.

Table 8: The prevalence rates of premature ejaculation [175]

Date Authors Method of Data Method of Operational Prevalence Rate Number of


Collection Recruitment Criteria Men
1998 Dunn et al. Mail General practice Having difficulty 14% (past 3 mo) 617
[267] registers - random with ejaculating 31% (lifetime) 618
stratification prematurely
1999 Laumann et al. Interview NA Climaxing/ 31% 1,410
(NHSLS) [181] ejaculating too
rapidly during the
past 12 months
2002 Fugl-Meyer Interview Population register NA 9% 1,475
and Fugl-
Meyer [268]
2004 Rowland et al. Mailed Internet panel DSM IV 16.3% 1,158
[269] questionnaire
2004 Nolazco et al. Interview Invitation to Ejaculating fast or 28.3% 2,456
[270] outpatient clinic prematurely
2005 Laumann et al. Telephone- Random (systematic) Reaching climax 23.75% 13,618
[193] personal sampling too quickly during (4.26% frequently)
interview/mailed the past 12
questionnaires months
2005 Basile Fasolo Clinician-based Invitation to DSM IV 21.2% 12,558
et al. [271] outpatient clinic
2005 Stulhofer et al. Interview Stratified sampling Often ejaculating 9.5% 601
[272] in less than < 2
minutes
2007 Porst et al. Web-based Internet panel Control over 22.7% 12,133
(PEPA) [182] survey ejaculation,
Self-report distress
2008 Shindel et al. Questionnaire Male partners of Self-report 50% 73
[273] infertile couples premature
under evaluation ejaculation
2009 Brock et al. telephone Web-based survey DSM III 16% 3,816
[274] interview Control 26%
Distress 27%
2010 Traeen and Mailed Web interview + 27% 11,746 + 1,671
Stigum [232] questionnaire + Randomisation
internet

SEXUAL AND REPRODUCTIVE HEALTH - MARCH 2021 41


2010 Son et al. [275] Questionnaire Internet panel DSM IV 18.3% 600
(age < 60 years)
2010 Amidu et al. Questionnaire NA NA 64.7% 255
[276]
2010 Liang et al. NA NA ISSM 15.3% 1,127
[277]
2010 Park et al. Mailed Stratified sampling Suffering from PE 27.5% 2,037
[278] questionnaire
2010 Vakalopoulos One-on-one Population-based EED 58.43% 522
et al. [279] survey cohort ISSM lifelong PE 17.7%
2010 Hirshfeld et al. Web-based Online advertisement Climaxing/ 34% 7,001
[229] survey in the United States ejaculating too
and Canada rapidly during the
past 12 months
2011 Christensen Interview + Population register NA 7% 5,552
et al. [280] questionnaire (random)
2011 Serefoglu et al. Interview Stratified sampling Complaining 20.0% 2,593
[183] about ejaculating
prematurely
2011 Son et al. [281] Questionnaire Internet panel Estimated 10.5% 334
IELT < 5 mins,
inability to control
ejaculation,
distress
2011 Tang and Interview Primary care setting PEDT > 9 40.6% 207
Khoo [282]
2012 Mialon et al. Mailed Convenience Control over 11.4% 2,507
[283] questionnaire sampling (age 18-25 ejaculation
years) Distress
2012 Shaeer and Web-based Online advertisement Ejaculate before 83.7% 804
Shaeer [284] survey in Arabic countries the person wishes
to ejaculate at
least sometimes
2012 Shindel et al. Web-based Online advertisement PEDT > 9 8-12% 1,769
[285] survey targeted to MSM
+ distribution
of invitation to
organisations
catering to MSM
2012 McMahon Computer NA PEDT > 11 16% 4,997
et al. [286] assisted Self-Reported 13%
interviewing, (always/nearly-
online, or always)
in-person self-
completed
2012 Lotti et al. Interview Men seeking medical PEDT > 9 15.6% 244
[287] care for infertility
2013 Zhang et al. Interview Random stratified Self-reported 4.7% 728
[288] sample of married premature
men aged 30-60 ejaculation
2013 Lee et al. [289] Interview Stratified random PEDT > 11 11.3% 2,081
sampling Self-Reported 19.5%
IELT<1 min 3% 1,035
2013 Gao et al. Interview Random stratified Self-reported 25.8% 3,016
[184] sample of premature
monogamous ejaculation
heterosexual men in
China
2013 Hwang et al. Survey of Married heterosexual Estimated IELT 21.7% 290
[290] married couples couples in Korea < 2 minutes
PEDT > 11 12.1%

42 SEXUAL AND REPRODUCTIVE HEALTH - MARCH 2021


2013 Vansintejan Web Based Online and flyer IPE score < 50% 4% 72
et al. [291] survey advertisements to of total possible
Belgian men who
have sex with men
(Only HIV+ men in
this study)
2013 Shaeer et al. Web Based Targeting English- ISSM definition 6.3% 1133
[292] survey speaking men aged [179]
> 18 years, living PEDT 49.6%
most of their lives in Unfiltered self- 77.6%
the USA, regardless reported
of personal interests Filtered self- 14.4%
and web browsing reported
preferences
2016 Karabakan Interview (heavy Targeting police PEDT > 10 9.2% 1000
[293] bias toward academy students
younger men) aged 24-30 years
who applied for
routine urological
examination
2017 Gao et al. Field survey with Comprising men Self-estimated Lifelong PE 1239
[294] face-to-face aged 20-68 years IELT 10.98%
interviews in five cities in the Acquired PE
Anhui province 21.39%

DMS = Diagnostic and Statistical Manual of Mental Disorders; NA = not applicable; ISSM = International Society
for Sexual Medicine; PEDT = Premature Ejaculation Diagnostic Tool; IELT = intravaginal ejaculatory latency time;
IPE = Index of Premature Ejaculation; mo = months.

5. MANAGEMENT OF ERECTILE DYSFUNCTION


5.1 Definition and classification
Penile erection is a complex physiological process that involves integration of both neural and vascular events,
along with an adequate endocrine milieu. It involves arterial dilation, trabecular smooth muscle relaxation and
activation of the corporeal veno-occlusive mechanism [295]. Erectile dysfunction is defined as the persistent
inability to attain and maintain an erection sufficient to permit satisfactory sexual performance [296]. Erectile
dysfunction may affect psychosocial health and have a significant impact on the QoL of patients and their
partner’s [170, 297-299].
There is established evidence that the presence of ED increases the risk of future CV events
including myocardial infarction, cerebrovascular events, and all-cause mortality, with a trend towards an
increased risk of cardiovascular mortality [300]. Therefore, ED can be an early manifestation of coronary artery
and peripheral vascular disease and should not be regarded only as a QoL issue, but also as a potential
warning sign of CVD [301-304]. A cost analysis showed that screening men presenting with ED for CVD
represents a cost-effective intervention for secondary prevention of both CVD and ED, resulting in substantial
cost savings relative to identification of CVD at the time of presentation [305].
Erectile dysfunction is commonly classified into three groups based on aetiology: organic,
psychogenic and mixed ED. However, this classification should be used, with caution as most cases are
actually of mixed aetiology. It has therefore been suggested to use the terms “primary organic” or “primary
psychogenic”.

5.2 Risk factors


Erectile dysfunction is associated with unmodifiable and modifiable common risk factors including
age, diabetes mellitus, dyslipidaemia, hypertension, CVD, BMI/obesity/waist circumference, MetS,
hyperhomocysteinemia, lack of exercise, and smoking (a positive dose-response association between quantity
and duration of smoking has been demonstrated) [298, 302, 306-313]. Furthermore, an association between
ED status and pharmaco-therapeutic agents for CVD (e.g., thiazide diuretics and β-blockers, except nebivolol),
exert detrimental effects on erectile function, whereas newer drugs (i.e., angiotensin-converting enzyme-
inhibitors, angiotensin-receptor-blockers and calcium-channel-blockers) have neutral or even beneficial effects

SEXUAL AND REPRODUCTIVE HEALTH - MARCH 2021 43


[302, 314, 315]. Atrial fibrillation [316], hyper

You might also like