You are on page 1of 21

PRIMER

Tension-type headache
Sait Ashina   1,2 ✉, Dimos D. Mitsikostas3, Mi Ji Lee   4, Nooshin Yamani5,
Shuu-Jiun Wang   6, Roberta Messina   7, Håkan Ashina8, Dawn C. Buse9,
Patricia Pozo-Rosich10,11, Rigmor H. Jensen   8, Hans-Christoph Diener   12 and
Richard B. Lipton9,13
Abstract | Tension-type headache (TTH) is the most prevalent neurological disorder worldwide
and is characterized by recurrent headaches of mild to moderate intensity, bilateral location,
pressing or tightening quality, and no aggravation by routine physical activity. Diagnosis is based
on headache history and the exclusion of alternative diagnoses, with clinical criteria provided
by the International Classification of Headache Disorders, third edition. Although the biological
underpinnings remain unresolved, it seems likely that peripheral mechanisms are responsible for
the genesis of pain in TTH, whereas central sensitization may be involved in transformation from
episodic to chronic TTH. Pharmacological therapy is the mainstay of clinical management and
can be divided into acute and preventive treatments. Simple analgesics have evidence-based
effectiveness and are widely regarded as first-line medications for the acute treatment of TTH.
Preventive treatment should be considered in individuals with frequent episodic and chronic
TTH, and if simple analgesics are ineffective, poorly tolerated or contraindicated. Recommended
preventive treatments include amitriptyline, venlafaxine and mirtazapine, as well as some
selected non-pharmacological therapies. Despite the widespread prevalence and associated
disability of TTH, little progress has been made since the early 2000s owing to a lack of attention
and resource allocation by scientists, funding bodies and the pharmaceutical industry.

Photophobia
Tension-type headache (TTH) is the most prevalent which may have implications for pathophysiology and
Sensitivity to light. neurological disorder worldwide1–3. This primary head- treatment4. TTH management can be conceptualized
ache is characterized by recurrent headaches of mild to within a biopsychosocial framework, which acknowl-
Phonophobia moderate pain intensity. The headaches tend to be bilat- edges that biological, psychological and environmental
Sensitivity to sound.
eral and are sometimes described as of a ‘hatband’-like factors all have substantial and interactive roles in the
pattern, although pain can also occur in the forehead, development, maintenance and successful management
posterior head regions and neck. The headaches are also of this disorder.
characterized by the absence of accompanying symptoms This Primer provides an overview of the epidemiol-
such as photophobia, phonophobia and nausea4. ogy, pathophysiology and diagnosis of TTH. In addition,
Diagnosis of TTH is based on clinical criteria in recommended practices for clinical management and
the third edition of the International Classification of future research directions are discussed.
Headache Disorders (ICHD-3), which divides TTH
into three subtypes based on headache frequency: Epidemiology
infrequent episodic TTH (ETTH; headaches for In epidemiological studies, a diagnosis of TTH is estab-
<1 day/month on average or <12 days/year), frequent lished by meeting the diagnostic criteria of the ICHD;
ETTH (headaches for 1–14 days/month or ≥12 and this information may be captured by interviews con-
<180 days/year) and chronic TTH (CTTH; headaches ducted by trained clinicians, in-person lay interviews
for ≥15 days/month or ≥180 days/year)4 (Box 1). These based on semi-structured diagnostic questionnaires,
distinctions are important, as infrequent ETTH has validated telephone interviews or self-reported question-
minimal effects on individuals and rarely requires use naires. A two-step survey can also be used, in which a
✉e-mail: sashina@ of health-care services, whereas CTTH is associated screening tool survey is followed by a case-ascertainment
bidmc.harvard.edu with considerable disability and is often difficult to interview with clinicians. Self-reported medical diagno-
https://doi.org/10.1038/ treat1–3,5. The ICHD-3 also subclassifies TTH as asso- sis is not used as a case definition in epidemiological
s41572-021-00257-2 ciated or not associated with pericranial tenderness, studies owing to under-diagnosis.


NATURE REVIEWS | DISEASE PRIMERS | Article citation ID: (2021) 7:24 1

0123456789();:
Primer

Author addresses Data on the prevalence of TTH in children and


adolescents are relatively sparse, with available studies
1
BIDMC Comprehensive Headache Center, Department of Neurology and Department reporting a wide range in estimates, from 9.8% in Sweden
of Anesthesia, Critical Care and Pain Medicine, Harvard Medical School, Beth Israel to 72.3% in Brazil14. However, one study in rural Tanzania
Deaconess Medical Center, Boston, MA, USA. found a very low prevalence for TTH of 0.04% in individ-
2
Department of Clinical Medicine, Faculty of Health Sciences, University of Copenhagen,
uals 0–10 years of age and 1.3% in people 11–20 years of
Copenhagen, Denmark.
3
First Neurology Department, Aeginition Hospital, Medical School, National and age16. The 2017 GBD study found a global prevalence
Kapodistrian University of Athens, Athens, Greece. of TTH in individuals aged 5–9, 10–14 and 15–19 of
4
Department of Neurology, Samsung Medical Center, Sungkyunkwan University 12.1%, 35.7% and 35.8% in women/girls and 11.7%,
School of Medicine, Seoul, South Korea. 34.5%, and 34.0% in men/boys, respectively17. In com-
5
Headache Department, Iranian Center of Neurological Research, Neuroscience parison with 2007, there was an increase in prevalence
Institute, Tehran University of Medical Sciences, Tehran, Iran. of 0.5–1.0% among women/girls and of 0.9–1.3% among
6
Department of Neurology, Neurological Institute, Taipei Veterans General Hospital, men/boys at different age groups17.
Brain Research Center, National Yang-Ming University, Taipei, Taiwan. The global prevalence of CTTH is ~2–3% in most
7
Neuroimaging Research Unit and Neurology Unit, IRCCS San Raffaele Scientific population studies13,18–22 although values range from
Institute, Milan, Italy.
0.2% to 4.8%23. This variation might be due to differ-
8
Danish Headache Center, Department of Neurology, University of Copenhagen,
Rigshospitalet Glostrup, Glostrup, Denmark. ences in study samples based on age profiles, ethnicity
9
Department of Neurology, Albert Einstein College of Medicine, Bronx, NY, USA. and environment as well as data collection methodology.
10
Headache Unit, Neurology Department, Vall d’Hebron University Hospital, In most studies, the prevalence of CTTH is higher in
Barcelona, Spain. women/girls than in men/boys (range 3.0–3.9% versus
11
Headache and Neurological Pain Research Group, Vall d’Hebron Research 1.1–2.0%, respectively)19,21. CTTH is rare in young ado-
Institute, Departament de Medicina, Universitat Autònoma de Barcelona, lescents and the prevalence increases until age 39 and
Barcelona, Spain. then declines23. One large survey of 7,900 children in
12
Faculty of Medicine, Institute for Medical Informatics, Biometry and Epidemiology, Taiwan found the prevalence of CTTH as 1.0% in ado-
University Duisburg-Essen, Essen, Germany. lescents aged 12–14 years (ref.24). However, one survey in
13
Montefiore Headache Center, Bronx, NY, USA.
an elderly cohort in Kinmen, Taiwan, found a prevalence
of 2.7% for chronic daily headache (CDH)20,25.
Incidence The prevalence of infrequent ETTH and probable TTH
Data from the Global Burden of Disease (GBD) study were rarely surveyed in populations until recently. The
estimated that there were 882.4 million new cases of average estimated prevalence of infrequent ETTH is
TTH globally in 2017 (ref.2). In this study, there were 8.4% (range 3.3–17.7%), and the average estimated
~63.6 million new cases in the European Union and prevalence of probable TTH is 9.9% (range 0.4–27.3%)23.
~44.5 million new cases in the USA2,3. In a 12-year lon- Most of these studies showed a similar prevalence in
gitudinal study in Denmark, the incidence of frequent male and female subjects.
ETTH and CTTH was 14.2 per 1,000 person-years in
a random sample of the general population6. Incidence Risk factors
was 2.6 times higher in women than in men6. In a similar Risk factors for TTH have been studied in cross-sectional
study in Taiwan, in which adolescents aged 13–14 years and longitudinal studies. A positive association between
were followed for 12–24 months, the reported incidence educational level and prevalence of ETTH has been
was 3.9 per 1,000 person-years, with a 4.6-fold higher reported12, although an inverse relationship was found
incidence in female than in male subjects7. for CTTH, which was more prevalent in groups with
lower levels of education. In another study, TTH was
Prevalence associated with fatigue and, among women, a lack of
Prevalence estimates vary between studies and regions ability to relax after work26.
of the world owing, in part, to differences in case defi- In a longitudinal study from Denmark, using multi-
nitions and study methods as well as differences in variable models, young age, female sex, poor self-rated
Chronic daily headache ethnicity and other demographic characteristics of health, not being able to relax after work and sleep-
Primary or secondary populations1–3. Data from the GBD study estimated ing few hours per night were associated with incident
headache, defined as the that in 2017 there were ~2.33 billion people with TTH TTH6. Risk factors for incident CTTH in a study of
presence of a headache worldwide2,8, with 121.6 million in the USA3,9. Prevalence adolescents in Taiwan were baseline frequent headache
on ≥15 days/month for
at least 3 months, which may
increased by 31.7% in the USA from 1990 to 2017 (ref.3). (that is, headache for ≥7 days/month) and female sex7.
include chronic tension-type In the European Union, the prevalence of TTH was In an elderly cohort in Taiwan (≥65 years with mean age
headache, chronic migraine, ~173.7 million in 2017 (ref.2). 75.0 ± 7.9 years), overuse of acute headache medications,
new daily persistent headache, The estimated 1-year prevalence of TTH in popu- a history of migraine and depression severity were risk
medication overuse headache
lation studies was 10.8% in China10, 36.1% in Jordan11, factors for CTTH, and medication overuse was a poor
or other disorders with
headaches. 38.3% in the USA12, ~80% in European studies and prognostic factor for predicting the persistence of CDH
86.6% in Denmark13,14. In a population-based Danish in 2 years20.
Probable TTH twin registry, the 1-year prevalence of TTH among indi- In a 2-year follow-up study in adolescents with CDH,
Tension-type headache viduals aged 12–41 was 86%15. In another Danish study, mostly CTTH (61%), factors associated with persistence
(TTH) that is missing one of
the features required to fulfil
59% of individuals had infrequent ETTH. TTH has a of CDH in 2 years were baseline major depression and
all ICHD-3 criteria for a type female preponderance with a sex prevalence ratio of medication overuse27. In a Danish study, poor progno-
or subtype of TTH. 1.2:1, far lower than the 3:1 ratio that typifies migraine14. sis of TTH (defined as ≥180 days with TTH per year at

2 | Article citation ID: (2021) 7:24 www.nature.com/nrdp

0123456789();:
Primer

follow-up due either to increased frequency from ETTH work was estimated to be 820 days per 1,000 patients32.
at baseline into CTTH or to persistent CTTH) was asso- The corresponding amount for migraine headache was
ciated with baseline CTTH, coexisting migraine, being 270 work days per 1,000 employed people with migraine.
unmarried and sleeping problems28. Another study with Sickness absence rate in both groups was higher in
similar definitions for poor outcome29 found that unilat- women than in men. Approximately 60% of patients
eral headache, nausea and individual headache attack with TTH reported impaired social activities and work-
duration of ≥72 h were nonsignificantly associated with ing ability33. ETTH in Brazilian university students was
poor outcome of individuals with TTH alone. noted to have significant impact on physical activity and
productivity34.
Burden and impact
Based on the GBD study, TTH is ranked as the second Limitations of epidemiological studies
most common cause of chronic disease and injury The definition of TTH in epidemiological studies is
globally30. The most burdened age groups in terms of challenging for several reasons. First, infrequent ETTH
years of life lived with disability (YLD) for TTH were is common and may not be accurately recalled and
between 15 and 49 years. However, as TTH had a reported in epidemiological studies. Second, the inci-
much lower disability weight than migraine, overall, it dence of frequent ETTH or CTTH may represent an
caused 7.1 million YLD compared with 47.2 million for exacerbation of a pre-existing disorder rather than new
migraine globally in 2017 (ref.2). In a clinic-based study onset of a distinct condition. Third, TTH can coexist
on disease burden, patients with ETTH reported an with migraine in the same person and diagnostic rules
average of 2.3 work days and 1.6 household work days for coding migraine and TTH vary between studies.
lost in the last 3 months, whereas patients with CTTH When this occurs, suggestions from the ICHD-3 are to
reported a mean of 8.9 work days and 10.3 days of house code all present conditions; however, attacks of pheno­
activity lost in the last 3 months31. In the Danish popu- typic TTH in patients with migraine may be mild
lation, 12% of employed participants with TTH, repre- migraine, and coding these headaches as TTH may be
senting 9% of the general employed population, lost one incorrect35,36. By contrast, other studies may code these
or more working days in the preceding year owing to individuals as having migraine only. Biomarkers to dif-
headache, and the mean number of absence days from ferentiate between phenotypic TTH and mild migraine
are not available, but would be useful for this purpose.

Box 1 | Diagnostic criteria for tension-type headache according to the ICHD-3 Comorbidities
TTH can co-occur with several disorders that may
Infrequent ETTH
Meeting the following criteria A and B, in addition to criteria C–E, below.
affect the choice of therapy37. Population-based data
have shown that anxiety, depression and sleep distur-
A. Frequency: at least 10 headache episodes occurring on average <1 day/month
(that is, <12 days/year)
bances (including insomnia) are more prevalent in
B. Duration: 30 min to 7 days individuals with TTH than in the non-headache general
population38–42. Depression and anxiety are also associ-
Frequent ETTH
ated with the frequency and severity of TTH attacks43.
Meeting the following criteria A and B, in addition to criteria C–E, below.
However, it is not possible to ascertain a causal relation-
A. Frequency: at least 10 headache episodes occurring on average 1–14 days/month
ship based on cross-sectional data. One longitudinal
for >3 months (≥12 and <180 days/year)
B. Duration: 30 min to 7 days
study showed possible bidirectional effects between
pain and emotional factors on the burden of CTTH44.
CTTH However, these associations could be influenced by
Meeting the following criteria A and B, in addition to criteria C–E, below.
coexisting migraine, which is highly comorbid with
A. Frequency: headache occurring on ≥15 days/month on average for >3 months depression and anxiety45,46.
(≥180 days/year)
TTH is also associated with other pain disorders,
B. Duration: hours to days, or unremitting
including migraine. Indeed, one population-based study
C. At least two of the following four characteristics:
found that 83% of people with migraine within the past
1. bilateral location
year also had TTH13. TTH has been found to be asso-
2. pressing or tightening (non-pulsating) quality ciated with comorbid neck pain and low back pain47–49.
3. mild or moderate intensity Almost 90% of people with TTH have comorbid neck
4. not aggravated by routine physical activity such as walking or climbing stairs pain49, and ~80% of individuals with TTH have low
D. Both of the following: back pain48. Moreover, TTH frequency positively cor-
for infrequent and frequent ETTH: relates with the frequency of comorbid neck pain and
1. no nausea or vomiting low back pain48,49.
2. no more than one of photophobia or phonophobia
for CTTH: Mechanisms/pathophysiology
1. no more than one of photophobia, phonophobia or mild nausea The potential pathophysiological mechanisms of TTH
2. neither moderate or severe nausea nor vomiting can be divided into three broad categories: genetic
factors, myofascial mechanisms (including myofascial
E. Not better accounted for by another ICHD-3 diagnosis
nociception) and mechanisms of chronification (includ-
CTTH, chronic tension-type headache; ETTH, episodic tension-type headache; ICHD-3, ing central sensitization and altered descending pain
International Classification of Headache Disorders, third edition. Data from ICHD-3 (ref.4).
modulation) (Fig. 1).


NATURE REVIEWS | DISEASE PRIMERS | Article citation ID: (2021) 7:24 3

0123456789();:
Primer

Myofascial mechanisms Genetic disposition


• Increased tenderness • 5-HTTLPR polymorphism
• Increased hardness • Val158Met COMT polymorphism
• Local inflammation • APOE-ε4 polymorphism
• Local ischaemia

Tension-type
Peripheral headache
sensitization

Vascular mechanisms Central mechanisms


• Increased blood flow in cerebral arteries • Central sensitization
• Abnormal carotid artery blood flow • Dysfunction in descending
• Abnormal extracranial vascular response pain modulation

Fig. 1 | Pathophysiology of tension-type headache. The potential pathophysiological mechanisms of tension-type


headache involve possible genetic factors, peripheral and central mechanisms. Peripheral mechanisms, including
myofascial and to a lesser degree vascular factors, may lead to peripheral sensitization of nociceptors. Central
mechanisms involve central sensitization of nociceptive pathways and altered descending pain modulation. 5-HTTLPR,
5-hydroxytryptamine (serotonin)-transporter-linked-gene-linked polymorphic region; APOE, apolipoprotein;
COMT, catechol-O-methyltransferase.

Genetic factors Peripheral mechanisms


Two registry-based twin studies have investigated the Myofascial mechanisms. Tenderness of pericranial
influence of genetic factors on TTH50,51 and found that muscles is common in patients with TTH during the
heritability estimates vary with the presence of concom- acute headache phase and, in some patients, outside
itant migraine. In those with migraine, the heritability of headache periods, pointing to a potential role of
of TTH was 19% in one study, whereas heritability of myofascial tissues in the pathophysiology of TTH58,59.
TTH was 48% in male twins and 44% in female twins Manual palpation is the most common method used for
in those without migraine50,51. The difference in con- the evaluation of tenderness, and tenderness is quan-
cordance rates of monozygotic twins and same-sex dizy- tified using a subjective pain scoring system60. Both
gotic twins was ~5% and ~10% in infrequent ETTH and pericranial muscles and tendon insertions have higher
frequent ETTH, respectively. This finding may suggest tenderness scores in patients with ETTH and CTTH
that the genetic effects were more substantial in frequent than in controls60,61. Whether muscular tenderness is a
ETTH than in infrequent ETTH, although a direct com- consequence of the pain or contributes to the develop-
parison between the two conditions was not carried ment of TTH attacks is still debated62. In one study with
out52. For CTTH, heritability could not be estimated 30 min tooth clenching (a stimulus known to induce
from twin studies owing to small sample sizes, but a headache), muscle tenderness preceded the headache
family aggregation study found a threefold increased in patients with ETTH or CTTH63. Similarly, increased
risk of CTTH in first-degree relatives of patients with muscle tenderness anticipated a lower pain threshold in
CTTH53. Of note, this study should be interpreted patients who developed frequent ETTH over 12 years
with caution because concomitant migraine was not of follow-up in a population-based study64. Pericranial
evaluated. tenderness is exacerbated during the acute headache
The specific genes that are causative of TTH are phase and increases with the severity and frequency of
unknown. Studies have found a possible contribu- TTH attacks, supporting the presence of more severe
tion of the 5-HTT-gene-linked polymorphic region tenderness in individuals with CTTH than in those
(5-HTTLPR) genotype and Val158Met COMT (encoding with ETTH61,63,65. By contrast, other studies have found
catechol‐O‐methyltransferase) polymorphism in the risk increased muscular tenderness and stiffness in patients
of CTTH or its phenotype54–56. By contrast, the APOE-ε4 with either ETTH or CTTH when they are headache-free
gene was suggested as protective against TTH57. compared with healthy individuals66,67. Along with

4 | Article citation ID: (2021) 7:24 www.nature.com/nrdp

0123456789();:
Primer

increased tenderness, pericranial muscles are ‘harder’ ETTH have increased pain sensitivity after intramuscu-
in patients with ETTH and CTTH66,68. Muscle hardness lar infusion of inflammatory substances compared with
can be investigated using the hardness meter, a reliable, controls87, which could be related to central sensitiza-
quantitative measurement method. Muscle hardness tion (increased excitability of the central nervous system)
was increased in patients with CTTH both on days with and peripheral sensitization (reduced thresholds and
and on days without TTH66. Of note, inhibiting nitric increased responsiveness of peripheral nociceptors)88 of
oxide synthase (NOS) with NG-monomethyl-l-arginine myofascial sensory afferents, and is likely caused by the
hydrochloride (l-NMMA) leading to decreased nitric release of endogenous inflammatory mediators in tender
oxide (NO) levels reduced pain and muscle hardness in muscles87. Further studies are needed to clarify the rela-
patients with CTTH69,70, supporting a role for NO in pain tionship between muscle tender points and myofascial
modulation in TTH. trigger points, and their role in TTH pathophysiology78.
Myofascial trigger points, hyperirritable spots asso- Some evidence suggests an association between myo-
ciated with a taut band (a band of muscle fibres with fascial pain and local muscle ischaemia (likely caused
increased tone) in skeletal muscles that elicit local and by sustained muscle contraction, alterations in metab-
referred pain when compressed, may be important in olism, microcirculation and mitochondrial function) in
the pathophysiology of TTH. The referred pain from the tender areas in TTH and in other myofascial pain
muscles in the head, neck and shoulder regions can disorders89. A few studies have reported increased lev-
mimic the pain pattern of TTH71,72. Several studies have els of substances involved in pain and inflammation,
found a high relative frequency of myofascial trigger such as IL-6, bradykinin and serotonin in the blood
points in patients with ETTH or CTTH compared with and myofascial trigger points of patients with TTH73,90.
controls73,74. In addition, increased nociception at myo- By contrast, one study found normal interstitial levels
fascial trigger points could lead to sensitization of cen- of inflammatory mediators (such as bradykinin, aden-
tral nociceptive pathways in individuals with ETTH and osine 5-triphosphate, glutamate and prostaglandin E2)
CTTH75,76. However, data on the association of myofas- in tender points of the trapezius muscle in patients with
cial trigger points and the severity and frequency of TTH CTTH, suggesting that these points do not have per-
attacks have been inconsistent74. Moreover, prospective sistent inflammation in patients with TTH91. Using the
studies evaluating the role of myofascial trigger points in 133xenon clearance technique, one study92 found nor-
the evolution of ETTH to CTTH are needed71. Methods mal resting blood flow and relative flow increase dur-
used to investigate the role of peripheral mechanisms in ing isometric work in large areas of temporal muscle in
TTH pathophysiology (such as manual palpation and patients with CTTH92. In addition, normal muscle blood
palpometer for measuring muscle tenderness, and sur- flow and lactate levels were found in tender points in
face versus needle electromyography (EMG) for meas- patients with CTTH during static exercise, suggesting
uring muscle activity) have varied, and might contribute that muscle ischaemia may not contribute to the onset
to variation in results between studies77. of TTH93.
Myofascial nociception is mediated by thin myeli-
nated (Aδ) and unmyelinated (C) fibres that are activated Vascular factors
by noxious and innocuous stimuli, including muscle In contrast to migraine, cerebral vascular input is of
strain or contraction, inflammation and ischaemia62. minor importance in TTH94. Haemodynamic changes
Sustained contraction of head, shoulder and neck mus- in the cerebral vasculature could activate trigeminal
cles was believed to contribute to the onset of pain in nociceptors and lead to headache pain95. In support
TTH by the sensitization of peripheral nociceptors.78 of this, patients with ETTH have increased blood flow
However, interestingly, many studies have found normal velocities in the anterior, middle and posterior cerebral
global EMG activity in most patients with ETTH, with arteries, whereas patients with CTTH had alterations of
small areas, such as trigger points, with increased EMG the blood flow velocities of the basilar and middle cere-
activities in patients with CTTH62,79,80. bral arteries, compared with controls96–98. These findings
Physiologically, stress can trigger or aggravate the could be explained by the vasodilator activity of CGRP
headache by increasing muscle contraction, by releas- and NO, or by impaired sympathetic function; how-
ing catecholamines and cortisol, by peripheral sensi- ever, although increased NO levels have been found in
tization and/or by affecting central pain processing81. patients with CTTH, further studies are needed to clarify
Long-term corticosteroid release in those with stress the role of CGRP and the sympathetic system in TTH
may cause tissue damage leading to pain and increased pathophysiology99. Cranial haemodynamics in patients
pain perception82,83. In TTH, stress has been shown with CTTH has been studied during nitroglycerin and
to trigger higher pericranial muscle pain ratings in head down tilt headache provocation100, both of which
patients with TTH than in controls84. However, there induced a significant increase in the severity of head-
is no evidence supporting a link between stress and ache in patients with CTTH, but not in controls. Both
increased peri­cranial EMG activity, and its role in TTH patients with CTTH and controls had similar changes in
pathophysiology remains unclear84,85. the carotid artery blood flow after nitroglycerin admin-
Muscle nociceptors can be activated by several istration, but higher carotid artery blood flow was found
inflammatory mediators (such as serotonin, brady- in patients with CTTH compared with controls during
kinin, substance P and calcitonin gene-related peptide head down tilt, probably due to distention of intracranial
(CGRP)), resulting in activation and sensitization of veins100. Some evidence suggests a failure of dilatation of
peripheral sensory afferents86. Patients with frequent temporal arteries in response to exercise in patients with


NATURE REVIEWS | DISEASE PRIMERS | Article citation ID: (2021) 7:24 5

0123456789();:
Primer

Motor cortex (M1/M2) Somatosensory cortex (S1/S2)


Descending input from the motor cortex modulates motor Ascending input from thalamo-cortical projections
neurons leading to maladaptive changes in muscle fibres increases somatosensory cortical levels of glutamate

Thalamus
Ascending inputs
from spinal trigeminal
nucleus and dorsal
horns at C3–C4 levels
lead to sensitization
Limbic system of third order neurons
The limbic system in thalamus mediated by
modulates the nitric oxide and glutamate
motor cortex,
decending
pathways modulate
the periaqueductal
grey and rostral
ventromedial
medulla PAG

PAG and RVM


Decreased inhibition
(PAG) and increased
facilitation (RVM) of
nociceptive transmission
Muscle fibres Sensory afferents at the level of trigeminal
Increased muscle hardness and C fibres, Aδ fibres nucleus caudalis and
sustained contraction leads to and Aβ fibres (become dorsal horns at C3–C4
local ischaemia and inflammation pro-nociceptive RVM levels lead to sensitization
or pain simulatory)
transmit peripheral
nociceptive input STN
from muscles and
blood vessels to the
Blood vessels DH C3–C4
trigeminal and
Increased blood flow in cerebral
C1–C4 dorsal
arteries, abnormal cartoid artery
root ganglions
blood flow, and abnormal
extracranial vascular response Medulla and upper cervical spinal cord
Trigeminal Increased afferent nociceptive input leads
ganglion Dorsal root to sensitization of second order neurons
ganglion in spinal trigeminal nucleus and
of C1–C4 dorsal horns at C3–C4 levels mediated by
nitric oxide and glutamate

Fig. 2 | Proposed mechanisms of chronification of tension-type headache. Abnormal input from myofascial structures
in the head and neck and, to a lesser degree, from vascular structures, results in sensitization of peripheral nociceptors.
This increased nociceptive input leads to sensitization of second order neurons in the spinal trigeminal nucleus and dorsal
horns of the spinal cord at levels C1 to C4. Aβ fibres, which normally only respond to innocuous stimuli, become pro-
nociceptive or pain stimulatory. Dysfunction of descending modulatory pain pathways results in decreased inhibition and
increased facilitation of nociceptive transmission. This leads to increased nociceptive input to the thalamus and sensitization
of third order neurons, which send impulses to the cortex. The result is increased pain perception and chronification of
tension-type headache. DH, dorsal horn; M1, primary motor cortex; M2, supplementary motor area; PAG, periaqueductal
grey; RVM, rostral ventromedial medulla; S1, primary somatosensory area; S2, secondary somatosensory area; STN, spinal
trigeminal nucleus (comprising pars caudalis, pars interpolaris and pars oralis; pars caudalis extends from C2 to the obex).

CTTH, supporting an abnormal extracranial vascular thalamus, limbic system, motor cortex and somatosensory
response in TTH101. cortex (Fig. 2).

Central factors Central sensitization. Several early studies found nor-


The transformation of ETTH into CTTH remains mal pressure pain detection thresholds in patients with
incompletely understood. The mechanism of chronifi- ETTH61,102–104; however, these studies were performed
cation of TTH may include sensitization of second-order before the separation of infrequent and frequent TTH in
neurons in the dorsal horn of the spinal cord or the ICHD-2 (ref.105), which may explain these data. Indeed,
spinal trigeminal nucleus, sensitization of supraspinal when infrequent ETTH was separated from frequent
neurons, and decreased antinociceptive or modulat- ETTH, abnormal or low pain detection thresholds were
ing activity from supraspinal structures, such as the found in patients with frequent ETTH in the cephalic

6 | Article citation ID: (2021) 7:24 www.nature.com/nrdp

0123456789();:
Primer

region but not in patients with infrequent ETTH61,106,107. headaches only119. This biphasic response occurred in
In addition, patients with CTTH had lower pain detec- a similar manner in patients with migraine, and where
tion thresholds and tolerance thresholds to pressure, migraine-like attacks were induced. These findings
electrical and thermal stimuli in cephalic and extra- may suggest that NO-related central sensitization is an
cephalic regions compared with controls77,102,103,108–111. important common denominator of primary headache
In patients with CTTH, alterations in pain sensitiv- disorders118.
ity were more evident using suprathreshold stimuli, Glutamate, a neuroexcitatory amino acid, has been
which may be more clinically relevant than threshold investigated in several studies on migraine but only in
measurements112,113. two studies in patients with CTTH. Data from these
Studies of temporal summation with repetitive pain- studies are conflicting: normal plasma glutamate levels
ful stimuli in patients with CTTH provided inconsistent were found in one study and increased platelet glutamate
results, probably owing to the application of different levels were found in the other99,120.
types of stimulation (electrical versus mechanical stim- Comorbid conditions may alter pain sensitivity
ulation). Compared with controls, patients with CTTH in individuals with TTH. In patients with TTH and
had higher pain ratings from repeated pressure stim- migraine, the presence of back pain was associated with
uli, and nonsignificant temporal summation changes increased central sensitization and muscular tender-
to electrical stimuli have been observed112,114. In con- ness compared with individuals without headache48.
trast to electrical stimuli, which stimulate nerve fibres Whether central sensitization facilitates the onset of
directly, pressure pain stimulation activates both noci- back pain in patients with TTH or vice versa or whether
ceptors and central mechanisms. Abnormal temporal it might be the consequence of comorbidity is unclear48.
summation in patients with CTTH might, therefore, Fibromyalgia is also an important comorbidity of TTH,
result from an altered response in both peripheral and with a twofold increased prevalence in patients with
central mechanisms. CTTH compared with those with ETTH, implying a
Overall, pain perception studies support the pres- shared mechanism between the two conditions, such as
ence of hyperalgesia (increased sensitivity to painful central sensitization121. In addition, the increased prev-
stimuli) and allodynia (pain elicited by stimuli that are alence of temporomandibular disorder symptoms in
normally not painful) in patients with frequent ETTH patients with TTH also suggests increased pain percep-
or CTTH48,62,75,102,106,110,112. The nonspecific and wide- tion associated with TTH, although it is unclear whether
spread nature of pain hypersensitivity in these patients the mechanisms linking the two conditions is central
suggests the involvement of sensitization of central pain sensitization from TTH or peripheral sensitization from
processing regions62,108,115. temporomandibular disorder122.
In support of a role of central sensitization in TTH Personality traits and psychological comorbidities
chronification, several studies have found an associa- may affect patients’ vulnerability to and coping strat-
tion between central pain sensitization and more fre- egy for pain. For example, high levels of depression and
quent TTH attacks77,116. In addition, one longitudinal, neuro­ticism were associated with increased pain sensi-
population-based study found normal pain thresholds tivity in cephalic and extracephalic regions in individuals
at baseline in patients with ETTH, but reduced thresh- with TTH and migraine, suggesting that pain percep-
olds in patients who developed CTTH after 12 years64. tion and modulation may be influenced by negative
Sensitization of both supraspinal and spinal nociceptive appraisals of pain38. Sleep disturbances are common in
pathways might explain the increased muscle tenderness patients with TTH. This association might be explained
in patients with TTH. The pain response to mechanical by the involvement of common neurotransmitters and
pressure applied to tender muscles is quantitatively and brain networks, including the hypothalamus and brain-
qualitatively different in patients with CTTH from that stem, in the pathophysiology of the two diseases123. The
in controls106,113. Palpation pressure elicited a stronger inadequate sleep may lead to central sensitization as
pain response in patients with CTTH, as represented by demonstrated by lower pain thresholds in individuals
a steeper stimulus–response function, compared with with CTTH124,125.
controls113. These findings seem to be explained by cen- Neuroimaging studies have found functional126 and
tral sensitization at the dorsal horn of the spinal cord and structural127,128 alterations in cortical regions involved
trigeminal nucleus108,113. A moderate correlation between in pain processing in patients with TTH. For example,
cortical pain hypersensitivity and myofascial tenderness decreased grey matter volume of areas of the limbic
has been found in patients with CTTH117. system, such as the insula and anterior cingulate cor-
NO may be involved in the pain pathways of the tex, were found in patients with CTTH and those with
spinal cord and the central sensitization at the spinal chronic migraine, compared with patients with episodic
level118. The role of NO has been investigated in CTTH. TTH and patients with episodic migraine128. In addition,
Increased NOS activity in platelets was found in patients decreased grey matter volume of limbic areas in patients
with CTTH, reflecting central upregulation of NOS with ETTH and CTTH was significantly associated with
activity in neurons in the spinal dorsal horn, trigeminal longer disease duration. Furthermore, structural brain
nucleus and supraspinal structures99. In a provocation alterations of the insula, cingulate and somatosen-
study using glyceryl trinitrate (GTN), which is an NO sory cortex change dynamically between the pain and
donor, patients with CTTH had an immediate headache pain-free phases in patients with ETTH127,129. Thus,
and a delayed headache the latter of which resembled imaging findings might be a consequence of frequent
TTH, whereas controls developed immediate mild peripheral nociceptive inputs and central sensitization.


NATURE REVIEWS | DISEASE PRIMERS | Article citation ID: (2021) 7:24 7

0123456789();:
Primer

Some central mechanisms related to CTTH are simi- (citalopram) did not achieve better treatment outcome
lar to those found in chronic migraine (such as increased compared with non-selective agents (amitriptyline)139,140.
pain perception, NO-related mechanisms, comorbidi-
ties and brain structural and functional abnormalities). Endogenous opioids. Endogenous opioids have an
Whether chronic migraine and CTTH are two different important role in modulating nociceptive information141.
conditions or a single process in which two primary Opioid receptors are found in central or peripheral
headaches converge is debated. Considering the shared nociceptive afferent fibres, and in regions involved in
pathophysiological mechanisms, it can be suspected that descending pain modulation, such as the periaqueductal
central sensitization is a key factor in the chronification grey, nucleus raphe magnus and dorsal raphe in the
of the two disorders, whereas the difference in clinical rostral ventral medulla141. There is no compelling evi-
phenotypes between migraine and TTH suggest that dence of impaired endogenous opioid system in TTH,
they are two different entities. although the number of studies is small. Two studies
found normal β-endorphin levels in peripheral blood
Dysfunction in descending pain modulation. Deficiency and cerebrospinal fluid (CSF) in patients with ETTH or
of the descending pain modulatory system may con- CTTH142,143. In ETTH, reduced β-endorphin levels were
tribute to the increased pain sensitivity and central observed and correlated with pressure pain threshold144.
sensitization in CTTH. Neurophysiological studies In addition, plasma and CSF met-enkephalin levels are
investigating the activity of inhibitory brainstem neu- increased in patients with CTTH145,146, with a small but
rons in patients with TTH have provided inconsistent significant decrease in CSF dynorphin levels in patients
results, perhaps owing to methodological issues such as with CTTH compared with controls147.
the use of different recording systems117,130. Conversely,
many studies have pointed to dysfunction of the dif- Neuropeptides
fuse noxious inhibitory controls (DNICs)77, which com- CGRP is one of the important neurotransmitters of
prises the inhibition of spinal and trigeminal neurons trigeminal nociceptive signalling148. In migraine, the
by nociceptive stimulation applied to spatially distant finding of increased plasma levels of CGRP during
regions of the body. This phenomenon is usually regu- and outside of migraine attacks has led to the success
lated by supraspinal descending projections. Compared of modern CGRP-targeting treatments149,150. However,
with controls, patients with CTTH had facilitation in TTH, studies have consistently found normal CGRP
of the nociceptive flexion reflex, which is a polysyn- concentrations in cranial and peripheral blood and CSF
aptic spinal withdrawal reflex that may be depressed in patients with ETTH and CTTH147,151,152, which did not
by supraspinal descending projections and DNICs131. increase during spontaneous or induced attacks151,153.
Impaired inhibition of repeated nociceptive inputs Of note, owing to methodological issues of in vitro meas-
was also found in patients with CTTH117,132. Negative urement of CGRP, it is difficult to draw a firm conclu-
findings were found regarding a possible influence of sion that CGRP is not involved in the pathophysiology
stress on temporal summation or DNICs114. In contrast, of TTH.
greater levels of physical activity have been shown to In addition to CGRP, nerve fibres containing sub-
be associated with less pain facilitation and more pain stance P, neuropeptide Y (NPY) and vasoactive intesti-
inhibition133. nal peptide (VIP) are found in extracranial cutaneous,
muscular and perivascular tissues, suggesting a poten-
Serotonin. The monoamine serotonin (5-hydroxy­ tial role of these peptides in the pathophysiology of
tryptamine; 5-HT) has been investigated in several headaches, including TTH154. However, plasma levels
early studies of TTH. Although these studies used of substance P, NPY and VIP were normal in cranial
hetero­geneous methods and yielded conflicting results, and peripheral blood of patients with CTTH, irrespec-
it is widely accepted that patients with ETTH have tive of headache status155. In a study of individuals with
increased platelet and plasma serotonin levels, and ETTH, substance P was reduced in platelets from patients
decreased platelet serotonin uptake compared with with ETTH compared with controls, and levels of sub-
controls134,135. By contrast, platelet and plasma serotonin stance P were negatively correlated with pressure pain
levels were either normal or decreased in patients with threshold144. Normal plasma levels of NPY were noted
CTTH, and platelet serotonin uptake was also normal during and between attacks in patients with ETTH156.
compared with controls136,137. In addition, 5-HTTLPR Of note, assay methodology should be considered when
polymorphism may be associated with CTTH risk54. interpreting these study results.
When the increased levels and decreased uptake of ser-
otonin are regarded as a physiological antinociceptive Diagnosis, screening and prevention
reaction to pain, those findings may suggest that patients Clinical presentation
with CTTH have a deficiency in the ability to increase Episodic tension-type headache. TTH is sometimes
plasma serotonin in response to peripheral nociceptive described as a ‘featureless headache’ because it lacks the
inputs, which may predispose to dysfunctional pain associated features of migraine (such as nausea, photo­
processing or modulation108. Serotonin acts on central phobia and phonophobia) or trigeminal autonomic
antinociceptive mechanisms along with other neuro- cephalalgias (such as lacrimation and conjunctival
transmitters such as noradrenaline and adenosine138, injection (enlargement of conjunctival blood vessels))5.
which may explain why stronger serotonin reuptake inhi- Most individuals with ETTH rarely, if ever, seek medical
bition by a selective serotonin reuptake inhibitor (SSRI) care despite the widespread prevalence in the general

8 | Article citation ID: (2021) 7:24 www.nature.com/nrdp

0123456789();:
Primer

Medication overuse
population32, which has resulted in a scarcity of litera- data gathered by prospective diary entries161. In addition,
headache ture related to clinical features of ETTH157,158. Much of diagnostic headache diaries can help to differentiate
Secondary headache the available data come from a Danish population-based CTTH from other headache disorders, mainly migraine
characterized as headache cohort13. Based on these data, ETTH is bilateral in 90% and medication overuse headache (MOH)162.
occurring on ≥15 days/month
of occurrences, of pressing or tightening quality in 78% of Physical examination is carried out in all patients
that is attributed to regular
overuse of analgesics or occurrences, of mild or moderate pain intensity in 99% with suspected TTH and includes fundoscopic, general
migraine-specific acute of instances and is not aggravated by routine physical and neurological examinations (comprising mental sta-
medications (on ≥10 or activity in 74% of instances13. In another report from the tus, cranial nerve, motor and sensory function, balance
≥15 days/month, depending
same cohort, episodes of headache in individuals with and coordination, reflexes and gait assessment). This
on the medication) in
individuals with a pre-existing
frequent ETTH lasted between 30 min and 24 h in 83% is often normal aside from possible pericranial ten-
headache disorder such of instances29. derness, which is assessed by manual palpation61,157,158.
as tension-type headache Fundoscopic examination is performed to rule out pap-
or migraine. Chronic tension-type headache. TTH is often described illoedema, which can be a sign of idiopathic intracranial
as ‘graded’, as the intensity of pain increases with the hypertension, or other causes of increased intracranial
frequency of headache158. In CTTH, headaches are pressure.
more frequent, more severe and more likely to require Comorbid disorders should also be assessed by
medical care than in ETTH19,32. CTTH typically devel- detailed medical history and physical examination at the
ops gradually in individuals with ETTH4. Clinical fea- initial clinical evaluation. Such assessment is likely to
tures of CTTH have been examined in only a very few inform better practices and individualize approaches
studies13,19 but are similar to those reported in individ- to clinical management.
uals with ETTH, except for an increased intensity of
pain13. In a population-based sample of persons with Differential diagnosis
CTTH, pain is of bilateral location in 88% of instances, of A wide range of headache disorders can mimic the
pressing or tightening quality in 83% of occurrences, clinical features of TTH160; careful exclusion of these
of mild or moderate pain intensity in 96% of instances disorders relies on a thorough review of medical his-
and is not aggravated by routine physical activity in 71% tory and an adequate physical examination. On this
of cases13. Furthermore, 58% of individuals with CTTH basis, further diagnostic tests might be required if a
did not report nausea, photophobia or phonophobia as patient has red flags that are suggestive of secondary
accompanying symptoms13. Pericranial tenderness is headache causes163,164 (Box 2). Further tests may include
the most common abnormal finding with CTTH, and
is more common in those with CTTH than in patients
Box 2 | Secondary headache disorders: red flags
with ETTH or migraine or in healthy volunteers without
headache61,159. Intracranial space-occupying lesion
Red flags:
Diagnostic work-up • progressively worsening headache
TTH is diagnosed on the basis of headache history and • headache brought on by sneezing, coughing or exercise
the exclusion of other disorders that may mimic TTH.
Subdural haematoma
Diagnostic criteria are provided by ICHD-3 (ref.4) (Box 1). Red flags:
TTH should be suspected in any individual who
• head trauma
reports recurrent headache of mild or moderate inten-
sity. The diagnosis is most probable in patients whose Secondary headache
headache is bilateral and tightening or pressing in qual- Red flags:
ity. A typical description of TTH is the sensation of a • headache associated with weight loss and/or change
tightening rubber band around the head. Appropriate in memory or personality
clinical evaluation is needed to establish diagnosis, and • headache onset >50 years of age
a review of medical history must account for onset, dura- • focal neurological symptoms
tion and frequency of headache, as well as pain features, • weight loss
accompanying symptoms and aggravating and relieving • impaired memory and/or altered consciousness
factors157,158. Assessment of accompanying symptoms or personality
are mostly used to differentiate TTH from migraine,
which is the most common and important differential Meningitis
diagnosis160. Red flags:
Diagnostic headache diaries are the best available • unexplained fever
assessment tool for TTH and can be used to support • neck stiffness
diagnosis and guide clinical decision-making. Headache Intracranial hypertension or hypotension
diaries enable patients to report clinical headache features Red flag:
(such as frequency and characteristics) and accompany- • headache aggravated by posture or manoeuvres that
ing symptoms. The benefits of headache diaries are multi­ raise intracranial pressure
fold and include better differentiation of ETTH from
Subarachnoid haemorrhage
CTTH compared with clinical evaluation alone. Indeed,
Red flag:
patients tend to over-report their headache frequency
• neck stiffness
during an initial clinical evaluation when compared with


NATURE REVIEWS | DISEASE PRIMERS | Article citation ID: (2021) 7:24 9

0123456789();:
Primer

New daily persistent


neuroimaging, blood sampling, lumbar puncture and Brain tumours often present with headache that most
headache polysomnography. Systematic testing of all patients with often has features of TTH rather than migraine173. Other
Primary headache disorder TTH (for example, neuroimaging) should be avoided differential diagnosis of TTH includes a range of sec-
characterized by unremitting and does not appear to improve quality of life or mitigate ondary headache disorders, such as headache attributed
pain within 24 h of a clearly
remembered onset and
health concerns in the long term165. to traumatic brain injury, idiopathic intracranial hyper-
for >3 months. Migraine is the common and important differential tension, low CSF pressure, sleep apnoea and recurrent
diagnosis of TTH30,160,166. It is not uncommon for clini- rhinosinusitis160. These disorders may have features sug-
cians to conflate TTH and migraine, as individuals with gestive of TTH and clinicians should, therefore, become
migraine have coexisting TTH with a similar prevalence suspicious if red flags are raised by the medical history
to that seen in the general non-migraine population167. and/or physical examination (Box 2). In such instances,
Indeed, the 1-year prevalence of frequent ETTH was 56% neuroimaging is often recommended to confirm or
in individuals with migraine in one population-based exclude underlying causes.
study168. Differentiating migraine and CTTH is difficult,
as CTTH often manifests with headache episodes of at Management
least moderate pain intensity19 that are comparable to Multimodal treatment should be tailored for each
those seen in individuals with migraine19. Migraine may patient with TTH and may include pharmacotherapy,
also be conflated with CTTH in those with overuse of interventional medicine, behavioural therapies, physi-
analgesics. One point that merits special emphasis is the cal and occupational therapies, exercise, healthy lifestyle
exclusion of a secondary diagnosis of TTH in individuals habits and stress management (Fig. 3). Infrequent TTH
who meet the diagnostic criteria for chronic migraine, as can be managed primarily with acute pharmacological
defined by the ICHD-3 (ref.4). Indeed, it is very common treatment and lifestyle modifications but frequent ETTH
for individuals with chronic migraine to experience at or CTTH may require preventive pharmacotherapy
least some headache days with clinical features that fulfil and/or behavioural interventions.
the diagnostic criteria for TTH. A mainstay of clinical management of TTH is phar-
No biomarkers can differentiate between TTH, macotherapy. Educating patients about TTH triggers,
migraine and other headache disorders, therefore, cli- natural history and treatment options and goals is essen-
nicians can benefit from the use of diagnostic head- tial. Decision-making requires consideration of patients’
ache diaries, in which clinical features are prospectively preferences and comorbidities. As the evidence base is
recorded. less well developed for TTH than for migraine, clinical
Clinicians should also consider MOH as an addi- experience has a larger role in making treatment choices
tional diagnosis in individuals with CTTH4,169. Another for TTH. In this context, comorbidities should also be
differential diagnosis to CTTH is new daily persistent considered in the decision-making. Patients should
headache (NDPH)4. NDPH is a rare headache disorder in also be educated to know that the goal of the medical
the general population, with an estimated prevalence of interventions is not to cure TTH but to help patients
0.03–0.1%170. Of note, NDPH is reported by 11% of head- to manage the condition. Alignment on both the goals
ache centre patients with CDH171 and in 21% of adolescent of treatment and the approach to treatment improves
patients with CDH172. outcomes174. In addition, unless the patient’s individual

Management of TTH

Patient education and trigger management


Prophylactic:
Prophylactic: pharmacological
Symptomatic non-pharmacological
In case of high-frequency 1st option:
1st option: amitriptyline
ETTH, or CTTH, or when • Bio-behavioural therapies
aspirin and/or
acetaminophen symptomatic treatment (CBT, biofeedback,
fails, or comorbidities relaxation)
• Physical and/or 2nd option:
require chronic
mirtazapine
2nd option: management of the occupational therapy
caffeine condition, initiate • Acupuncture
combinations prophylactic treatment • Lifestyle modification 3rd option:
or NSAIDs (including sleep hygiene) venlafaxine

Fig. 3 | Management of tension-type headache. Management of tension-type headache (TTH) starts with patient education
and management of known headache triggers. Symptomatic treatment is required for mild to severe headaches, or when
headaches affect function. Non-pharmacological treatments can be used concurrently with pharmacological therapies
(including symptomatic and/or prophylactic) or as stand-alone therapies. Of note, multiple non-pharmacological
therapies can be combined. When prophylactic treatment is needed, starting with non-pharmaceutical interventions is
recommended unless the patient prefers pharmacological therapy. After 6 months of successful management (defined as a
≥50% reduction in headache frequency), treatment may be paused to see whether TTH relapses or worsens. CBT, cognitive
behavioural therapy; CTTH, chronic TTH; ETTH, episodic TTH.

10 | Article citation ID: (2021) 7:24 www.nature.com/nrdp

0123456789();:
Primer

Box 3 | Management of TTH in specific populations the efficacy of the analgesics for the acute treatment of
TTH189. Indeed, in a meta-analysis of four trials189, the
Children and adolescents fixed combination of aspirin, acetaminophen and caf-
There is some evidence only for acetaminophen267 in the acute treatment of tension- feine was superior to paracetamol alone or placebo, with
type headache (TTH) and amitriptyline268,269 for prevention, whereas non-pharmacological limited adverse effects reported. The most common
and multidisciplinary interventions216,270–272 have stronger evidence and should be
adverse events of these combinations were nervousness
preferable or used in combination with amitriptyline when needed.
(in 6.5% of participants), nausea (4.3%), abdominal
Elderly individuals pain or discomfort (4.1%) and dizziness (3.2%)190,191. In
First-choice treatment is non-pharmacological intervention (such as biofeedback or addition, rebound headache after caffeine withdrawal
relaxation techniques) whenever accessible and accepted by patients. Acetaminophen
may occur192. Caffeine may increase the antinocice-
is the drug of first choice for acute treatment (NSAIDs should be avoided) and
amitriptyline for prevention of TTH, followed by venlafaxine and mirtazapine.
ptive effects of analgesics by promoting their gastric
Close monitoring is required owing to the risk of adverse effects181,199,273. Notably, TTH absorption190. EFNS-TF recommends the oral use of
may be a risk factor for dementia274. caffeine combinations for the acute treatment of TTH
with level B recommendation181. There is some evidence
Pregnancy and lactation
Pharmacological treatment should be offered and not be postponed when needed275.
that local application of peppermint oil in an ethanol
Acetaminophen is the first-choice drug for symptomatic treatment but combinations solution may be helpful in TTH relief193,194.
with caffeine or NSAIDs should be avoided275. Amitriptyline is preferred for the
prevention of TTH, followed by venlafaxine and mirtazapine (same FDA classification Medications to avoid. Opioids are best avoided owing
as for pregnancy — C class)275. to the risk of MOH and the potential for misuse169,195.
Triptans have not shown efficacy for symptomatic
treatment of TTH and might be avoided unless head-
needs and preferences are considered, treatment adher- ache attacks have features of migraine181. Of note,
ence is very low175; thus, shared decision-making is sumatriptan gave only modest headache relief for
indispensable176. Potential nocebo-producing behav- CTTH compared with placebo but the study may have
iours (such as frequent previous experiences of severe been underpowered196. Muscle relaxants are not recom­
subjective adverse events or frequent previous treatment mended for symptomatic TTH treatment owing to low
discontinuations owing to safety reasons) should also be efficacy. In some individuals, such as children, preg-
taken into account, in order to recruit tailored strategies nant or breast-feeding women and elderly patients,
to limit it177. In addition, the potential role of medication the use of NSAIDs or combinations of caffeine with
overuse in exacerbating headache should be emphasized aspirin or acetaminophen should be avoided for safety
along with clear guidance about limits on weekly or reasons (Box 3).
monthly use of acute treatments178 (Box 3).
Preventive treatment
Acute treatment Preventive treatment is recommended only for patients
Simple analgesics. Both acetaminophen and aspirin with frequent ETTH and CTTH. Infrequent TTH does
have been investigated for the symptomatic treatment of not require preventive treatment unless there are specific
TTH179,180 (Table 1). Both treatments have a good safety comorbidities (such as depression or fibromyalgia), if
profile. The serious adverse events include liver dys- symptomatic treatment fails or daily functioning is
function for acetaminophen and bleeding with aspirin. affected by TTH.
The European Federation of Neurological Societies Task Deciding the duration of TTH prophylaxis remains an
Force (EFNS-TF) recommends the oral use of acetami- active challenge. All studies lasted 2–6 months, but many
nophen and aspirin for the acute treatment of TTH with patients may require longer treatment. Comorbidity,
level A recommendation181. Notably, the effectiveness of response to treatment, patient characteristics, previ-
symptomatic medication is associated with better head- ous headache history, patient preferences and lifestyle
ache parameters (history, frequency and duration) and choices should be considered in deciding how long to
lower emotional burden182. continue treatment. In those with an excellent response
Of the NSAIDs, ibuprofen183 and ketoprofen184 have (such as 50–75% reduction in monthly headache days)
the highest efficacy for the acute treatment of TTH. pausing the treatment after 3 or 6 months and monitor-
Other NSAIDs that have a lower evidence of efficacy ing for any recurrence of the headache is a widely used
contain naproxen185 and diclofenac186,187, yet are common approach, unless other comorbidities, such as depression
in practice (Table 1). NSAIDs inhibit the cyclooxygen- or anxiety disorders, require longer treatment.
ase pathway and may be associated with gastric, hepatic
and kidney adverse events, blood pressure elevation and Amitriptyline. Level of evidence of efficacy of amitrip-
increased risk of bleeding. EFNS-TF recommends tyline (tricyclic antidepressant) in TTH prevention is
the oral use of ibuprofen, ketoprofen, naproxen and moderate to high in several studies supporting its level A
diclofenac for the acute treatment of TTH with level A rating by EFNS-TF181,197 (Table 2). The analgesic effect of
recommendation181. amitriptyline is related to, for example, noradrenaline
reuptake inhibition, N-methyl-d-aspartate (NMDA)
Combination analgesics. Although caffeine alone was receptor antagonism, blockade of muscarinic recep-
not effective for TTH treatment in a placebo-controlled tors and ion channels140, or modulation of the seroto-
study188, there is evidence that the combination of caf- ninergic and noradrenergic descending pain inhibitory
feine with acetaminophen, aspirin or ibuprofen improves system198. Common tolerability issues include drowsiness,


NATURE REVIEWS | DISEASE PRIMERS | Article citation ID: (2021) 7:24 11

0123456789();:
Primer

Table 1 | A selection of acute medications for TTH tested in randomized placebo-controlled trials
Drug by type of study Route NNT NNH Quality of Ref.
evidence
Meta-analysis
Aspirin PO 6 (4.2–12)a No difference Low 180

Paracetamol/acetaminophen PO 22 (15–40)a No difference High 179

Ibuprofen PO 14 (8.4–47)a NA Moderate 183

Ketoprofen PO 9.0 (4.8–72)a 15 (8.7–15) Low 184

Aspirin + paracetamol + caffeine PO 6 (4.8–6.5)b 41 High 189

Single studies
Diclofenac sodium PO 7 (4.5–14.7)a 50 Low 186

Diclofenac sodium PΟ 10 (5.7–33.9)a 122 Low 186

Naproxen PO 18c 53 Low 185

Metoclopramide + diphenhydramine IV 6 (3–66)d NA Low 264

Pethidine IM 7 (4.5–14.7)a NA Low 265

Chlorpromazine IV 2 (1.4–3.5)e NA Low 265

Metoclopramide IV 2 (1.3–2.6)e No difference Low 265

Quality of evidence is based on the results of the meta-analyses, otherwise calculated as follows: one positive RCT = low, two
positive RCTs = moderate, three or more positive RCTs = high-quality evidence). IM, intramuscular; IV, intravenous; NA, not
applicable; NNT, number of patients needed to treat for a patient to have a beneficial outcome; NNH, number of patients needed
to treat for a patient to have an adverse event; PO, per os; RCT, randomized controlled trial; TTH, tension-type headache. aPain
free at 2 h post-dose was the beneficial outcome. bNNT is the number of headache attacks that are needed to treat for a headache
attack to have a complete pain-free outcome. cWhen NNTs or NNHs extend from a negative absolute risk reduction the 95%
confidence intervals are not presented. dNNT is the number of patients who would need to be treated with the more efficacious
medication (metoclopramide + diphenhydramine) instead of the ketorolac for a single patient to achieve headache freedom in
the emergency department without requiring rescue medication and maintain headache freedom for 24 h. ePrevention of the
use of a rescue medication 2 h post-dose was the beneficial outcome.

weight gain, dry mouth, dizziness, sweating, constipation effects include drowsiness, weight gain and somnolence
and increased appetite. Amitriptyline is the first-choice but mirtazapine has a better tolerability profile than ami-
drug for the prevention of TTH in elderly and younger triptyline or other tricyclic antidepressants and may be
patients, following the same guidelines, although indi- recommended in specific subpopulations, such as elderly
viduals with cardiac arrhythmias may be excluded. All patients.
elderly patients need to be monitored with electrocardi-
ography at baseline, with every dosage adjustment and Venlafaxine. The evidence for efficacy of venlafaxine
at their yearly examination, along with monitoring for (serotonin–noradrenaline reuptake inhibitor) for TTH
potential anticholinergic adverse effects199. In addition, prevention is low, with only one small RCT evaluating
amitriptyline may cause cognitive impairment or urinary this drug for TTH203 and supporting its level B rating
retention. An observational clinic-based study found that by EFNS-TF181. As with mirtazapine, the mechanism of
patients using prophylactic amitriptyline for TTH had action of venlafaxine is uncertain, but the drug works
higher headache frequency, higher headache burden, independently of depression comorbidity and the
worse sleep quality and more severe depression than mechanism is likely similar to that of of mirtazapine181.
those who did not receive medicinal prevention, because Nausea, vomiting, dizziness and abdominal pain are the
patients with these particular features considered them- most common adverse events, although this drug has
selves to benefit more200. By comparison, lower effec- better tolerability than amitriptyline181.
tiveness of prophylactic amitriptyline was reported by
patients with TTH with widespread pain hyperalgesia200. Other agents. Third-line treatments with label B reco­
mmendation by EFNS-TF are clomipramine (tricyclic
Mirtazapine. Two randomized controlled trials (RCTs)201,202 antidepressant), maprotiline (tetracyclic antidepressant)
have evaluated the efficacy of mirtazapine (a serotoninergic/ and mianserin (atypical antidepressant)181. One system-
noradrenergic antidepressant) for TTH prevention, sup- atic meta-analysis revealed limited scientific evidence
porting its level B rating by EFNS-TF181. The efficacy of for the use of SSRIs and venlafaxine for the prevention
mirtazapine was comparable to that of amitriptyline in of TTH in adults197,204. In addition, tizanidine (a muscle
one study202 and significantly superior to placebo in the relaxant) has been tested for CTTH prophylaxis in two
other one201. The mechanism of action remains unclear placebo-controlled studies205,206, but contradictory results
and may involve the inhibitory descending noradren- were reported, probably because of the strong placebo
ergic and serotoninergic pathways201. Common adverse response in the negative study206. Thus, tizanidine might

12 | Article citation ID: (2021) 7:24 www.nature.com/nrdp

0123456789();:
Primer

Cognitive behavioural
be useful in practice for the prevention of CTTH when preventive treatment at the start of withdrawal therapy
therapy other agents fail. increases the chance of MOH cure214. Regardless of
A time-limited One trial207 evaluated the combination of pindolol the type of treatment, 25–35% of patients with MOH
psychotherapeutic approach (a β-blocker) and amitriptyline compared with placebo relapse, requiring special care and careful monitoring
that focuses on assessing
or amitriptyline alone, and found a higher effectiveness for prevention215. Patients who relapse, those with exces-
and modifying dysfunctional
moods and thoughts (cognitive) with pindolol and amitriptyline combination therapy sive medication overuse or those with use of opioid or
and actions and behaviours than with placebo, but no higher effectiveness than opioid-containing analgesics should be managed by a
(behavioural) and replacing amitriptyline207. Topiramate, valproate and buspirone multidisciplinary team including headache specialists,
them with strategies to help
were effective in individual open label studies208–210, and physiotherapists and psychiatrists or psychologists,
patients reach their goals.
may serve as alternative medicinal options for CTTH while hospitalization for detoxification may be needed
Biofeedback prevention when other agents have failed, were poorly in some severe cases169,195.
A psychotherapeutic approach tolerated or were contraindicated. Evidence suggests that
in which physiological data are botulinum toxin type A and ibuprofen are ineffective or Non-pharmacological treatments
fed back to the user in the form
even harmful for the prevention of CTTH181. All individuals with TTH benefit from education includ-
of visual and auditory stimuli,
and in the case of tension-type ing education about healthy lifestyle habits that help
headache there is a focus on Medication overuse TTH management. Additional non-pharmacological
recognizing activation of the Medication overuse and MOH can occur in those with therapies with good evidence and safety for the man-
nervous system and learning
TTH, although at a lower prevalence than in those agement of TTH include physical and occupational ther-
techniques to calm and relax
the nervous system.
with migraine, complicating the headache features apies, behavioural therapies (cognitive behavioural therapy
and worsening the prognosis. Up to 20% of patients (CBT), biofeedback and relaxation therapy), complemen-
Relaxation therapy with MOH have TTH as underlying primary headache tary and integrative medicine (acupuncture and massage)
A psychological intervention disorder28,169,211,212. In those with MOH and TTH, the key and lifestyle enhancements, including management of
that teaches various
elements of management may include patient education sleep, healthy diet and hydration, stress management and
approaches to relaxing the
nervous system including and cognitive behavioural approaches both to educate regular exercise216. EFNS guidelines recommend that
paced diaphragmatic the patient and to help manage anxiety, management non-pharmacological treatments are considered before
breathing, imagery, of risk factors for MOH (such as high frequency of pharmacological therapy181. Non-pharmacological ther-
distraction and meditation.
headache attacks, comorbidity with anxiety disorders apies should be considered as appropriate based on the
and depression, and the use of specific medications for evidence, patient preference and in situations where
acute treatment)169, withdrawal of overused medica- some pharmacological options are contraindicated
tions and prophylactic treatment along with manage- (such as during pregnancy or lactation). They can be
ment of the potential comorbidities195,213. Withdrawal combined with pharmacotherapy or administered on
of the overused medications can be abrupt or gradual their own, and multiple non-pharmacological therapies
depending on the drug used (for example, opioids and can be administered concurrently or sequentially.
opioid-containing analgesics should be gradually dis-
continued whereas simple analgesics and NSAIDs can be Behavioural therapies. Biobehavioural therapies (that
abruptly discontinued), patient preferences and idiosyn- is, CBT, biofeedback and relaxation therapy) are effec-
crasy, and comorbidities211. The use of pharmaceutical tive and safe treatment options for TTH217. Indeed,

Table 2 | A selection of prophylactic treatments for TTH tested in randomized placebo- or sham-controlled trials
Treatment by study Route NNT NNH Overall quality Refs
of evidence
EFNS Task Force study
Amitriptyline PO NA (30% reduction)a 3 (1.6–3.9) High 181

Amitriptyline PO 12b 2 (1.6–2.6) High 139,181

Single studies
Mirtazapine PO NA (34% reduction)a 4 (2.4–13.0) Moderate 201,266

SMT NA 18b No difference Low 218

Amitriptyline + SMT NA 3 (1.9–6.0)c 2 (1.6–2.6) Low 218

Venlafaxine PO 4 (2.0–14.2)d 6b Low 203

Meta-analysis
Acupuncture NA 11 (5.8–41.5)d 20b High 238

Quality of evidence is based on the results of the meta-analyses, otherwise calculated as follows: one positive RCT = low, two positive
RCTs = moderate, three or more positive RCTs = high quality of evidence. EFNS, European Federation of Neurological Societies;
NA, not applicable; NNH, number of patients needed to treat for a patient to have an adverse event; NNT, number of patients needed
to treat for a patient to have a beneficial outcome; PO, per os; RCT, randomized controlled trial; SMT: stress management treatment,
or cognitive behavioural treatment; TTH, tension-type headache. aNNT cannot be calculated, reduction of the area under the curve
vs placebo was the beneficial outcome. bWhen NNTs or NNHs extend from a negative absolute risk reduction the 95% confidence
intervals are not presented. c>50% reduction of the headache index after treatment was the beneficial outcome. d>50% reduction
in monthly headache days after treatment was the beneficial outcome.


NATURE REVIEWS | DISEASE PRIMERS | Article citation ID: (2021) 7:24 13

0123456789();:
Primer

Stress management therapy


one meta-analysis including 53 studies of biofeedback their patients that excessive reliance on trigger factors
A combination of relaxation for TTH among adults, adolescents and children found is not evidence based and could result in unnecessary
with cognitive coping skills for a significant medium-to-large effect size (Cohen’s d 0.73; avoidance behaviour.
preventing and managing 95% CI 0.61–0.84) that improved the frequency of head- Advice about trigger management (which is different
stress and headaches.
ache episodes, muscle tension, self-efficacy, symptoms from the traditional advice to avoid all triggers), as well
Headache triggers of anxiety and depression, and analgesic medication as therapy with graduated exposure to selected triggers to
Factors associated with an intake over an average follow-up phase of 15 months. promote desensitization (known as learning to cope with
increased probability of Stress management therapy showed strong evidence of triggers, or LCT therapy) are also effective in reducing
headache over a relatively
efficacy in a randomized and placebo-controlled study, monthly headache days in patients with TTH221.
short, clinically relevant
time period.
and seemed to have equivalent efficacy to amitripty- Growing evidence highlights the association of
line or nortriptyline for CTTH prevention218. A com- TTH with various sleep disorders, including exces-
bination of the two therapies (stress management plus sive daytime sleepiness, insomnia, poor sleep quality,
either amitriptyline or nortriptyline) may improve the insufficient sleep and shift work28,123,227,228. This comor-
outcome of monotherapy218. Maladaptive coping strat- bidity increases the risk of TTH chronification and it
egies such as catastrophizing and avoidance are more is more prevalent in CTTH than in ETTH123. Thus,
common in patients with TTH than in controls, with it is plausible that improving sleep quality will subse-
a higher frequency in patients with CTTH than in quently reduce headache activity. One meta-analysis
those with ETTH219,220. These are addressed by CBT221. of three trials found that psychological sleep interven-
Additionally, long-term group behavioural treatment tions (for example, sleep restriction/bed restriction,
(psychoeducation, progressive muscle relaxation, cop- stimulus control and sleep hygiene, which encom-
ing strategies for pain and stress, and goal-setting skills) passes optimal lifestyle and environmental factors
with pharmacotherapy showed promising results in for sleep) improve headache frequency and sleep,
reducing headache frequency and related disability although the effect on headache intensity differs
in patients with TTH222. Evidence is growing for the use between studies229. In addition to these interventions,
of mindfulness-based therapies (including mindfulness medications that improve sleep — such as amitriptyline
based cognitive therapy (MBCT) and mindfulness based or mirtazapine — may be considered for patients who
stress management, which combine the mindful practice report sleep disturbances200.
of focused attention without judgement, mindfulness
meditation and cognitive therapy, relaxation therapy Physical therapies. Physical therapy, aerobic exercise,
and stress management therapy techniques) and accept- strength training, massage and cervical manipulation
ance and commitment therapy (a form of behavioural in particular may be beneficial in the management
therapy that combines mindfulness skills with the devel- of TTH230–233. In an open randomized trial, strength
opment of psychological flexibility and the practice of training with patient education decreased the average
self-acceptance)223. monthly headache frequency from 24.5 to 20.5 days in
female adolescents with TTH, but no significant dif-
Trigger management/prevention ferences between treatment groups were detected230.
Some patients and clinicians search for headache triggers In another controlled trial, treatment with manipula-
that can be used as therapeutic strategy by recommend- tion and massage or massage alone significantly reduced
ing avoidance behaviour. Triggers are typically identi- the Headache Disability Inventory score (a 25-item
fied by patient self-report. One study of 344 people with headache disability inventory assessing the impact
TTH in China found that 67.4% of patients reported of headache in daily living) in patients with TTH233,
trigger factors, of which the most common were sleep and combined treatment significantly decreased head-
disturbance, negative affect and sunlight exposure224. ache frequency compared with the massage-only group.
Other frequently reported triggers are stress, distur- Several meta-analyses have found that patients with
bances in self-care and routines, dehydration, certain TTH benefit from physical therapy234–237.
foods and drinks (such as alcohol), travel, high alti-
tude, activity or exertion (for example, various specific Acupuncture. One systematic meta-analysis found
movements or postures) and other activities that disrupt that acupuncture is an effective and safe treatment for
healthy equilibrium225. frequent ETTH or CTTH (leading to a reduction of
Patient-reported triggers rely on patient beliefs but monthly headaches by >50% reported by 51% of patients
not on an actual association between the trigger and the treated with acupuncture versus 43% of patients treated
onset of headache. Headache triggers are best studied with sham acupuncture), although further trials com-
using electronic diaries or in randomized trials226. With paring acupuncture with other treatment options are
electronic diaries, exposure to triggers is recorded before needed238. Despite several methodological shortcom-
the onset of headache, and the risk of headache follow- ings related to placebo effect and blinding of the trials,
ing the trigger is compared with headache risk at other which are applicable to all RCTs to some extent, the qual-
times. In randomized trials, the probability of headache ity of evidence for acupuncture is moderate to high, as
following blinded exposure to the trigger is compared with amitriptyline. Acupuncture has a clear advantage
with blinded exposure to the placebo. However, there is over amitriptyline in terms of the safety profile, with the
a scarcity of studies that have ascertained whether some number of patients needed to treat for a patient to have
factors are headache triggers in susceptible individuals an adverse event of 20 with acupuncture versus 2–4 for
with TTH. Thus, clinicians are encouraged to inform amitriptyline (Table 2).

14 | Article citation ID: (2021) 7:24 www.nature.com/nrdp

0123456789();:
Primer

Transformed migraine
Interventional therapies. Anaesthetic injections at peric- The consequences of TTH have received less attention
Subtype of the chronic daily ranial myofascial trigger points may also be effective for than those of migraine owing to the greater burden of
headache, not defined by TTH in terms of reducing monthly painful days239. In migraine at the individual level, although the very high
ICHD, that initially begins as addition, there is moderate evidence that myofascial prevalence of TTH amplifies its effects from a soci-
episodic migraine and then
trigger point dry needling is effective in patients with etal perspective248. While migraine-specific HRQOL
evolves over months or years
to daily or near-daily chronic TTH (in terms of a reduction in headache scales have been developed to measure the effect of
headaches resembling a intensity, frequency and duration, or improvement in migraine on a patient’s life, no disease-specific HRQOL
mixture of tension-type functional and sensory outcomes)240,241. Dry needling, measures for TTH are available. The consequences of
headache and migraine.
or intramuscular stimulation, is an alternative medicine TTH on the individual include impaired HRQOL,
technique, adapted from acupuncture, that is designed psychological comorbidities, loss of productivity and
to stimulate trigger points or muscles that are irrita- economic costs249,250.
ble by inserting needles into the skin and thus treat Previous studies have found that patients with
dysfunction of skeletal muscle and connective tissue CTTH have a poor quality of life with a generalized
(American Physical Therapy Association). impairment in functioning compared with the gen-
eral population and patients with episodic migraine251.
Pharmacoeconomics Similarly, global impairment measured by General
The indirect and direct costs of TTH owing to medical Health Questionnaire-28 was greater for CTTH than for
services and medications have been investigated in only ETTH31. Moreover, headache frequency is associated with
a few studies. In the Eurolight project study, the indirect reduced HRQOL and disability248. A population-based
and direct average annual costs per person were €303 for survey in the Republic of Georgia, using the Short Form
TTH in Europe (of which indirect costs contributed to (36) Health Survey (SF-36), found that people with-
92% of the estimate) and the total annual cost of TTH out headache had higher scores on all eight subscales
amongst adults aged 18–65 years was €21 billion242. indicating Physical Component Summary (PCS) and
Compared with migraine, patients with TTH may Mental Component Summary (MCS) than those with
seek medical attention less frequently to manage their episodic headache (coexistent migraine and TTH) or
headache. Indeed, the consultation rate (the proportion chronic headache (unspecified headache for ≥15 days/
of individuals with a given disease consulting or seek- month)252. However, no differences were found between
ing care from a physician) in TTH is between 16% and respondents with episodic headache and those with
44%32,243–246. In a Danish population study, only 16% chronic headache or between individuals with migraine
of patients with TTH contacted their general practi- and those with TTH252. In another population-based
tioner, compared with 56% of people with migraine32. study from Spain, using SF-36, no difference in HRQOL
However, when corrected for the higher prevalence was found between CTTH and transformed migraine25,253.
of TTH, the total health-care utilization was found to Of note, as HRQOL declines with age and the CTTH
be 54% higher for TTH compared with migraine32,247. group was older, age could confound this finding. In a
In a community-based population study from Chile, the population study from Denmark, using the SF-12 sur-
consultation rate in individuals with TTH was 39% com- vey, both PCS and MCS scores were lower for individ-
pared with 63% in those with migraine245. Younger age uals with coexistent headache (TTH and migraine),
and moderate to severe headache intensity were associ- followed by pure TTH and pure migraine than for those
ated with likelihood of consultation245. Of note, the head- without headache when adjusted for age, sex and educa-
ache consultation rate in individuals with pure frequent tion level254. Moreover, both PCS and MCS scores were
TTH (not coexistent with migraine) in 2001 (44%) was lower for patients with CTTH followed by ETTH, than
higher than in 1989 (33%) in a longitudinal Danish popu­ for the no headache group. Lower PCS scores in TTH
lation study244. The proportion of individuals with TTH were associated with age, female sex and poor self-rated
using prescription medications increased from 2.2% in health, and lower MCS-12 scores in TTH were associated
1989 to 7% in 2001, but the proportion of individuals with depression254.
using over-the-counter medications (OTC) remained Patients with CTTH had numerically lower SF-36
high (93% in 1989 versus 90.4% in 2001) 244. Only scores in six out of eight domains than those with
~1% of individuals with TTH were given preventive migraine, indicating worse HRQOL, in a large study
medications244. OTC simple analgesics and NSAIDs are in an outpatient tertiary headache clinic in Taiwan255.
the most frequently used acute headache-abortive med- HRQOL scores were generally higher for CTTH than for
ications in TTH owing to the absence of disease-specific transformed migraine (patients who met the criteria
treatments32,243,244,247. Although these therapies can pro- for both CTTH and transformed migraine were classi-
vide symptomatic relief, they may lead to MOH when fied as transformed migraine in the study). In another
used frequently, and subsequently increase health-care study of 209 patients from a tertiary headache clinic256
utilization and medical costs for the individual with using the SF-20 survey, ‘poor health’ was associated with
TTH and indirect costs for society247,248. social functioning, and mental health scales were greater
for TTH than for migraine257.
Quality of life
Health-related quality of life Disability
Although TTH is the most common neurologi- Owing to the moderate severity of headache in the
cal disorder worldwide, its effects on health-related majority of patients with CTTH, impaired function
quality of life (HRQOL) have been under-studied249. may not lead to absenteeism; only a minority of patients


NATURE REVIEWS | DISEASE PRIMERS | Article citation ID: (2021) 7:24 15

0123456789();:
Primer

Box 4 | Patient advocate perspective frequently coexists with TTH, and this interaction
needs to be studied further in epidemiological studies167.
A lack of understanding is the lived reality of most people who are disabled by headache Moreover, more studies on the incidence of TTH and
disorders, such as tension-type headache (TTH). Why would you want to tell your family associative and influencing factors are warranted. The
and co-workers that you are disabled (temporarily or more persistently), if you have identification of risk factors for transformation from
come to learn that you will be immediately dismissed as a complainer? This stigma has
ETTH to CTTH is important. In particular, modifi­able
a profound negative effect and often negates the plea from patients that more research
and better care are warranted. We do not want to miss out on life. We do not get to risk factors are of great value and can help to inform
choose when to feel well and when to spend our days in bed. We dream of contributing better clinical practice260. Together, incidence and prog-
to society without having to be disabled by recurrent episodes of headache. nostic longitudinal studies can help us to develop pre-
Thus, we need more research into the biological underpinnings of TTH. This will ventive measures for TTH. Epidemiological studies
facilitate identification of targeted therapies that are effective, well-tolerated and safe. should also quantify indirect consequences of TTH to
We need policy-makers and funding bodies to acknowledge individuals living with provide accurate total cost estimates, which may include
headache diseases and to better appreciate the considerable socioeconomic impact effect on family life (for example, partner relation-
of TTH. As a patient advocate myself, I try to impart to people who are disabled by ships and childcare) and career potential (such as sick
headache disorders the importance of speaking up and speaking out. The negative days and productivity losses at work).
effects and the personal stories must be shared to break down the ingrained inter-personal
and structural stigma. This will ultimately result in allocation of more resources for
research, which is integral to the development of better therapies, pharmacological as Pathophysiology
well as lifestyle-based. Future work should also target candidate genes asso-
Katie MacDonald, Alliance for Headache Disorders Advocacy ciated with various TTH phenotypes. Studies of larger
samples of carefully phenotyped groups with TTH
using unbiased approaches such as whole-genome
with CTTH report their work performance and social sequencing, whole-exome sequencing or genome-wide
functioning to be severely impaired258,259. Accordingly, association studies may generate additional insights
although patients with migraine probably report into the genetic basis of TTH. Roles of other neuro-
more actual lost workdays than those with TTH, TTH chemical peptides in the pathophysiology of TTH are
accounts for a considerable loss of work effectiveness still inconclusive. Future investigations should take
due to headache32,247. the type of TTH (infrequent ETTH, frequent ETTH
The effect of TTH is higher for those with TTH or CTTH) and the effect of comorbid migraine into
with substantial and complicating comorbidities such consideration. How emotional factors (such as psycho-
as anxiety and depression than for patients without logical stress) can activate the mechanisms of head pain
comorbidities37,248,250. Although comorbid depression is unclear, and future studies are warranted. Moreover,
can generally be mild to moderate, patients with TTH firm evidence is needed to better understand the role
might have high levels of anxiety, which substantially of peripheral myofascial factors in the genesis of pain
impairs functioning258. Indeed, in a small study with in TTH. Further research, including larger samples of
25 patients with CTTH referred from a neurology clinic, patients with TTH, is also needed to better elucidate
anxiety had a mediating effect and depression had a the role of vascular inputs in TTH pathophysiology.
modulating effect between headache frequency and These studies should include provocation studies
intensity, and reduced QOL (mental health and social with neurochemical signalling compounds, MRI and
functioning) measured with SF-36 (ref.250). However, angiography to explore differences between those
a major limitation of this study was the highly selective with frequent ETTH and those with migraine, as well
study samples and the small sample size. In another as healthy non-headache controls. This may help us
small study of 89 patients with TTH, depression scores to determine whether some patients have a predom-
positively correlated with poor QOL in patients with inantly myofascial TTH pathophysiology, while oth-
CTTH. In addition, patients with CTTH scored lower ers have a predominantly vascular/neuronal TTH
on the Nottingham Health Profile (NHP), an everyday pathophysiology.
life questionnaire than patients with ETTH220. However,
patients in the CTTH group in this study were older than Diagnosis
those in the ETTH group. The diagnosis of TTH is still purely clinical, as reliable
and robust biomarkers are lacking. Similar to migraine,
Outlook there is a growing urge to identify reliable serum and
Despite advances in basic and clinical research in TTH, clinical biomarkers useful to diagnose primary head-
this primary headache remains neglected (Box 4). There aches including TTH, monitor their activity and deter-
are many unanswered questions about the epidemiol- mine the response to treatment. Circulating serum
ogy, impact, diagnosis, pathophysiology and treatment biomarkers such as CGRP have been proposed as
of TTH. Answering these questions may lead to the diagnostic and therapeutic tools in primary headaches.
development of more effective and mechanism-based CGRP is thought to be a potential and promising bio-
treatments. marker of migraine and cluster headache261. Prospective
daily headache diary studies in patients with TTH are
Epidemiology needed for better characterization of the disease and
Future epidemiological studies should use rigorous identification of clinical biomarkers, and would also be
methodology, and better and uniform definitions of valuable in ascertaining how often patients with CTTH
TTH, including the use of ICHD criteria. Migraine have a secondary diagnosis of MOH.

16 | Article citation ID: (2021) 7:24 www.nature.com/nrdp

0123456789();:
Primer

Management pharmacological and non-pharmacological therapies


The treatment of TTH remains mainly disease non- owing to cost, availability of providers and time to par-
specific. There has been interest in the design of NOS ticipate. Telemedicine consultation for non-acute head-
inhibitors with therapeutic purposes and these drugs ache is useful and non-inferior to traditional in-person
may be tested in CTTH. Future studies should eluci- clinic evaluation263. Using mobile health platforms and
date the role of triptans in CTTH, as triptans may block telehealth may increase access to these therapies by low-
central sensitization262. CGRP monoclonal antibodies ering cost and increasing convenience. Research needs
should be investigated for TTH prevention. Behavioural to be conducted to determine the efficacy and feasibility
and physical therapies have good evidence for their use of alternative modes of delivery.
alone or in combination with pharmacotherapy. There
Published online xx xx xxxx
are significant problems with access to some of the

1. GBD 2016 Headache Collaborators. Global, regional, 16. Winkler, A. et al. The prevalence of headache and episodic tension-type headache on the quality
and national burden of migraine and tension-type with emphasis on tension-type headache in rural of life and performance of a university student
headache, 1990-2016: a systematic analysis for the Tanzania: a community-based study. Cephalalgia 29, population. Headache 41, 710–719 (2001).
Global Burden of Disease Study 2016. Lancet Neurol. 1317–1325 (2009). 35. Lipton, R. B., Cady, R. K., Stewart, W. F., Wilks, K. &
17, 954–976 (2018). 17. Leonardi, M. et al. Global Burden of Headache Hall, C. Diagnostic lessons from the Spectrum study.
An important study that used data from the Disorders in Children and Adolescents 2007-2017. Neurology 58, S27–S31 (2002).
GBD 2016 study to provide global, regional, Int. J. Environ Res. Public Health 18, 250 (2020). 36. Lipton, R. B. et al. 2000 Wolfe Award. Sumatriptan
and national estimates for prevalence and years 18. Pascual, J., Colas, R. & Castillo, J. Epidemiology of for the range of headaches in migraine sufferers:
of life lived with disability for migraine and TTH. chronic daily headache. Curr. Pain Headache Rep. 5, results of the Spectrum Study. Headache 40,
2. Deuschl, G. et al. The burden of neurological diseases 529–536 (2001). 783–791 (2000).
in Europe: an analysis for the Global Burden of 19. Gobel, H., Petersen-Braun, M. & Soyka, D. The 37. Jensen, R. & Stovner, L. J. Epidemiology and
Disease Study 2017. Lancet Public. Health 5, epidemiology of headache in Germany: a nationwide comorbidity of headache. Lancet Neurol. 7, 354–361
e551–e567 (2020). survey of a representative sample on the basis of (2008).
Updated data from the GBD 2017 providing the headache classification of the International 38. Ashina, S. et al. Neuroticism, depression and pain
global and European estimates for incidence, Headache Society. Cephalalgia 14, 97–106 (1994). perception in migraine and tension-type headache.
prevalence and years of life lived with disability 20. Wang, S. J. et al. Chronic daily headache in Chinese Acta Neurol. Scand. 136, 470–476 (2017).
for migraine and TTH. elderly: prevalence, risk factors, and biannual 39. Ødegård, S. S., Sand, T., Engstrøm, M., Zwart, J. A.
3. GBD 2017 US Neurological Disorders Collaborators. follow-up. Neurology 54, 314–319 (2000). & Hagen, K. The impact of headache and chronic
Burden of neurological disorders across the US from 21. Lavados, P. M. & Tenhamm, E. Epidemiology of musculoskeletal complaints on the risk of insomnia:
1990-2017: a Global Burden Of Disease Study. tension-type headache in Santiago, Chile: a prevalence longitudinal data from the Nord-Trøndelag health
JAMA Neurol. 78, 165–176 (2020). study. Cephalalgia 18, 552–558 (1998). study. J. Headache Pain 14, 24 (2013).
Updated data from the GBD 2017 providing 22. Pryse-Phillips, W. et al. A Canadian population survey 40. Ødegård, S. S. et al. Associations between sleep
estimates for incidence, prevalence and years on the clinical, epidemiologic and societal impact disturbance and primary headaches: the third
of life lived with disability for migraine and of migraine and tension-type headache. Can. J. Nord-Trøndelag Health Study. J. Headache Pain
TTH in the United States. Neurol. Sci. 19, 333–339 (1992). 11, 197–206 (2010).
4. Headache Classification Committee of 23. Ferrante, T. et al. The PACE study: past-year 41. Kim, J. et al. Insomnia in tension-type headache:
the International Headache Society (IHS). The prevalence of tension-type headache and its subtypes a population-based study. J. Headache Pain 18,
International Classification of Headache Disorders, in Parma’s adult general population. Neurol. Sci. 36, 95 (2017).
3rd edition. Cephalalgia 38, 1–211 (2018). 35–42 (2015). 42. Song, T. J. et al. Anxiety and depression in
This is the third and the latest edition of the 24. Wang, S. J., Fuh, J. L., Lu, S. R. & Juang, K. D. tension-type headache: a population-based study.
ICHD-3, which is used as a diagnostic tool for Chronic daily headache in adolescents: prevalence, PLoS ONE 11, e0165316 (2016).
headache disorders including TTH. impact, and medication overuse. Neurology 66, 43. Mitsikostas, D. D. & Thomas, A. M. Comorbidity
5. Jensen, R. H. Tension-type headache - the normal and 193–197 (2006). of headache and depressive disorders. Cephalalgia
most prevalent headache. Headache 58, 339–345 25. Silberstein, S. D., Lipton, R. B. & Sliwinski, M. 19, 211–217 (1999).
(2018). Classification of daily and near-daily headaches: 44. Fuensalida-Novo, S. et al. The burden of headache
6. Lyngberg, A. C., Rasmussen, B. K., Jorgensen, T. & field trial of revised IHS criteria. Neurology 47, is associated to pain interference, depression and
Jensen, R. Incidence of primary headache: a Danish 871–875 (1996). headache duration in chronic tension type headache:
epidemiologic follow-up study. Am. J. Epidemiol. 161, 26. Rasmussen, B. K. Migraine and tension-type headache a 1-year longitudinal study. J. Headache Pain 18,
1066–1073 (2005). in a general population: precipitating factors, female 119 (2017).
One of the few population-based studies on the hormones, sleep pattern and relation to lifestyle. 45. Minen, M. T. et al. Migraine and its psychiatric
incidence of TTH in the general population. Pain 53, 65–72 (1993). comorbidities. J. Neurol. Neurosurg. Psychiatry 87,
7. Lu, S. R. et al. Incidence and risk factors of chronic 27. Wang, S. J., Fuh, J. L., Lu, S. R. & Juang, K. D. 741–749 (2016).
daily headache in young adolescents: a school cohort Outcomes and predictors of chronic daily headache 46. Buse, D. C., Silberstein, S. D., Manack, A. N.,
study. Pediatrics 132, e9–e16 (2013). in adolescents: a 2-year longitudinal study. Neurology Papapetropoulos, S. & Lipton, R. B. Psychiatric
8. United Nations, Department of Economic and 68, 591–596 (2007). comorbidities of episodic and chronic migraine.
Social Affairs Population Division. World population 28. Lyngberg, A. C., Rasmussen, B. K., Jorgensen, T. & J. Neurol. 260, 1960–1969 (2013).
prospects 2017 – data booklet (ST/ESA/SER.A/401), Jensen, R. Prognosis of migraine and tension-type 47. Yoon, M. S. et al. Chronic migraine and chronic
https://www.un.org/development/desa/pd/sites/www. headache: a population-based follow-up study. tension-type headache are associated with concomitant
un.org.development.desa.pd/files/files/documents/ Neurology 65, 580–585 (2005). low back pain: results of the German Headache
2020/Jan/un_2017_world_population_prospects- 29. Ashina, S., Lyngberg, A. & Jensen, R. Headache Consortium study. Pain 154, 484–492 (2013).
2017_revision_databooklet.pdf (2017). characteristics and chronification of migraine and 48. Ashina, S. et al. Increased pain sensitivity in migraine
9. U.S. Census Bureau, https://www.census.gov (2021). tension-type headache: a population-based study. and tension-type headache coexistent with low back
10. Yu, S. et al. The prevalence and burden of primary Cephalalgia 30, 943–952 (2010). pain: a cross-sectional population study. Eur. J. Pain
headaches in China: a population-based door-to-door 30. GBD 2015 Disease and Injury Incidence and 22, 904–-914 (2018).
survey. Headache 52, 582–591 (2012). Prevalence Collaborators. et al. Global, regional, 49. Ashina, S. et al. Prevalence of neck pain in migraine
11. Alzoubi, K. H. et al. Prevalence of migraine and and national incidence, prevalence, and years and tension-type headache: a population study.
tension-type headache among adults in Jordan. lived with disability for 310 diseases and injuries, Cephalalgia 35, 211–219 (2015).
J. Headache Pain 10, 265–270 (2009). 1990-2015: a systematic analysis for the Global 50. Ulrich, V., Gervil, M. & Olesen, J. The relative influence
12. Schwartz, B. S., Stewart, W. F., Simon, D. & Burden of Disease Study 2015. Lancet 388, of environment and genes in episodic tension-type
Lipton, R. B. Epidemiology of tension-type headache. 1545–1602 (2016). headache. Neurology 62, 2065–2069 (2004).
JAMA 279, 381–383 (1998). 31. Cassidy, E. M., Tomkins, E., Hardiman, O. & O’Keane, V. 51. Russell, M. B., Levi, N. & Kaprio, J. Genetics of
13. Rasmussen, B. K., Jensen, R., Schroll, M. & Olesen, J. Factors associated with burden of primary headache in tension-type headache: a population based twin
Epidemiology of headache in a general population - a a specialty clinic. Headache 43, 638–644 (2003). study. Am. J. Med. Genetics. B Neuropsychiatr. Genet.
prevalence study. J. Clin. Epidemiol. 44, 1147–1157 32. Rasmussen, B. K., Jensen, R. & Olesen, J. Impact 144b, 982–986 (2007).
(1991). of headache on sickness absence and utilisation of 52. Russell, M. B., Saltyte-Benth, J. & Levi, N. Are
14. Stovner, L. J. et al. The global burden of headache: medical services: a Danish population study. infrequent episodic, frequent episodic and chronic
a documentation of headache prevalence and J. Epidemiol. Community Health 46, 443–446 tension-type headache inherited? A population-based
disability worldwide. Cephalalgia 27, 193–210 (1992). study of 11 199 twin pairs. J. Headache Pain 7,
(2007). 33. Rasmussen, B. K. Migraine and tension-type 119–126 (2006).
15. Russell, M. B., Levi, N., Saltyte-Benth, J. & headache in a general population: psychosocial 53. Ostergaard, S., Russell, M. B., Bendtsen, L. &
Fenger, K. Tension-type headache in adolescents factors. Int. J. Epidemiol. 21, 1138–1143 (1992). Olesen, J. Comparison of first degree relatives
and adults: a population based study of 33,764 twins. 34. Bigal, M. E., Bigal, J. M., Betti, M., Bordini, C. A. and spouses of people with chronic tension headache.
Eur. J. Epidemiol. 21, 153–160 (2006). & Speciali, J. G. Evaluation of the impact of migraine Br. Med. J. 314, 1092–1093 (1997).


NATURE REVIEWS | DISEASE PRIMERS | Article citation ID: (2021) 7:24 17

0123456789();:
Primer

54. Park, J. W., Kim, J. S., Lee, H. K., Kim, Y. I. & Lee, K. S. 76. Arendt-Nielsen, L., Castaldo, M., Mechelli, F. & 102. Langemark, M., Jensen, K., Jensen, T. S. & Olesen, J.
Serotonin transporter polymorphism and harm Fernandez-de-Las-Penas, C. Muscle triggers as a Pressure pain thresholds and thermal nociceptive
avoidance personality in chronic tension-type possible source of pain in a subgroup of tension-type thresholds in chronic tension-type headache. Pain 38,
headache. Headache 44, 1005–1009 (2004). headache patients? Clin. J. Pain 32, 711–718 (2016). 203–210 (1989).
55. Fernández-de-las-Peñas, C. et al. Genetic contribution 77. Bezov, D., Ashina, S., Jensen, R. & Bendtsen, L. 103. Gobel, H., Weigle, L., Kropp, P. & Soyka, D. Pain
of catechol-O-methyltransferase polymorphism Pain perception studies in tension-type headache. sensitivity and pain reactivity of pericranial muscles in
(Val158Met) in children with chronic tension-type Headache 51, 262–271 (2011). migraine and tension-type headache. Cephalalgia 12,
headache. Pediatr. Res. 70, 395–399 (2011). 78. Bendtsen, L. & Fernandez-de-la-Penas, C. The role 142–151 (1992).
56. Fernández-de-las-Peñas, C. et al. Catechol-O- of muscles in tension-type headache. Curr. Pain 104. Metsahonkala, L. et al. Extracephalic tenderness
methyltransferase (COMT) rs4680Val158Met Headache Rep. 15, 451–458 (2011). and pressure pain threshold in children with headache.
polymorphism is associated with widespread pressure 79. Hubbard, D. R. & Berkoff, G. M. Myofascial trigger Eur. J. Pain 10, 581–585 (2006).
pain sensitivity and depression in women with chronic, points show spontaneous needle EMG activity. Spine 105. Headache Classification Committee of the
but not episodic, tension-type headache. Clin. J. Pain 18, 1803–1807 (1993). International Headache Society. The international
35, 345–352 (2019). 80. Jensen, R., Fuglsang-Frederiksen, A. & Olesen, J. classification of headache disorders: 2nd edition.
57. Gupta, R., Kumar, V., Luthra, K., Banerjee, B. & Quantitative surface EMG of pericranial muscles in Cephalalgia 24 (Suppl. 1), 9–160 (2004).
Bhatia, M. S. Polymorphism in apolipoprotein E headache. A population study. Electroencephalogr. 106. Buchgreitz, L., Lyngberg, A. C., Bendtsen, L. &
among migraineurs and tension-type headache Clin. Neurophysiol. 93, 335–344 (1994). Jensen, R. Frequency of headache is related to
subjects. J. Headache Pain 10, 115–120 (2009). 81. Cathcart, S., Winefield, A. H., Lushington, K. & sensitization: a population study. Pain 123, 19–27
58. Jensen, R. Peripheral and central mechanisms in Rolan, P. Stress and tension-type headache (2006).
tension-type headache: an update. Cephalalgia 23 mechanisms. Cephalalgia 30, 1250–1267 (2010). 107. Fernández-de-Las-Peñas, C. et al. Evidence of localized
(Suppl. 1), 49–52 (2003). 82. Melzack, R. From the gate to the neuromatrix. and widespread pressure pain hypersensitivity in
59. Jensen, R. Pathophysiological mechanisms of Pain Suppl. 6, S121–S126 (1999). patients with tension-type headache: A systematic
tension-type headache: a review of epidemiological 83. Choi, J. C., Chung, M. I. & Lee, Y. D. Modulation of review and meta-analysis. Cephalalgia 41, 265–273
and experimental studies. Cephalalgia 19, 602–621 pain sensation by stress-related testosterone and (2021).
(1999). cortisol. Anaesthesia 67, 1146–1151 (2012). 108. Bendtsen, L. Central sensitization in tension-type
60. Langemark, M. & Olesen, J. Pericranial tenderness 84. Leistad, R. B., Sand, T., Westgaard, R. H., Nilsen, K. B. headache–possible pathophysiological mechanisms.
in tension headache. A blind, controlled study. & Stovner, L. J. Stress-induced pain and muscle Cephalalgia 20, 486–508 (2000).
Cephalalgia 7, 249–255 (1987). activity in patients with migraine and tension-type A comprehensive review on the important
61. Jensen, R., Rasmussen, B. K., Pedersen, B. & headache. Cephalalgia 26, 64–73 (2006). role and mechansms of central sensitization
Olesen, J. Muscle tenderness and pressure pain 85. Hatch, J. P. et al. Electromyographic and affective in the pathophysiology of CTTH.
thresholds in headache. A population study. responses of episodic tension-type headache 109. Castien, R. F., van der Wouden, J. C. & De Hertogh, W.
Pain 52, 193–199 (1993). patients and headache-free controls during stressful Pressure pain thresholds over the cranio-cervical
62. Ashina, S., Bendtsen, L. & Ashina, M. Pathophysiology task performance. J. Behav. Med. 15, 89–112 region in headache: a systematic review and
of tension-type headache. Curr. Pain Headache Rep. 9, (1992). meta-analysis. J. Headache Pain 19, 9 (2018).
415–422 (2005). 86. Mense, S. The pathogenesis of muscle pain. 110. Bendtsen, L., Jensen, R. & Olesen, J. Decreased
63. Jensen, R. & Olesen, J. Initiating mechanisms of Curr. Pain Headache Rep. 7, 419–425 (2003). pain detection and tolerance thresholds in chronic
experimentally induced tension-type headache. 87. Mork, H., Ashina, M., Bendtsen, L., Olesen, J. & tension-type headache. Arch. Neurol. 53, 373–376
Cephalalgia 16, 175–182 (1996). Jensen, R. Possible mechanisms of pain perception (1996).
64. Buchgreitz, L., Lyngberg, A. C., Bendtsen, L. in patients with episodic tension-type headache. 111. Ashina, S. et al. Increased muscular and cutaneous
& Jensen, R. Increased pain sensitivity is not a A new experimental model of myofascial pain. pain sensitivity in cephalic region in patients with
risk factor but a consequence of frequent headache: Cephalalgia 24, 466–475 (2004). chronic tension-type headache. Eur. J. Neurol. 12,
a population-based follow-up study. Pain 137, 88. Bendtsen, L. Central and peripheral sensitization in 543–549 (2005).
623–630 (2008). tension-type headache. Curr. Pain Headache Rep. 7, 112. Ashina, S., Bendtsen, L., Ashina, M., Magerl, W.
65. Fernandez-de-Las-Penas, C., Cuadrado, M. L., 460–465 (2003). & Jensen, R. Generalized hyperalgesia in patients
Arendt-Nielsen, L., Ge, H. Y. & Pareja, J. A. Increased 89. Henriksson, K. & Lindman, R. in Tension-Type with chronic tension-type headache. Cephalalgia 26,
pericranial tenderness, decreased pressure pain Headache: Classification, Mechanisms and Treatment 940–948 (2006).
threshold, and headache clinical parameters in (eds Olesen, J. & Schoenen, J.) 97–107 (Raven Press, 113. Bendtsen, L., Jensen, R. & Olesen, J. Qualitatively
chronic tension-type headache patients. Clin. J. Pain New York, 1993). altered nociception in chronic myofascial pain. Pain
23, 346–352 (2007). 90. Kocer, A., Kocer, E., Memisogullari, R., Domac, F. M. 65, 259–264 (1996).
66. Ashina, M., Bendtsen, L., Jensen, R., Sakai, F. & & Yuksel, H. Interleukin-6 levels in tension headache 114. Cathcart, S., Winefield, A. H., Lushington, K. &
Olesen, J. Muscle hardness in patients with chronic patients. Clin. J. Pain 26, 690–693 (2010). Rolan, P. Noxious inhibition of temporal summation is
tension-type headache: relation to actual headache 91. Ashina, M. et al. Tender points are not sites of ongoing impaired in chronic tension-type headache. Headache
state. Pain 79, 201–205 (1999). inflammation -in vivo evidence in patients with chronic 50, 403–412 (2010).
67. Bendtsen, L., Ashina, S., Moore, A. & Steiner, T. J. tension-type headache. Cephalalgia 23, 109–116 115. Bendtsen, L. & Jensen, R. Tension-type headache.
Muscles and their role in episodic tension-type (2003). Neurol. Clin. 27, 525–535 (2009).
headache: implications for treatment. Eur. J. Pain 20, 92. Langemark, M., Jensen, K. & Olesen, J. Temporal 116. Chen, W. T. et al. Comparison of somatosensory cortex
166–175 (2016). muscle blood flow in chronic tension-type headache. excitability between migraine and “strict-criteria”
A review on peripheral and muscular factors Arch. Neurol. 47, 654–658 (1990). tension-type headache: a magnetoencephalographic
involved in the pathophysiology of TTH and 93. Ashina, M. et al. In vivo evidence of altered skeletal study. Pain 159, 793–803 (2018).
implications for the treatment of episodic TTH. muscle blood flow in chronic tension-type headache. 117. Rossi, P., Vollono, C., Valeriani, M. & Sandrini, G.
68. Sakai, F., Ebihara, S., Akiyama, M. & Horikawa, M. Brain 125, 320–326 (2002). The contribution of clinical neurophysiology to the
Pericranial muscle hardness in tension-type headache. 94. Olesen, J. Clinical and pathophysiological observations comprehension of the tension-type headache
A non-invasive measurement method and its clinical in migraine and tension-type headache explained by mechanisms. Clin. Neurophysiol. 122, 1075–1085
application. Brain 118, 523–531 (1995). integration of vascular, supraspinal and myofascial (2011).
69. Ashina, M., Lassen, L. H., Bendtsen, L., Jensen, R. & inputs. Pain 46, 125–132 (1991). 118. Ashina, M. Neurobiology of chronic tension-type
Olesen, J. Effect of inhibition of nitric oxide synthase 95. Schytz, H. W., Amin, F. M., Selb, J. & Boas, D. A. headache. Cephalalgia 24, 161–172 (2004).
on chronic tension-type headache: a randomised Non-invasive methods for measuring vascular This review contributes to our understanding of the
crossover trial. Lancet 353, 287–289 (1999). changes in neurovascular headaches. J. Cereb. mechanisms leading to CTTH, and explains the role
70. Ashina, M., Bendtsen, L., Jensen, R., Sakai, F. & Blood Flow. Metab. 39, 633–649 (2019). NO and neuropeptides have in the pathophysiology
Olesen, J. Possible mechanisms of glyceryl-trinitrate- 96. Wallasch, T. M. Transcranial Doppler ultrasonic of this primary headache.
induced immediate headache in patients with chronic features in episodic tension-type headache. 119. Ashina, M., Bendtsen, L., Jensen, R. & Olesen, J.
tension-type headache. Cephalalgia 20, 919–924 Cephalalgia 12, 293–296 (1992). Nitric oxide-induced headache in patients with chronic
(2000). 97. Sliwka, U. et al. Spontaneous oscillations in tension-type headache. Brain 123, 1830–1837 (2000).
71. Fernandez-De-Las-Penas, C. & Arendt-Nielsen, L. cerebral blood flow velocity give evidence of different 120. Alam, Z., Coombes, N., Waring, R. H., Williams, A. C.
Improving understanding of trigger points and autonomic dysfunctions in various types of headache. & Steventon, G. B. Plasma levels of neuroexcitatory
widespread pressure pain sensitivity in tension-type Headache 41, 157–163 (2001). amino acids in patients with migraine or tension
headache patients: clinical implications. Expert Rev. 98. Karacay Ozkalayci, S., Nazliel, B., Batur Caglayan, H. Z. headache. J. Neurol. Sci. 156, 102–106 (1998).
Neurother. 17, 933–939 (2017). & Irkec, C. Cerebral blood flow velocity in migraine 121. de Tommaso, M. et al. Clinical features of headache
72. Schmidt-Hansen, P. T., Svensson, P., Jensen, T. S., and chronic tension-type headache patients. patients with fibromyalgia comorbidity. J. Headache
Graven-Nielsen, T. & Bach, F. W. Patterns of J. Pain Res. 11, 661–666 (2018). Pain 12, 629–638 (2011).
experimentally induced pain in pericranial muscles. 99. Sarchielli, P., Alberti, A., Floridi, A. & Gallai, V. 122. Gonçalves, D. A., Bigal, M. E., Jales, L. C.,
Cephalalgia 26, 568–577 (2006). L-Arginine/nitric oxide pathway in chronic tension- Camparis, C. M. & Speciali, J. G. Headache
73. Fernandez-de-Las-Penas, C. Myofascial head pain. type headache: relation with serotonin content and and symptoms of temporomandibular disorder:
Curr. Pain Headache Rep. 19, 28 (2015). secretion and glutamate content. J. Neurol. Sci. 198, an epidemiological study. Headache 50, 231–241
74. Do, T. P., Heldarskard, G. F., Kolding, L. T., 9–15 (2002). (2010).
Hvedstrup, J. & Schytz, H. W. Myofascial trigger 100. Hannerz, J. & Jogestrand, T. Is Chronic tension- 123. Cho, S. J., Song, T. J. & Chu, M. K. Sleep and
points in migraine and tension-type headache. type headache a vascular headache? The relation tension-type headache. Curr. Neurol. Neurosci. Rep.
J. Headache Pain 19, 84 (2018). between chronic tension-type headache and cranial 19, 44 (2019).
75. Palacios-Cena, M. et al. Trigger points are hemodynamics. Headache 38, 668–675 (1998). 124. Engstrøm, M. et al. Sleep quality, arousal and pain
associated with widespread pressure pain sensitivity 101. Drummond, P. D. & Lance, J. W. Extracranial thresholds in tension-type headache: a blinded
in people with tension-type headache. Cephalalgia 38, vascular reactivity in migraine and tension headache. controlled polysomnographic study. Cephalalgia 34,
237–245 (2018). Cephalalgia 1, 149–155 (1981). 455–463 (2014).

18 | Article citation ID: (2021) 7:24 www.nature.com/nrdp

0123456789();:
Primer

125. Fernández-de-Las-Peñas, C. et al. Sleep disturbances humans during migraine headache. Ann. Neurol. 28, 172. Bigal, M. E., Lipton, R. B., Tepper, S. J., Rapoport, A. M.
in tension-type headache and migraine. Ther. Adv. 183–187 (1990). & Sheftell, F. D. Primary chronic daily headache and
Neurol. Disord. 11, 1756285617745444 (2017). 150. Ashina, M., Bendtsen, L., Jensen, R., Schifter, S. & its subtypes in adolescents and adults. Neurology 63,
126. Wang, P. et al. Regional homogeneity abnormalities Olesen, J. Evidence for increased plasma levels of 843–847 (2004).
in patients with tension-type headache: a resting-state calcitonin gene-related peptide in migraine outside 173. Forsyth, P. A. & Posner, J. B. Headaches in
fMRI study. Neurosci. Bull. 30, 949–955 (2014). of attacks. Pain 86, 133–138 (2000). patients with brain tumors. Neurology 43, 1678
127. Schmidt-Wilcke, T. et al. Gray matter decrease 151. Ashina, M. et al. Plasma levels of calcitonin (1993).
in patients with chronic tension type headache. gene-related peptide in chronic tension-type 174. Buse, D. C. & Lipton, R. B. Facilitating communication
Neurology 65, 1483–1486 (2005). headache. Neurology 55, 1335–1340 (2000). with patients for improved migraine outcomes.
128. Chen, W. T. et al. Comparison of gray matter volume 152. Gupta, R., Ahmed, T., Banerjee, B. & Bhatia, M. Curr. Pain Headache Rep. 12, 230–236 (2008).
between migraine and “strict-criteria” tension-type Plasma calcitonin gene-related peptide concentration 175. Gaul, C. et al. Clinical outcome of a headache-specific
headache. J. Headache Pain 19, 4 (2018). is comparable to control group among migraineurs multidisciplinary treatment program and adherence
129. Chen, B., He, Y., Xia, L., Guo, L. L. & Zheng, J. L. and tension type headache subjects during inter-ictal to treatment recommendations in a tertiary headache
Cortical plasticity between the pain and pain-free period. J. Headache Pain 10, 161–166 (2009). center: an observational study. J. Headache Pain 12,
phases in patients with episodic tension-type 153. Ashina, M., Bendtsen, L., Jensen, R., Schifter, S. & 475–483 (2011).
headache. J. Headache Pain 17, 105 (2016). Olesen, J. Calcitonin gene-related peptide levels 176. Joseph-Williams, N. et al. Implementing shared
130. Fumal, A. & Schoenen, J. Tension-type headache: during nitric oxide-induced headache in patients decision making in the NHS: lessons from the MAGIC
current research and clinical management. Lancet with chronic tension-type headache. Eur. J. Neurol. programme. Br. Med. J. 357, j1744 (2017).
Neurol. 7, 70–83 (2008). 8, 173–178 (2001). 177. Mitsikostas, D. D. Nocebo in headache. Curr. Opin.
131. Sandrini, G. et al. Abnormal modulatory influence 154. Uddman, R. et al. Peptide-containing nerve fibres in Neurol. 29, 331–336 (2016).
of diffuse noxious inhibitory controls in migraine and human extracranial tissue: a morphological basis for 178. Hagen, K., Jensen, R., Bøe, M. G. & Stovner, L. J.
chronic tension-type headache patients. Cephalalgia neuropeptide involvement in extracranial pain? Pain Medication overuse headache: a critical review of end
26, 782–789 (2006). 27, 391–399 (1986). points in recent follow-up studies. J. Headache Pain
132. Buchgreitz, L., Egsgaard, L. L., Jensen, R., 155. Ashina, M., Bendtsen, L., Jensen, R., Ekman, R. & 11, 373–377 (2010).
Arendt-Nielsen, L. & Bendtsen, L. Abnormal pain Olesen, J. Plasma levels of substance P, neuropeptide Y 179. Stephens, G., Derry, S. & Moore, R. A. Paracetamol
processing in chronic tension-type headache: and vasoactive intestinal polypeptide in patients with (acetaminophen) for acute treatment of episodic
a high-density EEG brain mapping study. Brain chronic tension-type headache. Pain 83, 541–547 tension-type headache in adults. Cochrane Database
131, 3232–3238 (2008). (1999). Syst. Rev. 6, CD011889 (2016).
133. Law, L. F. & Sluka, K. A. How does physical activity 156. Gallai, V. et al. Neuropeptide Y in juvenile migraine 180. Derry, S., Wiffen, P. J. & Moore, R. A. Aspirin for
modulate pain? Pain 158, 369–370 (2017). and tension-type headache. Headache 34, 35–40 acute treatment of episodic tension-type headache
134. D’Andrea, G. et al. Metabolism and menstrual cycle (1994). in adults. Cochrane Database Syst. Rev. 1, CD011888
rhythmicity of serotonin in primary headaches. 157. Jensen, R. & Becker, W. J. in The Headaches (2017).
Headache 35, 216–221 (1995). (eds Olesen, J. et al.) 685-692 (Lippincott, Williams 181. Bendtsen, L. et al. EFNS guideline on the treatment
135. Marazziti, D. et al. Platelet 3H-imipramine binding & Wilkins, 2005). of tension-type headache - report of an EFNS task
and sulphotransferase activity in primary headache. 158. Jensen, R. & Becker, W. J. in The Headaches force. Eur. J. Neurol. 17, 1318–1325 (2010).
Cephalalgia 14, 210–214 (1994). (eds Olesen, J. et al.) 693-699 (Lippincott, Williams Evidence-based or expert recommendations for
136. Bendtsen, L. & Mellerup, E. T. The platelet serotonin & Wilkins, 2005). the management of the TTH.
transporter in primary headaches. Eur. J. Neurol. 5, 159. Aaseth, K., Grande, R. B., Lundqvist, C. & Russell, M. B. 182. Fernández-de-Las-Peñas, C. et al. Variables
277–282 (1998). Pericranial tenderness in chronic tension-type associated with use of symptomatic medication during
137. Bendtsen, L., Jensen, R., Hindberg, I., Gammeltoft, S. headache: the Akershus population-based study of a headache attack in individuals with tension-type
& Olesen, J. Serotonin metabolism in chronic chronic headache. J. Headache Pain 15, 58 (2014). headache: a European study. BMC Neurol. 20, 43–43
tension-type headache. Cephalalgia 17, 843–848 160. Schoenen, J. & Jensen, R. in The Headaches (2020).
(1997). (eds Olesen, J. et al.) 701-709 (Lippincott, Williams 183. Derry, S., Wiffen, P. J., Moore, R. A. & Bendtsen, L.
138. Sawynok, J. & Reid, A. Interactions of descending & Wilkins, 2005). Ibuprofen for acute treatment of episodic tension-type
serotonergic systems with other neurotransmitters 161. Rasmussen, B. K., Jensen, R. & Olesen, J. headache in adults. Cochrane Database Syst. Rev. 7,
in the modulation of nociception. Behav. Brain Res. Questionnaire versus clinical interview in the diagnosis CD011474 (2015).
73, 63–68 (1996). of headache. Headache 31, 290–295 (1991). 184. Veys, L., Derry, S. & Moore, R. A. Ketoprofen for
139. Bendtsen, L., Jensen, R. & Olesen, J. A non-selective 162. Steiner, T. J. et al. Aids to management of headache episodic tension-type headache in adults. Cochrane
(amitriptyline), but not a selective (citalopram), disorders in primary care (2nd edition): on behalf of Database Syst. Rev. 9, CD012190 (2016).
serotonin reuptake inhibitor is effective in the the European Headache Federation and Lifting The 185. Prior, M. J., Cooper, K. M., May, L. G. & Bowen, D. L.
prophylactic treatment of chronic tension-type Burden: the Global Campaign against Headache. Efficacy and safety of acetaminophen and naproxen in
headache. J. Neurol. Neurosurg. Psychiatry 61, J. Headache Pain 20, 57 (2019). the treatment of tension-type headache. A randomized,
285–290 (1996). This publication by experts convened by Lifting The double-blind, placebo-controlled trial. Cephalalgia 22,
140. Ashina, S., Bendtsen, L. & Jensen, R. Analgesic effect Burden: the Global Campaign against Headache in 740–748 (2002).
of amitriptyline in chronic tension-type headache is collaboration with the European Headache 186. Kubitzek, F., Ziegler, G., Gold, M. S., Liu, J. M.
not directly related to serotonin reuptake inhibition. Federation provides educational materials & Ionescu, E. Low-dose diclofenac potassium in
Pain 108, 108–114 (2004). with practical aids for primary care physicians the treatment of episodic tension-type headache.
141. Kumamoto, E., Mizuta, K. & Fujita, T. Opioid actions and non-specialist health-care providers for Eur. J. Pain 7, 155–162 (2003).
in primary-afferent fibers — involvement in analgesia management of headache disorders. 187. Pini, L. A. et al. Tolerability and efficacy of a
and anesthesia. Pharmaceuticals 4, 343–365 (2011). 163. Do, T. P. et al. Red and orange flags for secondary combination of paracetamol and caffeine in the
142. Bach, F. W., Langemark, M., Secher, N. H. & Olesen, J. headaches in clinical practice: SNNOOP10 list. treatment of tension-type headache: a randomised,
Plasma and cerebrospinal fluid beta-endorphin in Neurology 92, 134–144 (2019). double-blind, double-dummy, cross-over study versus
chronic tension-type headache. Pain 51, 163–168 164. Evans, R. W. et al. Neuroimaging for migraine: the placebo and naproxen sodium. J. Headache Pain 9,
(1992). American Headache Society systematic review and 367–373 (2008).
143. Leone, M., Sacerdote, P., D’Amico, D., Panerai, A. E. evidence-based guideline. Headache 60, 318–336 188. Diener, H. C., Pfaffenrath, V., Pageler, L., Peil, H. &
& Bussone, G. Beta-endorphin concentrations in the (2020). Aicher, B. The fixed combination of acetylsalicylic acid,
peripheral blood mononuclear cells of migraine and 165. Howard, L. et al. Are investigations anxiolytic paracetamol and caffeine is more effective than single
tension-type headache patients. Cephalalgia 12, or anxiogenic? A randomised controlled trial of substances and dual combination for the treatment
154–157 (1992). neuroimaging to provide reassurance in chronic of headache: a multicentre, randomized, double-blind,
144. Mazzotta, G., Sarchielli, P., Gaggioli, A. & Gallai, V. daily headache. J. Neurol. Neurosurg. Psychiatry 76, single-dose, placebo-controlled parallel group study.
Study of pressure pain and cellular concentration of 1558 (2005). Cephalalgia 25, 776–787 (2005).
neurotransmitters related to nociception in episodic 166. Ashina, M. Migraine. N. Engl. J. Med. 383, 189. Diener, H.-C., Gold, M. & Hagen, M. Use of a fixed
tension-type headache patients. Headache 37, 1866–1876 (2020). combination of acetylsalicylic acid, acetaminophen
565–571 (1997). 167. Rasmussen, B. K., Jensen, R., Schroll, M. & Olesen, J. and caffeine compared with acetaminophen alone
145. Langemark, M., Bach, F. W., Ekman, R. & Olesen, J. Interrelations between migraine and tension-type in episodic tension-type headache: meta-analysis of
Increased cerebrospinal fluid Met-enkephalin headache in the general population. Arch. Neurol. four randomized, double-blind, placebo-controlled,
immunoreactivity in patients with chronic tension-type 49, 914–918 (1992). crossover studies. J. Headache Pain 15, 76 (2014).
headache. Pain 63, 103–107 (1995). 168. Lyngberg, A. C., Rasmussen, B. K., Jorgensen, T. 190. Lipton, R. B., Diener, H. C., Robbins, M. S., Garas, S. Y.
146. de Lourdes Figuerola, M., Leston, J. & Barontini, M. & Jensen, R. Has the prevalence of migraine and & Patel, K. Caffeine in the management of patients
Plasma met-enkephalin levels: its relationship with tension-type headache changed over a 12-year with headache. J. Headache Pain 18, 107 (2017).
age and type of headache. Arch. Gerontol. Geriatr. 22, period? A Danish population survey. Eur. J. Epidemiol. 191. Migliardi, J. R., Armellino, J. J., Friedman, M.,
137–143 (1996). 20, 243–249 (2005). Gillings, D. B. & Beaver, W. T. Caffeine as an analgesic
147. Bach, F. W. et al. Effect of sulpiride or paroxetine 169. Diener, H. C. et al. Pathophysiology, prevention, adjuvant in tension headache. Clin. Pharmacol. Ther.
on cerebrospinal fluid neuropeptide concentrations and treatment of medication overuse headache. 56, 576–586 (1994).
in patients with chronic tension-type headache. Lancet Neurol. 18, 891–902 (2019). 192. Silverman, K., Evans, S. M., Strain, E. C. &
Neuropeptides 27, 129–136 (1994). 170. Yamani, N. & Olesen, J. New daily persistent Griffiths, R. R. Withdrawal syndrome after the
148. Tajti, J., Uddman, R., Moller, S., Sundler, F. & headache: a systematic review on an enigmatic double-blind cessation of caffeine consumption.
Edvinsson, L. Messenger molecules and receptor disorder. J. Headache Pain 20, 80 (2019). N. Engl. J. Med. 327, 1109–1114 (1992).
mRNA in the human trigeminal ganglion. J. Auton. 171. Bigal, M. E., Sheftell, F. D., Rapoport, A. M., 193. Göbel, H., Fresenius, J., Heinze, A., Dworschak, M.
Nerv. Syst. 76, 176–183 (1999). Tepper, S. J. & Lipton, R. B. Chronic daily headache: & Soyka, D. Effectiveness of Oleum menthae piperitae
149. Goadsby, P. J., Edvinsson, L. & Ekman, R. Vasoactive identification of factors associated with induction and and paracetamol in therapy of headache of the tension
peptide release in the extracerebral circulation of transformation. Headache 42, 575–581 (2002). type. Nervenarzt 67, 672–681 (1996).


NATURE REVIEWS | DISEASE PRIMERS | Article citation ID: (2021) 7:24 19

0123456789();:
Primer

194. Göbel, H., Heinze, A., Heinze-Kuhn, K., Göbel, A. 215. Chiang, C. C., Schwedt, T. J., Wang, S. J. & Dodick, D. W. 238. Linde, K. et al. Acupuncture for the prevention of
& Göbel, C. Peppermint oil in the acute treatment Treatment of medication-overuse headache: a tension-type headache. Cochrane Database Syst. Rev.
of tension-type headache. Schmerz 30, 295–310 systematic review. Cephalalgia 36, 371–386 (2016). 4, CD007587 (2016).
(2016). 216. Andrasik, F. et al. Non-pharmacological approaches 239. Karadaş, Ö., Gül, H. L. & Inan, L. E. Lidocaine injection
195. Diener, H. C. et al. European Academy of Neurology for headaches in young age: an updated review. of pericranial myofascial trigger points in the
guideline on the management of medication-overuse Front. Neurol. 9, 1009 (2018). treatment of frequent episodic tension-type headache.
headache. Eur. J. Neurol. 27, 1102–1116 (2020). 217. Nestoriuc, Y., Rief, W. & Martin, A. Meta-analysis J. Headache Pain 14, 44 (2013).
An important guideline on the management of of biofeedback for tension-type headache: efficacy, 240. Gildir, S., Tüzün, E. H., Eroğlu, G. & Eker, L. A
MOH, a complication of primary headaches specificity, and treatment moderators. J. Consult. randomized trial of trigger point dry needling versus
including migraine and TTH. Clin. Psychol. 76, 379–396 (2008). sham needling for chronic tension-type headache.
196. Brennum, J., Kjeldsen, M. & Olesen, J. The 5-HT1-like 218. Holroyd, K. A. et al. Management of chronic Medicine 98, e14520 (2019).
agonist sumatriptan has a significant effect in chronic tension-type headache with tricyclic antidepressant 241. Vázquez-Justes, D., Yarzábal-Rodríguez, R.,
tension-type headache. Cephalalgia 12, 375–379 medication, stress management therapy, and their Doménech-García, V., Herrero, P. & Bellosta-López, P.
(1992). combination: a randomized controlled trial. JAMA Effectiveness of dry needling for headache: a
197. Banzi, R. et al. Selective serotonin reuptake inhibitors 285, 2208–2215 (2001). systematic review. Neurologia https://doi.org/
(SSRIs) and serotonin-norepinephrine reuptake 219. Ukestad, L. K. & Wittrock, D. A. Pain perception 10.1016/j.nrl.2019.09.010 (2020).
inhibitors (SNRIs) for the prevention of tension-type and coping in female tension headache sufferers and 242. Linde, M. et al. The cost of headache disorders in
headache in adults. Cochrane Database Syst. Rev. 4, headache-free controls. Health Psychol. 15, 65–68 Europe: the Eurolight project. Eur. J. Neurol. 19,
CD011681 (2015). (1996). 703–711 (2012).
198. Exposto, F. G., Bendixen, K. H., Ernberg, M., 220. Rollnik, J. D., Karst, M., Fink, M. & Dengler, R. 243. Edmeads, J. et al. Impact of migraine and tension-
Bach, F. W. & Svensson, P. Characterization and Coping strategies in episodic and chronic tension-type type headache on life-style, consulting behaviour,
predictive mechanisms of experimentally induced headache. Headache 41, 297–302 (2001). and medication use: a canadian population survey.
tension-type headache. Cephalalgia 39, 1207–1218 221. Martin, P. R. et al. Behavioral management of Can. J. Neurol. Sci. 20, 131–137 (1993).
(2019). the triggers of recurrent headache: a randomized 244. Lyngberg, A. C., Rasmussen, B. K., Jorgensen, T. &
199. Starling, A. J. Diagnosis and management of headache controlled trial. Behav. Res. Ther. 61, 1–11 (2014). Jensen, R. Secular changes in health care utilization
in older adults. Mayo Clin. Proc. 93, 252–262 (2018). 222. Christiansen, S., Jürgens, T. P. & Klinger, R. Outpatient and work absence for migraine and tension-type
200. Palacios-Ceña, M. et al. Variables associated with combined group and individual cognitive-behavioral headache: a population based study. Eur. J. Epidemiol.
the use of prophylactic amitriptyline treatment in treatment for patients with migraine and tension-type 20, 1007–1014 (2005).
patients with tension-type headache. Clin. J. Pain 35, headache in a routine clinical setting. Headache 55, 245. Lavados, P. M. & Tenhamm, E. Consulting behaviour
315–320 (2019). 1072–1091 (2015). in migraine and tension-type headache sufferers:
201. Bendtsen, L. & Jensen, R. Mirtazapine is effective 223. Gu, Q., Hou, J. C. & Fang, X. M. Mindfulness a population survey in Santiago, Chile. Cephalalgia
in the prophylactic treatment of chronic tension-type meditation for primary headache pain: a meta-analysis. 21, 733–737 (2001).
headache. Neurology 62, 1706–1711 (2004). Chin. Med. J. 131, 829–838 (2018). 246. Wang, S. J., Fuh, J. L., Young, Y. H., Lu, S. R. &
202. Martin-Araguz, A., Bustamante-Martinez, C. & 224. Wang, J. et al. Triggers of migraine and tension-type Shia, B. C. Frequency and predictors of physician
de Pedro-Pijoan, J. M. Treatment of chronic tension headache in China: a clinic-based survey. Eur. J. Neurol. consultations for headache. Cephalalgia 21, 25–30
type headache with mirtazapine and amitriptyline 20, 689–696 (2013). (2001).
[Spanish]. Rev. Neurol. 37, 101–105 (2003). 225. Pellegrino, A. B. W., Davis-Martin, R. E., Houle, T. T., 247. Jensen, R. & Rasmussen, B. K. Burden of headache.
203. Zissis, N. P. et al. A randomized, double-blind, Turner, D. P. & Smitherman, T. A. Perceived triggers Expert Rev. Pharmacoecon Outcomes Res. 4,
placebo-controlled study of venlafaxine XR in of primary headache disorders: a meta-analysis. 353–359 (2004).
out-patients with tension-type headache. Cephalalgia Cephalalgia 38, 1188–1198 (2018). 248. Lenaerts, M. E. Burden of tension-type headache.
27, 315–324 (2007). 226. Lipton, R. B., Pavlovic, J. M., Haut, S. R., Curr. Pain Headache Rep. 10, 459–462 (2006).
204. Jackson, J. L., Mancuso, J. M., Nickoloff, S., Grosberg, B. M. & Buse, D. C. Methodological issues This review article discusses the burden caused by
Bernstein, R. & Kay, C. Tricyclic and tetracyclic in studying trigger factors and premonitory features TTH on society including economic impact, loss of
antidepressants for the prevention of frequent of migraine. Headache 54, 1661–1669 (2014). productivity, increase in health-care utilization and
episodic or chronic tension-type headache in adults: 227. Kim, K. M. et al. Excessive daytime sleepiness in reduced quality of life.
a systematic review and meta-analysis. J. Gen. tension-type headache: a population study. Front. 249. Saylor, D. & Steiner, T. J. The global burden of
Intern. Med. 32, 1351–1358 (2017). Neurol. 10, 1282 (2019). headache. Semin. Neurol. 38, 182–190 (2018).
205. Fogelholm, R. & Murros, K. Tizanidine in chronic 228. Benito-González, E. et al. Variables associated 250. Penacoba-Puente, C., Fernandez-de-Las-Penas, C.,
tension-type headache: a placebo controlled with sleep quality in chronic tension-type headache: Gonzalez-Gutierrez, J. L., Miangolarra-Page, J. C. &
double-blind cross-over study. Headache 32, A cross-sectional and longitudinal design. PLoS ONE Pareja, J. A. Interaction between anxiety, depression,
509–513 (1992). 13, e0197381 (2018). quality of life and clinical parameters in chronic tension-
206. Murros, K. et al. Modified-release formulation 229. Sullivan, D. P., Martin, P. R. & Boschen, M. J. type headache. Eur. J. Pain 12, 886–894 (2008).
of tizanidine in chronic tension-type headache. Psychological sleep interventions for migraine 251. Abu Bakar, N. et al. Quality of life in primary headache
Headache 40, 633–637 (2000). and tension-type headache: a systematic disorders: A review. Cephalalgia 36, 67–91 (2016).
207. Agius, A. M., Jones, N. S. & Muscat, R. A randomized review and meta-analysis. Sci. Rep. 9, 6411 (2019). 252. Lampl, C. et al. Will (or can) people pay for headache
controlled trial comparing the efficacy of low-dose 230. Tornøe, B. et al. Specific strength training compared care in a poor country? J. Headache Pain 13, 67–74
amitriptyline, amitriptyline with pindolol and with interdisciplinary counseling for girls with (2012).
surrogate placebo in the treatment of chronic tension-type headache: a randomized controlled trial. 253. Guitera, V., Munoz, P., Castillo, J. & Pascual, J. Quality
tension-type facial pain. Rhinology 51, 143–153 J. Pain Res. 9, 257–270 (2016). of life in chronic daily headache: a study in a general
(2013). 231. Andersen, C. H. et al. Effect of resistance training on population. Neurology 58, 1062–1065 (2002).
208. Mitsikostas, D. D., Gatzonis, S., Thomas, A. & Ilias, A. headache symptoms in adults: secondary analysis of 254. Ashina, S. et al. Health-related quality of life
Buspirone vs amitriptyline in the treatment of chronic a RCT. Musculoskelet. Sci. Pract. 32, 38–43 (2017). in tension-type headache: a population-based
tension-type headache. Acta Neurol. Scand. 96, 232. Krøll, L. S., Hammarlund, C. S., Linde, M., Gard, G. study. Scand. J. Pain https://doi.org/10.1515/
247–251 (1997). & Jensen, R. H. The effects of aerobic exercise for sjpain-2020-0166 (2021).
209. Lenaerts, M., Bastings, E., Sianard, J. & Schoenen, J. persons with migraine and co-existing tension-type 255. Wang, S. J., Fuh, J. L., Lu, S. R. & Juang, K. D.
Sodium valproate in severe migraine and tension-type headache and neck pain. A randomized, controlled, Quality of life differs among headache diagnoses:
headache: an open study of long-term efficacy and clinical trial. Cephalalgia 38, 1805–1816 (2018). analysis of SF-36 survey in 901 headache patients.
correlation with blood levels. Acta Neurol. Belg. 96, 233. Espí-López, G. V., Zurriaga-Llorens, R., Monzani, L. Pain 89, 285–292 (2001).
126–129 (1996). & Falla, D. The effect of manipulation plus massage 256. Solomon, G. D., Skobieranda, F. G. & Gragg, L. A.
210. Lampl, C., Marecek, S., May, A. & Bendtsen, L. A therapy versus massage therapy alone in people with Does quality of life differ among headache diagnoses?
prospective, open-label, long-term study of the efficacy tension-type headache. A randomized controlled clinical Analysis using the medical outcomes study
and tolerability of topiramate in the prophylaxis of trial. Eur. J. Phys. Rehabil. Med. 52, 606–617 (2016). instrument. Headache 34, 143–147 (1994).
chronic tension-type headache. Cephalalgia 26, 234. Falsiroli Maistrello, L., Geri, T., Gianola, S., Zaninetti, M. 257. Stewart, A. L., Hays, R. D. & Ware, J. E. Jr. The MOS
1203–1208 (2006). & Testa, M. Effectiveness of Trigger point manual short-form general health survey. Reliability and
211. Bottiroli, S. et al. Psychological, clinical, and treatment on the frequency, intensity, and duration validity in a patient population. Med. Care 26,
therapeutic predictors of the outcome of detoxification of attacks in primary headaches: a systematic review 724–735 (1988).
in a large clinical population of medication-overuse and meta-analysis of randomized controlled trials. 258. Palermo, T. M., Putnam, J., Armstrong, G. &
headache: a six-month follow-up of the COMOESTAS Front. Neurol. 9, 254 (2018). Daily, S. Adolescent autonomy and family functioning
Project. Cephalalgia 39, 135–147 (2018). 235. Falsiroli Maistrello, L., Rafanelli, M. & Turolla, A. are associated with headache-related disability.
212. Monteith, T. S. & Oshinsky, M. L. Tension-type Manual therapy and quality of life in people with Clin. J. Pain 23, 458–465 (2007).
headache with medication overuse: pathophysiology headache: systematic review and meta-analysis of 259. Nachit-Ouinekh, F. et al. Use of the headache
and clinical implications. Curr. Pain Headache Rep. randomized controlled trials. Curr. Pain Headache impact test (HIT-6) in general practice: relationship
13, 463–469 (2009). Rep. 23, 78 (2019). with quality of life and severity. Eur. J. Neurol. 12,
213. Grande, R. B., Aaseth, K., Benth, J., Lundqvist, C. 236. Bot, M. N. et al. Physical treatments reduce pain 189–193 (2005).
& Russell, M. B. Reduction in medication-overuse in children with tension-type headache: a systematic 260. Bigal, M. E. & Lipton, R. B. Modifiable risk factors for
headache after short information. The Akershus study review and meta-analysis. J. Oral. Facial Pain migraine progression (or for chronic daily headaches)–
of chronic headache. Eur. J. Neurol. 18, 129–137 Headache 34, 240–254 (2020). clinical lessons. Headache 46 (Suppl. 3), S144–S146
(2011). 237. Kamonseki, D. H., Lopes, E. P., van der Meer, H. A. & (2006).
214. Carlsen, L. N. et al. Comparison of 3 treatment Calixtre, L. B. Effectiveness of manual therapy in patients 261. Durham, P. & Papapetropoulos, S. Biomarkers
strategies for medication overuse headache: with tension-type headache. A systematic review associated with migraine and their potential role in
a randomized clinical trial. JAMA Neurol. 77, and meta-analysis. Disabil. Rehabil. https://doi.org/ migraine management. Headache 53, 1262–1277
1069–1078 (2020). 10.1080/09638288.2020.1813817 (2020). (2013).

20 | Article citation ID: (2021) 7:24 www.nature.com/nrdp

0123456789();:
Primer

262. Burstein, R., Collins, B. & Jakubowski, M. Defeating Acknowledgements honoraria as a consultant and/or speaker for Allergan/AbbVie,
migraine pain with triptans: a race against the The authors thank T. P. Do, for skilful assistance and provid- Almirall, Biohaven, Chiesi, Eli Lilly, Lundbeck, Medscape,
development of cutaneous allodynia. Ann. Neurol. ing the initial sketches for Figs. 1 and 2. They also thank Neurodiem, Novartis and Teva Pharmaceuticals. Her research
55, 19–26 (2004). R. Burstein for helping with editing and advising on structural group has received research grants from AGAUR, ERANet
263. Müller, K. I., Alstadhaug, K. B. & Bekkelund, S. I. and functional neuroanatomy in Fig. 2. They also thank Neuron, la Caixa foundation, International Headache Society,
A randomized trial of telemedicine efficacy and safety K. MacDonald from Alliance for Headache Disorders Advocacy Migraine Research Foundation, FEDER RISC3CAT, Instituto
for nonacute headaches. Neurology 89, 153–162 for providing her perspective on tension-type headache. Investigación Carlos III, PERIS, Novartis and Teva; and has
(2017). received funding for clinical trials from AbbVie, Alder,
264. Friedman, B. W. et al. A randomized trial of intravenous Author contributions Electrocore, Eli Lilly, Lundbeck, Novartis and Teva. P.P.-R. is a
ketorolac versus intravenous metoclopramide plus Introduction (H.A., R.B.L. and S.A.); Epidemiology (S.J.W., trustee member of the board of the International Headache
diphenhydramine for tension-type and all nonmigraine, S.A. and R.B.L.); Mechanisms/pathophysiology (R.M., Society and a member of the Council of the European
noncluster recurrent headaches. Ann. Emerg. Med. 62, M.J.L., S.A. and R.H.J.); Diagnosis, screening and prevention Headache Federation, and is on the editorial board of Revista
311–318.e314 (2013). (H.A., H.C.D. and S.A.); Management (D.D.M, D.C.B. and de Neurologia; is an associate editor for Cephalalgia,
265. Cicek, M. et al. Prospective, randomised, double R.B.L.); Quality of life (N.Y., P.P.R. and R.B.L.); Outlook Headache, Frontiers in Neurology, Neurologia and is on the
blind, controlled comparison of metoclopramide (S.A., H.-C.D., R.H.J. and R.B.L.); Overview of Primer (S.A.) Scientific Advisory Board of the Journal of Headache and
and pethidine in the emergency treatment of acute Pain. P.P.-R. does not own stocks from any pharmaceutical
primary vascular and tension type headache episodes. Competing interests company. R.H.J. has received honoraria for lectures and
Emerg. Med. J. 21, 323–326 (2004). S.A. received honoraria for consulting from Allergan/AbbVie, patient leaflets from MSD, Berlin-Chemie Menarini, ATI,
266. Bendtsen, L., Buchgreitz, L., Ashina, S. & Jensen, R. Amgen, Biohaven, Eli Lilly, Impel NeuroPharma, Novartis, Novartis, Teva, Allergan and Pfizer and is conducting clinical
Combination of low-dose mirtazapine and ibuprofen Satsuma, Supernus, Theranica and Percept. S.A. is an associ- trials for Eli-Lilly and Lundbeck. H.-C.D. received honoraria for
for prophylaxis of chronic tension-type headache. ate editor for Neurology Reviews and BMC Neurology, serves participation in clinical trials, contribution to advisory boards
Eur. J. Neurol. 14, 187–193 (2007). on an Advisory Board for the Journal of Headache and Pain, or oral presentations from Allergan/AbbVie, Amgen,
267. Pothmann, R. & Lobisch, M. Acute treatment and is a member of the Education Committee of the Electrocore, Lilly, Medtronic, Novartis, Pfizer, Teva and Weber
of episodic childhood tension-type headache International Headache Society. D.D.M. has received hono- & Weber. Electrocore provided financial support for research
with flupirtine and paracetamol - a double-blind raria as a consultant and/or speaker from Allergan/AbbVie, projects. The German Research Council (DFG), the German
crossover-study [German]. Schmerz 14, 1–4 Amgen, Biogen, Cefaly, Eli Lilly, Genesis Pharma, Medscape, Ministry of Education and Research (BMBF) and the
(2000). Merz, Mylan, Novartis, Roche, Sanofi, Specifar and Teva European Union support his headache research. H.-C.D.
268. Hershey, A. D., Powers, S. W., Bentti, A. L. & Pharmaceuticals; has received funding for clinical trials from serves on the editorial boards of Cephalalgia and Lancet
Degrauw, T. J. Effectiveness of amitriptyline in the Alder, Cefaly, Electrocore, Biogen, Eli Lilly, Genesis Pharma, Neurology, chairs the Clinical Guidelines Committee of the
prophylactic management of childhood headaches. Merz, Novartis and Teva. D.D.M. is also past president of German Society of Neurology and is a member of the Clinical
Headache 40, 539–549 (2000). the European Headache Federation, current co-chair of the Trials Committee of the IHS. R.B.L. is the Edwin S. Lowe
269. Andrasik, F., Grazzi, L., Usai, S. & Bussone, G. Headache Panel in the European Academy of Neurology and Professor of Neurology at the Albert Einstein College of
Pharmacological treatment compared to behavioural president of the Hellenic Headache Society, and is an associ- Medicine in New York. He receives research support from the
treatment for juvenile tension-type headache: results ate editor for Journal of Headache and Pain. D.D.M. does not NIH. R.B.L. also receives support from the Migraine Research
at two-year follow-up. Neurol. Sci. 28 (Suppl. 2), own stocks from any pharmaceutical company. M.J.L. has Foundation and the National Headache Foundation. R.B.L.
S235–S238 (2007). received honoraria as a consultant and/or speaker for Eli Lilly, serves on the editorial board of Neurology, is senior adviser
270. Balottin, U. et al. Psychotherapy versus usual care in Sanofi-Aventis and YuYu Pharma; has been the PI or to Headache, associate editor for Cephalalgia and has
pediatric migraine and tension-type headache: a single- co-investigator in trials sponsored by Eli Lilly, Novartis, Teva reviewed for the NIA and NINDS, holds stock options in
blind controlled pilot study. Ital. J. Pediatr. 40, 6 (2014). (Otsuka), Allergan/AbbVie, Yuhan Company, Samjin Pharm Biohaven Holdings and CntrlM; serves as consultant, advisory
271. Klausen, S. H., Rønde, G., Tornøe, B. & Bjerregaard, L. and DongA ST; received research support from the National board member, or has received honoraria or conducted
Nonpharmacological interventions addressing pain, Research Foundation of Korea; is a trustee member of the research funded by Allergan/Abbvie, American Academy of
sleep, and quality of life in children and adolescents board of the International Headache Society and junior editor Neurology, American Headache Society, Amgen, Biohaven,
with primary headache: a systematic review. of Cephalalgia. S-J.W. has served on the advisory boards of Biovision, Dr. Reddy’s (Promius), Electrocore, Eli Lilly, eNeura
J. Pain Res. 12, 3437–3459 (2019). Daiichi-Sankyo, Eli Lilly and Taiwan Novartis; has received Therapeutics, GlaxoSmithKline, Grifols, Lundbeck (Alder),
272. Soee, A. B., Skov, L., Skovgaard, L. T. & honoraria as a moderator from Allergan/AbbVie, Pfizer, Eli Merck, Pernix, Pfizer, Teva, Trigemina, Vector, Vedanta. R.L.
Thomsen, L. L. Headache in children: effectiveness Lilly and Eisai and has been the PI in trials sponsored by receives royalties from Wolff’s Headache seventh and eighth
of multidisciplinary treatment in a tertiary paediatric Eli Lilly, Novartis, and Allergan/AbbVie. S-J.W. has received editions, Oxford University Press, 2009, Wiley and Informa.
headache clinic. Cephalalgia 33, 1218–1228 research grants from the Taiwan Minister of Technology and All other authors declare no competing interests.
(2013). Science (MOST), Brain Research Center, National Yang-Ming
273. Berk, T., Ashina, S., Martin, V., Newman, L. & University from The Featured Areas Research Center Program Peer review information
Vij, B. Diagnosis and treatment of primary headache within the framework of the Higher Education Sprout Project Nature Reviews Disease Primers thanks V. Gómez-Mayordomo,
disorders in older adults. J. Am. Geriatr. Soc. 66, by the Ministry of Education (MOE) in Taiwan, Taipei Veterans who co-reviewed with M. Cuadrado, H. Göbel, P. Martin and
2408–2416 (2018). General Hospital and Taiwan Headache Society. R.M. has L. Stovner for their contribution to the peer review of this work.
274. Yang, F. C. et al. Increased risk of dementia in patients received honoraria as speaker for Eli Lilly, Novartis and Teva
with tension-type headache: a nationwide retrospective Pharmaceuticals. D.B. received honoraria for consulting for Publisher’s note
population-based cohort study. PLoS ONE 11, Allergan, Amgen, Biohaven, Eli Lilly, Novartis and Teva, and Springer Nature remains neutral with regard to jurisdictional
e0156097 (2016). has received grant support from Amgen, the National claims in published maps and institutional affiliations.
275. Negro, A. et al. Headache and pregnancy: a systematic Headache Foundation and the FDA. D.B. is an editor for
review. J. Headache Pain 18, 106 (2017). Current Pain and Headache Reports. P.P.-R. has received © Springer Nature Limited 2021


NATURE REVIEWS | DISEASE PRIMERS | Article citation ID: (2021) 7:24 21

0123456789();:

You might also like