You are on page 1of 14

REVIEWS

M E TA B O L I C S I G N A L L I N G

Cytosolic lipolysis and lipophagy:


two sides of the same coin
Rudolf Zechner1,2,, Frank Madeo1,2 and Dagmar Kratky1,3
Abstract | Fatty acids are the most efficient substrates for energy production in vertebrates
and are essential components of the lipids that form biological membranes. Synthesis of
triacylglycerols from non-esterified free fatty acids (FFAs) combined with triacylglycerol storage
represents a highly efficient strategy to stockpile FFAs in cells and prevent FFA-induced
lipotoxicity. Although essentially all vertebrate cells have some capacity to store and utilize
triacylglycerols, white adipose tissue is by far the largest triacylglycerol depot and is uniquely
able to supply FFAs to other tissues. The release of FFAs from triacylglycerols requires their
enzymatic hydrolysis by a process called lipolysis. Recent discoveries thoroughly altered and
extended our understanding of lipolysis. This Review discusses how cytosolic ‘neutral’ lipolysis
and lipophagy, which utilizes ‘acid’ lipolysis in lysosomes, degrade cellular triacylglycerols as well
as how these pathways communicate, how they affect lipid metabolism and energy homeostasis
and how their dysfunction affects the pathogenesis of metabolic diseases. Answers to these
questions will likely uncover novel strategies for the treatment of prevalent metabolic diseases.

Fatty acids
Fatty acids are pivotal for life. They constitute integral During feeding, WAT stores excessive calories as fat.
Monocarboxylic acids with parts of biological membranes, harbour the highest Conversely, during fasting, these fat stores are degraded
long saturated or unsaturated energy density of all known energy substrates, are precur­ to supply the body with FFAs. The release of FFAs
aliphatic carbon chains. They sors for numerous lipid mediators and modulate protein from triacylglycerols requires ester hydrolysis by an
are either unesterified, in free
function through posttranslational acylation of target enzymatic process called lipolysis. Given that triacyl­
form (free fatty acids) or
esterified to various alcohols. proteins. Therefore, it is not surprising that fatty acids glycerol molecules are themselves unable to cross bio­
in dietary fats and oils are key nutritional components, logical membranes, their transport in and out of cells
Acid–base homeostasis accounting for 20–35% of the daily calorie intake in a also requires lipolysis and resynthesis. In physiological
Maintenance of a constant pH normal human diet. Additionally, all other major diet­ terms, lipolysis fulfils three major functions in verte­
in extracellular body fluids by
balanced concentrations of
ary components, such as monosaccharides (which are brates: gastro­intestinal lipolysis mediates the absorp­
acids and bases. derived from dietary carbohydrates) and amino acids tion of nutritional fats and oils in the stomach and gut 2,
(which are derived from diet­ary protein), can be con­ vascular lipolysis enables the cellular uptake of FFAs
verted into free fatty acids (FFAs). Despite their role in and mono­acylglycerols from circulating triacylglycerol-
essential cellular functions, high concentrations of FFAs rich plasma lipoproteins (very-low-density lipoproteins
are toxic because of their limited solubility and amphi­ (VLDLs) and chylomicrons)3, and intracellular l­ipolysis
1
BioTechMed-Graz, pathic nature; their adverse impact on cellular acid–base accounts for the release of FFAs from cytoplasmic lipid
Mozartgasse 12, 8010 Graz, homeostasis; and their ready transformation into highly droplets (cLDs) and from endocytosed lipoprotein-­
Austria. bioactive, cytotoxic lipid species. These destructive effects associated triacylglycerols3. Cytoplasmic lipases hydrolyse
2
Institute of Molecular
Biosciences, University of
of FFAs — collectively referred to as ‘lipotoxicity’ — can cLD-associated triacylglycerols at pH ~7 in a process
Graz, Heinrichstrasse 31, lead to cellular d­ ysfunction and cell death1. called ‘neutral’ lipolysis 4. By contrast, lipoprotein-­
8010 Graz, Austria. To avoid lipotoxicity, FFAs are detoxified by their three­ associated triacylglycerols are transported by early and
3
Institute of Molecular fold esterification to the trivalent alcohol glycerol to form late endosomes to the lysosome, where they undergo
Biology and Biochemistry,
triacylglycerols. Triacylglycerols (colloquially called ‘fat’ ‘acid’ lipolysis — hydrolysis by lysosomal acid lipase
Medical University of Graz,
Neue Stiftingtalstrasse 6/6, or ‘oil’, depending on their aggregation state) are inert and (LAL) at pH 4.5–5 (REF. 5). Historically, these distinct
8010 Graz, Austria. sufficiently hydrophobic to provide optimal packaging of pathways were considered separate and were thought to
Correspondence to R.Z. FFAs for efficient transport and storage. Essentially all cells have little functional overlap. However, the discovery of
rudolf.zechner@uni-graz.at are able to incorporate fatty acids into triacylglycerols. lipophagy in 2009 (REF. 6) suggested that acid lipoly­sis
doi:10.1038/nrm.2017.76 In vertebrates, the vast majority (>90%) of ­triacylglycerols is not restricted to the catabolism of endocytosed lipo­
Published online 30 Aug 2017 are deposited in white adipose tissue (WAT). proteins. Lipophagy is a subtype of macroautophagy

NATURE REVIEWS | MOLECULAR CELL BIOLOGY VOLUME 18 | NOVEMBER 2017 | 671


©
2
0
1
7
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
.
REVIEWS

• Membrane integrity • Growth and


• Storage and detoxification and fluidity development
of FFAs in lipid droplets • Biosynthesis of steroid • Vision
• Emulsification in lipoproteins Membrane-lipid hormones, bile acids, • Glycoprotein Endocannabinoid • Glycolysis
for transport synthesis vitamin D synthesis signalling • Gluconeogenesis
Acid lipolysis Neutral lipolysis

CE, RE Cholesterol, retinol


TAG DAG MAG Glycerol
O
O H2O HO H O HO H O HO
O ATGL O 2
HSL 2
MGL
O O HO HO
O LAL O LAL O
O O O HO
CE, RE
• Lipid signalling, such as via PPARs
• Lipotoxicity
• Activation to acyl-CoA
- mitochondrial transport and β-oxidation
- protein acylation and acetylation FFA
O
- synthesis of membrane lipids,
OH
prostaglandins, leukotrienes, ceramides

Figure 1 | Overview of the enzymes involved in lipolysis and the potential functions of their enzymatic products.
Neutral lipolysis (upper panel): adipose triglyceride lipase (ATGL), hormone-sensitive lipase (HSL) and monoacylglycerol
lipase (MGL) consecutively hydrolyse triacylglycerol (TAG), diacylglycerol (DAG) and Nature Reviews | Molecular
monoacylglycerol (MAG),Cell Biology
respectively, to generate non-esterified free fatty acids (FFAs) and glycerol. Acid lipolysis (lower panel): lysosomal acid
lipase (LAL) hydrolyses both TAG and DAG. Whether LAL is also able to hydrolyse MAG is controversial. HSL and LAL
additionally catalyse the hydrolysis of cholesteryl ester (CE) to cholesterol and retinyl ester (RE) to retinol. The potential
utilization of lipolytic intermediates and products is indicated. PPARs, peroxisome proliferator-activated receptors.

that facilitates the transport of cLD components, such a severe lipolytic defect and leads to systemic triacyl­
as triacylglycerols or cLD proteins, to lysosomes for glycerol accumulation in WAT and non-adipose tissues.
White adipose tissue
(WAT). A loose connective ­subsequent degradation by lysosomal enzymes. Triacylglycerol accrual in Atgl−/− animals is most pro­
tissue that predominantly Key discoveries during the last decade have reshaped nounced in cardiac muscle, resulting in mitochondrial
consists of adipocytes our understanding of many aspects of fat catabolism in dysfunction and lethal cardiomyopathy 3 months after
and that stores excess cells. This Review focuses on the principal pathways of birth. Global ATGL deficiency also causes severe cold
nutrients as triacylglycerols.
intracellular lipolysis — neutral lipolysis of cLDs and sensitivity. Whether defective thermoregulation results
Very-low-density acid lipolysis of cLDs (lipophagy) as well as lipo­protein- from brown adipose tissue dysfunction or impaired
lipoproteins associated lipids — and their crosstalk. Also discussed ­cardiac and haemodynamic function is not known.
(VLDL). Liver-derived plasma are new insights that are particularly relevant to both Many proteins regulate neutral lipolysis either directly
lipoproteins of very low density
understanding of the pathogenesis of metabolic dis­ or indirectly (FIG. 2). Two proteins directly interact with
that transport lipids
(predominantly eases and the exploration of new treatment strategies the patatin domain of ATGL and control its enzymatic
triacylglycerols) from the liver for lipid-associated disorders. activity in opposite ways: lipid droplet-binding protein
to non-hepatic tissues. CGI‑58 (also known as ABHD5) activates ATGL12 by an
Neutral lipolysis in the cytoplasm unknown mechanism, whereas G0/G1 switch pro­
Chylomicrons
Intestine-derived plasma
The discovery of adipose triglyceride lipase (ATGL; also tein 2 (G0S2) inhibits the enzyme13 (FIG. 2a). Consistent
lipoproteins that transport known as patatin-like phospholipase domain-­containing with a role for CGI‑58 in ATGL activation, CGI‑58−/−
dietary lipids (predominantly ­protein 2 (PNPLA2))7–9 in 2004, and the informative mice develop a lipid storage phenotype similar to the
triacylglycerols) from the phenotype of ATGL-deficient mice10 established the one observed in Atgl−/− mice, albeit less pronounced14.
digestive tract (intestine)
current concept of lipolysis in adipose and non-adipose In sharp contrast to Atgl−/− mice, however, CGI‑58−/− mice
to the liver and other tissues.
tissues. ATGL initiates triacylglycerol hydrolysis to form additionally exhibit severe ichthyosis, trans-­epidermal
Lipases diacylglycerol and FFAs. Hormone-sensitive lipase water loss and postnatal death around 16 hours after
Enzymes that hydrolyse fatty (HSL) and monoacylglycerol lipase (MGL)11 complete birth. The epidermal skin defects can be restored by
acid–glycerol esters. the process by consecutively hydrolysing diacyl­glycerols the exclusive expression of CGI‑58 in keratinocytes15.
Patatin-like phospholipase
into monoacylglycerols and FFAs and hydrolysing Although the ATGL-independent biochemical func­
domain-containing protein 2 ­monoacylglycerols into glycerol and FFAs (FIG. 1). tion of CGI‑58 in the epidermis remains unknown, it is
(PNPLA2). A protein with a interesting to note that deficiencies in PNPLA1 (REF. 16)
patatin domain that hydrolyses ATGL initiates triacylglycerol hydrolysis. ATGL belongs (another PNPLA family member, of unknown biochem­
neutral lipids, phospholipids
to a family of nine PNPLA genes and is the only robust ical function) and diacylglycerol O-acyltransferase 2
and retinyl esters.
triacylglycerol hydrolase within the family. ATGL has (DGAT2, which is an enzyme essential for the final step
Patatin domain an unusual Ser–Asp catalytic dyad in its patatin domain of triacylglycerol synthesis)17 in mice cause a very similar
A protein domain of instead of the more classical Ser–Asp–His catalytic skin phenotype. These observations suggest that all three
approximately 180 amino triad present in other lipid hydrolases. The pheno­ proteins act within the same, currently undefined bio­
acids in length that was
originally discovered
type of Atgl−/− mice clearly highlighted the crucial role chemical pathway in the epidermis. ATGL and CGI‑58
in the potato tuber storage of this enzyme in mammalian lipolysis and energy deficiencies in humans manifest clinical features similar
protein patatin. homeostasis10. Global ATGL deficiency in mice causes to those observed in the genetic mouse models (BOX 1).

672 | NOVEMBER 2017 | VOLUME 18 www.nature.com/nrm


©
2
0
1
7
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
.
REVIEWS

Brown adipose tissue Human G0S2 was discovered in expression studies response protein 1 (EGR1), leading to the repression of
Adipose tissue involved in in blood mononuclear cells during cell cycle re‑entry ATGL expression29, another argued that FSP27 directly
thermoregulation, uncoupling (G0‑to‑G1 transition)18. G0S2 suppresses lipolysis by interacts with ATGL to inhibit its hydrolytic activity 30.
mitochondrial electron inhibiting ATGL13 (FIG. 2a). Subsequent studies showed
transport from ATP synthesis
and thereby generating heat
that a 32‑amino-acid fragment of G0S2 is sufficient to HSL and MGL complete the lipolytic enzyme trio.
during chronic cold exposure. interact with the patatin domain of ATGL and non- ATGL possesses narrow substrate specificity for triacyl­
competitively inhibit its activity 19. CGI‑58 and G0S2 glycerols containing long-chain fatty acids, preferentially
sn bind to different sites on ATGL20. Overexpression and cleaving ester bonds in the sn‑1 or sn‑2 position31. The
A notation that stands
knockout studies in mutant mouse models demon­ enzyme poorly hydrolyses diacylglycerols and mono­
for ‘stereospecific numbering’
and describes the
strated that G0S2 also regulates lipolysis in various acylglycerols, so other lipases are needed to release the
stereochemical configuration ­tissues in vivo21–24. remaining fatty acids from the glycerol backbone. The
of chiral glycerol derivatives. Other proteins regulate ATGL indirectly. FFA- main diacylglycerol hydrolase is HSL (FIG. 2b), which
binding proteins (FABPs) expressed in the liver, intes­ predominantly cleaves fatty acid residues in the sn‑1 or
tine, muscle, adipose tissue and skin (FABPs 1–5) sn‑3 position of diacyl­glycerols32. HSL exhibits much
interact with CGI‑58 (REF. 25) (FIG. 2a). This interaction broader substrate specificity than ATGL and can hydro­
activates ATGL by a currently unknown mechanism. lyse ester bonds in triacylglycerols, diacylglycerols,
Pigment epithelium-derived factor (PEDF) also activ­ monoacyl­glycerols, cholesteryl esters, retinyl esters and
ates ATGL, although the mechanism by which this short-chain carbonic acid esters33 (FIG. 1). The findings
occurs is controversial. One study showed that secreted that HSL cannot compensate for an absence of ATGL
PEDF binds to ATGL and unleashes the phospho­lipase to prevent systemic lipid accumulation and that HSL
activity of the enzyme at the plasma membrane 26, deficiency results in diacyl­glycerol, rather than triacyl­
whereas other studies concluded that the ATGL–PEDF glycerol, accumulation in various tissues of ATGL-
interaction occurs on cLDs and induces triacylglycerol deficient mice and humans10,34 indicated that HSL has
hydrolysis27,28. Fat-specific protein FSP27 (also known a predominant role in diacylglycerol catabolism within
as CIDE‑C) inhibits ATGL by a mechanism that is also the lipolytic ­cascade. Lack of hydrolysis of cholesteryl
controversial. Whereas one study claimed that FSP27 esters and retinyl esters in the testes of HSL-deficient
interacts with the transcription factor early growth mice leads to defective spermatogenesis and infertility 35.

a b c d
Starvation Starvation NPs Starvation

cAMP
TFEB TFEB TFEB
PKA PKG cAMP cGMP
PPARs– PPARs– PPARs– PPARs–
cAMP SIRT1 FABPs PGC1α PGC1α PGC1α SIRT1 PGC1α
P P
PKA PKG ERK FABP4
Perilipin 1 PEDF
Ac Ac
FOXO1 AMPK FOXO1
P Ser659 Ser660
P
CGI-58 P P P P
Ser600
ATGL Ser406 Ser563 HSL MGL LAL
PDE3B G0S2
Ser565 P
cAMP
CAMK2
mTORC1, EGR1 GSK4
mTORC2 OxLDL AggLDL
PDE3B AMPK
FSP27 Lysosomes

Insulin, IGF Nutrient abundance Insulin, IGF Nutrient abundance

Figure 2 | Direct and indirect regulation of the lipases involved in neutral starvation, protein kinase A (PKA) activates HSL by phosphorylation of Ser
Nature Reviews | Molecular Cell Biology
and acid lipolysis. a | Transcription of adipose triglyceride lipase (ATGL) residues 563, 659 and 660. Natriuretic peptides (NPs) stimulate lipolysis
is controlled by sirtuin 1 (SIRT1)-mediated deacetylation of forkhead box through PKG and phosphorylation of HSL at Ser660. AMPK, glycogen
protein O1 (FOXO1) and by peroxisome proliferator-activated receptor synthase kinase 4 (GSK4) and calcium/calmodulin-­dependent protein kinase
(PPAR) and PPARγ co-activator 1α (PGC1α). ATGL enzyme activity is regulated type II (CAMK2) inhibit HSL by phosphorylating Ser565. FABP4 interacts
by direct interaction of ATGL with its co-activator, lipid droplet-binding with HSL and stimulates its activity. c | Monoacylglycerol lipase (MGL)
protein CGI‑58, and with its inhibitor, G0/G1 switch protein 2 (G0S2), as well expression is responsive to PPAR–PGC1α activation. d | Similar to
as by phosphorylation of Ser406 by AMP-activated protein kinase (AMPK). ATGL expression, lysosomal acid lipase (LAL) expression is also positively
Other factors, including pigment epithelium-derived factor (PEDF), regulated by transcription factor EB (TFEB), PGC1α and SIRT1‑dependent
fat-specific protein FSP27, fatty acid-binding proteins (FABPs), mTOR deacetylation of FOXO1. Oxidized (Ox) or aggregated (Agg) low-density
complex 1 (mTORC1) and mTORC2 and early growth response protein 1 lipoprotein (LDL) inhibits LAL activity by changing the lysosomal pH.
(EGR1), also regulate ATGL activity. b | Transcription of hormone-sensitive Ac, acetylation; cAMP, cyclic adenosine monophosphate; GF, insulin-like
lipase (HSL) is also regulated by the PPAR–PGC1α axis. In response to growth factor; P, phosphorylation; PDE3B, phosphodiesterase 3B.

NATURE REVIEWS | MOLECULAR CELL BIOLOGY VOLUME 18 | NOVEMBER 2017 | 673


©
2
0
1
7
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
.
REVIEWS

Endocannabinoid signalling Gene mutations leading to HSL deficiency in humans fasted state, catecholamines (in WAT, muscle and the
Signalling that comprises were recently reported36 (BOX 1). Even though HSL- liver) and glucagon (in the liver) increase cellular cAMP
cannabinoid receptors and deficient mice and humans exhibit similar phenotypes, concentrations by activating a Gs protein-dependent
their endogenous ligands, unlike mice, HSL-deficient human males are fertile36. adenylyl cyclase. Next, cAMP activates protein kinase
which regulate numerous
physiological processes,
The final step in the lipolytic cascade is catalysed by A (PKA), which phosphorylates a number of cLD-­
including energy homeostasis, MGL, which cleaves both triacylglycerol-derived and associated proteins, including perilipin 1, HSL and
pain perception, inflammation glycerophospholipid-derived monoacylglycerols in all CGI‑58. Unphosphorylated perilipin 1 binds CGI‑58
and tumorigenesis. three stereospecific positions11. Owing to the increased (REF. 38) and thereby prevents CGI‑58‑mediated ATGL
solubility of monoacylglycerols compared with triacyl­ activation. Perilipin 1 phosphorylation at multiple
Catecholamines
Tyrosine derivatives of
glycerols and diacylglycerols, the enzyme is cyto­plasmic residues leads to the release of CGI‑58, which is then
catechol, including adrenaline, and, unlike ATGL or HSL, does not act on cLDs. MGL able to activate ATGL39. Simultaneously, phosphoryl­
noradrenaline and dopamine, fulfils a crucial role in endocannabinoid signalling by ated HSL translocates from the cytosol to cLDs. HSL
that act as neuromodulators hydrolysing and inactivating 2­ ‑arachidonylglycerol, regulation by enzyme phosphorylation is complex 40
and hormones.
which is the most abundant endocannabinoid in (FIG.  2b) . Five distinct Ser residues are phosphoryl­
Natriuretic peptides mammals, regulating a wide variety of biological pro­ ated by either activating kinases (PKA, PKG and
Peptides that control the cesses, including energy homeostasis, pain perception, extracellular-­signal-regulated kinases (ERKs)) or inhib­
homeostasis of water, sodium ­inflammation and tumorigenesis37. itory kinases (AMP-activated protein kinase (AMPK),
and potassium in the body.
calcium/calmodulin-­dependent protein kinase type II
In adipocytes, these peptides
also regulate lipolysis.
Direct and indirect regulation of neutral lipolysis. and glycogen synthase kinase 4), which respectively
Endocrine regulation of neutral lipolysis is complex trigger and prevent HSL translocation and activation
and involves numerous hormones, growth factors (FIG. 2b). In addition to catecholamines and glucagon,
and (adipo)cytokines that are linked to diverse signal other effectors and metabolites are also able to activ­
transduction pathways33 (FIG. 2). The majority of stud­ ate neutral lipolysis via the cAMP–PKA pathway,
ies addressed the control of lipolysis by catecholamines including thyroid-stimulating hormone and melano­
and insulin in WAT during fasting and feeding. In the cortins33. Natriuretic peptides induce lipolysis through

Box 1 | Pathophysiology of neutral lipolysis


Rare mutations in the genes encoding adipose triglyceride lipase (ATGL), lipid droplet-binding protein CGI‑58 and
hormone-sensitive lipase (HSL) cause distinct metabolic disorders in humans.
• ATGL. Individuals with mutations in the gene encoding ATGL develop neutral lipid storage disease with myopathy
(NLSDM)34. Genetically confirmed NLSDM has been diagnosed in over 40 individuals with approximately 30 different
mutations. NLSDM is a rare autosomal disorder characterized by systemic triacylglycerol accumulation in multiple tissues,
including cardiac and skeletal muscles; the liver, skin and pancreas; and blood leukocytes (Jordan anomaly)164. Cardiac
steatosis is associated with severe dilated cardiomyopathy in 44% of patients, and this condition often requires heart
transplantation to avert cardiac death. Progressive skeletal myopathy is observed in the majority of patients and often
leads to loss of muscle strength and difficulties in walking. Other more moderate and rare clinical features include
hepatosteatosis and hepatomegaly, pancreatitis and cognitive impairment. Unlike Atgl−/− mice, ATGL-deficient humans
are mostly normoglycaemic and show no signs of increased insulin sensitivity or glucose tolerance165.
• CGI‑58. Human CGI‑58 deficiency causes neutral lipid storage disease with ichthyosis (NLSDI; also designated
Dorfman-Chanarin syndrome). All individuals with NLSDI exhibit severe congenital ichthyosiform erythroderma166,167.
More infrequent clinical features include hepatosplenomegaly, myopathy, hearing loss and mental retardation.
Although patients with NLSDI also exhibit systemic triacylglycerol accumulation and Jordan anomaly, lipid
accumulation in muscle and the heart does not lead to severe skeletal or cardiac myopathy. In 2001, mutations in
CGI‑58 were found to cause NLSDI168. Unlike NLSDM, NLSDI always occurs in conjunction with a severe skin phenotype,
supporting an ATGL-independent function of CGI‑58 in the epidermis. Similar to CGI‑58−/− mice14, the epidermal
keratinocytes of humans with NLSDI are unable to synthesize acylceramides or to form a functional corneocyte lipid
envelope169. Although the pathophysiological consequences of CGI‑58 deficiency are well established, the actual
biochemical function of CGI‑58 remains a subject of debate. The postulated ATGL-independent functions of CGI‑58
include enzyme activity as an acyltransferase and/or involvement in autophagy145,170,171. To date, however, a link between
these activities and the skin pathology observed in patients with NLSDI remains elusive.
• HSL. Mutations causing HSL deficiency were first reported only recently36,172. Individuals with homozygous mutations
in HSL exhibit a relatively benign phenotype. Similar to Hsl−/− and Atgl−/− mice, HSL-deficient patients develop partial
lipodystrophy owing to a reduction in peroxisome proliferator-activated receptor γ (PPARγ)-driven triacylglycerol
synthesis in HSL-deficient adipocytes, reinforcing the links among lipolysis, PPAR signalling and lipid synthesis. Unlike
Hsl−/− mice, however, all HSL-deficient humans develop insulin resistance and type 2 diabetes, fatty liver disease and
dyslipidaemia, characterized by increased plasma triacylglycerol and low-density lipoprotein concentrations and
decreased plasma high-density lipoprotein concentrations. It is conceivable that this metabolic syndrome phenotype is
a consequence of the partial lipodystrophy affecting HSL-deficient individuals. A similar phenotype in knockout mice
may have been missed because plasma lipid and lipoprotein parameters were studied in young mice, before the onset
of lipodystrophy. It is also interesting to note that HSL-deficient human males are fertile, whereas male Hsl−/− mice are
sterile35. Functional spermatogenesis and fertility in mice require the presence of enzymatically active HSL in the testes,
which is apparently not the case in human males with HSL deficiency173.

674 | NOVEMBER 2017 | VOLUME 18 www.nature.com/nrm


©
2
0
1
7
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
.
REVIEWS

the synthesis of cyclic guanosine monophosphate adipose tissue and muscle56 may at least partially under­
and the activation of PKG, which phosphorylates peri­ lie the confusion. It is also uncertain whether AMPK
lipin 1 and HSL41 (FIG. 2b). Finally, pituitary growth hor­ or PKA ­phosphorylates ATGL at Ser406 to regulate
mone (somato­tropin) activates ATGL by inducing its its activity 57,58.
gene transcription via signal transducer and activator Whereas ATGL and HSL are regulated by multi­
of ­transcription 5 (STAT5) signalling 42. ple mechanisms, MGL regulation is less complex
Inhibitors of lipolysis include the classical ­hormone (FIG. 2c). Tissue-specific isoforms of the enzyme exist 59,
inhibitors insulin and insulin-like growth factors43 but whether this variation affects activity is not clear.
as well as non-hormone inhibitors like lactate, adeno­ Posttranslational modifications have not been linked to
sine, β‑hydroxybutyrate and nicotinic acid (niacin), MGL function. Transcriptionally, MGL is a target for
which act through G-protein-coupled receptors 33. the nuclear receptor peroxisome proliferator-activated
Although known for decades, the complexity of insulin-­ receptor α (PPARα)60.
mediated inhibition of lipolysis remains incompletely
understood. According to the established model, Neutral lipolysis regulates key signalling pathways in
insulin induces PKB-dependent phosphorylation and energy homeostasis. In addition to having a ­crucial role
activation of phosphodiesterase 3B (PDE3B), which as energy substrates, the products and inter­mediates
subsequently hydrolyses cAMP (FIG. 2a,b). This event in of neutral lipolysis impact lipid signalling and energy
turn leads to decreased phosphorylation of perilipin 1 homeo­stasis. For example, a functional link exists
and HSL by the inhibition of PKA as well as the induc­ between lipolysis, PPAR signalling and oxidative ­capacity
tion of protein phosphatase 1 and downregulation of in various tissues. Defective lipolysis in the heart of Atgl−/−
lipolysis. Consistent with this model, insulin-mediated (REF. 61) and CGI‑58−/− (REF. 62) mice not only causes
inhibition of lipolysis was recently shown to require severe fat accumulation but also leads to impaired mito­
the activation of PDE3B44. However, PDE3B activation chondrial respiratory function. This insufficiency results
did not depend on PDE3B phosphorylation by PKB44. from pronounced downregulation of PPARα ­target
Accordingly, the mech­anism of PDE3B activation by genes involved in mitochondrial substrate oxid­ation
insulin remains unknown. Insulin additionally inhibits and oxid­ative phosphorylation. PPARα agonists such as
ATGL mRNA synthesis in WAT and the liver by exerting WY14643 and fenofibrate reverse the phenotype, con­
an inhibitory effect on the transcription factor forkhead sistent with the concept that mitochondrial dysfunction
box protein O1 (FOXO1), which is a direct activator of contributes to the lethal cardiomyopathy in Atgl−/− mice.
ATGL transcription45 (FIG. 2). Conversely, FOXO1 activ­ This conclusion is further supported by the observation
ation by deacetylation mediated by NAD+-dependent that cardiomyocyte-exclusive expression of ATGL pre­
protein deacetylase sirtuin 1 (SIRT1) induces ATGL vents severe cardiomyopathy and premature death63.
expression46, thereby providing an explanation for the Decreased expression of PPARα target genes was also
pro-lipolytic effect of SIRT1 (REF. 47) and the SIRT1 observed in the liver 64,65, intestinal cells66, macrophages67
­activator resveratrol48. and brown adipocytes57,68 of Atgl−/− mice. Hepatic PPARα
Other major insulin-stimulated kinases that regulate signalling not only is affected by tissue-autonomous
neutral lipolysis include mTOR complex 1 (mTORC1), lipolysis in the liver but also depends on lipolysis in
mTORC2 (REF. 49) and AMPK50 (FIG. 2). The inhibition and FFA delivery from WAT. Conditional knockout of
of mTORC1 by rapamycin or inactivation by raptor the genes encoding ATGL69 or CGI‑58 (REF. 70) in WAT
deletion increase neutral lipolysis and FFA release from drastically lowers hepatic PPARα activity and target
adipocytes51; conversely, mTORC1 activation through gene expression. In pancreatic β-cells, ATGL regulates
overexpression of GTP-binding protein Rheb decreases mitochondrial function and glucose-­stimulated insulin
lipolysis, supporting an antilipolytic role for mTOR secretion through PPARδ71, and β-cell-specific deletion
signalling 51. The signalling mechanism is complex and of ATGL causes hyperglycaemia, which is reversible by
incompletely understood but involves transcriptional the ­administration of PPARδ (but not PPARα) agonists.
regulation of ATGL by EGR152. Similarly, mTORC2 ATGL-mediated lipolysis also affects PPARγ sig­
inactivation in WAT due to adipose-specific loss of the nalling and lipid synthesis in WAT. Mice with global or
mTORC2 subunit rictor also resulted in unrestrained adipose tissue-specific ATGL deficiency are resistant to
lipolysis, supporting an important antilipolytic role for obesity induced by a high-fat diet owing to decreased
mTOR signalling during nutrient abundance53. food intake and downregulation of FFA uptake and
The role of AMPK in the regulation of lipolysis is triacyl­g lycerol synthesis in WAT69,72. The defect in
less well defined. AMPK is activated during fasting and PPARγ signalling and lipid synthesis in WAT can be
exercise, when cellular AMP concentrations increase, abrogated by the PPARγ agonist rosiglitazone72. This
but whether or not this induction contributes to the finding is reminiscent of the impaired PPARγ signal­
upregulation of lipolysis is controversial. Contradictory ling and lipodystrophy that occurs in HSL-deficient
reports claim that the kinase induced, inhibited or mice and humans36,73 and supports a more general func­
did not affect neutral lipolysis in WAT50. The findings tion for lipolysis in PPAR signalling, beyond the activity
that the lipolytic response depends on the duration of of ­individual lipases.
experi­mental AMPK inhibition or activation54, that The signal transduction pathways linking lipolysis
AMPK stimulates ATGL but inhibits HSL55 (FIG. 2a,b) and to PPAR activity remain unclear. The evidence that
that the HSL phosphorylation pattern differs between FABPs bind to CGI‑58, thereby affecting PPARα and

NATURE REVIEWS | MOLECULAR CELL BIOLOGY VOLUME 18 | NOVEMBER 2017 | 675


©
2
0
1
7
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
.
REVIEWS

Starvation acetyl-proteome82–84. Histone acetylation at specific Lys


residues is a highly regulated process that requires, in
addition to acetyl-CoA, the activity of more than a
P
AMPK ULK1
Autophagy, dozen histone acetyltransferases. The abundance of
mTOR lipophagy cyto­plasmic acetyl-CoA affects histone acetylation and
P regu­lates the transcription of numerous genes85,86, but
TFEB P SIRT1
Ac the specific effect of lipolysis on histone acetylation
PGC1α
­patterns remains to be elucidated.
P Ac Unlike histone acetylation, protein acetylation out­
TFEB PGC1α PPARα FOXO1 side the nucleus occurs at low stoichiometry, likely by a
non-enzymatic mechanism that is mostly driven by
PGC1α local acetyl-CoA concentrations82,83. Consistent with
this concept, non-nuclear protein acetylation is highest
PPARα
RXR

in mitochondria and is markedly affected by fasting-­


TFEB FOXO1 induced lipolysis and FFA oxidation. Fasting increases
mitochondrial protein acetylation, but this effect is
completely attenuated in Atgl−/− mice84 and in mice that
lack very-long-chain acyl-CoA dehydrogenase, which
ATGL LAL
is a key enzyme for the initiation of FFA ­β-oxidation87.
HSL
This finding supports the view that lipolysis and
MGL acetyl-CoA production drive mitochondrial protein
acetylation. During fasting or calorie restriction, when
mitochondrial acetyl-CoA concentrations are high,
Neutral Acid
lipolysis lipolysis ­target sites in multiple proteins are hyperacetylated and
require subsequent deacetylation (‘repair’) by the mito­
Figure 3 | Common regulation of neutral lipolysis and lipophagy. During nutrient
Nature Reviews | Molecular Cell Biology
chondrial deacetylase SIRT3 (REFS 84,88). Consistent
starvation, inhibition of the mTOR signalling pathway allows the dephosphorylation
and nuclear translocation of transcription factor EB (TFEB). Starvation also activates
with the increased demand for such repair, SIRT3 is
AMP-activated protein kinase (AMPK), which phosphorylates peroxisome proliferator- upregulated during fasting and calorie restriction88,89.
activated receptor γ (PPARγ) co-activator 1α (PGC1α), thereby activating gene Increased hepatic acetyl-CoA concentrations owing to
transcription mediated by PPARα–PGC1α–retinoid X receptor (RXR). AMPK enhances increased adipose lipolysis adversely affect insulin con­
sirtuin 1 (SIRT1) activity, leading to the deacetylation of downstream SIRT1 targets, trol of hepatic glucose production90. Whether this effect
including PGC1α and the forkhead box O (FOXO) transcription factors. After involves differences in protein acetylation in response to
translocation to the nucleus, TFEB, PPARα–PGC1α–RXR and the FOXO transcription acetyl-CoA concentrations is not known.
factors initiate the transcription of adipose triglyceride lipase (ATGL) and lysosomal
acid lipase (LAL), resulting in increased neutral lipolysis and acid lipolysis, respectively. Acid lipolysis in lysosomes
AMPK induces lipophagy by phosphorylating and activating serine/threonine-protein
In addition to cLDs, endosomes and lysosomes are the
kinase ULK1.
only cell organelles where intracellular triacyl­glycerol
hydrolysis occurs. Neutral lipolysis was thought to be
PPARγ signalling in hepatocytes and fat cells25, supports restricted to the catabolism of cLD-associated triacyl­
lipolysis-­coupled transport of FFAs into the cell nucleus, glycerols, whereas lysosomal (acid) lipolysis was
where they serve as activating ligands for PPARs. assumed to be exclusively responsible for the degrad­
Consistent findings show that PPAR activity requires the ation of exogenous plasma lipoprotein-associated lipids,
nuclear translocation of FFA-loaded FABPs74–76. FABP4 including triacylglycerols. The discovery of lipophagy 6
also interacts with HSL and stimulates its enzyme activ­ challenged this view and demonstrated that lyso­somal
ity, but it is not known whether this interaction regulates triacylglycerol degradation also contributes to the
PPAR activity77,78. Additionally, lipolysis activates SIRT1 ­turnover of c­ ytosolic lipid stores.
(REF. 79), which deacetylates and thereby activates PPARγ Lysosomal triacylglycerol degradation is carried out
co-activator 1α (PGC1α). SIRT proteins also induce the by LAL owing to its optimal activity at the lysosomal pH
activity of FOXO transcription factors, which regu­ of 4.5–5. LAL is highly glycosylated and exists in vari­
late multiple processes in adipogenesis and adipose ous tissue-specific isoforms. It exhibits broad substrate
metabolism, including autophagy 80 and the expression specificity, hydrolysing triacylglycerols, diacyl­glycerols,
of ATGL46 and LAL81. Accordingly, ATGL, sirtuins and cholesteryl esters91,92 and retinyl esters93 (FIG. 1). Whether
FOXO1 form an autoregulatory loop that coordinates LAL also hydrolyses monoacylglycerols is contro­
lipid metabolism and links neutral lipolysis and acid versial91,92. LAL localization is not restricted to lyso­
lipolysis (FIG. 3). somes, as it can be secreted from cells via the classical
endo­plasmic reticulum (ER)–Golgi secretory pathway
Protein acetylation. Lipolysis may also regulate meta­ and can subsequently re-enter cells and lysosomes by
bolic pathways via protein acetylation. The oxidation endocytosis94. Furthermore, LAL can be secreted from
of lipolysis-derived FFAs generates acetyl-CoA, which cells by a process called exophagy, in which lysosomes
is the sole cofactor for protein acetylation. It is there­ fuse with the plasma membrane and release their content
fore not surprising that lipolysis affects the cellular into the extracellular space95.

676 | NOVEMBER 2017 | VOLUME 18 www.nature.com/nrm


©
2
0
1
7
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
.
REVIEWS

Scavenger receptors Exogenous triacylglycerols, cholesteryl esters and the postnatal state, and the lipid accumulation pheno­
Lipoprotein receptors that r­ etinyl esters destined for LAL-mediated degradation type develops more gradually in these mice than in
remove modified lipoproteins enter cells by receptor-dependent uptake of plasma lipo­ humans98,99. Notably, Kupffer cells, and not hepatocytes,
(for example, acetylated proteins. Lipoprotein binding by various receptors, such are the main liver cell type in which cholesteryl esters
or oxidized low-density
lipoproteins) and other
as the low-density lipoprotein (LDL) receptor, LDL and triacylglycerols accumulate, which raises both the
negatively charged receptor-related proteins or scavenger receptors; subse­ question of how LAL-deficient hepatocytes manage
macromolecules from quent lipoprotein internalization; and the recycling of to cope with constant encounters with lipoprotein-­
the blood. endosomal receptors have been extensively reviewed else­ associated lipids and the possibility that LAL is not the
where96,97. Although the intracellular sorting mechanisms sole enzyme responsible for degrading triacylglycerols
Kupffer cells
Specialized macrophages
for all lipoprotein receptors are not equally well studied, and choles­teryl esters in lysosomes. Additionally, Lal−/−
in the liver. it is generally understood that when late endosomes even­ mice become lipodystrophic99,100.
tually fuse with lysosomes, lysosomal hydrolases degrade Lysosomes are unable to store any degradation prod­
Foam cell all lipoprotein components, including apolipoproteins, ucts, so the catabolic machinery, including LAL, and
A lipid-filled macrophage that
is present in atherosclerotic
phospholipids and neutral lipids. To date, the only known lysosomal export mechanisms are constitutively active.
lesions and plaques. enzyme involved in the hydrolysis of neutral lipids is LAL. Therefore, the regulation of acid lipolysis, and specifically
LAL, is less complex than the regulation of neutral lipolysis
LAL regulation. An absence or dysfunction of LAL (FIG. 2). LAL regulation occurs predominantly at the gene
results in excessive lysosomal lipid accumulation in transcription stage. Similar to the regulation of neutral
humans and mice, consistent with the notion that LAL lipases, fasting promotes FOXO1‑dependent LAL gene
is the sole enzyme responsible for degrading triacyl­ transcription following the SIRT1‑mediated deacetylation
glycerols and cholesteryl esters in lysosomes. LAL defi­ and nuclear translocation of FOXO1 (REF. 81). LAL gene
ciency in humans causes Wolman disease and cholesteryl transcription is also activated by transcription factor EB
ester storage disease (CESD) (BOX 2). Lal−/− mice survive (TFEB)101. Whether transcription factor E3 (TFE3) also
regulates LAL gene transcription has not been directly
tested but seems likely considering the highly redun­
Box 2 | Pathophysiology of acid lipolysis
dant activities of TFEB and TFE3 (REFS 102,103) and the
Mutations in the lysosomal acid lipase (LAL)-encoding gene (LIPA) cause two diseases in fact that in Caenorhabditis elegans, the TFE3 orthologue
humans that differ in severity. Wolman disease is a lipid storage disorder174 characterized HLH‑30 activates lysosomal lipolysis104. When nutrients
by a failure to thrive, progressive hepatosplenomegaly, adrenal calcification and death are abundant, the mTOR complex is active and phos­
within the first year after birth. Cholesterol ester storage disease (CESD) has a similar phorylates TFEB and TFE3, which remain cytosolic and
phenotype but a later onset and a much slower progression175. Both conditions are
inactive under this condition. Upon fasting, the proteins
caused by reduced LAL activity, giving rise to excessive cholesteryl ester and
triacylglycerol accumulation in lysosomes176,177. Whereas individuals with Wolman
are dephosphorylated, translocate to the nucleus and
disease invariably lack the enzyme and require liver transplantation to survive, regulate the transcription of numerous genes (FIG. 3).
individuals with CESD frequently, but not always, exhibit some remnant enzyme In oxidative tissues such as the liver and muscle, TFEB
activity178. The spectrum of disease manifestation and progression in CESD ranges from regulates the expression of LAL and induces the tran­
severe liver disease and liver failure in children to no apparent clinical features until scription of other transcription factors that are involved
adulthood. The outcome is often independent of the magnitude of remnant LAL activity. in catabolic pathways, such as PPARα and its c­ o-activator,
Therefore, the more recently used designation ‘LAL deficiency’ (LAL‑D) seems more PGC1α101,105, which further augment the expression of
appropriate for both conditions. Unlike Lal−/− mice, which develop severe steatohepatitis, LAL and the neutral lipases ATGL, HSL and MGL60.
characterized by massive macrophage infiltration and inflammation, humans with LAL‑D Few factors are known to regulate LAL post-­
develop liver steatosis and fibrosis but show no signs of liver inflammation178,179.
transcriptionally, but oxidized and aggregated LDLs are
Individuals with CESD develop dyslipidaemia characterized by increased plasma
low-density lipoprotein (LDL) cholesterol concentrations and decreased plasma
important inhibitors (FIG. 2d). Although the details of
high-density lipoprotein (HDL) cholesterol concentrations, resulting in increased related mechanisms are unknown, modified lipoproteins
atherosclerosis178. LAL has a crucial role in cholesterol homeostasis by linking appear to cause an increase in luminal pH in lysosomes
exogenously internalized cholesterol and cholesteryl esters to the regulation of de novo owing to decreased vacuolar H+-ATPase-mediated pro­
cholesterol synthesis pathways (see main text). Excessive uptake of modified LDL ton pumping and lysosomal-to-­membrane leakiness106,107,
in macrophages via the scavenger receptor pathway has been shown to inhibit LAL in leading to decreased LAL activity. The modified LDL-
lysosomes and extracellular compartments. In macrophages, this derepresses de novo induced hydrolytic defect results in cholesteryl ester
cholesterol synthesis, increases LDL receptor activity and attenuates HDL-mediated retention in lysosomes108, induction of de novo choles­
reverse cholesterol transport. Together, these events lead to increased foam cell terol synthesis and inhibition of reverse cholesterol trans­
and coronary lesion formation and an increased risk of coronary artery disease.
port 109. Taken together, these processes may contribute
Enzyme replacement therapy was recently approved for the treatment of individuals
with LAL-D180. Similar treatment strategies have been previously approved for other
to foam cell formation and increased a­ therosclerosis in
lysosomal storage diseases, such as Fabry disease and Pompe disease. In a individuals with CESD (BOX 2).
placebo-controlled trial, 66 patients were treated with an enzymatically active LAL
preparation (sebelipase)180. Although the mechanism of cellular enzyme uptake LAL-generated lipids regulate cellular lipid homeostasis.
remains unknown, sebelipase treatment was beneficial for many patients with LAL-D. The products of LAL are FFAs, unesterified cholesterol
It improved liver function, normalized plasma transaminase activities and decreased and retinol (FIG. 1). The lysosomal transport and exit strat­
hepatic triglyceride content. Whether this (very costly) treatment will affect the egies for these highly hydrophobic compounds are only
long-term outcome or end points of the disease in more severely affected individuals partially understood. The best-studied lysosomal lipid
requires further investigation. Sebelipase may also be a treatment strategy for export process is the transport and secretion of unesteri­
atherosclerosis, which is commonly observed in individuals with LAL-D179.
fied cholesterol by Niemann–Pick disease type C (NPC)

NATURE REVIEWS | MOLECULAR CELL BIOLOGY VOLUME 18 | NOVEMBER 2017 | 677


©
2
0
1
7
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
.
REVIEWS

proteins, which are involved in lysosomal cholesterol neurological disorders, cancer, cardiomyopathies and
traffick­ing 110. According to the current model, NPC2 metabolic dis­orders116. The major variants of autophagy
binds unesterified cholesterol and shuttles it to the lyso­ include macro­autophagy, chaperone-mediated auto­
somal membrane, where NPC1 facilitates transmembrane phagy (CMA) and microautophagy 117. One form of
passage. Deficiencies in either NPC1 or NPC2 cause NPC, macroautophagy, designated ‘lipophagy’, was discovered
which is a severe lysosomal cholesterol storage disorder 110. in 2009 and was shown to contribute to the hydrolysis of
In addition to binding NPC proteins, cholesterol binds to ­triacylglycerols stored in cLDs6.
lysosome-associated membrane glycoprotein 1 (LAMP1) Lipophagy relies on the same general mechanisms
and LAMP2, and similarly to what has been observed as macroautophagy, which involves more than 30 ATG-
for NPC, the absence of these proteins results in the encoding genes118 (FIG. 4a). Macroautophagy in general
accumu­lation of unesterified cholesterol in lysosomes111. and lipophagy in particular are strongly induced by the
How other LAL products are exported remains largely major metabolic kinases mTORC1 and AMPK during
unknown. For example, it is unclear whether lysosomal lengthy fasting 119. The activity of these kinases depends
monoacylglycerol hydrolysis is needed or whether mono­ on growth factor signalling, the cellular energy status
acylglycerols per se exit lysosomes for further hydrolysis by (ATP:AMP ratio) and nutrient availability 120. Nutrient-
MGL in the cytoplasm. Similarly, the mechanisms of the mediated transcriptional regulation of hepatic autophagy
transport and release of FFAs and retinol remain elusive. also occurs through the nuclear receptors PPARα and the
Accumulation of FFAs in NPC1‑deficient late endosomes liver X receptors121.
originally indicated a role for NPC1 in endosomal FFA The large size of cLDs impedes their recruitment into
release112. However, normal FFA flux in the fibroblasts lipoautophagosomes6; therefore, lipophagy recruits only
of individuals lacking NPC1 did not support a role for parts of cLDs (FIG. 4b). The potential mechanisms of and
NPC1 in FFA translocation from lysosomes113, leaving the the recruitment factors involved in cLD fragmentation as
­mechanisms of FFA export essentially uncharacterized. well as the specific cargo targets on cLDs remain unidenti­
After their release from lysosomes, LAL-generated fied. Lipoautophagosomes eventually fuse with late
lipids regulate numerous metabolic processes and serve endosomes or lysosomes to form auto­lysosomes (FIG. 4b).
as crucial precursors for the de novo synthesis of lipid The fusion mechanism is insufficiently understood but
species (FIG.  1). Unesterified cholesterol becomes an may involve soluble N‑ethylmaleimide-sensitive factor
important membrane constituent or, upon membrane attachment protein receptors, microtubule-associated
saturation, is re-acylated and deposited as cholesteryl protein light chain 3 (LC3), LAMP1, LAMP2B and
esters in cLDs. The cholesterol content of the ER in turn LAMP2C as well as small GTPases, such as Ras-related
regulates the processing, localization and activity of sterol protein Rab7a122–125. Within autolysosomes, LAL presum­
regulatory element-binding protein 2, which is an impor­ ably hydrolyses triacylglycerols and cholesteryl esters,
tant transcriptional activator of cholesterol biosynthesis whereas acid phospholipases and peptidases degrade
enzymes and the LDL receptor 96. Furthermore, the oxid­ cLD-­associated phospholipids and proteins, respectively.
ation products of unesterified cholesterol (for example, A recent study presented an alternative mech­anism,
27‑OH‑cholesterol) are ligands of nuclear receptors in which late endosomes and lysosomes directly bind to
in the liver X receptor family 114. These receptors affect ‘primed’ cLDs in a process that requires Rab7a (REF. 123)
­cholesterol homeostasis by inducing the conversion (FIG. 4b). This interaction is a transient ‘kiss and run’-like
of cholesterol into bile acids in the liver, upregulating process that resembles the microautophagy that occurs
cholesterol efflux from peripheral cells and increasing in yeast, in which catabolic organelles directly bind to
­intestinal ­cholesterol excretion. target structures and hydrolyse them after their incorpor­
Similar to FFAs produced by neutral lipolysis, LAL- ation126. How this process would allow for the inter­action
generated FFAs are able to regulate PPARα and PPARα of LAL with triacylglycerols within the core of cLDs
target gene expression100. However, the consequences of remains to be elucidated.
decreased PPARα signalling in LAL deficiency may be A substantial contribution of lipophagy to the catabo­
less severe than in humans or mice with ATGL deficiency lism of cellular triacylglycerols and cholesteryl esters
and do not lead to the development of cardiomyopathies. has been demonstrated in various cell types, including
Notably, a study in C. elegans demonstrated that LAL hepato­cytes, enterocytes, macrophages, brown adipo­
is involved in the generation of a specific lipid ligand cytes and neurons127. Autophagy inhibitors, knockdown
(oleoylethanolamine) for nuclear receptors that regu­ of ATG5 in liver cells and the phenotype of mice with
late the expression of metabolic genes and longevity 115. liver-­specific ATG7 deficiency substantiated the crucial
Whether LAL-generated retinol and monoacylglycerols role of lipophagy in the breakdown of hepatic triacyl­
contribute to nuclear receptor and endocannabinoid glycerols6. Inhibition of neutral lipases or autophagy
­signalling, respectively, is not known. showed that both processes are activated in hepatocytes
during fasting 6. The quantitative contribution of each
Lipophagy of the lipolytic pathways (cytosolic lipolysis and lipo­
Steatosis In addition to the ubiquitin–proteasome protein phagy) to overall lipid catabolism is unknown and may
A process describing the degrad­ation system, autophagy, which is less substrate vary consider­ably between different cell types, such as
abnormal retention of neutral
lipids (triacylglycerols and
selective, represents the most important catabolic path­ hepatocytes, macrophages and adipocytes. Generally,
cholesteryl esters) within cells way for vari­ous cellular components. Defective auto­ the magnitude of s­ teatosis in mice with knockout of
and tissues. phagy is associated with numerous diseases, including autophagy-­specific genes is variable and more moderate

678 | NOVEMBER 2017 | VOLUME 18 www.nature.com/nrm


©
2
0
1
7
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
.
REVIEWS

a Autophagy pro-LC3-I
ATG4
• Insulin
• Growth factors LC3-I
• Amino acids
ATG12 ATG7
ATG7 ATG3

ATG10 LC3-II

mTORC1 ATG5
ATG16 p62
ATG5 Cargo
AMPK p62
ATG16 ATG12

ULK1/2-ATG13–FIP200 WIPIs
LC3 lipidation and
Cytoplasm complex PI3P PE autophagosome formation
Beclin 1 VPS34
ER Phagophore
initiation and elongation

b Lipophagy LAL
Lysosome

cLD Rab7 p62


cLD-cargo p62
LC3-II
Cytoplasm
Lipoautophagosome Autolysosome
ER Phagophore

Lipophagy–lipolysis Perilipin 3
crosstsalk” Perilipin 2 HSC70
Perilipin 2,3
LAMP2A
cLD FFA
LC3-II Acid lipolysis
ATGL– CGI-58
LDL

Receptor-mediated endocytosis
Figure 4 | Autophagy, lipophagy, lipolysis and their crosstalk. a | Autophagy. Nutrient uptake and signalling through
insulin, growth factors and amino acids activate mTOR complex 1 (mTORC1), which inhibits autophagosome assembly by
phosphorylating and inactivating a complex composed of serine/threonine-protein kinase ULK1 or ULK2, ATG13 and FAK
family-interacting protein of 200 kDa (FIP200; also known as RB1CC1). Upon nutrient deprivation, in the absence of growth
factors and in conditions of increasing energy demand, AMP-activated protein kinase (AMPK) becomes active, inhibits
Nature Reviews | Molecular Cell Biology
mTORC1 and activates ULK1. Activated ULK1 phosphorylates ATG13 and FIP200 in ULK1–ATG13–FIP200 and ULK2–
ATG13–FIP200 complexes, which then activate beclin 1. Activated beclin 1 interacts with phosphatidylinositol 3‑kinase
catalytic subunit type 3 (PIK3C3; also known as VPS34). Beclin 1–VPS34 heterodimers generate PI‑3,4,5‑phosphate (PI3P),
which nucleates the formation of autophagosomal membranes. Subsequent autophagosome elongation facilitated by
the binding of PI3P to WD repeat domain phosphoinositide-interacting proteins (WIPIs) and autophagosome maturation
involves the successive activity of two ubiquitin-like conjugation systems. The conjugation of ATG12 to ATG5 by the ligases
ATG7 and ATG10 and subsequent binding of the complex to the multimeric protein ATG16 and conjugation of
proteolytically processed microtubule-associated protein light chain 3 (pro‑LC3‑I) to phosphatidylethanolamine (PE)
by the ligases ATG7 and ATG3 lead to LC3‑II formation. The incorporation of LC3‑II into the growing membrane and the
recruitment of cargo adaptor proteins such as p62 lead to the enclosure of cargo (cytoplasmic components or cell
organelles) within autophagosomes. b | Lipophagy involves LC3‑II‑positive membranes engulfing small cytosolic lipid
droplets (cLDs) or sequestering portions of large cLDs. Lipoautophagosomes deliver cLD cargo to lysosomes, wherein
lysosomal acid lipase (LAL) degrades the lipid cargo. Subsequently, non-esterified free fatty acids (FFAs) are released into
the cytosol. Alternatively, lysosomes directly bind to ‘primed’ cLDs via a ‘kiss and run’ mechanism in a process that requires
Ras-related protein Rab7a. Crosstalk between lipophagy and neutral lipolysis is established when LC3‑II‑positive
membranes direct adipose triglyceride lipase (ATGL) to cLDs to increase lipolysis. Lipophagy–lipolysis crosstalk. Activation
of chaperone-mediated autophagy degrades the cLD coat proteins perilipin 2 and perilipin 3 through the coordinated
action of heat-shock cognate 71 kDa protein (HSC70; also known as HSPA8) and the receptor lysosome-associated
membrane glycoprotein 2A (LAMP2A). Removal of perilipins from the cLD surface allows the docking of autophagy
proteins, and the cLD surface becomes accessible to neutral lipolysis by ATGL in complex with lipid droplet-binding protein
CGI‑58, which hydrolyses the cLD triacylglycerols to generate FFAs. Acid lipolysis. After receptor-mediated endocytosis
of lipoproteins, cholesteryl esters and triacylglycerols are hydrolysed by LAL in lysosomes. ER, endoplasmic reticulum.

NATURE REVIEWS | MOLECULAR CELL BIOLOGY VOLUME 18 | NOVEMBER 2017 | 679


©
2
0
1
7
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
.
REVIEWS

than in mice lacking ATGL or its co-activator, CGI‑58. A third form of crosstalk involves the consecutive
For example, liver-­specific ATG7 deficiency, which is action of autophagy and neutral lipolysis147. In the first
expected to completely abolish hepatic autophagy, caused step, autophagic digestion of membrane material and sub­
a marked increase in hepatic fat content in some, but not sequent triacylglycerol synthesis lead to the formation of
all, studies101,128–131. Conflicting results were also reported cLDs. In the second step, cLD-associated triacylglycerols
for the role of autophagy in fatty liver disease. The majority are hydrolysed by ATGL. This process is closely linked to
of studies132–134 provided evidence for a protective role of mitochondrial FFA uptake and oxidation.
lipophagy against hepatosteatosis, but others concluded Importantly, neutral lipolysis may be directly involved
that hepatocyte-specific autophagy deficiency per se does in the regulation of autophagy, specifically through
not ­exacerbate hepatic steatosis131,135. the effect of ATGL on the hepatic function of PPARα
and SIRT1 (described above), both of which are well-­
Lipophagy and lipid synthesis. In addition to lipid catabo­ established activators of autophagy in the liver 119,121.
lism, lipophagy has important roles in adipogenesis, lipid Together with the co‑regulation of lipophagy and n ­ eutral
synthesis and cLD biogenesis. LC3‑II binding to cLDs lipolysis by the major metabolic hormones and their
is required for lipid synthesis and cLD growth in liver associ­ated regulatory hubs (mTOR, AMPK and the FOXO
cells and cardiomyocytes, and defective LC3 conjugation transcription factors) (FIG. 3), these data support the view
in mice with liver-specific ATG7 deficiency inhibited that neutral lipolysis and lipophagy should not be con­
cLD formation129,136. Similar results were observed in sidered distinct, but instead should be seen as two sides
Atg5−/− mouse embryonic fibroblasts137. Consistent with of the same coin.
a role for autophagy in adipogenesis, mice with adipose-­
specific ATG7 deficiency and newborn mice with global Therapeutic potential
ATG5 deficiency exhibit decreased adipose mass138–140, The question of whether the pharmacological inhib­
the opposite of what would have been predicted from a ition of lipolysis in WAT and the concomitant decline in
defect in lipophagic triacylglycerol catabolism. Autophagy plasma FFA concentrations represent a beneficial strat­
was found to be crucial for adipocyte differentiation at egy for preventing and treating ectopic lipid accumulation
early stages but dispensable later during differentiation141. and lipotoxicity has been widely discussed. The concept
Finally, as mentioned above, autophagy-defective livers appears attractive with regard to achieving a lasting reduc­
are unable to store lipids in response to fasting-induced tion of hepatosteatosis and preventing its progression to
WAT lipolysis131,135,136,142. Together, these data suggest that more morbid pathologies, such as steatohepatitis, cirrho­
besides having a role in catabolic pathways, autophagy sis and liver cancer 148. Decreased FFA concentrations
has a function in anabolic lipid metabolism. It is conceiv­ in the plasma may also increase insulin sensitivity and
able that phosphatidylethanolamine-conjugated LC3‑II decrease the concentration of atherogenic lipoproteins
(FIG. 4a) binds to cLDs and acts as a platform for both (including VLDL)149. The antilipolytic action of nico­
­anabolic and catabolic processes. tinic acid (­niacin), a drug long known to lower plasma
triacyl­glycerol levels and increase the concentrations of
Lipolysis–lipophagy crosstalk high-density lipoprotein (HDL) cholesterol, provided
The recent findings that metabolic kinases such as ERK partial proof of this concept 150. Niacin binds to the adi­
interact with LC3‑II143 and that LC3‑II binds ATGL on pose tissue-specific G-protein coupled receptor 109A
cLDs in brown adipocytes after cold exposure to promote (GPR109A; also known as HCAR2), leading to decreased
triacylglycerol hydrolysis144 strengthened the possibility cellular cAMP concentrations and reduced lipolysis151.
of a functional link between autophagy and cytosolic The antilipolytic effect decreases plasma FFA concentra­
lipolysis. Clear discrimination between and elucid­ tions, which in turn results in a favourable plasma lipo­
ation of the autophagy-related and autophagy-­unrelated protein profile owing to decreased VLDL production in
functions of ATGL will require systematic epistatic the liver and increased HDL-mediated reverse cholesterol
experiments combined with autophagic flux analyses. transport. Unfortunately, niacin has several adverse side
In addition to ATGL, CGI‑58 may be directly involved effects that restrict its broad application152. Additionally, in
in the regulation of autophagy by preventing cleavage of Gpr109a−/− mice, the decrease in hepatic VLDL synthesis
beclin 1 by caspase 3 (REF. 145). and the lowering of plasma triacylglycerol concentrations
Further evidence for a functional link between auto­ by ­niacin may not result from the inhibition of adipose
phagy and lipolysis came from the discovery that CMA lipo­lysis, but rather from a direct effect of niacin on hepatic
degrades cLD-associated proteins and thereby regulates ­triacylglycerol and VLDL synthesis in hepatocytes152,153.
neutral lipolysis146. In CMA, heat-shock cognate 71 kDa Another strategy for inhibiting lipolysis in WAT
protein recognizes and binds cytosolic proteins and inter­ focused on chemically synthesizing small-molecule inhib­
acts with LAMP2A and other chaperones to form a trans­ itors of HSL154,155; these compounds were developed
Fatty liver disease
A reversible condition location complex, which is transferred to lysosomes for before the discovery of ATGL. HSL inhibitors moderately
characterized by the excessive degradation (FIG. 4b). Perilipin 2 and perilipin 3, which are decreased plasma FFA concentrations in treated rats and
accumulation of neutral lipids abundant cLD-associated proteins that shield cLDs from mice. Importantly, however, HSL inhibition improved
in the liver. This condition lipases and lipolysis, are CMA targets146. Consequently, insulin sensitivity, decreased plasma glucose levels and
can be subdivided into
alcoholic fatty liver disease
the removal of perilipin 2 and perilipin 3 by CMA ­enables reduced glucose-induced insulin secretion from pan­
and non-alcoholic fatty ATGL to efficiently access the cLD surface, thereby creatic β-cells, in line with the concept that inhibition of
liver disease. increasing lipolytic rates. lipolysis may beneficially affect glucose homeostasis156.

680 | NOVEMBER 2017 | VOLUME 18 www.nature.com/nrm


©
2
0
1
7
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
.
REVIEWS

Box 3 | Lipolysis in cancer and cancer-associated cachexia


The roles of lipolysis and lipophagy in tumorigenesis and the development Another study identified CGI‑58 as a potential tumour suppressor because
of cancer-associated cachexia (CAC) have only recently attracted loss of CGI‑58 increased the propensity for tumour growth in a mouse
substantial attention. In 2010, monoacylglycerol lipase (MGL) was described intestinal cancer model188. Several, mostly ATGL-independent mechanisms
as a potential oncogene181. High MGL expression levels in cancer cells were held responsible for various CGI‑58‑dependent cancer phenotypes,
promote cell proliferation and tumour growth and strongly correlate with including the regulation of AMP-activated protein kinase (AMPK), the
malignancy. Conversely, inhibition of MGL reduces cancer cell proliferation production of cellular polyamine and inhibition of autophagy by the
and metastasis and induces cancer cell apoptosis182. The beneficial effects interaction of CGI‑58 with beclin 1 (REFS 145,187,188).
of MGL inhibition probably involve the induction of endocannabinoid Several recent studies reported a tumour suppressor function for lysosomal
signalling by inhibition of 2‑arachidonylglycerol hydrolysis and a decrease acid lipase (LAL). LAL deficiency causes tumour growth and metastasis in
in the levels of lipid mediators derived from free fatty acids (FFAs) owing to various cancer models189. LAL deficiency leads to mTOR-mediated expansion
impaired FFA production183. and organ infiltration of myeloid-derived suppressor cells (MDSCs), which
How triacylglycerol hydrolysis by adipose triglyceride lipase (ATGL) and suppress tumour immune surveillance. Re‑expression of LAL in hepatocytes
lipid droplet-binding protein CGI‑58 affects cancer cell metabolism or MDSCs reduces the abundance of MDSCs in the liver, improves hepatic
and proliferation is less clear. A robust correlation between low ATGL lipid metabolism, attenuates inflammation and decreases liver metastasis in
expression on the one hand and malignancy and reduced survival on a mouse xenotransplant melanoma model190,191. Similarly, LAL re‑expression
the other was reported for non-small-cell lung cancers, pancreatic in lung epithelial cells improved functional and metabolic parameters and
adenocarcinoma and leiomyosarcoma184. Similarly, the downregulation reduced the metastasis of lung cancer in mice192. Although still preliminary
of ATGL by epigenetic silencing of the gene encoding the E2 ligase UBCH8 and incomplete, these results highlight a previously underestimated role for
(also known as UBE2L6) correlated with poor prognosis in nasopharyngeal neutral and acid lipolysis in cancer cell metabolism and tumorigenesis.
cancer185. Decreased UBCH8 expression leads to reduced ATGL stability Lipases involved in neutral lipolysis may also have an essential role in CAC.
and activity and triacylglycerol accumulation185. In mice, ATGL deficiency Lipolysis is strongly induced in the white adipose tissue (WAT) of individuals
frequently leads to pulmonary neoplasia and adenocarcinoma184, and with CAC160–162, and deficiency of ATGL and HSL in mice partially protects
deficiency of both ATGL and hormone-sensitive lipase (HSL) causes against CAC in two xenotransplant cancer models161. CAC is accompanied
liposarcoma186, suggesting a tumour suppressor function for these lipolytic by the activation of thermogenesis in brown adipose tissue193 and increased
enzymes. The biological basis for the pro-tumorigenic activity of MGL ‘browning’ of subcutaneous WAT194. Anti-inflammatory treatment partially
and the antitumorigenic activity of ATGL is not known. reversed these metabolic aberrations, suggesting that CAC predominantly
The role of CGI‑58 in tumorigenesis also remains insufficiently resulted from systemic inflammation194. A recent study demonstrated that
understood. Overexpression of CGI‑58 in macrophages causes increased a pronounced downregulation of AMPK induced HSL and provoked CAC
tumour growth in a xenotransplant model of colorectal cancer187. in murine cancer models163.

Recently, an ATGL-specific inhibitor (atglistatin) attracted major interest (BOX 3). ATGL and HSL activity
was discovered and tested in mice157. Atglistatin treat­ is high in the WAT of cancer patients with CAC160–162 and
ment effectively lowered plasma FFA, triacylglycerol and contributes to the associated wasting process161,163. ATGL
choles­terol concentrations; improved insulin sensitivity; or HSL deficiency partially prevents CAC in murine
and attenuated hepatosteatosis in mice treated with a ­cancer models161,163. Whether pharmacological inhibition
high-fat diet158. Unexpectedly, atglistatin-treated a­ nimals of lipolysis can prevent CAC in mice is currently being
were resistant to obesity induced by a high-fat diet; investigated. The roles of autophagy and lipophagy in
showed no signs of triacylglycerol deposition in ectopic CAC are unknown.
tissues; and had normal or even improved cardiac func­
tion. The drastic reduction in liver fat, with no signs of Conclusion and future perspective
lipid accumulation, in skeletal and cardiac muscles differs Despite major progress in our understanding of how the
strikingly from observations in Atgl−/− mice, which exhibit different branches of lipolysis function, how these path­
systemic lipid accumulation10. Although the reasons for ways communicate with each other and how they affect
these contrasting results are not completely understood, the (patho)physiology of mammals, many key questions
it is conceivable that the competitive and reversible inhib­ remain unanswered. (i) The structural and functional
itory properties of atglistatin lead to oscillating inhibition characterization of the neutral lipolytic machinery
of ATGL, thereby preventing the detrimental effects of is incomplete, as the three-dimensional structures of
complete ATGL deficiency. The fact that atglistatin is enzymes and regulators are not available, and the ATGL-
not detectably absorbed by cardiac or skeletal muscle independent function(s) of CGI‑58 remain(s) elusive. (ii)
possibly contributes to the absence of muscle steatosis The extent to which lipophagy contributes to bulk triacyl­
in treated animals. Thus, the inhibition of lipolysis by glycerol catabolism in various cell types and tissues needs
ATGL and/or HSL inhibitors may serve as a powerful to be determined. (iii) The multitude of biological pro­
strategy in the prevention and treatment of metabolic cesses that are regulated by products of neutral lipolysis,
disease. In particu­lar, liver steatosis, with its tremendous lipophagy and acid lipolysis require better characteriza­
Cachexia
A wasting syndrome prevalence, represents an important target. Whether drug tion. (iv) The mechanistic link between neutral lipolysis
characterized by an treatment can reduce atherosclerotic plaque formation, as and lipophagy requires further interrogation. (v) Finally,
unintentional and nutritionally observed in Atgl−/− mice159, needs to be examined. we need to better understand the roles of neutral lipolysis,
irreversible loss of body mass Inhibitors of lipolysis may also be beneficial for lipophagy and acid lipolysis in the pathogenesis of meta­
(muscle and fat mass). Various
chronic diseases can lead to
the treatment of cancer-associated cachexia (CAC). bolic disorders and to evaluate whether the inhibition of
cachexia, but it is most In the past decade, the roles of neutral and acid lipases lipolysis can help to prevent insulin resistance and fatty
common in cancer. in tumori­genesis and the pathogenesis of CAC have liver disease in humans.

NATURE REVIEWS | MOLECULAR CELL BIOLOGY VOLUME 18 | NOVEMBER 2017 | 681


©
2
0
1
7
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
.
REVIEWS

1. Unger, R. H. Lipotoxic diseases. Annu. Rev. Med. 53, 26. Notari, L. et al. Identification of a lipase-linked cell 52. Chakrabarti, P. et al. Insulin inhibits lipolysis in
319–336 (2002). membrane receptor for pigment epithelium-derived adipocytes via the evolutionarily conserved
2. Armand, M. Lipases and lipolysis in the human factor. J. Biol. Chem. 281, 38022–38037 (2006). mTORC1‑Egr1‑ATGL-mediated pathway. Mol. Cell.
digestive tract: where do we stand? Curr. Opin. Clin. 27. Chung, C. et al. Anti-angiogenic pigment epithelium- Biol. 33, 3659–3666 (2013).
Nutr. Metab. Care 10, 156–164 (2007). derived factor regulates hepatocyte triglyceride 53. Kumar, A. et al. Fat cell-specific ablation of rictor
3. Young, S. G. & Zechner, R. Biochemistry and content through adipose triglyceride lipase (ATGL). in mice impairs insulin-regulated fat cell and whole-
pathophysiology of intravascular and intracellular J. Hepatol. 48, 471–478 (2008). body glucose and lipid metabolism. Diabetes 59,
lipolysis. Genes Dev. 27, 459–484 (2013). 28. Borg, M. L. et al. Pigment epithelium-derived factor 1397–1406 (2010).
4. Zechner, R. FAT FLUX: enzymes, regulators, and regulates lipid metabolism via adipose triglyceride 54. Gaidhu, M. P. et al. Prolonged AICAR-induced
pathophysiology of intracellular lipolysis. EMBO Mol. lipase. Diabetes 60, 1458–1466 (2011). AMP‑kinase activation promotes energy dissipation
Med. 7, 359–362 (2015). 29. Singh, M. et al. Fat-specific protein 27 inhibits lipolysis in white adipocytes: novel mechanisms integrating
5. Dubland, J. A. & Francis, G. A. Lysosomal acid lipase: by facilitating the inhibitory effect of transcription HSL and ATGL. J. Lipid Res. 50, 704–715 (2009).
at the crossroads of normal and atherogenic factor Egr1 on transcription of adipose triglyceride 55. Kim, S. J. et al. AMPK phosphorylates desnutrin/ATGL
cholesterol metabolism. Front. Cell Dev. Biol. 3, lipase. J. Biol. Chem. 289, 14481–14487 (2014). and hormone-sensitive lipase to regulate lipolysis
3 (2015). 30. Grahn, T. H. et al. Fat-specific protein 27 (FSP27) and fatty acid oxidation within adipose tissue.
6. Singh, R. et al. Autophagy regulates lipid metabolism. interacts with adipose triglyceride lipase (ATGL) to Mol. Cell. Biol. 36, 1961–1976 (2016).
Nature 458, 1131–1135 (2009). regulate lipolysis and insulin sensitivity in human 56. Watt, M. J. et al. Regulation of HSL serine
7. Zimmermann, R. et al. Fat mobilization in adipose adipocytes. J. Biol. Chem. 289, 12029–12039 phosphorylation in skeletal muscle and adipose tissue.
tissue is promoted by adipose triglyceride lipase. (2014). Am. J. Physiol. Endocrinol. Metab. 290, E500–E508
Science 306, 1383–1386 (2004). 31. Eichmann, T. O. et al. Studies on the substrate (2006).
8. Villena, J. A., Roy, S., Sarkadi-Nagy, E., Kim, K. H. and stereo/regioselectivity of adipose triglyceride 57. Ahmadian, M. et al. Desnutrin/ATGL is regulated by
& Sul, H. S. Desnutrin, an adipocyte gene encoding lipase, hormone-sensitive lipase, and diacylglycerol- AMPK and is required for a brown adipose phenotype.
a novel patatin domain-containing protein, is induced O‑acyltransferases. J. Biol. Chem. 287, 41446–41457 Cell Metab. 13, 739–748 (2011).
by fasting and glucocorticoids: ectopic expression (2012). 58. Pagnon, J. et al. Identification and functional
of desnutrin increases triglyceride hydrolysis. J. Biol. 32. Rodriguez, J. A. et al. In vitro stereoselective characterization of protein kinase A phosphorylation
Chem. 279, 47066–47075 (2004). hydrolysis of diacylglycerols by hormone-sensitive sites in the major lipolytic protein, adipose
9. Jenkins, C. M. et al. Identification, cloning, lipase. Biochim. Biophys. Acta 1801, 77–83 (2010). triglyceride lipase. Endocrinology 153, 4278–4289
expression, and purification of three novel human 33. Lafontan, M. & Langin, D. Lipolysis and lipid (2012).
calcium-independent phospholipase A2 family mobilization in human adipose tissue. Prog. Lipid Res. 59. Karlsson, M. et al. Exon-intron organization and
members possessing triacylglycerol lipase and 48, 275–297 (2009). chromosomal localization of the mouse monoglyceride
acylglycerol transacylase activities. J. Biol. Chem. 279, 34. Fischer, J. et al. The gene encoding adipose lipase gene. Gene 272, 11–18 (2001).
48968–48975 (2004). triglyceride lipase (PNPLA2) is mutated in neutral lipid 60. Rakhshandehroo, M. et al. Comprehensive analysis
10. Haemmerle, G. et al. Defective lipolysis and altered storage disease with myopathy. Nat. Genet. 39, of PPARα-dependent regulation of hepatic lipid
energy metabolism in mice lacking adipose triglyceride 28–30 (2007). metabolism by expression profiling. PPAR Res. 2007,
lipase. Science 312, 734–737 (2006). 35. Osuga, J. et al. Targeted disruption of hormone- 26839 (2007).
11. Vaughan, M., Berger, J. E. & Steinberg, D. Hormone- sensitive lipase results in male sterility and adipocyte 61. Haemmerle, G. et al. ATGL-mediated fat catabolism
sensitive lipase and monoglyceride lipase activities in hypertrophy, but not in obesity. Proc. Natl Acad. regulates cardiac mitochondrial function via PPAR-α
adipose tissue. J. Biol. Chem. 239, 401–409 (1964). Sci. USA 97, 787–792 (2000). and PGC‑1. Nat. Med. 17, 1076–1085 (2011).
12. Lass, A. et al. Adipose triglyceride lipase-mediated 36. Albert, J. S. et al. Null mutation in hormone-sensitive 62. Zierler, K. A. et al. Functional cardiac lipolysis in mice
lipolysis of cellular fat stores is activated by CGI‑58 lipase gene and risk of type 2 diabetes. N. Engl. critically depends on comparative gene
and defective in Chanarin-Dorfman Syndrome. J. Med. 370, 2307–2315 (2014). identification‑58. J. Biol. Chem. 288, 9892–9904
Cell Metab. 3, 309–319 (2006). 37. Petrosino, S. & Di Marzo, V. FAAH and MAGL (2013).
13. Yang, X. et al. The G0/G1 switch gene 2 regulates inhibitors: therapeutic opportunities from regulating 63. Schoiswohl, G. et al. Adipose triglyceride lipase plays
adipose lipolysis through association with adipose endocannabinoid levels. Curr. Opin. Investig. Drugs a key role in the supply of the working muscle with
triglyceride lipase. Cell Metab. 11, 194–205 (2010). 11, 51–62 (2010). fatty acids. J. Lipid Res. 51, 490–499 (2010).
14. Radner, F. P. et al. Growth retardation, impaired 38. Subramanian, V. et al. Perilipin A mediates the 64. Wu, J. W. et al. Deficiency of liver adipose triglyceride
triacylglycerol catabolism, hepatic steatosis, and lethal reversible binding of CGI‑58 to lipid droplets in 3T3‑L1 lipase in mice causes progressive hepatic steatosis.
skin barrier defect in mice lacking comparative gene adipocytes. J. Biol. Chem. 279, 42062–42071 Hepatology 54, 122–132 (2011).
identification‑58 (CGI‑58). J. Biol. Chem. 285, (2004). 65. Ong, K. T., Mashek, M. T., Bu, S. Y., Greenberg, A. S.
7300–7311 (2010). 39. Granneman, J. G. & Moore, H. P. Location, location: & Mashek, D. G. Adipose triglyceride lipase is a major
15. Grond, S. et al. Skin barrier development depends protein trafficking and lipolysis in adipocytes. hepatic lipase that regulates triacylglycerol turnover
on CGI‑58 protein expression during late-stage Trends Endocrinol. Metab. 19, 3–9 (2008). and fatty acid signaling and partitioning. Hepatology
keratinocyte differentiation. J. Invest. Dermatol. 137, 40. Watt, M. J. & Steinberg, G. R. Regulation and function 53, 116–126 (2011).
403–413 (2017). of triacylglycerol lipases in cellular metabolism. 66. Obrowsky, S. et al. Adipose triglyceride lipase is
16. Grond, S. et al. PNPLA1 deficiency in mice and Biochem. J. 414, 313–325 (2008). a TG hydrolase of the small intestine and regulates
humans leads to a defect in the synthesis of 41. Collins, S. A heart-adipose tissue connection in the intestinal PPARα signaling. J. Lipid Res. 54, 425–435
omega‑O‑Acylceramides. J. Invest. Dermatol. 137, regulation of energy metabolism. Nat. Rev. Endocrinol. (2013).
394–402 (2017). 10, 157–163 (2014). 67. Chandak, P. G. et al. Efficient phagocytosis requires
17. Stone, S. J. et al. Lipopenia and skin barrier 42. Kaltenecker, D. et al. Adipocyte STAT5 deficiency triacylglycerol hydrolysis by adipose triglyceride lipase.
abnormalities in DGAT2‑deficient mice. J. Biol. Chem. promotes adiposity and impairs lipid mobilisation J. Biol. Chem. 285, 20192–20201 (2010).
279, 11767–11776 (2004). in mice. Diabetologia 60, 296–330 (2016). 68. Mottillo, E. P., Bloch, A. E., Leff, T. & Granneman, J. G.
18. Russell, L. & Forsdyke, D. R. A human putative 43. Saltiel, A. R. Insulin signaling in the control of glucose Lipolytic products activate peroxisome proliferator-
lymphocyte G0/G1 switch gene containing a CpG-rich and lipid homeostasis. Handb Exp. Pharmacol. 233, activated receptor (PPAR) α and δ in brown adipocytes
island encodes a small basic protein with the potential 51–71 (2016). to match fatty acid oxidation with supply. J. Biol.
to be phosphorylated. DNA Cell Biol. 10, 581–591 44. DiPilato, L. M. et al. The role of PDE3B Chem. 287, 25038–25048 (2012).
(1991). phosphorylation in the inhibition of lipolysis by insulin. 69. Schoiswohl, G. et al. Impact of reduced ATGL-mediated
19. Cerk, I. K. et al. A peptide derived from G0/G1 switch Mol. Cell. Biol. 35, 2752–2760 (2015). adipocyte lipolysis on obesity-associated insulin
gene 2 acts as noncompetitive inhibitor of adipose 45. Chakrabarti, P. & Kandror, K. V. FoxO1 controls resistance and inflammation in male mice.
triglyceride lipase. J. Biol. Chem. 289, 32559–32570 insulin-dependent adipose triglyceride lipase (ATGL) Endocrinology 156, 3610–3624 (2015).
(2014). expression and lipolysis in adipocytes. J. Biol. Chem. 70. Jaeger, D. et al. Fasting-induced G0/G1 switch gene 2
20. Lu, X., Yang, X. & Liu, J. Differential control of 284, 13296–13300 (2009). and FGF21 expression in the liver are under regulation
ATGL-mediated lipid droplet degradation by CGI‑58 46. Chakrabarti, P. et al. SIRT1 controls lipolysis in of adipose tissue derived fatty acids. J. Hepatol. 63,
and G0S2. Cell Cycle 9, 2719–2725 (2010). adipocytes via FOXO1‑mediated expression of ATGL. 437–445 (2015).
21. Heckmann, B. L. et al. Defective adipose lipolysis and J. Lipid Res. 52, 1693–1701 (2011). 71. Tang, T. et al. Desnutrin/ATGL activates PPARδ to
altered global energy metabolism in mice with adipose 47. Picard, F. et al. Sirt1 promotes fat mobilization in promote mitochondrial function for insulin secretion
overexpression of the lipolytic inhibitor G0/G1 switch white adipocytes by repressing PPAR-γ. Nature 429, in islet beta cells. Cell Metab. 18, 883–895 (2013).
gene 2 (G0S2). J. Biol. Chem. 289, 1905–1916 771–776 (2004). 72. Schreiber, R. et al. Hypophagia and metabolic
(2014). 48. Lasa, A. et al. Resveratrol regulates lipolysis via adaptations in mice with defective ATGL-mediated
22. Zhang, X. et al. Targeted disruption of G0/G1 switch adipose triglyceride lipase. J. Nutr. Biochem. 23, lipolysis cause resistance to HFD-induced obesity.
gene 2 enhances adipose lipolysis, alters hepatic 379–384 (2012). Proc. Natl Acad. Sci. USA 112, 13850–13855
energy balance, and alleviates high-fat diet-induced 49. Lamming, D. W. & Sabatini, D. M. A. Central role for (2015).
liver steatosis. Diabetes 63, 934–946 (2014). mTOR in lipid homeostasis. Cell Metab. 18, 465–469 73. Haemmerle, G. et al. Hormone-sensitive lipase
23. Wang, Y. et al. The G0/G1 switch gene 2 is an (2013). deficiency in mice causes diglyceride accumulation in
important regulator of hepatic triglyceride 50. Ceddia, R. B. The role of AMP-activated protein kinase adipose tissue, muscle, and testis. J. Biol. Chem. 277,
metabolism. PLoS ONE 8, e72315 (2013). in regulating white adipose tissue metabolism. 4806–4815 (2002).
24. Heier, C. et al. G0/G1 switch gene 2 regulates cardiac Mol. Cell Endocrinol. 366, 194–203 (2013). 74. Armstrong, E. H., Goswami, D., Griffin, P. R., Noy, N.
lipolysis. J. Biol. Chem. 290, 26141–26150 (2015). 51. Chakrabarti, P., English, T., Shi, J., Smas, C. M. & Ortlund, E. A. Structural basis for ligand regulation
25. Hofer, P. et al. Fatty acid-binding proteins interact & Kandror, K. V. Mammalian target of rapamycin of the fatty acid-binding protein 5, peroxisome
with comparative gene identification‑58 linking complex 1 suppresses lipolysis, stimulates lipogenesis, proliferator-activated receptor beta/delta
lipolysis with lipid ligand shuttling. J. Biol. Chem. 290, and promotes fat storage. Diabetes 59, 775–781 (FABP5‑PPARβ/δ) signaling pathway. J. Biol. Chem.
18438–18453 (2015). (2010). 289, 14941–14954 (2014).

682 | NOVEMBER 2017 | VOLUME 18 www.nature.com/nrm


©
2
0
1
7
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
.
REVIEWS

75. Tan, N. S. et al. Selective cooperation between fatty hepatosplenomegaly, and shortened life span. J. Lipid 125. Saftig, P., Beertsen, W. & Eskelinen, E. L. LAMP‑2:
acid binding proteins and peroxisome proliferator- Res. 42, 489–500 (2001). a control step for phagosome and autophagosome
activated receptors in regulating transcription. 100. Radovic, B. et al. Lysosomal acid lipase regulates VLDL maturation. Autophagy 4, 510–512 (2008).
Mol. Cell. Biol. 22, 5114–5127 (2002). synthesis and insulin sensitivity in mice. Diabetologia 126. Mijaljica, D., Prescott, M. & Devenish, R. J.
76. Wolfrum, C., Borrmann, C. M., Borchers, T. 59, 1743–1752 (2016). Microautophagy in mammalian cells: revisiting a
& Spener, F. Fatty acids and hypolipidemic drugs 101. Settembre, C. et al. TFEB controls cellular lipid 40‑year-old conundrum. Autophagy 7, 673–682
regulate peroxisome proliferator-activated receptors metabolism through a starvation-induced (2011).
α- and γ-mediated gene expression via liver fatty acid autoregulatory loop. Nat. Cell Biol. 15, 647–658 127. Cingolani, F. & Czaja, M. J. Regulation and functions
binding protein: a signaling path to the nucleus. (2013). of autophagic lipolysis. Trends Endocrinol. Metab. 27,
Proc. Natl Acad. Sci. USA 98, 2323–2328 (2001). 102. Martina, J. A. et al. The nutrient-responsive 696–705 (2016).
77. Shen, W. J., Sridhar, K., Bernlohr, D. A. & Kraemer, F. transcription factor TFE3 promotes autophagy, 128. Komatsu, M. et al. Impairment of starvation-induced
B. Interaction of rat hormone-sensitive lipase with lysosomal biogenesis, and clearance of cellular debris. and constitutive autophagy in Atg7‑deficient mice.
adipocyte lipid-binding protein. Proc. Natl Acad. Sci. Signal 7, ra9 (2014). J. Cell Biol. 169, 425–434 (2005).
Sci. USA 96, 5528–5532 (1999). 103. Raben, N. & Puertollano, R. TFEB and TFE3: linking 129. Shibata, M. et al. The MAP1‑LC3 conjugation system
78. Smith, A. J. et al. Physical association between the lysosomes to cellular adaptation to stress. Annu. Rev. is involved in lipid droplet formation. Biochem.
adipocyte fatty acid-binding protein and hormone- Cell Dev. Biol. 32, 255–278 (2016). Biophys. Res. Commun. 382, 419–423 (2009).
sensitive lipase: a fluorescence resonance energy 104. O’Rourke, E. J. & Ruvkun, G. MXL‑3 and HLH‑30 130. Conlon, D. M. et al. Inhibition of apolipoprotein B
transfer analysis. J. Biol. Chem. 279, 52399–52405 transcriptionally link lipolysis and autophagy to synthesis stimulates endoplasmic reticulum
(2004). nutrient availability. Nat. Cell Biol. 15, 668–676 autophagy that prevents steatosis. J. Clin. Invest.
79. Khan, S. A. et al. ATGL-catalyzed lipolysis regulates (2013). 126, 3852–3867 (2016).
SIRT1 to control PGC‑1γ/PPAR-α signaling. Diabetes 105. Emanuel, R. et al. Induction of lysosomal biogenesis in 131. Kwanten, W. J. et al. Hepatocellular autophagy
64, 418–426 (2015). atherosclerotic macrophages can rescue lipid-induced modulates the unfolded protein response and fasting-
80. Ng, F. & Tang, B. L. Sirtuins’ modulation of autophagy. lysosomal dysfunction and downstream sequelae. induced steatosis in mice. Am. J. Physiol. Gastrointest.
J. Cell. Physiol. 228, 2262–2270 (2013). Arterioscler. Thromb. Vasc. Biol. 34, 1942–1952 Liver Physiol. 311, G599–G609 (2016).
81. Lettieri Barbato, D., Tatulli, G., Aquilano, K. (2014). 132. Mao, Y. et al. Ghrelin attenuated lipotoxicity via
& Ciriolo, M. R. FoxO1 controls lysosomal acid lipase 106. Cox, B. E., Griffin, E. E., Ullery, J. C. & Jerome, W. G. autophagy induction and nuclear factor-κB inhibition.
in adipocytes: implication of lipophagy during nutrient Effects of cellular cholesterol loading on macrophage Cell Physiol. Biochem. 37, 563–576 (2015).
restriction and metformin treatment. Cell Death Dis. foam cell lysosome acidification. J. Lipid Res. 48, 133. Wang, Y., Singh, R., Xiang, Y. & Czaja, M. J.
4, e861 (2013). 1012–1021 (2007). Macroautophagy and chaperone-mediated autophagy
82. Wagner, G. R. & Payne, R. M. Widespread and 107. Li, W., Yuan, X. M., Olsson, A. G. & Brunk, U. T. are required for hepatocyte resistance to oxidant
enzyme-independent Nepsilon-acetylation and Uptake of oxidized LDL by macrophages results in stress. Hepatology 52, 266–277 (2010).
Nepsilon-succinylation of proteins in the chemical partial lysosomal enzyme inactivation and relocation. 134. Papackova, Z., Dankova, H., Palenickova, E.,
conditions of the mitochondrial matrix. J. Biol. Chem. Arterioscler Thromb. Vasc. Biol. 18, 177–184 Kazdova, L. & Cahova, M. Effect of short- and long-
288, 29036–29045 (2013). (1998). term high-fat feeding on autophagy flux and lysosomal
83. Weinert, B. T. et al. Acetylation dynamics and 108. Griffin, E. E., Ullery, J. C., Cox, B. E. & Jerome, W. G. activity in rat liver. Physiol. Res. 61 (Suppl. 2),
stoichiometry in Saccharomyces cerevisiae. Mol. Syst. Aggregated LDL and lipid dispersions induce lysosomal S67–S76 (2012).
Biol. 10, 716 (2014). cholesteryl ester accumulation in macrophage foam 135. Ma, D. et al. Autophagy deficiency by hepatic FIP200
84. Weinert, B. T., Moustafa, T., Iesmantavicius, V., cells. J. Lipid Res. 46, 2052–2060 (2005). deletion uncouples steatosis from liver injury in
Zechner, R. & Choudhary, C. Analysis of acetylation 109. Bowden, K. L. et al. Lysosomal acid lipase deficiency NAFLD. Mol. Endocrinol. 27, 1643–1654 (2013).
stoichiometry suggests that SIRT3 repairs nonenzymatic impairs regulation of ABCA1 gene and formation 136. Shibata, M. et al. LC3, a microtubule-associated
acetylation lesions. EMBO J. 34, 2620–2632 (2015). of high density lipoproteins in cholesteryl ester protein1A/B light chain3, is involved in cytoplasmic
85. Eisenberg, T. et al. Nucleocytosolic depletion of the storage disease. J. Biol. Chem. 286, 30624–30635 lipid droplet formation. Biochem. Biophys. Res.
energy metabolite acetyl-coenzyme a stimulates (2011). Commun. 393, 274–279 (2010).
autophagy and prolongs lifespan. Cell Metab. 19, 110. Ikonen, E. Cellular cholesterol trafficking and 137. Yang, L., Li, P., Fu, S., Calay, E. S. & Hotamisligil, G. S.
431–444 (2014). compartmentalization. Nat. Rev. Mol. Cell Biol. 9, Defective hepatic autophagy in obesity promotes ER
86. Marino, G. et al. Regulation of autophagy by cytosolic 125–138 (2008). stress and causes insulin resistance. Cell Metab. 11,
acetyl-coenzyme A. Mol. Cell 53, 710–725 (2014). 111. Li, J. & Pfeffer, S. R. Lysosomal membrane 467–478 (2010).
87. Pougovkina, O. et al. Mitochondrial protein glycoproteins bind cholesterol and contribute to 138. Baerga, R., Zhang, Y., Chen, P. H., Goldman, S.
acetylation is driven by acetyl-CoA from fatty acid lysosomal cholesterol export. eLife 5, e21635 & Jin, S. Targeted deletion of autophagy-related 5
oxidation. Hum. Mol. Genet. 23, 3513–3522 (2014). (2016). (atg5) impairs adipogenesis in a cellular model
88. Hebert, A. S. et al. Calorie restriction and SIRT3 112. Chen, F. W., Gordon, R. E. & Ioannou, Y. A. NPC1 late and in mice. Autophagy 5, 1118–1130 (2009).
trigger global reprogramming of the mitochondrial endosomes contain elevated levels of non-esterified 139. Singh, R. et al. Autophagy regulates adipose mass
protein acetylome. Mol. Cell 49, 186–199 (2013). (‘free’) fatty acids and an abnormally glycosylated form and differentiation in mice. J. Clin. Invest. 119,
89. Hirschey, M. D. et al. SIRT3 regulates mitochondrial of the NPC2 protein. Biochem. J. 390, 549–561 3329–3339 (2009).
fatty-acid oxidation by reversible enzyme (2005). 140. Zhang, Y. et al. Adipose-specific deletion of
deacetylation. Nature 464, 121–125 (2010). 113. Passeggio, J. & Liscum, L. Flux of fatty acids through autophagy‑related gene 7 (atg7) in mice reveals a
90. Perry, R. J. et al. Hepatic acetyl CoA links adipose NPC1 lysosomes. J. Biol. Chem. 280, 10333–10339 role in adipogenesis. Proc. Natl Acad. Sci. USA 106,
tissue inflammation to hepatic insulin resistance (2005). 19860–19865 (2009).
and type 2 diabetes. Cell 160, 745–758 (2015). 114. Lee, S. D. & Tontonoz, P. Liver X receptors at the 141. Guo, L. et al. Transactivation of Atg4b by C/EBPbeta
91. Sheriff, S., Du, H. & Grabowski, G. A. Characterization intersection of lipid metabolism and atherogenesis. promotes autophagy to facilitate adipogenesis.
of lysosomal acid lipase by site-directed mutagenesis Atherosclerosis 242, 29–36 (2015). Mol. Cell. Biol. 33, 3180–3190 (2013).
and heterologous expression. J. Biol. Chem. 270, 115. Folick, A. et al. Aging. Lysosomal signaling molecules 142. Takagi, A. et al. Mammalian autophagy is essential
27766–27772 (1995). regulate longevity in Caenorhabditis elegans. Science for hepatic and renal ketogenesis during starvation.
92. Warner, T. G., Dambach, L. M., Shin, J. H. 347, 83–86 (2015). Sci. Rep. 6, 18944 (2016).
& O’Brien, J. S. Purification of the lysosomal acid 116. Levine, B. & Kroemer, G. Autophagy in the 143. Martinez-Lopez, N., Athonvarangkul, D., Mishall, P.,
lipase from human liver and its role in lysosomal lipid pathogenesis of disease. Cell 132, 27–42 (2008). Sahu, S. & Singh, R. Autophagy proteins regulate
hydrolysis. J. Biol. Chem. 256, 2952–2957 (1981). 117. Mizushima, N. & Komatsu, M. Autophagy: renovation ERK phosphorylation. Nat. Commun. 4, 2799
93. Grumet, L. et al. Lysosomal acid lipase hydrolyzes of cells and tissues. Cell 147, 728–741 (2011). (2013).
retinyl ester and affects retinoid turnover. J. Biol. 118. Lamb, C. A., Yoshimori, T. & Tooze, S. A. The 144. Martinez-Lopez, N. et al. Autophagy in the CNS
Chem. 291, 17977–17987 (2016). autophagosome: origins unknown, biogenesis and periphery coordinate lipophagy and lipolysis
94. Sando, G. N. & Henke, V. L. Recognition and receptor- complex. Nat. Rev. Mol. Cell Biol. 14, 759–774 in the brown adipose tissue and liver. Cell Metab.
mediated endocytosis of the lysosomal acid lipase (2013). 23, 113–127 (2016).
secreted by cultured human fibroblasts. J. Lipid Res. 119. Efeyan, A., Comb, W. C. & Sabatini, D. M. 145. Peng, Y. et al. ABHD5 interacts with BECN1 to
23, 114–123 (1982). Nutrient‑sensing mechanisms and pathways. regulate autophagy and tumorigenesis of colon cancer
95. Haka, A. S. et al. Macrophages create an acidic Nature 517, 302–310 (2015). independent of PNPLA2. Autophagy 12, 2167–2182
extracellular hydrolytic compartment to digest 120. Feng, Y., Yao, Z. & Klionsky, D. J. How to control (2016).
aggregated lipoproteins. Mol. Biol. Cell 20, self‑digestion: transcriptional, post-transcriptional, 146. Kaushik, S. & Cuervo, A. M. Degradation of lipid
4932–4940 (2009). and post-translational regulation of autophagy. droplet-associated proteins by chaperone-mediated
96. Goldstein, J. L. & Brown, M. S. A century of Trends Cell Biol. 25, 354–363 (2015). autophagy facilitates lipolysis. Nat. Cell Biol. 17,
cholesterol and coronaries: from plaques to genes 121. Lee, J. M. et al. Nutrient-sensing nuclear receptors 759–770 (2015).
to statins. Cell 161, 161–172 (2015). coordinate autophagy. Nature 516, 112–115 (2014). 147. Rambold, A. S., Cohen, S. & Lippincott-Schwartz, J.
97. Dieckmann, M., Dietrich, M. F. & Herz, J. Lipoprotein 122. Ao, X., Zou, L. & Wu, Y. Regulation of autophagy Fatty acid trafficking in starved cells: regulation by
receptors—an evolutionarily ancient multifunctional by the Rab GTPase network. Cell Death Differ. 21, lipid droplet lipolysis, autophagy, and mitochondrial
receptor family. Biol. Chem. 391, 1341–1363 (2010). 348–358 (2014). fusion dynamics. Dev. Cell 32, 678–692 (2015).
98. Du, H., Duanmu, M., Witte, D. & Grabowski, G. A. 123. Schroeder, B. et al. The small GTPase Rab7 as 148. Fabbrini, E., Sullivan, S. & Klein, S. Obesity and
Targeted disruption of the mouse lysosomal acid lipase a central regulator of hepatocellular lipophagy. nonalcoholic fatty liver disease: biochemical,
gene: long-term survival with massive cholesteryl ester Hepatology 61, 1896–1907 (2015). metabolic, and clinical implications. Hepatology 51,
and triglyceride storage. Hum. Mol. Genet. 7, 124. Itakura, E., Kishi-Itakura, C. & Mizushima, N. 679–689 (2010).
1347–1354 (1998). The hairpin-type tail-anchored SNARE syntaxin 17 149. Samuel, V. T. & Shulman, G. I. The pathogenesis of
99. Du, H. et al. Lysosomal acid lipase-deficient mice: targets to autophagosomes for fusion with insulin resistance: integrating signaling pathways and
depletion of white and brown fat, severe endosomes/lysosomes. Cell 151, 1256–1269 (2012). substrate flux. J. Clin. Invest. 126, 12–22 (2016).

NATURE REVIEWS | MOLECULAR CELL BIOLOGY VOLUME 18 | NOVEMBER 2017 | 683


©
2
0
1
7
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
.
REVIEWS

150. Ginsberg, H. N. & Reyes-Soffer, G. Niacin: a long 168. Lefevre, C. et al. Mutations in CGI‑58, the gene and correlates with poor prognosis in
history, but a questionable future. Curr. Opin. Lipidol encoding a new protein of the esterase/lipase/ nasopharyngeal carcinoma. Oncotarget 6,
24, 475–479 (2013). thioesterase subfamily, in Chanarin-Dorfman syndrome. 41077–41091 (2015).
151. Tunaru, S. et al. PUMA‑G and HM74 are receptors Am. J. Hum. Genet. 69, 1002–1012 (2001). 186. Wu, J. W. et al. Epistatic interaction between
for nicotinic acid and mediate its anti-lipolytic effect. 169. Uchida, Y. et al. Neutral lipid storage leads to the lipase-encoding genes Pnpla2 and Lipe causes
Nat. Med. 9, 352–355 (2003). acylceramide deficiency, likely contributing to the liposarcoma in mice. PLoS Genet. 13, e1006716
152. Lauring, B. et al. Niacin lipid efficacy is independent of pathogenesis of Dorfman-Chanarin syndrome. (2017).
both the niacin receptor GPR109A and free fatty acid J. Invest. Dermatol. 130, 2497–2499 (2010). 187. Miao, H. et al. Macrophage ABHD5 promotes
suppression. Sci. Transl. Med. 4, 148ra115 (2012). 170. Zhang, J. et al. Comparative gene identification‑58 colorectal cancer growth by suppressing spermidine
153. Ganji, S. H. et al. Niacin noncompetitively inhibits (CGI‑58) promotes autophagy as a putative production by SRM. Nat. Commun. 7, 11716 (2016).
DGAT2 but not DGAT1 activity in HepG2 cells. J. Lipid lysophosphatidylglycerol acyltransferase. J. Biol. 188. Ou, J. et al. Loss of abhd5 promotes colorectal tumor
Res. 45, 1835–1845 (2004). Chem. 289, 33044–33053 (2014). development and progression by inducing aerobic
154. Claus, T. H. et al. Specific inhibition of hormone- 171. Ghosh, A. K., Ramakrishnan, G., Chandramohan, C. glycolysis and epithelial-mesenchymal transition.
sensitive lipase improves lipid profile while reducing & Rajasekharan, R. CGI‑58, the causative gene Cell Rep. 9, 1798–1811 (2014).
plasma glucose. J. Pharmacol. Exp. Ther. 315, for Chanarin-Dorfman syndrome, mediates 189. Yan, C., Zhao, T. & Du, H. Lysosomal acid lipase in
1396–1402 (2005). acylation of lysophosphatidic acid. J. Biol. Chem. 283, cancer. Oncoscience 2, 727–728 (2015).
155. Ebdrup, S., Refsgaard, H. H., Fledelius, C. 24525–24533 (2008). 190. Zhao, T., Du, H., Ding, X., Walls, K. & Yan, C.
& Jacobsen, P. Synthesis and structure-activity 172. Farhan, S. M. et al. A novel LIPE nonsense mutation Activation of mTOR pathway in myeloid-derived
relationship for a novel class of potent and selective found using exome sequencing in siblings with late- suppressor cells stimulates cancer cell proliferation
carbamate-based inhibitors of hormone selective onset familial partial lipodystrophy. Can. J. Cardiol. and metastasis in lal−/− mice. Oncogene 34,
lipase with acute in vivo antilipolytic effects. J. Med. 30, 1649–1654 (2014). 1938–1948 (2015).
Chem. 50, 5449–5456 (2007). 173. Wang, S. P. et al. The catalytic function of hormone- 191. Du, H., Zhao, T., Ding, X. & Yan, C. Hepatocyte-specific
156. Girousse, A. et al. Partial inhibition of adipose tissue sensitive lipase is essential for fertility in male mice. expression of human lysosome acid lipase corrects
lipolysis improves glucose metabolism and insulin Endocrinology 155, 3047–3053 (2014). liver inflammation and tumor metastasis in lal−/− mice.
sensitivity without alteration of fat mass. PLoS Biol. 174. Abramov, A., Schorr, S. & Wolman, M. Generalized Am. J. Pathol. 185, 2379–2389 (2015).
11, e1001485 (2013). xanthomatosis with calcified adrenals. AMA J. Dis. 192. Zhao, T., Ding, X., Du, H. & Yan, C. Lung epithelial
157. Mayer, N. et al. Development of small-molecule Child 91, 282–286 (1956). cell-specific expression of human lysosomal acid lipase
inhibitors targeting adipose triglyceride lipase. 175. Sloan, H. R. & Fredrickson, D. S. Enzyme deficiency ameliorates lung inflammation and tumor metastasis
Nat. Chem. Biol. 9, 785–787 (2013). in cholesteryl ester storage disease. J. Clin. Invest. 51, in lipa−/− mice. Am. J. Pathol. 186, 2183–2192
158. Schweiger, M. et al. Pharmacological inhibition of 1923–1926 (1972). (2016).
adipose triglyceride lipase corrects high-fat diet- 176. Burke, J. A. & Schubert, W. K. Deficient activity of acid 193. Tsoli, M. et al. Activation of thermogenesis in brown
induced insulin resistance and hepatosteatosis in mice. lipase in cholesterol-ester storage disease. J. Lab Clin. adipose tissue and dysregulated lipid metabolism
Nat. Commun. 8, 14859 (2017). Med. 78, 988–989 (1971). associated with cancer cachexia in mice. Cancer Res.
159. Lammers, B. et al. Macrophage adipose triglyceride 177. Patrick, A. D. & Lake, B. D. Deficiency of an acid 72, 4372–4382 (2012).
lipase deficiency attenuates atherosclerotic lesion lipase in Wolman’s disease. Nature 222, 1067–1068 194. Petruzzelli, M. et al. A switch from white to brown fat
development in low-density lipoprotein receptor (1969). increases energy expenditure in cancer-associated
knockout mice. Arterioscler Thromb. Vasc. Biol. 31, 178. Bernstein, D. L., Hulkova, H., Bialer, M. G. cachexia. Cell Metab. 20, 433–447 (2014).
67–73 (2011). & Desnick, R. J. Cholesteryl ester storage disease:
160. Agustsson, T. et al. Mechanism of increased lipolysis in review of the findings in 135 reported patients with an Acknowledgements
cancer cachexia. Cancer Res. 67, 5531–5537 (2007). underdiagnosed disease. J. Hepatol. 58, 1230–1243 Financial support was provided by the European Research
161. Das, S. K. et al. Adipose triglyceride lipase (2013). Council ERC Grant Agreement 340896, LipoCheX (R.Z.), the
contributes to cancer-associated cachexia. Science 179. Hoffman, E. P., Barr, M. L., Giovanni, M. A. DKs Molecular Enzymology (W901) (R.Z.) and Metabolic and
333, 233–238 (2011). & Murray, M. F. Lysosomal Acid Lipase Deficiency Cardiovascular Disease (W1226) (F.M., D.K.), the SFB
162. Ryden, M. et al. Lipolysis—not inflammation, cell (eds Pagon, R. A. et al.) (University of Washington, Lipotox (F30) funded by the Austrian Science Fund (FWF)
death, or lipogenesis—is involved in adipose tissue loss 2016). (R.Z., F.M., D.K.), the Louis-Jeantet Foundation (R.Z.), the
in cancer cachexia. Cancer 113, 1695–1704 (2008). 180. Burton, B. K. et al. A phase 3 trial of sebelipase alfa Fondation Leducq (grant 12CVD04) (R.Z.) and the
163. Rohm, M. et al. An AMP-activated protein kinase- in lysosomal acid lipase deficiency. N. Engl. J. Med. BioTechMed-Graz flagship projects EPIAge (F.M.) and Lipid
stabilizing peptide ameliorates adipose tissue wasting 373, 1010–1020 (2015). Signalling (D.K.). F.M. acknowledges additional support
in cancer cachexia in mice. Nat. Med. 22, 1120–1130 181. Nomura, D. K. et al. Monoacylglycerol lipase regulates from NAWI Graz; BioTechMed-Graz (“EPIAge”); FWF grants
(2016). a fatty acid network that promotes cancer P29262, P29203 and P27893; and the BMWFW and
164. Massa, R. et al. Neutral lipid-storage disease with pathogenesis. Cell 140, 49–61 (2010). University of Graz grants “Unkonventionelle Forschung”
myopathy and extended phenotype with novel 182. Zhang, J. et al. Monoacylglycerol lipase: A novel and “Flysleep”.
PNPLA2 mutation. Muscle Nerve 53, 644–648 potential therapeutic target and prognostic indicator
(2016). for hepatocellular carcinoma. Sci. Rep. 6, 35784 Author contributions
165. Natali, A. et al. Metabolic consequences of adipose (2016). Researching the literature for the article: R.Z., F.M. and D.K.;
triglyceride lipase deficiency in humans: an in vivo 183. Nomura, D. K. et al. Monoacylglycerol lipase exerts substantial contributions to discussion of the content: R.Z.,
study in patients with neutral lipid storage disease dual control over endocannabinoid and fatty acid F.M. and D.K.; writing: R.Z. and D.K.; review and/or editing of
with myopathy. J. Clin. Endocrinol. Metab. 98, pathways to support prostate cancer. Chem. Biol. 18, the manuscript before submission: R.Z., F.M. and D.K.
E1540–E1548 (2013). 846–856 (2011).
166. Dorfman, M. L., Hershko, C., Eisenberg, S. 184. Al‑Zoughbi, W. et al. Loss of adipose triglyceride lipase Competing interests statement
& Sagher, F. Ichthyosiform dermatosis with systemic is associated with human cancer and induces mouse The authors declare no competing interests.
lipidosis. Arch. Dermatol. 110, 261–266 (1974). pulmonary neoplasia. Oncotarget 7, 33832–33840
167. Chanarin, I. et al. Neutral-lipid storage disease: a new (2016). Publisher’s note
disorder of lipid metabolism. Br. Med. J. 1, 553–555 185. Zhou, X. et al. Epigenetic downregulation of the Springer Nature remains neutral with regard to jurisdictional
(1975). ISG15‑conjugating enzyme UbcH8 impairs lipolysis claims in published maps and institutional affiliations.

684 | NOVEMBER 2017 | VOLUME 18 www.nature.com/nrm


©
2
0
1
7
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