You are on page 1of 18

PRIMER

Cancer-associated cachexia
Vickie E. Baracos1, Lisa Martin2, Murray Korc3, Denis C. Guttridge4
and †Kenneth C. H. Fearon5
Abstract | Cancer-associated cachexia is a disorder characterized by loss of body weight with specific
losses of skeletal muscle and adipose tissue. Cachexia is driven by a variable combination of reduced
food intake and metabolic changes, including elevated energy expenditure, excess catabolism and
inflammation. Cachexia is highly associated with cancers of the pancreas, oesophagus, stomach,
lung, liver and bowel; this group of malignancies is responsible for half of all cancer deaths
worldwide. Cachexia involves diverse mediators derived from the cancer cells and cells within
the tumour microenvironment, including inflammatory and immune cells. In addition, endocrine,
metabolic and central nervous system perturbations combine with these mediators to elicit catabolic
changes in skeletal and cardiac muscle and adipose tissue. At the tissue level, mechanisms include
activation of inflammation, proteolysis, autophagy and lipolysis. Cachexia associates with a
multitude of morbidities encompassing functional, metabolic and immune disorders as well as
aggravated toxicity and complications of cancer therapy. Patients experience impaired quality of life,
reduced physical, emotional and social well-being and increased use of healthcare resources.
To date, no effective medical intervention completely reverses cachexia and there are no approved
drug therapies. Adequate nutritional support remains a mainstay of cachexia therapy, whereas drugs
that target overactivation of catabolic processes, cell injury and inflammation are currently
under investigation.

It is with sadness that we learned of the passing of reduced food intake and metabolic changes, including
Professor Kenneth Fearon on 3 September 2016. Ken’s elevated energy expenditure, excess catabolism and
research spanned every aspect of cancer-associated inflammation. Cachexia is distinct from starvation
cachexia, from experimental models to clinical trials. and ­simple mal­nutrition, which are readily reversible by
His landmark paper (Definition and classification of the ­provision of adequate nutrients.
cancer cachexia, an international consensus, Lancet Consensus is needed regarding the definition of
Oncol. 12, 489–495 (2011)) will continue to serve as and the specific criteria to adequately describe cancer-­
a roadmap for the field and as a legacy for researchers associated cachexia, as multiple discordant definitions
seeking to mitigate cachexia-related suffering. of cachexia are used in the literature. A single defin­
ition widely accepted by clinicians and researchers will
aid in the identification and treatment of patients with
Cachexia is a disorder characterized by the involuntary cachexia as well as the development and approval of
loss of body weight in addition to loss of homeostatic potential therapeutic agents2. An international Delphi
control of both energy and protein balance1; it has consensus process in 2011 provided a definition and
Correspondence to V.E.B. 
been acknowledged since the earliest written medical conceptual framework specific to cancer-associated
Division of Palliative Care
Medicine, Department ­treatises. Cachexia occurs in association with malig- cachexia2, stating that it is a multifactorial syndrome
of Oncology, University of nant disease and with multiple chronic non-malignant defined by an ongoing loss of skeletal muscle mass (with
Alberta, Cross Cancer diseases, including heart failure, kidney disease, chronic or without loss of fat mass) that can be partially but not
Institute 11560 University obstructive pulmonary disease, neurological disease, entirely reversed by conventional nutritional support.
Avenue, Edmonton, T6G 1Z2
Alberta, Canada.
AIDS and rheumatoid arthritis. Cancer-associated Depletion of skeletal muscle is a key feature of cancer-­
vickie.baracos@ualberta.ca cachexia — the focus of this Primer — has distinctive associated cachexia 2 and its consequences include
†Deceased tumour-driven components and leads to progressive increased chemotherapy toxicity, complications from
functional impairment, treatment-related complica- cancer surgery and mortality 3.
Article number: 17105
doi:10.1038/nrdp.2017.105 tions, poor quality of life and cancer-related mortality 2. Half of all cancer deaths worldwide (~8.2 million
Published online 18 Jan 2018 The dis­order is driven by a variable combination of people per year)4 are attributed to the cancers most

NATURE REVIEWS | DISEASE PRIMERS VOLUME 4 | ARTICLE NUMBER 17105 | 1


©
2
0
1
8
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

Author addresses Epidemiology


Prevalence
1
Division of Palliative Care Medicine, Department of Cancer is a leading cause of morbidity and mortality
Oncology, University of Alberta, Cross Cancer Institute worldwide, with ~14 million new cases and ~8.2 million
11560 University Avenue, Edmonton, T6G 1Z2 Alberta, deaths in 2012 (REF. 4). Cancer-associated cachexia is not
Canada.
included in national cancer statistics in any country
2
Department of Agricultural, Food & Nutritional Science,
University of Alberta, Edmonton, Alberta, Canada. and is seldom listed as the cause of death. However, it is
3
Section of Endocrinology, Departments of Medicine mainly associated with incurable disease and is highly
and Biochemistry and Molecular Biology, Indiana prevalent at the end of life. Thus, the rate of cancer
University School of Medicine, Indianapolis, Indiana, USA. death is a plausible upper limit for the number of people
4
Department of Cancer Biology and Genetics, affected by cachexia. Cachexia can also occur in curable
The Ohio State University, Columbus, Ohio, USA. cancers and may be reversed by successful treatment of
5
Clinical and Surgical Sciences, School of Clinical Sciences the underlying cancer 13.
and Community Health, Royal Infirmary, University of The diagnosis of cancer cachexia is based on the
Edinburgh, Edinburgh, UK. rate of weight loss as well as attainment of a low body
mass index (BMI)2. Most prevalence data are derived
frequently associated with cachexia, namely, pancreatic from national point prevalence studies or from system-
(0.33 million deaths), oesophageal (0.40 million), gas- atic screening programmes in cancer centres14–17. The
tric (0.72 million), pulmonary (1.59 million), hepatic exact criteria used to define cachexia are not consist-
(0.75 million) and colorectal (0.69 million) cancers. ent across studies, making it difficult to aggregate data.
The association of these cancers with cachexia may Regardless of the criteria applied, certain cancers are
be due to their diagnosis at an advanced stage; their more prominently associated with cachexia14–17 (FIG. 1).
direct effects on ingestion, digestion and absorption of Additional factors that contribute to the variable preva­
nutrients; their specific tumour characteristics; and/or lence of cachexia include more-advanced cancer stage,
their high mortality. Available data from palliative care sex (men are more susceptible than women), advanced
settings suggest that rates of cachexia are uniformly age, genetic risk factors, comorbidities and treatment-­
very high at the end of life, regardless of cancer site5. related catabolic effects. For example, ~30% of patients
However, despite its clear association with advanced- with cancer have concurrent cardiac disorders with
stage disease, cachexia is not an inevitable consequence risk of cardiac cachexia; concurrent cancer cachexia
of cancer. Interindividual variation has been noted and cardiac cachexia are speculated to progressively
with regards to the prevalence and severity of cachexia exacerbate each other 18. Similarly, several drugs used
among patients with the same cancer diagnosis and in cancer therapy (such as sorafenib, a tyrosine kinase
stage. Indeed, some patients with advanced-stage dis- inhibitor)19 or in palliation of cancer symptoms (such
ease maintain or gain weight, skeletal muscle and fat as glucocorticoids) have specific catabolic effects
mass6,7. As the nutritional deficits that form an impor- on skeletal muscle. These treatments provide addi-
tant part of cachexia are preventable and at least par- tional ­impetus to the loss of muscle in patients who
tially reversible, patients with cancer can demonstrate receive them.
protein anabolic responses to feeding 8,9. Furthermore, Variation in the prevalence of cachexia might also be
some individuals might be less susceptible to the partly due to genotype. A candidate gene approach has
develop­ment of cachexia. For example, patients with a been used to explore inherited genetic variations that
loss of function mutation in the gene encoding the cell could explain interindividual variations in susceptibil-
adhesion molecule P selectin (SELP) have a reduced ity to cachexia10. However, this area of research is in its
likelihood of developing cachexia10. Experimental stud- early stages and genome-wide approaches are needed
ies in rodent models also show that, even in advanced- to fully appreciate heritable risk.
stage malignancies, cachexia can be substantially
mitigated independent of tumour progression11,12. Cachexia in the context of obesity
In this Primer, we describe the emerging mech­ Current WHO statistics indicate that >600 million
anistic insights into cancer-associated cachexia, includ- adults worldwide are obese (BMI of >30 kg per m2)20,
ing imbalances of proteolysis and protein synthesis; with national rates as high as 50% in some countries21.
­imbalances of lipolysis and lipogenesis; and the roles of Accordingly, given that cachexia is partly defined by low
stem cells, inflammation and the central nervous system BMI, contemporary patients with cancer are increas-
(CNS). Individual genetic and tumour-specific factors ingly less likely to reach the traditionally accepted
as well as variations in treatment type might explain clin­ically underweight BMI of <18.5 kg per m2. One-
the considerable interindividual variation in cachexia third of cancer diagnoses are attributed to behavi­
preva­lence, aetiology, severity and progression. Each of oural and dietary risks, including being overweight
the patient-specific and tumour-specific elements might or obese, which increases the likelihood of obesity
be clinically relevant for a small number of individ­ in patients with a cancer diagnosis. By contrast, rates
uals, as well as relevant at the population level when of underweight adults are generally <10% in western
considered altogether. An improved understanding of countries but 30–40% in developing countries. This
the speci­fic perturbations that occur in a given patient upward shift in BMI renders the diagnosis of cachexia
could guide patient-directed therapeutic approaches. increasingly unclear.

2 | ARTICLE NUMBER 17105 | VOLUME 4 www.nature.com/nrdp


©
2
0
1
8
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

For patients who are of normal or low BMI before 1,200 kcal per day 25,26. Decreased muscle-protein syn-
their cancer diagnosis, the effect of weight loss is magni­ thesis has also been documented in weight-losing
fied. Underweight and severely underweight (BMI of patients with cancer; the fact that protein synthesis can
<16 kg per m2) patients have an increased risk of morbid- be reactiv­ated by the provision of nutrients8,9,27,28 high-
ity and mortality 22. Additionally, large magnitude weight lights the importance of reduced dietary intake in the
losses can occur in obese individuals without achieving aetiology of c­ ancer-associated cachexia.
a low absolute BMI22. Importantly, severe depletion Measurements of whole-body energy expenditure
of skeletal muscle (sarcopenia) may go ­undetected in and metabolic fluxes, lipolysis, gluconeogenesis, protein
patients with obesity 23 (FIG. 2). Muscle loss can occur synthesis, protein degradation and substrate consump-
in the absence of fat loss and can, therefore, escape tion have been made in populations of patients with
detection in obese individuals. For example, patients cancer-associated cachexia29. An elevated resting energy
with breast cancer may gain weight following diagno- expenditure promotes negative energy balance and is
sis, sometimes in association with loss of muscle mass, related in part to tumour metabolism. Tumours compete
­leading to d
­ evelopment of sarcopenic obesity 24. with other organs and tissues for energy fuels and bio-
synthetic substrates and possess an intrinsic metabolic
Mechanisms/pathophysiology rate, which is related to their mass and degree of aero­
Here, we outline the key mechanisms in cancer-­ bic versus anaerobic energy metabolism30. Additional
associated cachexia, relying on animal data and pointing contributions to elevated energy expenditure include
out where findings have been recapitulated in patient-­ inflammation and metabolic cycling (that is, increased
derived samples. Clinical data are limited because rates of substrate metabolism involving ATP hydrolysis).
cachexia occurs at a stage in which patient vulner­ability For example, increased rates of whole-body glycolysis
limits the use of invasive metabolic tests and biopsies and and the concomitantly augmented rate of gluconeo­
disease progression limits the number of patients avail- genesis from the lactic acid cycle are increased >300%29,
able for follow‑up. Accordingly, additional mechanistic as is triacylglycerol or fatty acid cycling 31. It has also been
insights must be derived from a­ nimal models. However, suggested that futile cycling — whereby oxidative phos-
disparities between clinical and ­animal findings remain phorylation is uncoupled from ATP synthesis, resulting
difficult to reconcile. in only the production of heat — in brown or browned
adipose tissue elicits increased and inefficient energy
Altered energy balance expenditure32, contributing to cachexia32,33. Furthermore,
Cancer profoundly alters the normal homeostatic con- mitochondrial dysfunction in skeletal muscles might
trol of energy balance (BOX 1). Reduced food intake is also occur; however, our understanding of muscle
an important and in some cases predominant compo- mitochondrial respiration during cancer-­associated
nent of cancer-associated weight loss. Energy intake cachexia is extremely limited, with only few and hetero-
is typically lower than resting energy expenditure geneous data from animal models and a lack of studies
in the same patients and caloric deficits can exceed in human patients34.

a 70 b 30

60 25
Average weight loss (%)

50
20
Prevalence (%)

40
15
30

10
20

5
10

0 0
rs

er

er
er

er

rs

er

er

er

er

es eck cer

er

er
ce
er

ge cer
ce

ce
nc

nc

nc

nc

nc

nc

nc

nc

nc

nc
nc

an

n
an

n
ca

ca

ca

ca

ca

ca

ca

ca

ca

ca

ca
ca

ca

ca
lc
Co al c
te

st

al

ng

al

ic

st

ng

al

ic
ta
ck

al
at
at

at
ge
ea

ct

ea
ta

Lu

Lu
c
ic

ic
ne

t
re

re
re

re
os

os
Br

Br
ha

ha
n
og

og
lo

lo
nc

nc
d

tro nd
Pr

Pr
op

op
ol

ol
an

Co
Pa

Pa
a
at

at
es
ad

ad
em

em
-o

-o
He

He
ro
Ha

Ha
st

s
Ga

Ga

Figure 1 | Cancer cachexia by tumour site. The prevalence of cachexia (defined as >5% weight loss in the previous
Nature Reviews | Disease Primers
6 months) by cancer site (part a) and the average percentage of weight loss and its variation (error bars) by cancer site
(part b) are shown. Data from REFS 14,17.

NATURE REVIEWS | DISEASE PRIMERS VOLUME 4 | ARTICLE NUMBER 17105 | 3


©
2
0
1
8
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

a Visceral adipose tissue b Pro-cachexia cytokines and factors


Subcutaneous
adipose tissue A complex tumour secretome is an important ­factor
unique to cancer-associated cachexia (FIG. 3). Tumours
secrete molecules that directly elicit catabolism in
­target tissues, including a long list of pro-­inflammatory
cytokines, eicosanoids and other factors with tissue-­
specific effects, such as heat shock protein 70 (HSP70)
Intermuscular adipose tissue Skeletal muscle and HSP90 (REF.  35), members of the transforming
growth factor-β (TGFβ) superfamily (including activins,
c d
myostatin and TGFβ, which act on skeletal muscle)
or adrenomedullin (which acts on adipose tissue).
The increased inflammation elicited by the tumour
also partici­pates in the generation of catabolic pro-­
inflammatory factors. These effectors modulate
homeo­static controls in the CNS, prompting catabolic
neural outputs via the sympathetic nervous system,
as well as neuroendocrine outputs (such as the release
of adrenal corticosteroids) and sickness behaviour
(such as a­ norexia and fatigue). These humoral, neu-
ral and behavi­oural outputs directly activate proteolysis
and lipolysis in target organs, p
­ rimarily skeletal muscle,
e f ­adipose tissue and cardiac muscle36.
Pro-inflammatory factors with catabolic actions have
attracted the most attention as mediators of cachexia.
Prostaglandins (in particular, prostaglandin E2) are
known mediators of tumour-induced bone resorption
and paraneoplastic hypercalcaemia and have similarly
been documented in animal models to be mediators of
excess catabolism in skeletal muscle37. Peptide inflam-
matory mediators of cachexia include IL‑6, which is a
key regulator of skeletal muscle, IL‑1, tumour necro-
sis factor (TNF), IFNγ, leukaemia inhibitory factor
(LIF), growth/differentiation factor 15 (GDF15) and
TNF-related weak inducer of apoptosis (TWEAK;
also known as TNFSF12) (FIG. 3). Identified primarily
g 12
through cell culture conditions and tumour xenograft
10 models, these factors signal through their respective
Patients (%)

8 cell surface receptors and activate selective transcrip-


6 tion factors, which in turn promote the transcription
4 of ubiquitin–proteasome and autophagy components
2 (FIG. 4). These signalling molecules are synthesized by

0 80
tumour or immune cells and their activities are suffi-
0 20 30 40 50 60 70 cient to promote catabolism in target organs such as
Lumbar SMI (cm2 per m2) skeletal muscle, but c­ onfirmatory patient data have
lagged behind38.
Figure 2 | Severe muscle Nature
depletion can occur
Reviews in Primers
| Disease In addition to inflammatory cytokines, other cir-
patients with cachexia and/or obesity. CT images from
two female patients with sarcopenia are shown; the culating factors have been described that exhibit
images on the left (parts a, c, e) correspond to a woman pro-­cachectic activity towards skeletal muscle (FIG. 4).
with a body mass index (BMI) of 47 kg per m2, and the Activin  A is a member of the TGFβ superfamily
images on the right (parts b, d, f) correspond to a woman of growth factors that is produced by both tumour
with a BMI of 17 kg per m2. Sarcopenia is occult in the and immune cells39. In cultured myotubes, activin A
woman on the left but obvious in the woman on the right. ­promotes atrophy; when it is overexpressed in mice,
The axial plane (parts a, b), the sagittal plane (parts c, d) it promotes weight loss and skeletal muscle loss with
and the coronal plane (parts e, f) are shown. The higher potency than IL‑6 (REFS 40–42). Another pro-­
histogram (part g) shows the distribution of skeletal cachexia cytokine is TWEAK, which belongs to the
muscle index (SMI, a standardized unit of muscle
TNF family. TWEAK acts through TNF receptor
area normalized for stature) in female patients with
advanced-stage cancer3. For both women, the lumbar superfamily member 12A (TNFRSF12A), which, when
SMI is 36.8 cm2 per m2 (indicated by the dashed line). overexpressed in tumours, correlates with cachexia; its
Although both women are 60 years of age, this SMI value role was shown via neutraliza­tion with antibodies of
is typical for a female patient with cancer who is >80 years TNFRSF12A, which inhibited weight loss and increased
of age3. lifespan in a mouse model43. Similar to TNF and IL‑6,

4 | ARTICLE NUMBER 17105 | VOLUME 4 www.nature.com/nrdp


©
2
0
1
8
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

Box 1 | Energy intake and energy expenditure imbalance in cancer-associated cachexia


Body weight remains stable when there is balance between energy
Energy TEE
intake (that is, calories provided via oral, enteral or parenteral routes) intake (REE + AEE + TEF)
and the total energy expenditure (TEE) by the body (see the
illustration). Body weight loss occurs when there is a negative energy
balance, a state in which TEE exceeds energy intake. TEE is the sum of
Stable weight
resting energy expenditure (REE), activity-related energy expenditure
(AEE) and the thermic effect of food (TEF). REE is the amount of energy
↑ TEE + TEF)
expended by the body at rest and is the largest contributor to TEE. + AEE
(↑ RE E
Accordingly, as TEE is difficult to measure clinically in free-living
individuals, REE is typically assumed to represent energy metabolism. rgy
↓ Ene e
REE can be accurately measured using indirect calorimetry or intak
estimated using various equations, which have many limitations.
Tumour metabolism and inflammation might increase REE and Weight loss
simultaneously decrease energy intake (through, for example, loss of
appetite), shifting the scale towards negative energy balance187,188. Additionally, cancer treatments also influence energy
balance; for example, energy intake may fall by >50% (~1,200 kcal per day) during chemoradiotherapy for cancers of the
head and neck25. These factors contribute to the negative energy balance in cancer-associated cachexia.

activin A and TWEAK can promote muscle atrophy in Additionally, inflammatory factors such as cytokines
non-malignant conditions, making these factors and and angiotensin II reduce AKT activity, thereby caus-
the respective receptors through which they signal ing FOXO nuclear shuttling and induction of muscle-­
potentially interesting therapeutic targets44,45. Clinical protein catabolism52–55. In addition to the muscle E3
trials for intervention studies targeting activin A and ubiquitin ligase genes, FOXO transcription factors have
TWEAK have been initiated in both cancer and non-­ a vital role in transcribing genes involved in the auto-
cancer indications (NCT00771329 and NCT01604642). phagy system56. Under physiological conditions, skele-
Information gained from these human studies might tal muscle homeostasis requires autophagy to eliminate
determine whether single-line therapy against activin A, damaged proteins and organelles. However, in cachexia,
TWEAK or TNFRSF12A is sufficient to rescue muscle the upregulation of autophagy genes leads to excessive
atrophy in patients with cancer. activation of autophagy pathways that contribute to
In rodent models of cancer cachexia, expression increased breakdown of skeletal muscle. In muscles
of E3 ubiquitin-protein ligase Trim63 (also known as of patients with cancer-associated cachexia, increased
Murf1) and F‑box only protein 32 (Fbxo32, also known autophagy-related protein expression has been docu­
as atrogin 1))46,47, which are part of the ATP-dependent mented, including increased expression of beclin 1,
ubiquitin–proteasome pathway, is strongly upregu- autophagy protein 5 and microtubule-­associated pro-
lated. Expression of these elements is largely under the teins 1A/1B light chain 3B (MAP1LC3B)57. Additional
control of the transcription factors forkhead box pro- transcription factors such as nuclear factor‑­κB (NF‑κB),
tein O1 (Foxo1) and Foxo3, whose activities are post-­ signal transducer and activator of transcription 3
translationally regulated48,49 and seem to function as a (STAT3) and CCAAT/enhancer-binding protein‑β
regulatory node between anabolic and catabolic pro- (C/EBPβ) also contribute to the regulation of the E3
cesses. Under physiological conditions, RAC serine/­ ubiquitin ligases and autophagy genes58–67 (FIG. 4). Given
threonine-protein kinase (AKT) phosphorylates the that animal models do not always recapitulate com-
FOXO proteins, causing their cyto­plasmic localization. plex events that occur in cancer cachexia in humans68,
However, in cachexia, AKT activity is often suppressed it will be important moving forward to validate the
either owing to the influence of inflammatory cytokines importance of these transcription factors by measur-
or to the decline in levels of insulin-like growth factor I ing their activities in skeletal muscle in patients with
(IGFI), which stimulates muscle anabolism. Decreased ­cancer-associated cachexia.
AKT activity leads to the dephosphorylation and sub- More clinical research is required to understand
sequent nuclear translocation of the FOXO proteins, the respective roles of protein synthesis and degrad­
which in turn enables their nuclear translocation and ation, the ubiquitin–proteasome system, autophagy
the transcription of TRIM63 and FBXO32, the induc- and specific signalling pathways in muscle-protein
tion of which correlates with the degradation of myo­ loss. The whole-body proteolysis rate has been stud-
fibrillar proteins, in particular, thick filament proteins ied in patients using isotopic tracer approaches and is
such as myosin heavy chain50,51. AKT also activates increased by a mean of 40%29. However, whether skele­
serine/threonine-protein kinase mTOR complex 1 tal muscle proteo­lysis is similarly activated is debated.
(mTORC1), which in turn activates ribosomal pro- Steady-state amino acid flux measurements across the
tein S6 kinase‑β1 (S6K1), thereby exerting anabolic legs in patients with c­ ancer suggest that muscle loss is
effects on muscle ­tissue. Thus, in cancer, as in other not necessarily driven by increased protein degrad­
chronic illnesses associ­ated with cachexia, muscle atro- ation69. When targeted RNA and protein analysis was
phy is likely regulated by an imbalance of anabolic and performed on muscle biopsy specimens, components of
­catabolic processes. the ubiquitin–proteasome pathway were associated with

NATURE REVIEWS | DISEASE PRIMERS VOLUME 4 | ARTICLE NUMBER 17105 | 5


©
2
0
1
8
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

weight loss in patients with cancer 70. However, global including malignant disease. An increasing body of evi-
gene expression analysis studies have thus far not been dence suggests that the CNS exerts overarching control
able to recapitulate findings in the protein catabolism of the pathogenesis of cachexia through the recog­
pathways, which have been widely described in animal nition of cytokines as molecular signals of sickness78.
models10,71. Decreased muscle-protein synthesis rates Animal studies are particularly germane to gaining an
have been described in patients with cancer-associated understanding of CNS function in cancer-­associated
cachexia72,73, as have decreases in the AKT pathway cachexia. Existing data support a model wherein periph-
and depletion of myofibrillar proteins74,75. However, eral inflammation is amplified and modified within
­muscle AKT activity is not always reduced in models of the mediobasal hypothalamus, creating a paracrine
cancer cachexia76 or in patients with cancer 77. Causes inflammatory milieu that in turn initiates and sustains
of this variation are currently unknown. Additionally, alterations in the activity of neuronal populations that
repeated measurements over time are lacking, such that regulate appetite and metabolic processes, including
the evolution of whole-body and ­tissue-specific changes proteolysis and lipo­lysis79,80. Hypothalamic exposure
remains to be determined. to any one of numerous inflammatory stimuli (such
as IL‑1β and TNF) triggers an acute illness response,
Homeostatic control in the CNS leading to anorexia, weight loss and skeletal muscle
Sickness behaviours (which include anorexia and catabo­ atrophy. These molecules act acutely by binding to
lism of lean body tissues), fever and lethargy are classic receptors on hypothalamic neuronal populations, such
responses in multiple forms of acute and chronic illness, as pro-­opiomelanocortin and Agouti-related protein

Tumour with
Tumour-derived catabolic factors
immune infiltration Crosstalk • Activins • Serotonin
• Myostatin • Parathyroid hormone-related protein
• TGFβ • Adrenomedullin, miR-21, HSP70
and HSP90 (exported in exosomes)

Pro-inflammatory mediators arising from


tumour–immune system crosstalk
• IL-1α • TNF • GDF15 • TNFRSF12A
• IL-1β • IL-11 • TWEAK • PGE2
• IL-6 • IL-17 • TRAF6
• IFNγ • LIF • Oncostatin M

CNS Target organs

White ‘Browned’ Brown


Neuropeptide Y
Skeletal adipose adipose adipose
Melanocortins and cardiac tissue tissue tissue
Adrenal
Appetite gland muscle
centres
Crosstalk

Outcome

Catabolic sympathoadrenal
outputs Anorexia Fatigue Muscle Excess Excess Futile
deconditioning proteolysis lipolysis cycling
Catabolic behavioural outputs Reduced food intake
Catabolic sympathetic outputs

Figure 3 | Interorgan relationships in cancer-associated cachexia. On the basis of clinical Nature Reviews | Disease
and experimental Primers
findings,
tumour-derived catabolic factors have been shown to act on target tissues to elicit excess catabolism. Numerous
pro-inflammatory cytokines are generated through tumour crosstalk with associated stromal cells and the immune system,
which act directly on target tissues as well as through alteration of central nervous system (CNS) controls of energy intake
and expenditure. Mobilization of adipose tissue results from reduced food intake as well as specific tumour-derived
lipolytic molecules (such as adrenomedullin), tumour factors that induce uncoupling and futile cycling in this tissue (such as
parathyroid hormone-related protein) and/or induce lipolysis by activating the sympathetic neural output to adipocytes.
Skeletal and cardiac muscle mobilization is induced by multiple pro-inflammatory cytokines, eicosanoids and transforming
growth factor‑β (TGFβ) family effectors (such as activin A and myostatin). Inflammation in the CNS alters the balance of
orexigenic neuropeptide Y and anorexigenic melanocortins, resulting in reduced food intake. CNS inflammation evokes a
catabolic programme in muscle, rapidly inducing atrophy. This effect is dependent on the production of glucocorticoids
by the adrenal gland. The dashed arrow represents a new finding, the importance of which in patients is currently unclear.
GDF15, growth/differentiation factor 15; HSP, heat shock protein; LIF, leukaemia inhibitory factor; miR, microRNA;
PGE2, prostaglandin E2; TNF, tumour necrosis factor; TNFRSF12A, TNF receptor superfamily member 12A; TRAF6,
TNF‑receptor-associated factor 6; TWEAK, TNF-related weak inducer of apoptosis (also known as TNFSF12).

6 | ARTICLE NUMBER 17105 | VOLUME 4 www.nature.com/nrdp


©
2
0
1
8
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

IL-6 LIF Activins


TWEAK TGFβ
TNF
IL6ST Myostatin
IGFI
TNFR TNFRSF12A ACVR2A or
TGFBR2
IGFR
STAT3
SMAD2/
SMAD3
NF-κB C/EBPβ

AKT PI3K Ca2+ mishandling


↑ Transcription Disrupted sarcomere,
Inhibition of muscle atrophy
P nuclear import and weakness
FOXO1 Contractile
E3 ubiquitin dysfunction
and/or ligase genes
FOXO3 Autophagy Myofibrillar
genes protein
Myonucleus breakdown

Figure 4 | Signalling pathways involved in tumour-induced skeletal muscle atrophy. Skeletal Naturemuscle atrophy
Reviews in cancer
| Disease Primers
cachexia is regulated by signalling pathways that are activated by cytokines produced by the tumour and stromal cells
within the tumour microenvironment and by cells of the immune system. These factors signal through their respective
cell surface receptors, which activate selective transcription factors; these in turn bind to promoters of genes encoding
components of the ubiquitin–proteasome and autophagy systems. In general, activation of these systems leads to the
selective destruction of myofibrillar proteins that form sarcomeres and provide contractile function to skeletal muscles.
Loss of these myofibrillar proteins presumably results in muscle atrophy and weakness. Alternatively, growth factors such as
transforming growth factor‑β (TGFβ) can signal to alter Ca2+ handling, leading to the dysfunction of the sarcomere
independent of the loss sarcomeric proteins. In addition to cytokines, growth factors such as insulin-like growth factor I
(IGFI) signal through RAC serine/threonine-protein kinase (AKT) to mediate functional repression of the transcription
factors forkhead box protein O1 (FOXO1) and FOXO3 by inhibiting their nuclear translocation and overall levels, which
together inhibit the transcription of atrophy genes. In cachexia, this inhibitory activity is often suppressed (indicated by
dashed lines), leading to the transcription of genes that encode E3 ubiquitin ligases and autophagy components.
ACVR2A, activin receptor type 2A; C/EBPβ, CCAAT/enhancer-binding protein‑β; IGFR, insulin-like growth factor receptor;
LIF, leukaemia inhibitory factor; IL6ST, IL‑6 receptor subunit-β (also known as GP130); NF‑κΒ, nuclear factor‑κB; P, phosphate;
PI3K, phosphoinositide 3‑kinase; SMAD, mothers against decapentaplegic homologue; STAT3, signal transducer and
activator of transcription 3; TGFBR2, TGFβ receptor type 2; TNF, tumour necrosis factor; TNFR, TNF receptor; TNFRSF12A,
TNF receptor superfamily member 12A; TWEAK, TNF-related weak inducer of apoptosis (also known as TNFSF12).

neurons, which trigger a feedforward loop that involves to lipolysis rather than the irreversible degeneration of fat
skeletal muscle-protein catabolism and lipolysis81. CNS- cells owing to apoptosis84,85 and that the overall increase
delimited IL‑1β signalling alone can evoke a catabolic in whole-body lipolysis in patients with cachexia is
programme in muscle, rapidly inducing atrophy 81. ~50%29. Biopsy studies from patients have additionally
This effect is dependent on hypothalamic–pituitary–­ shown that in white adipocytes, the lipolytic effects of
adrenal axis activation, as CNS IL‑1β‑induced atrophy catecholamines and natriuretic peptide were increased
is blocked by adrenalectomy or by muscle-specific knock 2–3‑fold in patients with cancer-associated cachexia86.
out of glucocorticoid receptors81. Thus, the involvement Fat and muscle atrophy in cachexia have been consid-
of glucocorticoids in tumour-associated muscle wasting ered independent events owing to the fact that cytokines
seems likely. such as TNF can induce catabolism in both skeletal
Studies of CNS regulation in clinical cancer-­ ­muscle and adipose cells. However, this concept was
associated cachexia to date are limited to investigations ­challenged by genetic studies performed in mice in which
of the systemic levels or administration of neuromodu­ Pnpla2 (which encodes patatin-like phospholipase-­
latory peptides, such as appetite-regulating hormone domain-containing protein 2) was ablated in mice bear-
(also known as ghrelin)82. Approaches such as functional ing Lewis lung carcinoma xenografts. Pnpla2‑deficient
MRI brain studies have been used to investigate regional mice were resistant to the lipolysis of white adipose tissue
CNS activity in obesity and anorexia nervosa but remain but surprisingly also retained hindlimb muscle mass87.
to be used in cachexia research. This finding suggested that fat loss predisposes to muscle
loss in cancer-associated cachexia. Similar conclusions
Adipose tissue depletion were reached in a study in which secretion of para­
In addition to skeletal muscle, a substantial portion of thyroid hormone-related protein (PTHLH) from Lewis
weight loss in patients with cancer derives from the lung tumours in mice was shown to alter the thermo­
depletion of adipose tissue6,83. Studies have shown that genesis of adipose tissue via the ‘browning’ of white adi-
this depletion results from a reduction in fat mass owing pose cells88. Thermogenesis is regulated by mitochondrial

NATURE REVIEWS | DISEASE PRIMERS VOLUME 4 | ARTICLE NUMBER 17105 | 7


©
2
0
1
8
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

brown fat uncoupling protein 1 (UCP1), the expression including subset analyses93. Inclusion of contempor­
of which increases in various mouse models of cancer ary patients ensures representation from a variety of
cachexia as well as in white adipose tissue from patients popu­lations, body weight demographics and treatment
with cancer-associated cachexia33. Use of an anti-PTHLH plans. Until uniform criteria become available, diagno-
antibody inhibited adipose browning as well as the sis remains based on clinical experience and includes
loss of skeletal muscle mass, suggesting that altered fat assessing weight loss, food intake and ­abnormalities
­metabolism is a prerequisite for skeletal muscle atrophy. of metabolism.
How fat loss predisposes to skeletal muscle atrophy is
not known. Although several mechanisms have been pro- Diagnosis
posed to account for tumour-induced lipo­lysis — includ- Weight loss. The presence of weight loss is an impor-
ing the presence of inflammatory cytokines released from tant clinical sign that can even be the first detectable
infiltrating tissue macrophages89,90, induction of adipose manifesta­tion of the presence of cancer. After the possi-
triglyceride lipase91 and loss of 5ʹ‑AMP-activated pro- bility of intentional weight loss (for example, by dieting)
tein kinase (AMPK)92 — whether one or more paracrine has been excluded, alternative causes of weight loss of
factors are capable of crosstalk between fat and skeletal unknown origin are investigated. Weight loss is typi-
muscle to mediate the ­catabolism of myofibrillar proteins cally the first element of a cachexia diagnosis and has a
in unclear. distinctive course in each patient. For example, weight
loss can occur before or after the cancer diagnosis and
Cardiac muscle atrophy can be slow, intense, continuous or episodic; it should
Little research is available concerning the effects of be monitored over time and referenced to the patient’s
cachexia on vital organs in those with cancer. Cardiac pre-cancer body weight.
muscle performs an essential physiological role and was Weight loss varies in its severity: a 5% loss is con-
assumed to be spared in cachexia, as it cannot simply be sidered the threshold of major risk of poor clinical out-
exploited as a repository of amino acids as skeletal m
­ uscle come1,2, with increasing risk as weight loss cumulatively
can. Although cardiac atrophy remains to be evaluated reaches 10%, 15%, 20% or higher 86. Cancer-associated
in human cancer-associated cachexia, research in ­animal cachexia contributes to poor prognosis through progres-
models has shown substantial cardiac atrophy in multi- sive depletion of the body’s energy and protein reserves;
ple cachexia-inducing tumour models along with echo- thus, it is relevant to determine the impact of weight
cardiography-defined evidence of cardiac functional loss as a function of initial body reserves2. The prognos-
impairment 18. The mechanisms of cardiac muscle atro- tic importance of weight loss in patients who initially
phy are also described in animal models and are highly have a low, intermediate or high BMI was determined
similar to those proposed for skeletal muscle, involving within an international cancer cachexia data repository
­inflammation, proteolysis, apoptosis and autophagy 36. including >10,000 patients22. This multivariate analysis
of the association between BMI, weight loss and mortal-
Diagnosis, screening and prevention ity was controlled for age, sex, cancer site, cancer stage
Although our understanding of cancer-associated and performance status22. A grading system based on
cachexia has progressed, a single unified international combinations of BMI and weight loss was developed
set of diagnostic criteria is not available. Indeed, a host to differentiate groups with distinct median survival
of disparate diagnostic criteria have been reported2,93, dur­ations22 (FIG. 5). This grading system has been valid­
which are a detriment to the identification and treat- ated94 and included in current international clinical
ment of cachexia in clinical practice. Regardless of which practice guidelines13.
criteria are used, weight loss, either alone or in combin­
ation with one or several additional features (such as Other criteria. The diagnosis of cancer cachexia will
anorexia, reduced food intake, muscle loss, decreased inevitably include additional information beyond
strength, fatigue and biochemical markers93) is always weight loss. Although no consensus is available on the
included. In addition to the use of different combin­ definitions of and methodologies for measuring skele-
ations of diagnostic criteria, heterogeneity in data collec­ tal muscle depletion, reduced food intake and the bio­
tion and reporting of each individual criterion makes logical indicators of altered metabolism and catabolism,
defining cachexia for clinical use difficult. measure­ment of these elements is becoming increasingly
International consensus groups have begun to address ­specific, precise and clinically available.
these disparities and have provided a conceptual frame- A key driving mechanism of cachexia is reduced
work for the classification of cancer-associated cachexia2. food intake2. The gap between energy expenditure and
Any useful classification criteria will define definitive energy intake can be estimated from direct measures of
cut-off values for each diagnostic criterion from large resting energy expenditure (indirect calorimetry) (BOX 1)
contemporary data sets by determining the values that and records of dietary intake95. A variety of validated
relate optimally to meaningful patient-­centred out- questionnaires are also available to assess quantity and
comes2. Data sets that include information collected type of dietary intake96 (BOX 2).
in a standardized fashion are necessary and should be There is no set standard for clinical assessment of
large enough to capture representative distributions of skeletal muscle mass; however, most published data
candidate diagnostic criteria, relevant covariates and have been collected from axial lumbar CT images
outcomes for adequately powered statistical analyses, (FIG. 2). Standard oncological CT images collected for

8 | ARTICLE NUMBER 17105 | VOLUME 4 www.nature.com/nrdp


©
2
0
1
8
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

cancer diagnostic purposes offer a new opportunity to BMI (kg per m2) Median
precisely quantify skeletal muscle and fat and to evalu- 28 25 22 20 BMI–WL survival
grade (months)
ate their changes over time23. Because these images are 0 0 1 1 3
2.5 0 20.9
almost never of the whole body, a convention has been
1 2 2 2 3
adopted to use a single lumbar image anchored at the 1 14.6

WL (%)
6
third lumbar vertebra as the cross-sectional areas of 2 3 3 3 4 2 10.8
11
muscle, visceral fat and total fat in this area are highly 3 3 3 4 4 3 7.6
correlated with whole-body volumes of these tissues. 15
3 4 4 4 4 4 4.3
Using this approach, low levels of muscle mass associ-
ated with treatment complications and mortality have Figure 5 | Grading scheme Nature
for WL Reviews
on the| Disease
basis ofPrimers
risk
been characterized3,23. CT‑defined skeletal muscle mass of mortality in patients with advanced-stage cancer.
measurements have been increasingly reported in the The grading scheme was developed based on groupings
literature (including >20,000 patients with c­ ancer to of body mass index (BMI) and weight loss (WL), showing
date since 2008), with calls for this approach to be used distinct median survival durations. The analysis was laid
more widely 97. Data are available in different diseases, out in a 5 × 5 matrix representing 5 different WL categories
cancer sites, cancer stages and ethnic groups; some within each of the 5 different BMI categories, producing
25 possible combinations of WL and BMI. A multivariate
provisional sex-specific cut-off values are available as
survival model was adjusted for age, sex, cancer site,
benchmarks to identify patients with muscle deple- cancer stage and performance status. Grade 0 was
tion23,93,98. These cut-off values have been determined assigned to the subgroups in the matrix with the lowest risk
using statistical methods to identify risks (such as mor- (longest survival), and grades 1–4 were assigned to the
tality, toxicity and quality of life) that emerge at specific subgroups according to decreasing survival. Grades were
threshold levels of skeletal muscle loss. As CT data con- developed based on 8,160 patients and an external
tinue to accumulate, these can be aggregated to develop validation cohort of 2,683 patients. Reprinted with
sex-specific and age-specific reference values for skeletal permission. © 2017 American Society of Clinical Oncology.
muscle depletion3. All rights reserved. Martin, L. et al. J. Clin. Oncol. 33 (1),
The specific abnormalities of metabolism that define 90–99 (2015).
cachexia in a given individual are not routinely assessed.
Although pro-inflammatory cytokines signal catabolic Screening
effects via their tissue receptors in muscle, adipose ­tissue Cachexia screening is performed with the aim of increas-
and hypothalamic neurons, serum cytokine levels have ing awareness and enabling early recognition and treat-
been proven to be inconsistent as biological criteria ment. To detect cachexia at an early stage and to detect
in cachexia diagnosis. As a consequence, laboratory its acceleration, regular evaluation of weight change
measures of the acute phase response (which involves and BMI is needed, beginning at cancer diagnosis and
proteins that are part of the innate immune system repeated depending on the stability of the clinical situ­
response to neoplasia) are used as surrogate indices ation. Cancer sites, stages and treatment plans with
of the inflammation-associated catabolic drive2,95. The higher prevalence of cachexia are a priority for screening.
acute phase response is characterized by leukocytosis, Height and weight are routine, if not mundane, clin-
fever and changes in the plasma concentrations of pos- ical measures, but the continuity of these measures over
itive acute phase proteins (namely, fibrinogen, α1‑acid time is essential to avoid large cumulative weight loss
glycoprotein, serum amyloid A and C‑reactive protein going unnoticed. Screening for weight loss is performed
(CRP)) and negative acute phase proteins (namely, as part of an evaluation of nutritional risk within clin-
albumin and transferrin)99. Typical laboratory values ical nutritional services of cancer centres and is linked
associated with cachexia are albumin of <35 g per litre, with nutritional therapy and monitoring of outcomes.
transthyretin of <110 mg per litre and CRP of >10 mg Mandatory screening for weight loss in patients with
per litre. Used alone or in composite indices such as the cancer has been established in some countries13, with the
Glasgow Prognostic Score (which provides scores based intent of detecting in‑hospital malnutrition. Screening
on albumin (<35 g per litre), CRP (>10 mg per litre) can be efficient, brief and inexpensive. Patient-reported
or both), acute phase response proteins correlate with outcomes are of value in the assessment of various facets
weight loss and are powerful prognostic indicators of of cachexia; evidence supports the reliability of patient
tumour progression, survival and symptom burden in self-reported height, weight and weight history 13.
multiple cancers100,101. CRP testing is not routine every- Weight loss history and an index of food intake may be
where, but neutrophil to lymphocyte ratios offer similar obtained directly or via validated nutritional screening
prognostic information102,103. tools13,107 (BOX 2).
Various pro-cachectic mediators suggested by pre- Abnormal screening results alone do not provide
clinical investigations are being evaluated in patient enough information to design individualized cachexia
populations at risk of cachexia. For example, PTHLH care pathways. Patients with a history of substantial
was shown to be independently prognostic for weight weight loss, therefore, need to be followed-up with specific
loss104, whereas increased levels of GDF15, IL‑6 and IL‑8 assessments to determine the origin and severity of food
are correlated with weight loss105. In the future, changes intake impairment and metabolic derangements. Within
of inflammatory markers over time might also be useful the conventional organization of cancer care, clinical
as markers of the effectiveness of cachexia therapy 106. ­services might exist that have aspects of the management

NATURE REVIEWS | DISEASE PRIMERS VOLUME 4 | ARTICLE NUMBER 17105 | 9


©
2
0
1
8
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

of cachexia in their charge, but no set standard is avail­ cancers of the lung, pancreas, oesophagus, stomach,
able. Only a few institutions possess a dedicated cachexia bowel and liver will experience cachexia1,14–17, and this
clinic108–111. Otherwise, cachexia can fall into the purview foreknowledge should be a basis for early and system-
of symptom control or palliative care but may be equally atic attention to cachexia management. Although exact
well-attended to by clinical nutritional services, insofar diagnostic criteria for ‘pre-cachexia’ remain undeter-
as access to dietitians and medical nutritionists is often mined, this is a useful concept that enables preventive
available in cancer centres and hospitals. Because of the intervention at the onset of low-grade weight loss.
important role of reduced dietary intake in development Recent phase III trials have adopted a strategy of early
of cachexia, presence of this issue can be part of primary intervention, moving away from using 5% weight loss
screening and used to direct further assessments towards as an inclusion criterion and instead including patients
identifying needs for nutritional support. with either minimal weight loss (≥2%; for example,
NCT00467844)112 or removing a requirement for prior
Prevention weight loss altogether (for example, NCT01355484)113.
Prevention has not been the standard in the clinical This approach is based on a high and imminent
approach to cancer-associated cachexia. Although weight ­probability of patients in the trial developing cachexia.
loss can occur early in the natural history of the cancer, The catabolic sequelae of cancer treatments can
active treatment of cachexia has often been left to the end add substantially to overall weight loss. For example,
stages of the disease and the refractory stage of the cach­ mean weight losses in patients receiving neoadjuvant
exia. However, a shift towards considering preventive chemotherapy for oesophago-gastric cancer (~4.2 kg)117
cachexia therapy (see below) is apparent. Notably, recent or receiving chemoradiotherapy for head and neck
trials of cachexia therapeutics112,113 have been shifted ­cancer (~11.4  kg)25 are substantial, and these losses are
to an earlier time in the disease trajectory and include composed mostly of lean tissue26,117. Preventive measures
therapies that are delivered concurrently with first-line can also be deployed in anticipation of these losses.
chemo­therapy rather than in the end of life phase114–116.
The earliest possible approach would be contingent on Nutritional and metabolic treatment
developing a clear understanding of the predictors of Provision of adequate nutrition is a mainstay of cachexia
cachexia, including the pre-cachexia biomarkers. treatment, and up to date guidelines for clinical nutri-
tion in oncology are available13. First-line approaches
Management include oral nutritional supplements (sterile liquids,
Cancer-associated cachexia evolves over time, and goals semi-solids or powders that provide macronutrients
of care should be established at each stage of the evo- and micronutrients for individuals who are unable to
lution. The majority of patients with advanced-stage meet their nutritional requirements through oral diet)
and consultation with a nutritional healthcare pro-
fessional to increase the quantity and quality of the
Box 2 | Assessment of dietary intake in clinical practice patient’s food. In patients in whom the dominant cause
Food intake — assessed as a component of clinical questionnaires — is used to screen of weight loss is deficit of dietary intake (for example,
or assess nutritional status. those receiving high-dose chemo­t herapy ahead of
• Example tools include the Patient-Generated Subjective Global Assessment bone marrow transplantation, in whom deficits of oral
(PGSGA)189, Mini Nutrition Assessment (MNA)190,191, Malnutrition Screening Tool intake can exceed 1,200 kcal per day), active nutritional
(MST)192, Malnutrition Universal Screening Tool (MUST)193, Short Nutrition Assessment manage­ment (using enteral and/or parenteral nutri-
Questionnaire (SNAQ)194 and Nutrition Risk Screen (NRS 2002)195. tion) (BOX 3) leads to better treatment tolerance and
• Questionnaires are completed by a healthcare provider or the patient, with responses better quality of life13. For the majority of patients, and
being categorical in nature. Reductions in food intake are assessed as a ‘yes’ or ‘no’ particularly for those with advanced-stage cancer, the
response or from a categorical list describing the severity of the reduction presence of insufficient dietary intake is not always
(for example, ‘no decrease in food intake’, ‘moderate decrease in food intake’
identified and active nutritional management is not
or ‘severe decrease in food intake’).
always implemented17. Furthermore, compliance to
• The time frames for assessing reductions in food intake tend to be retrospective and
oral nutritional supplements is generally low, and in
are variable (for example, current intake, recent intake, intake in the past week, intake
some cases, nutritional supplements merely displace
in the past month or intake in the past 3 months).
food intake at mealtime118. If volitional food intake
• These tools are clinically practical and expedient and are obtained from patient reporting.
remains insufficient after dietary consultation and oral
• The information obtained identifies patients with reduced food intake who might nutritional supplements have been deployed, escalation
benefit from further dietary assessment and intervention.
to artificial (enteral or parenteral) nutritional support
Food intake can also be assessed from patient food records. is an option (BOX 3). Orexigenic drugs (appetite stimu-
• Example tools include 24‑hour recall (of food consumed the previous day) lants) have been developed to counteract low appetite
and collection of diaries of food consumption during a 1–7‑day period. in patients with cachexia. Cannabinoids, cortico­steroids
• Current food and fluid intakes are recorded by the patient, and a healthcare and progestogens have these actions, but adverse effects
professional enters the information into a country-specific nutrient database; constrain their use119. For example, corticosteroids
additionally, macronutrient and micronutrient estimates are calculated. increase appetite but result in skeletal muscle atrophy.
• Diet records are burdensome to the patient and to the healthcare provider who must Progestogens increase appetite but cause muscle atro-
process the collected information but are useful for determining food preferences and phy and additionally increase risk of thromboembo-
patterns of consumption, which can aid in the development of a nutritional intervention.
lism119. New therapies to increase food intake that are

10 | ARTICLE NUMBER 17105 | VOLUME 4 www.nature.com/nrdp


©
2
0
1
8
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

Box 3 | Options for nutritional support in patients with cancer cachexia and are drawn from every point of our under-
standing of cachexia physiology, including tumour-­
Volitional nutrition specific factors, pro-inflammatory cytokines and
Volitional nutrition refers to the oral ingestion of nutrients as normal food and/or oral eicosanoids, and mediators of organ-specific or tissue-­
nutritional supplements. Dietary modification (such as increased calorie or protein specific control systems (FIG. 6). Our understanding of
density or texture modifications) and oral nutritional supplements are typical
the underlying mechanisms of cachexia in individual
first-line interventions to improve intake in patients with cancer. Management
of pain and symptoms is essential to optimize volitional food intake. patients is crude at best; accordingly, further character­
ization of the clinical aetiologies is needed to assist
Artificial nutrition individuals whose cachexia might be strongly driven
Artificial nutrition is non-volitional feeding and is initiated when an individual cannot
by a tumour-secreted factor, eicosanoids and cytokines,
meet their nutritional requirements via oral intake. The indications to implement
artificial nutrition are either total inability to eat for >1 week or an energy intake <60%
endocrine deficits (such as i­ nsulin resistance or hypo-
of the requirements for >2 weeks13. Artificial nutrition is provided by the enteral route, gonadism), a comorbidity, a treatment, a psychosocial
unless the gastrointestinal tract is not functional, and includes the following modes: factor or a symptom (such as pain). One approach
• Enteral nutrition (tube feeding), which is any mode of feeding that uses the might be to assess and ‘rank’ different pro-­cachectic
gastrointestinal tract to deliver all or part of a patient’s nutritional requirements. A tube mechanisms to guide treatment for that individ­ual;
is used to access either the stomach or jejunum. This might be used, for example, in the however, routine management has not achieved this
case of tumour obstruction of the oesophagus or dysphagia in pharyngeal cancer. level of sophistication and aetio­logy-based diagnostic
• Parenteral nutrition (intravenous feeding), which is a mode of feeding that delivers all or criteria have not been standardized in ­clinical care or
part of a patient’s nutritional requirements intravenously via a central or peripheral vein, in clinical trials.
thereby completely bypassing the gastrointestinal tract. This might be used, for example,
in patients with multisite bowel obstruction owing to disseminated ovarian cancer. Symptom control
Associated risks of artificial nutrition include infections, gastrointestinal adverse effects Cachexia does not occur in isolation; it occurs within
(such as nausea, vomiting, diarrhoea and hepatic abnormalities), metabolic a variable terrain of comorbid conditions, cancer treat-
dysregularities (such as hyperglycaemia) and mechanical complications (such as ment response and toxicity and alongside pain and
blocked tubes). High-quality evidence is lacking for the use of artificial nutrition to treat other symptoms. Symptoms are a considerable source
cancer-associated cachexia; however, in settings in which intake is severely impaired of clinical heterogeneity in weight-losing patients, can
primarily owing to tumour location or symptoms of treatment, artificial nutrition can be
change rapidly over the course of the disease trajec­
partially effective and improve outcomes13.
tory and treatment plan 126 and are most common
among patients receiving treatment for advanced-stage
under investigation include growth hormone secreta- ­cancers127 — but they remain undetected by clinicians in
gogue receptor type 1 (ghrelin receptor) agonists82,112 up to 50% of patients128. Symptoms present at any point
and melanocortin receptor 4 antagonists120; these agents in time are only a snapshot of a longer story that can last
act on the hypothalamus, which regulates appetite and several years, with multiple treatments and complica-
satiety, but also have systemic effects to promote protein tions contributing to progressive weight loss. Cachexia
anabolism and energy storage. cannot be divorced from these circumstances; good
Anabolic deficit may be partly addressed by main- medical management of pain and symptoms is another
tenance of physical activity, a notion that is endorsed major principle of cachexia management. Often, multi­
within oncology nutrition clinical practice guidelines13 ple causes of potentially reversible weight loss must be
as well as in the design of clinical trials of multimodal assessed and appropriately managed; this is a point for
intervention (for example, NCT02330926). Patients action and of immediate benefit to the patient 121.
should be given support to enable them to exercise Causes of weight loss that are amenable to effec-
within their safe capacity 13,121. Cachexia in chronic tive management include, but are not limited to, pain,
non-malignant illnesses such as chronic obstructive ­nausea, vomiting, dental problems, dysphagia, early
pulmonary disease has long been managed by a multi­ satiety, oesophageal obstruction, malabsorption, endo-
modal approach (including nutrition and exercise)122. crine and metabolic disorders, anxiety, depression,
Indeed, a systematic review of 16 trials in patients with distress and inability to sleep. These issues should be
cancer who were undergoing active oncology treatment treated according to clinical practice guidelines for pain
showed that aerobic exercise, resistance exercise and a and symptom management 119. Clinical trials of cachexia
combination of the two improve upper and lower body therapy do not include a common standard of support-
muscle strength compared with usual care123. However, ive care e­ lements across centres, which has been sug-
it has been noted that patients with established cachexia gested to be a source of heterogeneity in individual
might lack the motivation and self-efficacy to undertake patient response2. A multidisciplinary team approach
regular structured exercise124. Others have proposed to supportive care is needed13, the benefits of which
interventions designed to provide exercise intervention have been reported from prospectively conducted non­
optimized for individual patient activity tolerance125. randomized studies129,130. Clinical services are emerging
Altered metabolism remains the most challenging that operate in partnership between pallia­tive care physi­
aspect of cancer-associated cachexia for therapeutic cians and the onco­logy community, as endorsed cur-
intervention. Specific therapeutic targets have been rently by many cancer agencies, including the American
proposed for testing in clinical trials (TABLE 1) on the Society for Clinical Oncology 131; these services will
basis of preclinical investigation and cover a broad range be an asset for efficient management of s­ ymptoms
of mechanisms. These targets reflect the complexity of ­contributing to cachexia.

NATURE REVIEWS | DISEASE PRIMERS VOLUME 4 | ARTICLE NUMBER 17105 | 11


©
2
0
1
8
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

Table 1 | Example phase III clinical trials for cancer-associated cachexia, anorexia and skeletal muscle loss
Study n Therapy (mechanism) Therapeutic Results
approach
Temel et al.112 979 Anamorelin (growth hormone Appetite- • Increased lean body mass
(NCT01395914 secretagogue receptor type 1 modifying • No effect on handgrip strength
and NCT01387269) (ghrelin receptor) agonist) and anabolic
Crawford et al.113 651 Enobosarm (selective androgen Anabolic • Increased lean body mass, stair climb power and stair climb
(NCT01355484 receptor agonist) speed in those receiving taxane chemotherapy
and NCT01355497) • Increased lean body mass in those receiving nontaxane
chemotherapy
Bourdel- 341 Dietary advice Nutritional • No change in survival at 1 year and 2 years
Marchasson et al.196 • No change in chemotherapy toxicity
(NCT00459589) • No change in body weight
Sánchez-Lara et al.197 96 Omega‑3 fatty acids (oral nutritional Nutritional • Increased weight
(NCT01048970) supplement) • Increased lean body mass
• No change in chemotherapy response
Madeddu et al.198 60 l‑Carnitine, celecoxib and megestrol Multimodal Noninferiority of two-agent versus three-agent combination
acetate versus l‑carnitine and with respect to lean body mass and total daily physical activity
celecoxib (nutritional supplement and
cyclooxygenase 2 (COX2) inhibitor
with or without appetite stimulant)
NCT02330926 240 Ibuprofen, physical activity, dietary Multimodal In progress; assessing body weight, muscle mass and physical
advice and omega‑3 fatty acids activity
NCT02138422 276 Xilonix (XBiotech; Anti- In progress; assessing disease response rate, muscle mass
anti‑IL‑1α antibody) inflammatory and appetite
NCT02553187 160 Kanglaite (coicis oil) Herbal or In progress; assessing body weight and lean body mass
alternative
NCT02802540 78 Nabilone (synthetic cannabinoid) Other In progress; assessing anorexia, weight loss and calorific intake

Attention to treatment-related adverse effects such the condition. Similarly, a multimodal approach target­
as nausea and vomiting according to clinical practice ing the multiple facets of cachexia is likely to be the
guidelines132 is an important aspect of management optimal approach135–138. The rationale is that address-
of patients with cachexia, as they show increased rates of ing food intake alone (for example) would be insuffi-
severe toxicity 3,133,134. In routine clinical practice, patients cient because this would not necessarily be expected to
of advanced age or who have reduced levels of fitness are mitigate excess catabolism (and vice versa). In cancer
often started on a lowered dose or provided a regimen surgery, a multi­modal approach is embraced by multi-­
with reduced toxicity, at the treating oncologist’s discre- component enhanced recovery after surgery proto-
tion. These approaches might be relevant for patients cols139. These multimodal perioperative care pathways
with cachexia. The hypothesized association between are designed to achieve early recovery by maintaining
cachexia and toxicity of systemic cancer therapy is that pre­operative organ function and reducing the pro-
patients with low muscle mass have a reduced volume of found stress response follow­ing surgery. Key elements
distribution in relation to the dose of chemotherapy that include preoperative counselling, optimization of nutri-
they receive133,134. For example, when body surface area tion, standardized analge­sic and anaesthetic regimens
is used as the basis for dosing cytotoxic chemotherapy, and early mobilization140. In current clinical practice,
the dose may distribute, be metabolized and be cleared multi­modal cachexia therapy is achieved by a collabor­
within a grossly depleted lean compartment. An associ­ ative approach that engages a multidisciplinary team of
ation of dose-limiting toxicity with sarco­penia in different health professionals as well as patients and their famil­
treatment settings has repeatedly been shown3. Further ies. Expertises in multi­modal therapy include clinical
pharmacokinetic investigations are eagerly awaited to onco­logy, clinical nutrition and palliative care teams as
clarify whether patients with sarcopenia experience well as the possibil­ity of specialist referral (for example,
greater drug exposure during cancer treatment and to to gastro­enterologists). Guidelines and protocols exist,
provide a cachexia-specific basis for dose modifications. and descriptions of dedicated cachexia services have been
reported108–111. In ongoing clinical trials (for example,
Multimodal care NCT02330926), multimodal intervention is formalized
The inherent complexity of cancer cachexia calls for a and the study design can include nutritional therapy, anti-­
multifaceted assessment strategy that focuses on food inflammatory interventions and/or exercise therapy.
intake, pain and symptoms, specific losses of muscle and However, such multimodal interventions remain the
fat, catabolic factors, tumour burden, systemic inflam- minority of a­ vailable studies (FIG. 6; TABLE 1).
mation and altered endocrine status as well as on the A critically important underlying concept is that the
clinical, functional and psychosocial consequences of driving forces advancing tumour growth and metastases

12 | ARTICLE NUMBER 17105 | VOLUME 4 www.nature.com/nrdp


©
2
0
1
8
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

are the same driving forces that underlie cachexia. assist decision-­making concerning the use of parenteral
Accordingly, a combined and collaborative approach ­nutrition in patients with advanced cancer 145.
to cancer and cachexia therapy, rather than sequen- A particular weakness of many previous clinical
tial (or parallel) approaches, would seem most logical. ­trials of cancer cachexia interventions was the inclu-
The potential for a downward spiral, in which cachexia sion of patients with refractory cachexia or a mixture of
exacer­bates treatment toxicity 141,142, which then fur- these with individuals at earlier stages. Not surprisingly,
ther exacerbates cachexia, should be acknowledged. a substantial proportion of patients died within just a few
This ­spiral can be interrupted by careful management of weeks of random assignment114–116. Cachexia is associated
cachexia throughout treatment and attention to provid- with mortality, and some of the suggested mechanisms
ing chemotherapy at doses that are within the limits of include cardiac arrhythmias, electrolyte abnormali-
tolerability for patients already affected by cachexia at the ties that increase the risk of developing arrhythmias,
time that treatment is initiated. thromboembolic events, respiratory difficulties due to
diaphragm muscle weakness, aspiration pneumonias
Cachexia at the end of life due to the bedridden state and swallowing difficulties,
At the distal end of the cachexia spectrum, the dis­order gastrointestinal mucosal atrophy leading to endotoxin
can become refractory to therapeutic intervention; absorption, poor wound healing and sepsis146. Clinical
when catabolism intensifies exponentially6,143, cachexia trials of cachexia therapy have often had inclusion ­criteria
is driven by progressive disease that no longer responds such as an expected survival of >6 months. However,
to antineoplastic therapy, patients frequently become in the absence of adequate prognostic algorithms, such
emaciated and death becomes imminent. These blatant ­criteria have not prevented inclusion of imminently dying
manifestations heighten the sense of anguish related to patients in trials. High rates of attrition and missing data
cachexia for patients and their families. Psychological have created great d ­ ifficulty in interpreting the results of
support is key at this point, with less emphasis on ­dietary many investigations.
intake as a therapeutic objective.
Cachexia therapies are associated with risks, burdens Quality of life
and costs that need to be weighed against the expected Cachexia-related quality of life is not currently evalu-
benefits with the knowledge and consent of the patient. ated by agreed instruments. The European Organization
In refractory cachexia, medical interventions may be for Research and Treatment of Cancer QLQ‑C30
futile or inappropriately invasive13,144. Artificial nutri- (EORTC‑QLQC30) and Functional Assessment of
tional support in the form of parenteral nutrition is a Anorexia Cachexia Therapy (FAACT) questionnaires
well-known example; clinical practice guidelines in are the most commonly used instruments in the liter-
nutrition, oncology and palliative care consistently ature, but the former is not specific to cachexia and the
agree that in patients expected to survive <2 months, latter has been criticized as being not fully representa-
initiation of parenteral feeding is not recommended13,144. tive of all relevant domains147,148. A new cachexia quality
Robust predictors of a patient’s entry into the end of life of life instrument is in development within the EORTC
phase are needed to limit the potential harm of aggres- paradigm: the QLQ‑CAX24 is a cancer cachexia-specific
sive anti-cancer therapy and anti-cachexia therapy. questionnaire comprising 24 items for health-related
Prognostic models for survival have been developed to quality of life assessment in clinical trials and clinical
practice149. It contains five multi-item scales (food aver-
sion, eating and weight loss worry, eating difficulties, loss
of control and physical decline).
Psychological dimensions of the experience of cancer-­
associated cachexia from the perspective of patients and
their family members are less well explored than the
biomedical aspects of the condition150–155. Negative inter­
actions with food and eating are described at every stage
of cachexia; because of alterations in the perception of
Therapeutic approach the taste, smell and texture of food, usual foods and even
Nutritional favourite foods can become unpleasant or even repul-
Herbal or alternative sive126,151,152. Chewing and swallowing can be painful and
Exercise an unpleasant sensation of early ­satiety makes patients feel
Multimodal too full to continue eating or repelled by the quantity of
Appetite-modifying drugs food served. Furthermore, patients express distress about
Anti-inflammatory drugs
weight loss and not eating enough and are highly aware of
Anabolic drugs
death as the ultimate result of the weight loss126. Patients
Other drug types
describe laborious efforts to maintain and increase food
Figure 6 | Proportional distribution of therapeutic approaches
Nature Reviews | Disease
in clinical trialsPrimers intake and are often frustrated by emergent pain and
of cancer-associated cachexia therapy. Although not exhaustive, this summary symptoms. The wasted appearance of cachectic patients
of 134 trials (including published works and ongoing investigations reported in with c­ ancer is a major source of concern for both
www.clinicaltrials.gov) of the major classes of cachexia therapies includes treatments patients and families150,154. This experience occurs within
in phase II–IV clinical trials. unique ­personal, social, historical and cultural contexts.

NATURE REVIEWS | DISEASE PRIMERS VOLUME 4 | ARTICLE NUMBER 17105 | 13


©
2
0
1
8
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

Loss of control is another major theme in patient nar- activation of the ubiquitin–proteasome and autophagy
ratives concerning the experience of cancer-­associated pathways that drive muscle loss from within the myo­
cachexia. Even the most successful cachexia therapies fibre158. Failure of muscle stem cells to undergo differ-
only slow the rate of weight loss and weight is lost entiation impairs muscle repair, whereas expansion of
regardless of the level of food intake. Fatigue, weakness non-muscle progenitor cells that express E3 ubiquitin-­
and loss of independence compound the sense of help- protein ligase RNF5 and platelet-derived growth factor
lessness in patients, as does feeling pressured by others receptor-α (PDGFRα) can lead to muscle fibrosis and
with respect to food and eating 126,153,154. If the patient fatty infiltration.
and his or her family members do not have an equal Another recent report in seven different mouse
degree of understanding of disruption in food con- ­models suggested a link between muscle weakness and
nections, conflict with family members over food and bone metastasis arising from solid tumours such as
social isolation can ensue150. For patients, the inabil­ breast cancer or from lytic bone lesions owing to multi­
ity to share food in the manner hoped for by f­ amily ple myeloma161. Lytic metastatic lesions release TGFβ
members can contribute to a larger concern about being from the bone matrix; circulating TGFβ was shown
a burden150. to signal to skeletal muscle through the transcription
Psychosocial support for patients and families is factors mothers against decapentaplegic homologue 2
recom­mended as part of cachexia management, espe- (Smad2) and Smad3 to induce the transcription of Nox4
cially in the refractory stage135,136. The components of (which encodes NADPH oxidase 4). NADPH oxidase 4
such support are under development 155 and include stabil­izes ryanodine receptor 1 (Ryr1) via oxidation,
reduced emphasis on preparing and serving food, causing aberrant Ca2+ leakage from the sarcoplasmic
­enabling the family to provide aid without applying pres- reticulum and resulting in muscle weakness161. Muscles
sure, and providing information about cachexia and how from patients with lung and prostate cancer with
it will affect the patient and their families. metasta­ses in bone also exhibited oxidized RYR1. Many
solid tumours, including cancers of the stomach, pan-
Outlook creas, breast, colon and rectum162–165, release TGFβ into
Cancer-associated cachexia represents a constellation the systemic circulation. Using data from The Cancer
of various aetiologies that require further study from Genome Atlas (TCGA), a recent study documented
many angles, as illustrated in this Primer. Encouragingly, the presence of a strong inflammatory gene signature
cachexia now has wider recognition as an unmet med­ in pancreatic cancer that includes IL‑6 and IL‑11 and
ical need156: its research is supported by an international that occurs in conjunction with a strong TGFβ gene
society — the Society on Sarcopenia, Cachexia and signature166. Thus, inflammatory cytokines such as IL‑6
Wasting Disorders (http://society-scwd.org/) — and it and IL‑11 possibly cause muscle loss that combines
has a dedicated research publication (namely, the with TGFβ-induced ­muscle weakness to accelerate
Journal of Cachexia, Sarcopenia and Muscle157) as well ­cachexia-associated ­muscle dysfunction.
as a series of international conferences. Cancer research Despite these advances, none of the suggested
has for many decades been advanced through a system ­targets has yet led to approval of a drug for the indica-
of cooper­ative research groups that conduct national tion of cancer-associated cachexia. A translational gap
or international multidisciplinary research for cancer between human and animal studies of cancer cachexia
­control; a timely addition to cancer cachexia research remains. A large proportion of prior research in rodent
would be a cooperative group of researchers, cancer ­models has been conducted on a rather limited reper-
­centres and community-­based physicians to mount toire, including models harbouring colon 26 adeno-
studies that test new ways to screen, prevent, diagnose carcinoma, Lewis lung carcinoma, Yoshida hepatoma
and treat cancer cachexia. and Walker 256 carcino­sarcoma167–170. The original cell
lines have become unavailable; cells have been passed
Mechanisms/pathophysiology between different laboratories and currently available
Mechanistic insights continue to emerge from studies subclones might have deviated such that they no longer
in animal models. For example, new data have surfaced provide consistent results88. The generation of geneti-
that place attention on skeletal muscle stem cells158. cally engineered models of cancer, aligned with the
Specifically, a resident stem cell population called satel- clinical cachexia subsets, should offer new avenues for
lite cells159, which is marked by the transcription factor preclinical investigations. For example, a murine model
paired box protein Pax7 (REF. 160), provides regener­ of pancreatic cancer, which uses Kras G12D/+Trp53 –/–
ative capacity to this tissue. However, in experimental Pdx–Cre (KPC) congenic allografts in C57Bl/6 mice171,
cancer-­associated cachexia, this programme lacks the as well as patient-derived orthotopic pancreatic cancer
key myogenic factors myoblast determination pro- xenografts172, come closer to recapitulating the cachexia
tein 1 (Myod1) and myogenin, which are needed to features specifically associated with pancreatic ductal
repair damaged myo­fibres158. Instead, the Pax7‑positive adenocarcinoma. Finally, cachexia models do not reflect
­muscle progenitor cells demonstrate activation of the clinical complexity of oncology patients, who are
NF‑κB, which serves to retain a stem cell fate rather generally older and often present with substantial con-
than proceed through a differentiating regeneration comitant comorbidities and prior, as well as concurrent,
programme. This block in regeneration that contributed treatment with systemic therapies. Solutions to these
to myofibre atrophy occurred independent from the issues are urgently needed.

14 | ARTICLE NUMBER 17105 | VOLUME 4 www.nature.com/nrdp


©
2
0
1
8
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

Diagnosis through Toll-like receptor 7 (REF. 183). These effectors


Aetiology-based diagnostic criteria for cancer-associated may prove to be actionable targets of cancer cachexia
cachexia would represent a major advance. Weight loss in patients whose tumours overexpress them. However,
‘not otherwise specified’ is traditionally used as the where complex inflammatory cascades are activated,
basis for a diagnosis of cachexia and the indication for redundancy between individual mediators is common.
treatment, regardless of aetiology. In past clinical t­ rials, Accordingly, the targeting of a single mediator is unlikely
unspecified weight loss was treated with a ­number of to cure the majority of patients with cancer-­associated
highly specific agents (including the monoclonal anti- cachexia. Points of targeted intervention must be chosen
body infliximab, which targets TNF173) without testing to ­maximize the impact on the overall syndrome.
whether the patients overexpressed the target. As the field Intriguing relationships between cachexia therapy
continues to advance and patients can be better classi­ and cancer therapy are also emerging. For example, stud-
fied and subclassified using biomarkers, genomics or ies suggest that tumour cell proliferation and the excess
metabo­lomics, we will be able to offer a more personal­ muscle-protein catabolism characteristic of cachexia
ized, mechanistically derived treatment approach to each have a common mechanism. Specifically, dual specific-
patient. Some emerging areas of the human biology of ity mitogen-activated protein kinase kinase 1 (MAP2K1)
cancer-associated cachexia include genetic risk vari­ and MAP2K2 act downstream of RAS GTPases and
ants10,174,175, transcriptional vari­ants71,176 and biological RAF proto-oncogene serine/threonine-protein kinase
profiling using ­high-dimensional omics approaches177–179. (RAF) to induce the phosphorylation of MAPK1 and
MAPK3, thereby communicating input from growth
Management factors to promote proliferation of tumour cells 184.
Proof in principle exists that cachexia can be divorced This signalling pathway also seems to be involved in
mechanistically from the underlying disease. Experi­ the activation of excess muscle-­protein catabolism
mental evidence demonstrates that targeted ablation in tumour-bearing organisms7,185. MAP2K inhibitors
of cachexia signalling permits extended survival even might, therefore, simultaneously have anti-cachexic and
though the tumour continues to grow 11,12. In patients, anti-tumour activity. Further ­exploration of these types
direct evidence supports that anabolic processes can be of interactions is warranted.
activated under appropriate conditions. For example, Finally, clinical trials of cancer-associated cachexia
patients with locally advanced or metastatic disease continue to evolve in their design and end points. Until
demonstrate activation of muscle-protein synthesis recently, the landscape of clinical trials for cachexia was
after intake of high-quality proteins9,27,28, and this target­ somewhat limited and the overall quality of the evidence
ing of anabolic processes could be further optimized by was low 13,119. However, trials are no longer conducted in
the addition of anabolic agents. Furthermore, drugs the refractory stage, avoiding issues of confounding by
that inhibit catabolic processes have been shown to death, dropout and missing data. Additionally, patient
increase both lean and fat mass in kilogram quanti- populations and concurrent anticancer treatments in
ties in patients with some of the most catabolically current cachexia clinical trials are more homogeneous
active diseases, including advanced-stage lung cancer than in the past. However, regulatory authorities are
and cholangio­carcinoma7,8,112. Optimal conditions for being challenged with issues in study design and the
exploiting this anabolic potential are currently under specific nature of approvable end points186. For example,
study, with the overall aim of net improvements in although the consensus is that cachexia therapies should
muscle mass, f­ unctionality, performance status and be expected to produce stabilization or gain of radio­
­treatment tolerance. logically defined lean mass, skeletal muscle and fat mass,
Another area of growing interest is the use of mech­ an independent form of clinical benefit associated with
anistic insights to develop biomarkers and targeted these tissue gains is mandated by regulatory authori­
thera­pies for cachexia. To achieve this, more information ties. For example, the US FDA and European Medicines
on the origins of heterogeneity in the patient-specific Agency required clinical benefit as a co‑primary end
aetiology of cachexia is needed. For example, tumour point in trials NCT01395914 and NCT01355484 for the
overproduction of specific individual mediators that approval of anti-cachexia therapeutics. The definition
are potently catabolic towards muscle or adipose tissue, of this clinical benefit is contentious, and several recent
such as PTHLH32,33,88 and adrenomedullin180 (both of randomized phase III studies failed to meet criteria for
which elicit lipolysis in white adipose tissue), is a poten- approval, which were based on the clinical benefit end
tial target. PTHLH is normally absent in the peripheral points of handgrip strength112 and the stair climb test 113.
blood of healthy individuals, but mutations that amplify Alternatives to these end points are under discussion
its expression in tumours181 can promote high systemic and might include patient-reported outcomes, health-
concentrations in patients with cancer; this expression care utilization, costs and/or survival. Although these
is associated with poor prognosis182. Adrenomedullin developments are ongoing, creating opportunities for
has been shown to be encapsulated in tumour-derived patient participation in clinical trials of emerging drug
exosomes in pancreatic cancer 180. It was also demon- therapies and nutritional interventions for the indica-
strated that microvesicles (which include exosomes) tion of cancer-associated cachexia provides a way to
derived from pancreatic cancer cell lines can induce access front-line treatments. These investigations are
myoblast apoptosis by activ­ating mitogen-activated pro- essential to advance the testing and approval of new
tein kinases (MAPKs) and that this effect is mediated cachexia therapies.

NATURE REVIEWS | DISEASE PRIMERS VOLUME 4 | ARTICLE NUMBER 17105 | 15


©
2
0
1
8
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

1. DeWys, W. D. Pathophysiology of cancer cachexia: and weight loss, providing a prognostic grading visceral obesity, insulin resistance, and metabolic
current understanding and areas for future research. scheme for cancer-associated weight loss. dysfunction. FASEB J. 29, 988–1002 (2015).
Cancer Res. 42 (Suppl.), 721s–726s (1982). 23. Martin, L. et al. Cancer cachexia in the age of obesity: 46. Bodine, S. C. et al. Identification of ubiquitin ligases
2. Fearon, K. et al. Definition and classification of cancer skeletal muscle depletion is a powerful prognostic required for skeletal muscle atrophy. Science 294,
cachexia: an international consensus. Lancet Oncol. factor, independent of body mass index. J. Clin. Oncol. 1704–1708 (2001).
12, 489–495 (2011). 31, 1539–1547 (2013). 47. Gomes, M. D., Lecker, S. H., Jagoe, R. T., Navon, A.
This Delphi consensus process provides a This paper describes the risk of mortality & Goldberg, A. L. Atrogin‑1, a muscle-specific F‑box
definition, provisional diagnostic criteria in patients with solid tumours, which is associated protein highly expressed during muscle atrophy.
and a roadmap for future development of clinical with reduced skeletal muscle mass. Proc. Natl Acad. Sci. USA 98, 14440–14445 (2001).
cachexia research. 24. Demark-Wahnefried, W., Campbell, K. L. 48. Stitt, T. N. et al. The IGF‑1/PI3K/Akt pathway prevents
3. Kazemi-Bajestani, S. M. R., Mazurak, V. C. & Hayes, S. C. Weight management and its role expression of muscle atrophy-induced ubiquitin ligases
& Baracos, V. Computed tomography-defined muscle in breast cancer rehabilitation. Cancer 118, by inhibiting FOXO transcription factors. Mol. Cell 14,
and fat wasting are associated with cancer clinical 2277–2287 (2012). 395–403 (2004).
outcomes. Semin. Cell Dev. Biol. 54, 2–10 (2016). 25. Kubrak, C. et al. Clinical determinants of weight loss in 49. Sandri, M. et al. Foxo transcription factors induce the
4. World Health Organization. Cancer fact sheet. WHO patients receiving radiation and chemoirradiation for atrophy-related ubiquitin ligase atrogin‑1 and cause
http://www.who.int/mediacentre/factsheets/fs297/en/ head and neck cancer: a prospective longitudinal view. skeletal muscle atrophy. Cell 117, 399–412 (2004).
(2017). Head Neck 35, 695–703 (2012). Together with reference 47, this study identifies
5. Amano, K. et al. C‑Reactive protein, symptoms and 26. Silver, H. J., Dietrich, M. S. & Murphy, B. A. Changes FOXO transcription factors as regulators of the
activity of daily living in patients with advanced cancer in body mass, energy balance, physical function, and ubiquitin–proteasome system.
receiving palliative care. J. Cachexia Sarcopenia inflammatory state in patients with locally advanced 50. Acharyya, S. et al. Cancer cachexia is regulated by
Muscle 8, 457–465 (2017). head and neck cancer treated with concurrent selective targeting of skeletal muscle gene products.
6. Prado, C. M. et al. Central tenet of cancer cachexia chemoradiation after low-dose induction J. Clin. Invest. 114, 370–378 (2004).
therapy: do patients with advanced cancer have chemotherapy. Head Neck 29, 893–900 (2007). 51. Clarke, B. A. et al. The E3 Ligase MuRF1 degrades
exploitable anabolic potential? Am. J. Clin. Nutr. 98, 27. Engelen, M. P. K. J., Klimberg, V. S., Allasia, A. myosin heavy chain protein in dexamethasone-treated
1012–1019 (2013). & Deutz, N. E. P. Presence of early stage cancer skeletal muscle. Cell Metab. 6, 376–385 (2007).
7. Prado, C. M. M. et al. Skeletal muscle anabolism does not impair the early protein metabolic response 52. Sandri, M. Protein breakdown in muscle wasting:
is a side effect of therapy with the MEK inhibitor: to major surgery. J. Cachexia Sarcopenia Muscle 8, role of autophagy-lysosome and ubiquitin-proteasome.
selumetinib in patients with cholangiocarcinoma. 447–456 (2017). Int. J. Biochem. Cell Biol. 45, 2121–2129 (2013).
Br. J. Cancer 106, 1583–1586 (2012). 28. Engelen, M. P. K. J., Safar, A. M., Bartter, T., 53. Sandri, M. Protein breakdown in cancer cachexia.
8. Baracos, V. E. Skeletal muscle anabolism in patients Koeman, F. & Deutz, N. E. P. High anabolic potential Semin. Cell Dev. Biol. 54, 11–19 (2016).
with advanced cancer. Lancet Oncol. 16, 13–14 of essential amino acid mixtures in advanced nonsmall 54. Milan, G. et al. Regulation of autophagy and the
(2015). cell lung cancer. Ann. Oncol. 26, 1960–1966 (2015). ubiquitin-proteasome system by the FoxO
9. van Dijk, D. P. et al. Effects of oral meal feeding on Together with reference 9, this study demonstrates transcriptional network during muscle atrophy.
whole body protein breakdown and protein synthesis robust protein synthetic responses to amino acid Nat. Commun. 6, 6670 (2015).
in cachectic pancreatic cancer patients. J. Cachexia. feeding in patients with advanced-stage cancer. 55. Song, Y.‑H. et al. Muscle-specific expression of IGF‑1
Sarcopenia Muscle 6, 212–221 (2015). 29. Hall, K. D. & Baracos, V. E. Computational modeling blocks angiotensin II‑induced skeletal muscle wasting.
10. Johns, N. et al. New genetic signatures associated of cancer cachexia. Curr. Opin. Clin. Nutr. Metab. Care J. Clin. Invest. 115, 451–458 (2005).
with cancer cachexia as defined by low skeletal muscle 11, 214–221 (2008). 56. Sartori, R. et al. BMP signaling controls muscle mass.
index and weight loss. J. Cachexia Sarcopenia Muscle 30. Friesen, D. E., Baracos, V. E. & Tuszynski, J. A. Nat. Genet. 45, 1309–1318 (2013).
8, 122–130 (2017). Modeling the energetic cost of cancer as a result of 57. Johns, N. et al. Clinical classification of cancer
11. Zhou, X. et al. Reversal of cancer cachexia and muscle altered energy metabolism: implications for cachexia. cachexia: phenotypic correlates in human skeletal
wasting by actriib antagonism leads to prolonged Theor. Biol. Med. Model. 12, 17 (2015). muscle. PLoS ONE 9, e83618 (2014).
survival. Cell 142, 531–543 (2010). 31. Beck, S. A. & Tisdale, M. J. Effect of cancer cachexia 58. Guttridge, D. C., Mayo, M. W., Madrid, L. V.,
This experimental study shows a survival benefit on triacylglycerol/fatty acid substrate cycling in white Wang, C. Y. & Baldwin, A. S. NF‑kappaB-induced loss
in tumour-bearing mice from an agent for which adipose tissue. Lipids 39, 1187–1189 (2004). of MyoD messenger RNA: possible role in muscle
essentially the sole action is to prevent loss of 32. Kir, S. & Spiegelman, B. M. Cachexia and brown fat: decay and cachexia. Science 289, 2363–2366
skeletal muscle mass. a burning issue in cancer. Trends Cancer 2, 461–463 (2000).
12. Tseng, Y.‑C. et al. Preclinical Investigation of the novel (2016). 59. Cai, D. et al. IKKbeta/NF‑kappaB activation causes
histone deacetylase inhibitor AR‑42 in the treatment 33. Petruzzelli, M. et al. A switch from white to brown fat severe muscle wasting in mice. Cell 119, 285–298
of cancer-induced cachexia. J. Natl. Cancer Inst. 107, increases energy expenditure in cancer-associated (2004).
djv274 (2015). cachexia. Cell Metab. 20, 433–447 (2014). This study provides in vivo proof of concept that
13. Arends, J. et al. ESPEN guidelines on nutrition in 34. VanderVeen, B. N., Fix, D. K. & Carson, J. A. Disrupted NF‑κB signalling regulates skeletal muscle atrophy.
cancer patients. Clin. Nutr. 36, 11–48 (2017). skeletal muscle mitochondrial dynamics, mitophagy, 60. Mourkioti, F. et al. Targeted ablation of IKK2
This paper describes evidence-based guidelines for and biogenesis during cancer cachexia: a role for improves skeletal muscle strength, maintains mass,
nutritional screening and intervention in patients inflammation. Oxid. Med. Cell. Longev. 2017, and promotes regeneration. J. Clin. Invest. 116,
with cancer. 3292087 (2017). 2945–2954 (2006).
14. Pressoir, M. et al. Prevalence, risk factors and clinical 35. Zhang, G. et al. Tumor induces muscle wasting in mice 61. Silva, K. A. S. et al. Inhibition of Stat3 activation
implications of malnutrition in French Comprehensive through releasing extracellular Hsp70 and Hsp90. suppresses caspase‑3 and the ubiquitin-proteasome
Cancer Centres. Br. J. Cancer 102, 966–971 (2010). Nat. Commun. 8, 589 (2017). system, leading to preservation of muscle mass in
15. Bozzetti, F. & SCRINIO Working Group. Screening 36. Murphy, K. T. The pathogenesis and treatment of cancer cachexia. J. Biol. Chem. 290, 11177–11187
the nutritional status in oncology: a preliminary cardiac atrophy in cancer cachexia. Am. J. Physiol. (2015).
report on 1,000 outpatients. Support. Care Cancer Heart Circ. Physiol. 310, H466–H477 (2016). 62. Zimmers, T. A., Fishel, M. L. & Bonetto, A. STAT3 in the
17, 279–284 (2009). 37. Tashjian, A. H. Role of prostaglandins in the systemic inflammation of cancer cachexia. Semin. Cell
16. Segura, A. et al. An epidemiological evaluation of the production of hypercalcemia by tumors. Cancer Res. Dev. Biol. 54, 28–41 (2016).
prevalence of malnutrition in Spanish patients with 38, 4138–4141 (1978). 63. Ma, J. F. et al. STAT3 promotes IFNγ/TNFα-induced
locally advanced or metastatic cancer. Clin. Nutr. 24, 38. Fearon, K. C. H., Glass, D. J. & Guttridge, D. C. muscle wasting in an NF‑κB‑dependent and
801–814 (2005). Cancer cachexia: mediators, signaling, and metabolic IL‑6‑independent manner. EMBO Mol. Med. 9,
17. Hébuterne, X. et al. Prevalence of malnutrition and pathways. Cell Metab. 16, 153–166 (2012). 622–637 (2017).
current use of nutrition support in patients with cancer. 39. Loomans, H. A. & Andl, C. D. Intertwining of activin A 64. Guo, D., Wang, C., Wang, Q., Qiao, Z. & Tang, H.
JPEN. J. Parenter. Enteral Nutr. 38, 196–204 (2014). and TGFβ signaling: dual roles in cancer progression Pantoprazole blocks the JAK2/STAT3 pathway to
18. Kazemi-Bajestani, S. M. R., Becher, H., Fassbender, K., and cancer cell invasion. Cancers 7, 70–91 (2014). alleviate skeletal muscle wasting in cancer cachexia
Chu, Q. & Baracos, V. E. Concurrent evolution of 40. Loumaye, A. et al. Role of activin A and myostatin by inhibiting inflammatory response. Oncotarget 8,
cancer cachexia and heart failure: bilateral effects in human cancer cachexia. J. Clin. Endocrinol. Metab. 39640–39648 (2017).
exist. J. Cachexia Sarcopenia Muscle 5, 95–104 100, 2030–2038 (2015). 65. Zhang, G., Lin, R.‑K., Kwon, Y. T. & Li, Y.‑P. Signaling
(2014). 41. Togashi, Y. et al. Activin signal promotes cancer mechanism of tumor cell-induced up‑regulation of E3
19. Antoun, S. et al. Association of skeletal muscle wasting progression and is involved in cachexia in a subset of ubiquitin ligase UBR2. FASEB J. 27, 2893–2901
with treatment with sorafenib in patients with pancreatic cancer. Cancer Lett. 356, 819–827 (2015). (2013).
advanced renal cell carcinoma: results from a placebo- 42. Chen, J. L. et al. Differential effects of IL6 and activin A 66. Marchildon, F., Lamarche, É., Lala-Tabbert, N.,
controlled study. J. Clin. Oncol. 28, 1054–1060 in the development of cancer-associated cachexia. St‑Louis, C. & Wiper-Bergeron, N. Expression of
(2010). Cancer Res. 76, 5372–5382 (2016). CCAAT/enhancer binding protein beta in muscle
20. World Health Organization. BMI classification. WHO 43. Johnston, A. J. et al. Targeting of Fn14 prevents satellite cells inhibits myogenesis in cancer cachexia.
http://apps.who.int/bmi/index.jsp?introPage= cancer-induced cachexia and prolongs survival. Cell PLoS ONE 10, e0145583 (2015).
intro_3.html (2017). 162, 1365–1378 (2015). 67. Sun, R. et al. Valproic acid attenuates skeletal muscle
21. World Health Organization. Obesity and overweight 44. Mittal, A. et al. The TWEAK‑Fn14 system is a critical wasting by inhibiting C/EBPβ-regulated atrogin1
fact sheet. WHO http://www.who.int/mediacentre/ regulator of denervation-induced skeletal muscle expression in cancer cachexia. Am. J. Physiol. Cell
factsheets/fs311/en/ (2016). atrophy in mice. J. Cell Biol. 188, 833–849 (2010). Physiol. 311, C101–C115 (2016).
22. Martin, L. et al. Diagnostic criteria for the classification This is an early study demonstrating the 68. Mueller, T. C., Bachmann, J., Prokopchuk, O., Friess, H.
of cancer-associated weight loss. J. Clin. Oncol. 33, requirement of the TWEAK ligand and TNFRSF12A & Martignoni, M. E. Molecular pathways leading to
90–99 (2015). in regulating muscle wasting due to disuse atrophy. loss of skeletal muscle mass in cancer cachexia — can
This paper describes the risk of mortality in an 45. Sato, S., Ogura, Y., Tajrishi, M. M. & Kumar, A. findings from animal models be translated to humans?
international cachexia data set stratified by BMI Elevated levels of TWEAK in skeletal muscle promote BMC Cancer 16, 75 (2016).

16 | ARTICLE NUMBER 17105 | VOLUME 4 www.nature.com/nrdp


©
2
0
1
8
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

69. Bennegård, K., Lindmark, L., Edén, E., Svaninger, G. analysis. J. Cachexia Sarcopenia Muscle 8, 789–797 114. Fearon, K. C. H. et al. Double-blind, placebo-controlled,
& Lundholm, K. Flux of amino acids across the leg (2017). randomized study of eicosapentaenoic acid diester
in weight-losing cancer patients. Cancer Res. 44, 95. Vazeille, C. et al. Relation between hypermetabolism, in patients with cancer cachexia. J. Clin. Oncol. 24,
386–393 (1984). cachexia, and survival in cancer patients: a prospective 3401–3407 (2006).
70. DeJong, C. H. C. et al. Systemic inflammation study in 390 cancer patients before initiation of 115. Bruera, E. et al. Effect of fish oil on appetite and other
correlates with increased expression of skeletal muscle anticancer therapy. Am. J. Clin. Nutr. 105, 1139–1147 symptoms in patients with advanced cancer and
ubiquitin but not uncoupling proteins in cancer (2017). anorexia/cachexia: a double-blind, placebo-controlled
cachexia. Oncol. Rep. 14, 257–263 (2005). 96. Read, J. A. et al. Nutritional assessment in cancer: study. J. Clin. Oncol. 21, 129–134 (2003).
71. Gallagher, I. J. et al. Suppression of skeletal muscle comparing the Mini-Nutritional Assessment (MNA) 116. Strasser, F. et al. Comparison of orally administered
turnover in cancer cachexia: evidence from the with the Scored Patient-Generated Subjective Global cannabis extract and delta‑9‑tetrahydrocannabinol in
transcriptome in sequential human muscle biopsies. Assessment (PGSGA). Nutr. Cancer 53, 51–56 (2005). treating patients with cancer-related anorexia-cachexia
Clin. Cancer Res. 18, 2817–2827 (2012). 97. Hubbard, J. M., Cohen, H. J. & Muss, H. B. syndrome: a multicenter, phase III, randomized,
72. Tisdale, M. J. Cachexia in cancer patients. Nat. Rev. Incorporating biomarkers into cancer and aging double-blind, placebo-controlled clinical trial from the
Cancer 2, 862–871 (2002). research. J. Clin. Oncol. 32, 2611–2616 (2014). Cannabis-In‑Cachexia-Study-Group. J. Clin. Oncol. 24,
73. MacDonald, A. J. et al. Habitual myofibrillar protein 98. Fujiwara, N. et al. Sarcopenia, intramuscular fat 3394–3400 (2006).
synthesis is normal in patients with upper GI cancer deposition, and visceral adiposity independently 117. Awad, S. et al. Marked changes in body
cachexia. Clin. Cancer Res. 21, 1734–1740 (2015). predict the outcomes of hepatocellular carcinoma. composition following neoadjuvant chemotherapy for
74. Schmitt, T. L. et al. Activity of the Akt-dependent J. Hepatol. 63, 131–140 (2015). oesophagogastric cancer. Clin. Nutr. 31, 74–77 (2012).
anabolic and catabolic pathways in muscle and liver This paper shows an association of muscle 118. Fearon, K. C. H. Effect of a protein and energy
samples in cancer-related cachexia. J. Mol. Med. 85, depletion with mortality in an Asian population. dense n-3 fatty acid enriched oral supplement
647–654 (2007). 99. Epstein, F. H., Gabay, C. & Kushner, I. Acute-phase on loss of weight and lean tissue in cancer cachexia:
75. Acharyya, S. et al. Dystrophin glycoprotein complex proteins and other systemic responses to a randomised double blind trial. Gut 52, 1479–1486
dysfunction: a regulatory link between muscular inflammation. N. Engl. J. Med. 340, 448–454 (1999). (2003).
dystrophy and cancer cachexia. Cancer Cell 8, 100. Douglas, E. & McMillan, D. C. Towards a simple 119. Baracos, V. E., Watanbe, S. & Fearon, K.
421–432 (2005). objective framework for the investigation and in The Oxford Textbook of Palliative Medicine 5th edn
76. Penna, F. et al. Muscle atrophy in experimental cancer treatment of cancer cachexia: The Glasgow Prognostic (eds Cherny, N., Fallon, M., Kaasa, S., Portenoy, R. K.
cachexia: is the IGF‑1 signaling pathway involved? Score. Cancer Treat. Rev. 40, 685–691 (2014). & Currow, D. C.) 702–712 (2015).
Int. J. Cancer 127, 1706–1717 (2010). 101. Laird, B. J. A. et al. Quality of life in patients with 120. Dallmann, R. et al. The orally active melanocortin‑4
77. Stephens, N. A. et al. Evaluating potential biomarkers advanced cancer: differential association with receptor antagonist BL‑6020/979: a promising
of cachexia and survival in skeletal muscle of upper performance status and systemic inflammatory candidate for the treatment of cancer cachexia.
gastrointestinal cancer patients. J. Cachexia response. J. Clin. Oncol. 34, 2769–2775 (2016). J. Cachexia Sarcopenia Muscle 2, 163–174 (2011).
Sarcopenia Muscle 6, 53–61 (2015). 102. Wei, B. et al. The neutrophil lymphocyte ratio is 121. MacDonald, N., Easson, A. M., Mazurak, V. C.,
78. Burfeind, K. G., Michaelis, K. A. & Marks, D. L. The associated with breast cancer prognosis: an updated Dunn, G. P. & Baracos, V. E. Understanding and
central role of hypothalamic inflammation in the acute systematic review and meta-analysis. Onco. Targets. managing cancer cachexia. J. Am. Coll. Surg. 197,
illness response and cachexia. Semin. Cell Dev. Biol. Ther. 9, 5567–5575 (2016). 143–161 (2003).
54, 42–52 (2016). 103. Dolan, R. D., McSorley, S. T., Horgan, P. G., Laird, B. 122. Garvey, C. et al. Pulmonary rehabilitation exercise
79. Grossberg, A. J. et al. Arcuate nucleus & McMillan, D. C. The role of the systemic prescription in chronic obstructive pulmonary disease.
proopiomelanocortin neurons mediate the acute inflammatory response in predicting outcomes in J. Cardiopulm. Rehabil. Prev. 36, 75–83 (2016).
anorectic actions of leukemia inhibitory factor via patients with advanced inoperable cancer: Systematic 123. Stene, G. B. et al. Effect of physical exercise on muscle
gp130. Endocrinology 151, 606–616 (2010). review and meta-analysis. Crit. Rev. Oncol. Hematol. mass and strength in cancer patients during treatment
80. Bodnar, R. J. et al. Mediation of anorexia by human 116, 134–146 (2017). — a systematic review. Crit. Rev. Oncol. Hematol. 88,
recombinant tumor necrosis factor through a peripheral 104. Hong, N. et al. Serum PTHrP predicts weight loss 573–593 (2013).
action in the rat. Cancer Res. 49, 6280–6284 (1989). in cancer patients independent of hypercalcemia, 124. Wasley, D. et al. Patients with established cancer
81. Braun, T. P. et al. Central nervous system inflammation, and tumor burden. J. Clin. Endocrinol. cachexia lack the motivation and self-efficacy to
inflammation induces muscle atrophy via activation Metab. 101, 1207–1214 (2016). undertake regular structured exercise. Psychooncology
of the hypothalamic-pituitary-adrenal axis. J. Exp. 105. Lerner, L. et al. Plasma growth differentiation factor 15 http://dx.doi.org/10.1002/pon.4512 (2017).
Med. 208, 2449–2463 (2011). is associated with weight loss and mortality in cancer 125. Mayo, N. E. et al. Pedometer-facilitated walking
82. DeBoer, M. D. Ghrelin and cachexia: will treatment patients. J. Cachexia Sarcopenia Muscle 6, 317–324 intervention shows promising effectiveness for reducing
with GHSR‑1a agonists make a difference for patients (2015). cancer fatigue: a pilot randomized trial. Clin. Rehabil.
suffering from chronic wasting syndromes? Mol. Cell. 106. Macciò, A. et al. A randomized phase III clinical trial 28, 1198–1209 (2014).
Endocrinol. 340, 97–105 (2011). of a combined treatment for cachexia in patients with 126. Shragge, J. E., Wismer, W. V., Olson, K. L.
83. Fouladiun, M. et al. Body composition and time course gynecological cancers: evaluating the impact on & Baracos, V. E. Shifting to conscious control:
changes in regional distribution of fat and lean tissue metabolic and inflammatory profiles and quality of life. psychosocial and dietary management of anorexia
in unselected cancer patients on palliative care — Gynecol. Oncol. 124, 417–425 (2012). by patients with advanced cancer. Palliat. Med. 21,
correlations with food intake, metabolism, exercise 107. Isenring, E. & Elia, M. Which screening method is 227–233 (2007).
capacity, and hormones. Cancer 103, 2189–2198 appropriate for older cancer patients at risk for 127. Reilly, C. M. et al. A literature synthesis of symptom
(2005). malnutrition? Nutrition 31, 594–597 (2015). prevalence and severity in persons receiving
84. Zuijdgeest-van Leeuwen, S. D. et al. Lipolysis and lipid 108. Scott, D., Reid, J., Hudson, P., Martin, P. & Porter, S. active cancer treatment. Support. Care Cancer 21,
oxidation in weight-losing cancer patients and healthy Health care professionals’ experience, understanding 1525–1550 (2013).
subjects. Metabolism 49, 931–936 (2000). and perception of need of advanced cancer patients 128. Pakhomov, S. V., Jacobsen, S. J., Chute, C. G.
85. Rydén, M. & Arner, P. Fat loss in cachexia — is there a with cachexia and their families: The benefits of a & Roger, V. L. Agreement between patient-reported
role for adipocyte lipolysis? Clin. Nutr. 26, 1–6 (2007). dedicated clinic. BMC Palliat. Care 15, 100 (2016). symptoms and their documentation in the medical
86. Agustsson, T. et al. Mechanism of increased lipolysis in 109. Dev, R. et al. Hypermetabolism and symptom burden record. Am. J. Manag. Care 14, 530–539 (2008).
cancer cachexia. Cancer Res. 67, 5531–5537 (2007). in advanced cancer patients evaluated in a cachexia 129. Gagnon, B. et al. A prospective evaluation of an
87. Das, S. K. et al. Adipose triglyceride lipase contributes clinic. J. Cachexia Sarcopenia Muscle 6, 95–98 interdisciplinary nutrition–rehabilitation program for
to cancer-associated cachexia. Science 333, 233–238 (2015). patients with advanced cancer. Curr. Oncol. 20, 310
(2011). 110. Vigano, A., Del Fabbro, E., Bruera, E. & Borod, M. (2013).
This paper provides evidence that lipolysis in cancer The cachexia clinic: from staging to managing 130. Chasen, M. R., Feldstain, A., Gravelle, D.,
predisposes skeletal muscle to undergo wasting. nutritional and functional problems in advanced MacDonald, N. & Pereira, J. An interprofessional
88. Kir, S. et al. Tumour-derived PTH-related protein cancer patients. Crit. Rev. Oncog. 17, 293–304 palliative care oncology rehabilitation program: effects
triggers adipose tissue browning and cancer cachexia. (2012). on function and predictors of program completion.
Nature 513, 100–104 (2014). 111. Blum, D., Hess, J., Omlin, A., Jurt, G. Curr. Oncol. 20, 301 (2013).
89. Neves, R. X. et al. White adipose tissue cells and & Strasser, A. B. H. P. M. Comprehensive cancer 131. Smith, T. J. et al. American Society of Clinical Oncology
the progression of cachexia: inflammatory pathways. cachexia staging and its impact in the outpatient provisional clinical opinion: the integration of palliative
J. Cachexia Sarcopenia Muscle 7, 193–203 (2015). oncology setting: a phase II study. J. Clin. Oncol. 27, care into standard oncology care. J. Clin. Oncol. 30,
90. Camargo, R. et al. NF‑κBp65 and expression of its e20530–e20530 (2009). 880–887 (2012).
pro-inflammatory target genes are upregulated in 112. Temel, J. S. et al. Anamorelin in patients with 132. Hesketh, P. J. et al. Antiemetics: American Society of
the subcutaneous adipose tissue of cachectic cancer non‑small-cell lung cancer and cachexia (ROMANA 1 Clinical Oncology clinical practice guideline update.
patients. Nutrients 7, 4465–4479 (2015). and ROMANA 2): results from two randomised, J. Clin. Oncol. 35, 3240–3261 (2017).
91. Zimmermann, R. Fat mobilization in adipose tissue is double-blind, phase 3 trials. Lancet Oncol. 17, 133. Prado, C. M. M. et al. Prevalence and clinical
promoted by adipose triglyceride lipase. Science 306, 519–531 (2016). implications of sarcopenic obesity in patients with solid
1383–1386 (2004). This paper describes two phase III clinical studies of tumours of the respiratory and gastrointestinal tracts:
92. Rohm, M. et al. An AMP-activated protein kinase– a low-molecular-mass growth hormone secretagogue a population-based study. Lancet Oncol. 9, 629–635
stabilizing peptide ameliorates adipose tissue wasting receptor type 1 agonist, demonstrating lean tissue (2008).
in cancer cachexia in mice. Nat. Med. 22, 1120–1130 gain in patients with non-small-cell lung cancer. 134. Sjøblom, B. et al. Drug dose per kilogram lean body
(2016). 113. Crawford, J. et al. Study design and rationale for the mass predicts hematologic toxicity from carboplatin-
93. Martin, L. Diagnostic criteria for cancer cachexia: phase 3 clinical development program of enobosarm, doublet chemotherapy in advanced non-small-cell lung
data versus dogma. Curr. Opin. Clin. Nutr. Metab. Care a selective androgen receptor modulator, for the cancer. Clin. Lung Cancer 18, e129–e136 (2017).
19, 188–198 (2016). prevention and treatment of muscle wasting in cancer 135. Fearon, K. C. H. Cancer cachexia: developing
94. Vagnildhaug, O. M. et al. The applicability of a weight patients (POWER Trials). Curr. Oncol. Rep. 18, 37 multimodal therapy for a multidimensional problem.
loss grading system in cancer cachexia: a longitudinal (2016). Eur. J. Cancer 44, 1124–1132 (2008).

NATURE REVIEWS | DISEASE PRIMERS VOLUME 4 | ARTICLE NUMBER 17105 | 17


©
2
0
1
8
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

136. Maddocks, M. et al. Practical multimodal care for 160. Seale, P. et al. Pax7 Is Required for the Specification 184. Samatar, A. A. & Poulikakos, P. I. Targeting RAS-ERK
cancer cachexia. Curr. Opin. Support. Palliat. Care 10, of myogenic satellite cells. Cell 102, 777–786 (2000). signalling in cancer: promises and challenges. Nat. Rev.
298–305 (2016). 161. Waning, D. L. et al. Excess TGF‑β mediates muscle Drug Discov. 13, 928–942 (2014).
137. Fearon, K., Arends, J. & Baracos, V. Understanding the weakness associated with bone metastases in mice. 185. Baracos, V. E. Mitogen-activated protein kinases
mechanisms and treatment options in cancer cachexia. Nat. Med. 21, 1262–1271 (2015). inhibitors: potential therapeutic agents for cancer
Nat. Rev. Clin. Oncol. 10, 90–99 (2012). 162. Sheen-Chen, S.‑M. Serum levels of transforming growth cachexia. Mol. Cancer Ther. 16, 263–264 (2017).
138. Aapro, M. et al. Early recognition of malnutrition factor β1 in patients with breast cancer. Arch. Surg. 186. Fearon, K. C. H. et al. Request for regulatory guidance
and cachexia in the cancer patient: a position paper 136, 937 (2001). for cancer cachexia intervention trials. J. Cachexia
of a European School of Oncology Task Force. 163. Huang, A. et al. Increased serum transforming growth Sarcopenia Muscle 6, 272–274 (2015).
Ann. Oncol. 25, 1492–1499 (2014). factor‑beta1 in human colorectal cancer correlates 187. Hall, K. D. et al. Energy balance and its components:
139. Lassen, K. et al. Pancreaticoduodenectomy: ERAS with reduced circulating dendritic cells and increased implications for body weight regulation. Am. J. Clin.
recommendations. Clin. Nutr. 32, 870–871 (2013). colonic Langerhans cell infiltration. Clin. Exp. Immunol. Nutr. 95, 989–994 (2012).
140. Nelson, G. et al. Implementation of enhanced recovery 134, 270–278 (2003). 188. Purcell, S. A., Elliott, S. A., Baracos, V. E., Chu, Q. S. C.
after surgery (ERAS) across a provincial healthcare 164. Li, X. et al. Elevated serum level and gene & Prado, C. M. Key determinants of energy
system: the ERAS Alberta Colorectal Surgery polymorphisms of TGF‑β1 in gastric cancer. J. Clin. expenditure in cancer and implications for clinical
Experience. World J. Surg. 40, 1092–1103 (2016). Lab. Anal. 22, 164–171 (2008). practice. Eur. J. Clin. Nutr. 70, 1230–1238 (2016).
141. Palmela, C. et al. Body composition as a prognostic 165. Poch, B. et al. Systemic immune dysfunction in 189. Jager-Wittenaar, H. & Ottery, F. D. Assessing
factor of neoadjuvant chemotherapy toxicity and pancreatic cancer patients. Langenbecks Arch. Surg. nutritional status in cancer: role of the Patient-
outcome in patients with locally advanced gastric 392, 353–358 (2007). Generated Subjective Global Assessment. Curr. Opin.
cancer. J. Gastr. Cancer 17, 74 (2017). 166. Craven, K. E., Gore, J., Wilson, J. L. & Korc, M. Clin. Nutr. Metab. Care 20, 322–329 (2017).
142. Daly, L. E. et al. The impact of body composition Angiogenic gene signature in human pancreatic cancer 190. Vellas, B. et al. Overview of the MNA — its history
parameters on ipilimumab toxicity and survival in correlates with TGF-beta and inflammatory and challenges. J. Nutr. Health Aging 10, 456–463;
patients with metastatic melanoma. Br. J. Cancer 116, transcriptomes. Oncotarget 1, 323–341 (2010). discussion 463–465 (2006).
310–317 (2017). This study establishes an association between 191. Rubenstein, L. Z., Harker, J. O., Salvà, A., Guigoz, Y.
143. Lieffers, J. R. et al. A viscerally driven cachexia increased TGFβ signalling in pancreatic ductal & Vellas, B. Screening for undernutrition in geriatric
syndrome in patients with advanced colorectal cancer: adenocarcinoma, increased expression of practice: developing the short-form mini-nutritional
contributions of organ and tumor mass to whole-body pro-inflammatory genes and increased tyrosine assessment (MNA‑SF). J. Gerontol. A. Biol. Sci. Med.
energy demands. Am. J. Clin. Nutr. 89, 1173–1179 kinase JAK signalling, which may combine to Sci. 56, M366–372 (2001).
(2009). worsen muscle loss. 192. Ferguson, M., Capra, S., Bauer, J. & Banks, M.
144. Bozzetti, F. ESPEN guideline on ethical aspects of 167. Morrison, S. D. Feeding response to change in Development of a valid and reliable malnutrition
artificial nutrition and hydration. Clin. Nutr. 35, 1577 absorbable food fraction during growth of Walker 256 screening tool for adult acute hospital patients.
(2016). carcinosarcoma. Cancer Res. 32, 968–972 (1972). Nutrition 15, 458–464 (1999).
145. Bozzetti, F. et al. Development and validation of a 168. Baccino, F. M., Tessitore, L., Bonelli, G. & Isidoro, C. 193. Stratton, R. J. et al. Malnutrition in hospital outpatients
nomogram to predict survival in incurable cachectic Protein turnover states of tumour cells and host tissues and inpatients: prevalence, concurrent validity and ease
cancer patients on home parenteral nutrition. in an experimental model. Biomed. Biochim. Acta 45, of use of the ‘malnutrition universal screening tool’
Ann. Oncol. 26, 2335–2340 (2015). 1585–1590 (1986). (’MUST’) for adults. Br. J. Nutr. 92, 799–808 (2004).
146. Kalantar-Zadeh, K. et al. Why cachexia kills: examining 169. Ohe, Y. et al. Interleukin‑6 cDNA transfected Lewis lung 194. Kruizenga, H. M., Seidell, J. C., de Vet, H. C. W.,
the causality of poor outcomes in wasting conditions. carcinoma cells show unaltered net tumour growth rate Wierdsma, N. J. & van Bokhorst‑de van der
J. Cachexia Sarcopenia Muscle 4, 89–94 (2013). but cause weight loss and shortened survival in Schueren, M. A. E. Development and validation of
147. Tarricone, R., Ricca, G., Nyanzi-Wakholi, B. syngeneic mice. Br. J. Cancer 67, 939–944 (1993). a hospital screening tool for malnutrition: the short
& Medina‑Lara, A. Impact of cancer anorexia-cachexia 170. Tanaka, Y. et al. Experimental cancer cachexia induced nutritional assessment questionnaire (SNAQ).
syndrome on health-related quality of life and resource by transplantable colon 26 adenocarcinoma in mice. Clin. Nutr. 24, 75–82 (2005).
utilisation: a systematic review. Crit. Rev. Oncol. Cancer Res. 50, 2290–2295 (1990). 195. Kondrup, J. & Rasmussen, H. H., Hamberg, O.,
Hematol. 99, 49–62 (2016). 171. Michaelis, K. A. et al. Establishment and Stanga, Z. & Ad Hoc ESPEN Working Group.
148. Wheelwright, S. et al. A systematic review of characterization of a novel murine model of pancreatic Nutritional risk screening (NRS 2002): a new method
health‑related quality of life instruments in patients cancer cachexia. J. Cachexia Sarcopenia Muscle 8, based on an analysis of controlled clinical trials.
with cancer cachexia. Support. Care Cancer 21, 824–838 (2017). Clin. Nutr. 22, 321–336 (2003).
2625–2636 (2013). 172. Go, K. L. et al. Orthotopic patient-derived pancreatic 196. Bourdel-Marchasson, I. et al. Nutritional advice in
149. Wheelwright, S. J. et al. Development of the EORTC cancer xenografts engraft into the pancreatic older patients at risk of malnutrition during treatment
QLQ‑CAX24, a questionnaire for cancer patients with parenchyma, metastasize, and induce muscle wasting for chemotherapy: a two-year randomized controlled
cachexia. J. Pain Symptom Manage. 53, 232–242 to recapitulate the human disease. Pancreas 46, trial. PLoS ONE 9, e108687 (2014).
(2017). 813–819 (2017). 197. Sánchez-Lara, K. et al. Effects of an oral nutritional
150. Hopkinson, J. B. Psychosocial impact of cancer 173. Wiedenmann, B. et al. A multicenter, phase II study supplement containing eicosapentaenoic acid on
cachexia. J. Cachexia Sarcopenia Muscle 5, 89–94 of infliximab plus gemcitabine in pancreatic cancer nutritional and clinical outcomes in patients with
(2014). cachexia. J. Support. Oncol. 6, 18–25 (2008). advanced non-small cell lung cancer: randomised trial.
151. Oberholzer, R. et al. Psychosocial effects of cancer 174. Tan, B. H. L. et al. P‑selectin genotype is associated Clin. Nutr. 33, 1017–1023 (2014).
cachexia: a systematic literature search and qualitative with the development of cancer cachexia. EMBO Mol. 198. Madeddu, C. et al. Randomized phase III clinical trial
analysis. J. Pain Symptom Manage. 46, 77–95 (2013). Med. 4, 462–471 (2012). of a combined treatment with carnitine + celecoxib ±
152. Maschke, J. et al. Nutritional care of cancer patients: 175. Johns, N. et al. Genetic basis of interindividual megestrol acetate for patients with cancer-related
a survey on patients’ needs and medical care in reality. susceptibility to cancer cachexia: selection of potential anorexia/cachexia syndrome. Clin. Nutr. 31, 176–182
Int. J. Clin. Oncol. 22, 200–206 (2016). candidate gene polymorphisms for association studies. (2012).
153. Hopkinson, J. B. Food connections: A qualitative J. Genet. 93, 893–916 (2014).
exploratory study of weight- and eating-related distress 176. Narasimhan, A. et al. Small RNAome profiling from Acknowledgements
in families affected by advanced cancer. Eur. J. Oncol. human skeletal muscle: novel mi­­­RNAs and their targets The authors thank R.J.E. Skipworth (University of Edinburgh)
Nurs. 20, 87–96 (2016). associated with cancer cachexia. J. Cachexia for his valuable input.
154. Wheelwright, S., Darlington, A.‑S., Hopkinson, J. B., Sarcopenia Muscle 8, 405–416 (2017).
Fitzsimmons, D. & Johnson, C. A systematic review 177. Twelkmeyer, B., Tardif, N. & Rooyackers, O. Omics Author contributions
and thematic synthesis of quality of life in the informal and cachexia. Curr. Opin. Clin. Nutr. Metab. Care 20, Introduction (V.E.B.); Epidemiology (L.M.); Mechanisms/
carers of cancer patients with cachexia. Palliat. Med. 181–185 (2017). pathophysiology (M.K. and D.C.G.); Diagnosis, screening and
30, 149–160 (2016). 178. Gallagher, I. J., Jacobi, C., Tardif, N., Rooyackers, O. prevention (V.E.B.); Management (K.C.H.F.); Quality of life
155. Hopkinson, J. B. & Richardson, A. A mixed-methods & Fearon, K. Omics/systems biology and cancer (V.E.B.); Outlook (V.E.B.); and Overview of the Primer (V.E.B.).
qualitative research study to develop a complex cachexia. Semin. Cell Dev. Biol. 54, 92–103 (2016).
intervention for weight loss and anorexia in advanced 179. Ebhardt, H. A. et al. Comprehensive proteome analysis Competing interests statement
cancer: The Family Approach to Weight and Eating. of human skeletal muscle in cachexia and sarcopenia: V.E.B. receives financial support from the Canadian Institutes
Palliat. Med. 29, 164–176 (2014). a pilot study. J. Cachexia Sarcopenia Muscle 8, of Health Research and The Alberta Cancer Foundation. L.M.
156. von Haehling, S. & Anker, S. D. Cachexia as major 567–582 (2017). is funded by Alberta Innovates, the Izaak Walton Killam
underestimated unmet medical need: Facts and 180. Sagar, G. et al. Pathogenesis of pancreatic cancer Foundation and the American Society of Parenteral and
numbers. Int. J. Cardiol. 161, 121–123 (2012). exosome-induced lipolysis in adipose tissue. Gut 65, Enteral Nutrition. M.K. is partially funded by National Cancer
157. Journal of Cachexia, Sarcopenia and Muscle. 1165–1174 (2015). Institute grant CA‑075059 and by the consortium for the
http://onlinelibrary.wiley.com/journal/10.1007/ This is a clinical study of a novel mechanism of study of Chronic Pancreatitis, Diabetes and Pancreatic Cancer
13539.2190-6009 (2017). tumour-induced lipolysis. (U01 DK108323). D.C.G. receives funding support from the
158. He, W. A. et al. NF‑kB mediated Pax7 dysregulation 181. Sidler, B. et al. Amplification of the parathyroid US NIH and from the Ohio State University Comprehensive
in the muscle microenvironment promotes cancer hormone-related peptide gene in a colonic carcinoma. Cancer Center.
cachexia. J. Clin. Invest. 123, 4821–4835 (2013). J. Clin. Endocrinol. Metab. 81, 2841–2847 (1996).
This study provides evidence that dysfunctional 182. Washam, C. L. et al. Identification of PTHrP(12–48) Publisher’s note
regeneration is a causal event in cancer-induced as a plasma biomarker associated with breast cancer Springer Nature remains neutral with regard to jurisdictional
muscle wasting. bone metastasis. Cancer Epidemiol. Biomarkers Prev. claims in published maps and institutional affiliations.
159. Viguie, C. A., Lu, D.‑X., Huang, S.‑K., Rengen, H. 22, 972–983 (2013).
& Carlson, B. M. Quantitative study of the effects of 183. He, W. A. et al. Microvesicles containing mi­­­RNAs How to cite this Primer
long-term denervation on the extensor digitorum longus promote muscle cell death in cancer cachexia via TLR7. Baracos, V. E. et al. Cancer-associated cachexia. Nat. Rev.
muscle of the rat. Anat. Rec. 248, 346–354 (1997). Proc. Natl Acad. Sci. USA 111, 4525–4529 (2014). Dis. Primers 4, 17105 (2018).

18 | ARTICLE NUMBER 17105 | VOLUME 4 www.nature.com/nrdp


©
2
0
1
8
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.

You might also like