ArticlePDF AvailableLiterature Review

Abstract and Figures

Lipolysis is defined as the catabolism of triacylglycerols stored in cellular lipid droplets. Recent discoveries of essential lipolytic enzymes and characterization of numerous regulatory proteins and mechanisms have fundamentally changed our perception of lipolysis and its impact on cellular metabolism. New findings that lipolytic products and intermediates participate in cellular signaling processes and that "lipolytic signaling" is particularly important in many nonadipose tissues unveil a previously underappreciated aspect of lipolysis, which may be relevant for human disease.
Content may be subject to copyright.
Cell Metabolism
Review
FAT SIGNALS - Lipases and Lipolysis
in Lipid Metabolism and Signaling
Rudolf Zechner,
1,
*Robert Zimmermann,
1
Thomas O. Eichmann,
1
Sepp D. Kohlwein,
1
Guenter Haemmerle,
1
Achim Lass,
1
and Frank Madeo
1
1
Institute of Molecular Biosciences, University of Graz, Austria
*Correspondence: rudolf.zechner@uni-graz.at
DOI 10.1016/j.cmet.2011.12.018
Lipolysis is defined as the catabolism of triacylglycerols stored in cellular lipid droplets. Recent discoveries of
essential lipolytic enzymes and characterization of numerous regulatory proteins and mechanisms have
fundamentally changed our perception of lipolysis and its impact on cellular metabolism. New findings
that lipolytic products and intermediates participate in cellular signaling processes and that ‘‘lipolytic
signaling’’ is particularly important in many nonadipose tissues unveil a previously underappreciated aspect
of lipolysis, which may be relevant for human disease.
From Fat Ferments to Lipases: A Short Historical
Overview
Lipolysis describes the hydrolysis of triacylglycerols (TGs),
commonly referred to as fat. In the mid-19
th
century the great
French physiologist Claude Bernard (Bernard, 1856) noted that
the pancreatic juice of mammals was able to efficiently degrade
fat in the form of butter and oil. His observation led to the partial
characterization of a pancreatic fat-splitting ferment by Balser
(Balser, 1882), Langerhans (Langerhans, 1890), and Flexner
(Flexner, 1897). The importance of lipolysis to general metabo-
lism became apparent when Whitehead (Whitehead, 1909)
discovered that fat (TGs) could not enter cells in its unhydrolyzed
form. The absolute requirement for TG hydrolysis for the cellular
uptake or release of fatty acids (FAs) and glycerol defines three
processes in vertebrate physiology where lipolysis is essential:
gastrointestinal lipolysis mediates the catabolism of dietary fat;
vascular lipolysis is responsible for the hydrolysis of lipopro-
tein-associated TGs in the blood; and intracellular lipolysis cata-
lyzes the breakdown of TGs stored in intracellular lipid droplets
(LDs) for subsequent export of FAs (from adipose tissue) or their
metabolism (in nonadipose tissues).
Although fundamental aspects of lipolysis were understood
early on, it took more than a century to identify, isolate, and char-
acterize the main fat-splitting ferments, which have been called
lipases since 1900. The chief lipolytic enzymes of the gastroin-
testinal tract are lingual lipase, gastric lipase, pancreatic lipase,
and pancreatic lipase-related proteins 1, 2, and 3. Vascular TG
hydrolysis depends on lipoprotein lipase (LPL) and hepatic TG
lipase. Pancreatic and vascular lipases are structurally related
and prototypic for the pancreatic lipase family. Intracellular
lipolysis of TGs involves neutral (pH-optimum around pH 7)
and acid lipases present in lysosomes (pH optimum between
pH 4–5). Well-characterized neutral TG hydrolases include
adipose triglyceride lipase (ATGL) and hormone-sensitive lipase
(HSL), whereas lysosomal acid lipase (LAL) is the most important
lipase in lysosomes. Besides an active serine site, these
enzymes share no obvious structural homologies and are unre-
lated to the pancreatic lipase family.
Lipolysis is the catabolic branch of the FA cycle that provides
FAs in times of metabolic need and removes them when they are
present in excess. FAs are essential as substrates for energy
production and the synthesis of most lipids, including membrane
lipids and lipids involved in cellular signaling. Accordingly, all uni-
and multicellular organisms are able to synthesize FAs de novo
(endogenous FAs) from carbohydrate and/or protein metabo-
lites. Despite their fundamental physiological importance, an
oversupply of FAs is highly detrimental. Increased concentra-
tions of nonesterified FAs disrupt the integrity of biological
membranes, alter cellular acid-base homeostasis, and elicit the
generation of harmful bioactive lipids. These effects, in turn,
impair membrane function and induce endoplasmic reticulum
(ER) stress, mitochondrial dysfunction, inflammation, and cell
death. Collectively, these deleterious effects are subsumed
under the term lipotoxicity (Unger et al., 2010). As a countermea-
sure, essentially all cells are able to detoxify nonesterified FAs by
esterification with glycerol to yield inert TGs. Additionally, higher
organisms store FAs in a specialized organ (i.e., adipose tissue),
which supplies FAs to other high-demand tissues, such as
liver and muscle (exogenous FAs). The carefully regulated
balance of FA esterification and TG hydrolysis creates an effi-
cient buffer system, allowing sufficient FA flux without nonphy-
siological increases in cellular nonesterified FA concentrations.
Moreover, FA cycling creates metabolic intermediates that can
be utilized in anabolic processes or as extra- or intracellular
signaling molecules.
This review focuses on the catabolic branch of the FA cycle.
We summarize recent advances in understanding the enzymatic
mechanisms of the lipolytic process and the (patho)physiological
impact of lipolysis on energy homeostasis and cellular signaling.
Neutral Lipolysis: Three Steps, Three Enzymes
Neutral hydrolysis of TGs to FAs and glycerol requires three
consecutive steps that involve at least three different enzymes:
ATGL catalyzes the initial step of lipolysis, converting TGs to
diacylglycerols (DGs); HSL is mainly responsible for the hydro-
lysis of DGs to monoacylglycerols (MGs) and MG lipase (MGL)
hydrolyzes MGs. In adipose tissue, ATGL and HSL are respon-
sible for more than 90% of TG hydrolysis (Schweiger et al.,
2006). Although most nonadipose tissues also express ATGL
and HSL, expression levels are low in some tissues, raising the
Cell Metabolism 15, March 7, 2012 ª2012 Elsevier Inc. 279
question of whether other lipases are additionally required for
efficient lipolysis.
ATGL: the Initial Step of Lipolysis
ATGL is the newest member of the lipolytic enzyme trio. The
enzyme, first described in 2004 (Jenkins et al., 2004; Villena
et al., 2004; Zimmermann et al., 2004), belongs to the family
patatin domain-containing proteins including nine human and
eight murine members. The patatin domain was originally
discovered in lipid hydrolases of the potato and other plants
and named after the most abundant protein of the potato tuber,
patatin. Because some members of the family act as phospholi-
pases, the proteins were officially named patatin-like phospholi-
pase domain-containing protein A1 to A9 (PNPLA1–9) (Wilson
et al., 2006). Unfortunately, this name is misleading because it
implies that all members of this family are phospholipases.
However, several PNPLA proteins have only minor or no phos-
pholipase activity. A more appropriate designation for the family
should be considered.
ATGL (officially annotated as PNPLA2) preferentially hydro-
lyzes TGs. Orthologous enzymes are found in essentially all
eukaryotic species, including vertebrates, invertebrates, plants,
and fungi. The human protein comprises 504 amino acids (AA).
In analogy with patatin, the active site of the enzyme contains
an unusual catalytic dyad (S47 and D166) within the patatin
domain (Rydel et al., 2003). This domain comprises 180 AA
and is embedded within a 250 AA abasandwich structure
at the protein’s NH
2
-terminal half. The COOH-terminal half has
a predominantly regulatory function and contains a predicted
hydrophobic region for LD binding (Duncan et al., 2010;
Schweiger et al., 2008). Loss of this region increases the specific
in vitro activity of ATGL against artificial TG substrates but blunts
the intracellular activity due to the inability of the truncated
enzyme to bind to cellular LDs.
Regulation of ATGL
Regulation of ATGL expression and enzyme activity is complex
(Lass et al., 2011). ATGL mRNA expression is elevated by perox-
isome proliferator-activated receptor (PPAR) agonists, glucocor-
ticoids, and fasting, whereas insulin and food intake decrease
expression. More recently, it was shown that mTOR complex
1-dependent signaling reduces ATGL mRNA levels (Chakrabarti
et al., 2010). Conversely, activation of FoxO1 by SIRT1-mediated
deacetylation activates lipolysis by increasing ATGL mRNA
levels. SIRT1 silencing has the opposite effect (Chakrabarti
et al., 2011). The abundance of ATGL (and HSL) mRNA does
not always correlate with cellular lipase activity. For example,
isoproterenol and tumor necrosis factor-areduce ATGL (and
HSL) mRNA levels in adipocytes (Kralisch et al., 2005) but,
conversely, stimulate lipase activities and FA and glycerol
release. The discrepancy between enzyme mRNA levels and
activities is explained by the extensive posttranslational regula-
tion of ATGL and HSL (discussed below). Thus, cellular lipase
mRNA levels are inadequate as indicators of enzyme activities.
At least two serine residues in ATGL can be phosphorylated
(S406 and S430 in the murine enzyme) (Bartz et al., 2007). In
contrast to HSL (see below), ATGL modification is not protein
kinase A (PKA)-dependent (Zimmermann et al., 2004). Instead,
the laboratory of H.-S. Sul recently showed that AMP-activated
kinase (AMPK) phosphorylates S406, leading to increased
hydrolytic activity of murine ATGL (Ahmadian et al., 2011). The
role of AMPK in the regulation of lipolysis has been controversial,
with data showing that AMPK induces (Gaidhu et al., 2009; Yin
et al., 2003), inhibits (Daval et al., 2005; Gauthier et al., 2008),
or has no effect (Chakrabarti et al., 2011) on lipolysis. Interest-
ingly, in the nematode Caenorhabiditis (C.) elegans, AMPK-
mediated phosphorylation of ATGL-1 (the worm ortholog of
mammalian ATGL) at different phosphorylation sites inhibits TG
hydrolysis (Narbonne and Roy, 2009), delays the consumption
of TG stores, and prolongs life span during the stress-induced
dauer larval stage. The functional role of the second phosphory-
lation site (S430) is not known.
ATGL requires a coactivator protein, comparative gene
identification-58 (CGI-58), for full hydrolase activity (Lass et al.,
2006). CGI-58 was originally discovered in a screen comparing
the proteomes of humans and C. elegans. The gene is now
officially named a/bhydrolase domain-containing protein-5
(ABHD5), owing to the presence of an a/bhydrolase domain
commonly found in esterases, thioesterases, and lipases. How-
ever, CGI-58 is unlikely to exhibit hydrolase activity due to the
fact that asparagine-153 (N153) in CGI-58 replaces a nucleophilic
serine residue that is required in the active site of enzymatically
functional members of the esterase/thioesterase/lipase family.
Interestingly, substitution of N153 by serine does not convert
CGI-58 into a lipid hydrolase for TGs, DGs, MGs, or short
FA-glycerol esters (A. Lass et al., unpublished). In the epidermis
of the skin, CGI-58 may also stimulate another, currently
unknown, TG hydrolase distinct from ATGL (Radner et al.,
2010). Importantly, two laboratories showed that CGI-58, at least
in vitro, exhibits enzymatic activity as an acylCoA-dependent
acylglycerol-3-phosphate acyltransferase (AGPAT) (Ghosh
et al., 2008; Montero-Moran et al., 2010). The physiological role
of this activity requires clarification. It may affect phosphatidic
acid or lysophosphatidic acid signaling (Brown et al., 2010).
Recently, Yang et al. (Yang et al., 2010b) discovered a specific
peptide inhibitor for ATGL. The protein was originally identified in
blood mononuclear cells to act at the G0 to G1 transition of the
cell cycle. Consistent with this function, it was named G0G1
switch protein 2 (G0S2). Human and murine G0S2 have a pre-
dicted primary structure of 103 AA. The protein is found in
many tissues, with highest concentrations in adipose tissue
and liver. Consistent with lipase activities, G0S2 expression is
very low in adipose tissue during fasting but increases after
feeding (Yang et al., 2010b). Conversely, fasting or PPARa-agon-
ists increase hepatic G0S2 expression (Zandbergen et al.,
2005). The protein localizes to multiple cellular compartments,
including LDs, cytoplasm, ER, and mitochondria. The different
localizations may reflect multiple functions for G0S2 in regulating
lipolysis, the cell cycle, and, possibly, apoptosis via its ability
to interact with the mitochondrial antiapoptotic factor Bcl2
(Welch et al., 2009). In vitro, LD binding and inhibition of ATGL
depend on a physical interaction between the NH
2
-terminal
region of G0S2 (involving AA 27–42) and the patatin domain of
ATGL (Lu et al., 2010). Elucidation of whether G0S2 also regu-
lates tissue-specific lipolysis in vivo will require the characteriza-
tion of G0S2 transgenic and knockout mouse models.
LD-associated proteins participate in the CGI-58-mediated
regulation of ATGL (Figure 1, left). In hormonally nonstimulated
white and brown adipocytes, perilipin-1 interacts with CGI-58,
preventing its binding to and, thus, induction of ATGL. Upon
280 Cell Metabolism 15, March 7, 2012 ª2012 Elsevier Inc.
Cell Metabolism
Review
b-adrenergic stimulation, protein kinase A (PKA) phosphorylates
perilipin-1 at multiple sites, including the critical serine-517
residue, causing the release of CGI-58. The effector then binds
and stimulates ATGL (Granneman et al., 2009; Miyoshi et al.,
2007). Thus, b-adrenergic stimulation of PKA induces ATGL
activity by phosphorylation of perilipin-1 and not by direct
enzyme phosphorylation. Consistent with this model, frameshift
mutants of perilipin-1 in humans (L404fs and V398fs) fail to bind
CGI-58, leading to unrestrained lipolysis, partial lipodystrophy,
hypertriglyceridemia, and insulin resistance (Gandotra et al.,
2011). ATGL-mediated TG hydrolysis in nonadipose tissues
with high FA oxidation rates, such as muscle and liver, follows
another mechanism. In these tissues, perilipin-1 is replaced by
perilipin-5 (Figure 1, right). During fasting, perilipin-5 recruits
both ATGL and CGI-58 to LDs by direct binding of the enzyme
and its coactivator. Formation of the ternary complex involves
the COOH-terminal region of perilipin-5 (AA 200–463) (Granne-
man et al., 2011). The role of perilipin-5 within this complex is still
a matter of discussion. Recent data suggest that it is involved in
the interaction of LDs with mitochondria and inhibits ATGL-
mediated TG hydrolysis (Wang et al., 2011). Whether perilipins-2,
-3, and -4 also interact with either ATGL or CGI-58 is disputed.
Overexpression of perilipin-2 in hepatocyte cell lines inhibits
ATGL activity by restricting its physical access to LDs. Direct
protein-protein interaction is probably not required (Bell et al.,
2008; Listenberger et al., 2007).
Recently, several groups reported that pigment epithelium-
derived factor (PEDF) induces TG hydrolysis in adipose tissue,
muscle, and liver, via ATGL (Borg et al., 2011; Chung et al.,
2008). PEDF is a widely expressed 50 kD protein of the noninhi-
bitory serpin family of serine protease inhibitors (Filleur et al.,
2009). It exhibits a large spectrum of bioactivities, including anti-
angiogenic, antitumorigenic, neuroprotective, antioxidative, and
antiinflammatory effects. PEDF binds to ATGL and activates its
enzymatic activity (Borg et al., 2011; Notari et al., 2006). Although
the mechanism remains to be clarified, ATGL activation by PEDF
may be involved in the pathogenesis of insulin resistance and the
development of hepatosteatosis.
Another important aspect of ATGL regulation was identified by
genetic analyses of LD formation in Drosophila (D.) melanogaster
L2 cells (Beller et al., 2008; Guo et al., 2008). Compelling results
showed that ATGL delivery to LDs requires functional vesicular
transport. In the absence of essential protein components of
the transport machinery, such as ADP-ribosylation factor 1
(ARF1), small GTP-binding protein 1 (SAR1), the guanine-nucle-
otide exchange factor Golgi-Brefeldin A resistance factor
(GBF1), or deficiency of the coatamer protein coat-complex I
and II, ATGL translocation to LDs is blocked and the enzyme
remains associated with the ER (Soni et al., 2009). The process
requires physical binding of ATGL to GBF1 (Ellong et al., 2011).
HSL: the Main DG Lipase
In the early 1960s, it was noted that a lipolytic activity present in
adipose tissue was induced by hormonal stimulation. A land-
mark paper published by D. Steinberg’s group (Vaughan et al.,
1964) described the isolation and characterization of both HSL
and MGL. Although this classic work originally noted that HSL
is a much better DG hydrolase than TG hydrolase, standard
textbook knowledge perpetuated the conclusion that HSL was
rate-limiting for the catabolism of fat stores in adipose and
many nonadipose tissues. This view required revision when
HSL-deficient mice efficiently hydrolyzed TGs (Osuga et al.,
2000). HSL-deficient mice showed no signs of TG accumulation
in either adipose or nonadipose tissues; instead, they accumu-
alted large amounts of DGs in many tissues, suggesting that
in vivo the enzyme was more important as a DG- than a TG-
hydrolase (Haemmerle et al., 2002). Although originally some-
what controversial (Ryde
´n et al., 2007), it is now accepted that
ATGL is responsible for the initial step of lipolysis in human
adipocytes, and that HSL is rate-limiting for the catabolism of
DGs (Bezaire et al., 2009). In addition to DGs, HSL also hydro-
lyzes ester bonds of many other lipids (e.g., TGs, MGs, choles-
teryl esters, and retinyl esters) and short-chain carbonic acid
esters (Fredrikson et al., 1986).
The HSL expression profile essentially mirrors that of ATGL.
Highest mRNA and protein concentrations are found in white
adipose tissue (WAT) and brown adipose tissue (BAT); low
expression is detected in many other tissues, including muscle,
testis, steroidogenic tissues, and pancreatic islets (Holm et al.,
2000). Alternative exon usage leads to tissue-specific differ-
ences in mRNA and protein size (Holst et al., 1996). In adipose
tissue, the HSL protein comprises 768 AA. Unlike ATGL, with
orthologous enzymes found across all eukarya, HSL is less ubiq-
uitous phylogenetically. For example, no HSL ortholog is known
in birds, C. elegans,D. melanogaster, and Saccharomyces (S.)
cerevisiae. Interestingly, the closest structural relatives to HSL
are found in prokaryotes (e.g., lipase 2 in Moraxella TA144)
(Langin et al., 1993). Functional studies have delineated in HSL
an NH
2
-terminal lipid-binding region, the a/bhydrolase fold
Figure 1. Lipolysis in Adipose and Oxidative Tissues during Fasting
In adipose tissues, beta-adrenergic stimulation of lipolysis leads to the
consecutive hydrolysis of TG and the formation of FAs and glycerol. The
process requires three enzymes: ATGL cleaves the first esterbond in TGs, HSL
hydrolyzes DGs, and MGL MGs. For full hydrolytic activity, ATGL interacts with
its coactivator protein CGI-58, whereas HSL is phosphorylated, translocates
to the LD, and interacts with phosphorylated PLIN-1. Expression of the ATGL
inhibitor G0S2 during fasting is low in adipose and high in oxidative tissues
(e.g., liver). In oxidative tissues PLIN-1 is not present on LDs. Instead, PLIN-5 is
expressed and interacts with both ATGL and CGI-58, facilitating LD localiza-
tion of these proteins. ATGL, adipose triglyceride lipase; CGI-58, comparative
gene identification-58; DG, diacylglycerol; FA, fatty acid; G, glycerol; G0S2,
G0/G1 switch gene 2; HSL, hormone-sensitive lipase; MG, monoacylglycerol;
MGL, monoglyceride lipase; PLIN-1, perilipin-1; PLIN-5, perilipin-5; TG, tri-
acylglycerol.
Cell Metabolism 15, March 7, 2012 ª2012 Elsevier Inc. 281
Cell Metabolism
Review
domain including the catalytic triad, and the regulatory module
containing all known phosphorylation sites important for regula-
tion of enzyme activity (Holm et al., 2000).
Regulation of HSL
Since ATGL and HSL hydrolyze TGs in a coordinated manner, it is
not unexpected that they share many regulatory similarities. In
adipose tissue, HSL enzyme activity is strongly induced by
b-adrenergic stimulation, whereas insulin has a strong inhibitory
effect. The mechanisms of enzyme regulation, however, differ
markedly between the two lipases. While b-adrenergic stimula-
tion regulates ATGL primarily via recruitment of the coactivator
CGI-58 (see above), HSL is a major target for PKA-catalyzed
phosphorylation (Stra
˚lfors and Belfrage, 1983). Other kinases,
including AMPK, extracellular signal-regulated kinase, glycogen
synthase kinase-4, and Ca
2+
/calmodulin-dependent kinase,
also phosphorylate HSL to modulate its enzyme activity (Lass
et al., 2011). The enzyme has at least five potential phosphoryla-
tion sites, of which S660 and S663 appear to be particularly
important for hydrolytic activity (Anthonsen et al., 1998). Enzyme
phosphorylation affects enzyme activity moderately (an approxi-
mate 2-fold induction). For full activation, HSL must gain access
to LDs, which, in adipose tissue, is mediated by perilipin-1. Simul-
taneously with HSL, PKA also phosphorylates perilipin-1 on six
consensus serine residues. As a result, HSL binds to the NH
2
-
terminal region of perilipin-1, thereby gaining access to LDs
(Miyoshi et al., 2007; Shen et al., 2009; Wang et al., 2009).
Together, HSL-phosphorylation and enzyme translocation to
LDs coupled with ATGL activation by CGI-58 result in a more
than 100-fold increase in TG hydrolysis in adipocytes (Figure 1).
This activation process is modulated by other factors. For
example, receptor-interacting protein 140 (RIP-140) was shown
to induce lipolysis by binding to perilipin-1, increasing HSL
translocation to LDs, and activating ATGL via CGI-58 dissociation
from perilipin-1 (Ho et al., 2011). In nonadipose tissues, such as
skeletal muscle, HSL is activated by phosphorylation in response
to adrenaline and muscle contraction (Watt et al., 2006). These
tissues lack perilipin-1, and it remains to be determined which
alternative mechanisms regulate HSL access to LDs.
Insulin-mediated deactivation of lipolysis is associated with
transcriptional downregulation of ATGL and HSL expression
(Kershaw et al., 2006; Kralisch et al., 2005). Additionally, insulin
signaling results in phosphorylation and activation of various
phosphodiesterase (PDE) isoforms by PKB/AKT (Enoksson
et al., 1998), PDE-catalyzed hydrolysis of cAMP, and inhibition
of PKA. These actions halt lipolysis by preventing phosphoryla-
tion of both HSL and perilipin-1, activation and translocation
of HSL, and activation of ATGL by CGI-58. In addition to its
peripheral action, insulin also acts centrally via the sympathetic
nervous system to inhibit lipolysis in WAT. Elegant studies by
Scherer and coworkers (Scherer et al., 2011) showed that
increased insulin levels in the brain inhibit HSL and perilipin
phosphorylation, leading to reduced HSL and ATGL activities.
MGL: the Final Step in Lipolysis
MGL is considered to be the rate-limiting enzyme for the break-
down of MGs derived from extracellular TG hydrolysis (by LPL),
intracellular TG hydrolysis (by ATGL and HSL), and intracellular
phospholipid hydrolysis (by phospholipase C and membrane-
associated DG lipase aand b). The enzyme localizes to cell
membranes, cytoplasma (Sakurada and Noma, 1981), and LDs
(unpublished data). MGL received significant attention following
the realization that glycerophospholipid-derived MG 2-arachido-
nylglycerol (2-AG) is a major agonist for endocannabinoid
signaling and is inactivated by the hydrolytic activity of MGL
(see MG-signaling, below). The enzyme is ubiquitously ex-
pressed with highest expression levels in adipose tissue. The
importance of MGL for efficient degradation of MGs was recently
confirmed in mutant mouse models (Chanda et al., 2010; Schlos-
burg et al., 2010; Taschler et al., 2011). Lack of MGL impairs
lipolysis and is associated with increased MG levels in adipose
and nonadipose tissues alike.
MGL shares homology with esterases, lysophospholipases,
and haloperoxidases, and contains a consensus GXSXG motif
within a catalytic triad (S122, A239, and H269 for mouse MGL)
that is typical of lipases and esterases. Very recently, the crystal
structure of MGL was solved (Bertrand et al., 2010; Labar et al.,
2010). The enzyme exhibits the classic fold of the a/bhydrolases,
crystallizes as a dimer, and exhibits a wide, hydrophobic access
to the catalytic site. An apolar helix-domain lid covers the active
site and mediates the interaction of MGL with membrane struc-
tures and the recruitment of substrate.
Other Lipases Implicated in TG Catabolism
Experiments with ATGL-deficient mice and small-molecule HSL
inhibitors revealed that ATGL and HSL are responsible for more
than 90% of the lipolytic activity in WAT and cultured adipocytes
(Schweiger et al., 2006). In nonadipose tissues, the contribution
of other neutral lipases to the catabolism of stored TGs may be
more prominent. For example, in the liver of fasted mice, ATGL
accounts for less than 50% of neutral TG hydrolase activity
(Reid et al., 2008). This activity is physiologically important
because ATGL-deficient mice develop hepatosteatosis (Haem-
merle et al., 2006; Wu et al., 2011). However, the pronounced
remaining activity of hepatic TG hydrolase(s) in ATGL-deficient
mice indicates that other lipases contribute to the highly dynamic
turnover of TGs. This view is supported by the observation that
mice lacking ATGL in the liver have no apparent defect in VLDL
biogenesis (Wu et al., 2011), although assembly and secretion
of hepatic VLDL particles require substantial mobilization of
hepatic TG stores. Because HSL is also poorly expressed in
hepatocytes, the existence of alternative hepatic DG hydrolases
seems likely.
Several members of the carboxylesterase/lipase family and
the PNPLA family have been suggested as potential TG hydro-
lases. One of them, carboxyl esterase-3/triglyceride hydrolase-1
(Ces-3/Tgh-1, ortholog of human Ces-1), has gained major
interest because the recent characterization of Ces-3/Tgh-1-
deficient mice provided compelling evidence that the enzyme
participates in the assembly and secretion of hepatic VLDL
(Wei et al., 2010). How this biological function conforms to
the strict luminal localization of Tgh-1 in the ER remains to be
elucidated.
Structural relatives of ATGL within the PNPLA family were also
considered as potential TG hydrolases. For example, PNPLA4
and -5 exhibit TG-hydrolase, DG transacylase, and retinylester
hydrolase activity in vitro (Kienesberger et al., 2009b). Whether
these activities are also relevant in vivo remains to be deter-
mined. The member with highest homology to ATGL is adiponu-
trin (PNPLA3), with over 50% AA identity within the patatin
domain. Adiponutrin was originally discovered as a nutritionally
282 Cell Metabolism 15, March 7, 2012 ª2012 Elsevier Inc.
Cell Metabolism
Review
regulated adipose-specific transcript of unknown function
(Baulande et al., 2001). Interest in adiponutrin increased tremen-
dously when H. Hobbs and colleagues found a strong genetic
association between a nonsynonymous AA change (I148M) in
adiponutrin and susceptibility to develop nonalcoholic fatty liver
disease (NAFLD) (Romeo et al., 2008). Several other groups
confirmed and extended this important finding by showing
robust associations of I148M with alcoholic- and nonalcoholic
liver disease, hepatic fibrosis, and liver cirrhosis (Krawczyk
et al., 2011; Tian et al., 2010; Yuan et al., 2008). The close simi-
larity to ATGL as well as the presence of conserved structural
motifs typical for lipases/esterases (abasandwich structure
and the GXSXG motif within a catalytic dyad) suggest a lipase
function for adiponutrin. In accordance with this assumption,
several groups reported that adiponutrin acts as a TG hydrolase
and additionally exhibits DG transacylase activity (He et al.,
2010; Huang et al., 2011; Jenkins et al., 2004; Lake et al.,
2005). However, the expression profile generated in response
to fasting/feeding and the induction of adiponutrin gene expres-
sion by SREBP1a/c and CHREBP argued against an in vivo role
of adiponutrin in lipolysis (Dubuquoy et al., 2011; He et al., 2010;
Kershaw et al., 2006; Lake et al., 2005). Pinpointing the functional
role of adiponutrin was also confounded when PNPLA3-deficient
mice exhibited no detectable phenotype in lipid, lipoprotein, or
energy metabolism (Basantani et al., 2011; Chen et al., 2010).
Overexpression of the I148M variant of adiponutrin caused TG
accumulation (He et al., 2010), whereas overexpression of
wild-type adiponutrin in the liver created no obvious phenotype
(He et al., 2010). Overall, the biochemical and (patho)physiolog-
ical function of adiponutrin remains unclear.
Autophagy and Acid Lipolysis
In addition to classical lipolysis by extralysosomal neutral
lipases, TGs and cholesteryl esters can also be hydrolyzed by
LAL. LAL is thought to catabolize primarily lipoprotein-associ-
ated lipids subsequent to their receptor-mediated endocytosis
and fusion with lysosomes. Accordingly, the contribution of
LAL activity to lipolysis of intracellular LDs was not considered
relevant. Given the lysosomal localization of LAL, it was not intu-
itively obvious how lipids from LDs would enter lysosomes.
Addressing this, Singh and colleagues reported compelling
evidence linking lipolysis to macroautophagy (Singh et al.,
2009a). Macroautophagy is a lysosomal pathway that degrades
superfluous or damaged organelles as well as cytoplasmic inclu-
sions, such as misfolded protein aggregates (Levine and
Kroemer, 2008). These cytoplasmic cargos are trapped inside
double-membrane vesicles (autophagosomes) that ultimately
fuse with lysosomes, where their contents are degraded.
Subsequently, lipids and AA are released into the cytosol and
contribute to energy and AA supply in times of starvation.
Thus, along with lipolysis, macroautophagy is one of two
conserved responses to organismal and cellular fasting.
Singh and coworkers (Singh et al., 2009a) found that, in addi-
tion to conventional neutral lipases, autophagy of LDs is required
for fasting-induced lipolysis in murine liver and cultured hepato-
cytes. They showed that recruitment of microtubule-associated
protein 1 light chain 3 (LC-3) and formation of a regional
membrane through conjugation of autophagy-related protein 7
(Atg7) gives rise to double-membrane vesicles that engulf
portions of cytoplasmic LDs (autolipophagosomes). The autoli-
pophagosomes ultimately fuse with lysosomes, where their lipid
content is degraded by LAL. Consistent with a role for lipo-
autophagy during starvation, Singh et al. found that deletion of
Atg7 causes lipid accumulation in the liver. Chronic fat feeding
impairs the autophagic removal of lipid stores in the liver,
prompting excessive hepatic lipid deposition. Hotamisligil and
colleagues (Yang et al., 2010a) strengthened these findings by
showing that hepatic Atg7 expression is severely impaired in
ob/ob mice, contributing to the hepatosteatosis in these animals.
Liver fat accumulation was reduced by liver-specific restoration
of Atg7 expression.
While the effects of autophagy in genetically obese mice and in
mice fed high-fat diets seem consistent, a role for autophagy in
lipid metabolism of normal mice remains controversial. First,
autophagy appears to be predominantly relevant in the liver after
abnormally long fasting periods. Normally, during shorter fasting
periods the hepatic fat content increases due to the induction of
adipose tissue lipolysis and increased FA supply to the liver. It is
unlikely that autophagy would be induced under this condition.
Consistent with this prediction, Hotamisligil and colleagues did
not observe hepatic steatosis or changes in serum TG or FA
levels in lean mice following siRNA-mediated suppression of
Atg7, arguing that lipoautophagy is not involved (Yang et al.,
2010a). Second, hepatosteatosis as a result of Atg7 deficiency
was observed in some but not all studies. Although Atg7 defi-
ciency leads to severe liver enlargement, it is controversial
whether TGs accumulate. In fact, Uchiyama and colleagues
reported that hepatic Atg7 deletion upon starvation inhibits LD
formation both in vivo and in vitro, which leads to a lower hepatic
TG content (Shibata et al., 2009, 2010). Consistent with a role
for LC3 in LD formation, RNAi-mediated suppression of LC3-
expression prevented LD formation in a panel of different
(hepatic and nonhepatic) cell lines. Interestingly, LC3 localized
to the surface of LDs, and the authors argued that lipidation of
LC3 by phosphatidylethanolamine (formation of LC3-II), which
is the initial step in autophagosome formation during autophagy,
is also required for LD formation. A potential role of autophagy in
lipogenesis but not in lipolysis is consistent with the finding that
external administration of FAs induces, rather than inhibits, auto-
phagy when LDs are formed in the fasting liver (Tang et al., 2011).
A role for autophagy in lipogenesis also became evident from
analysis of adipose tissue in Atg7-deficient mice. Assuming
a lipolytic defect, we speculated that ablation of autophagy in
adipocytes would result in an obese phenotype (Zechner and
Madeo, 2009). In contrast, however, adipose-specific knockout
of Atg7 resulted in lean mice with reduced adipose mass,
enhanced insulin sensitivity, and an elevated rate of b-oxidation
(Singh et al., 2009b). The adipocytes contained smaller, multiloc-
ular LDs and exhibited normal basal lipolysis. In line with this,
inhibition of autophagy in cultured adipocytes using Atg7 siRNA
blocked TG accumulation (Singh et al., 2009b). Taken together,
these data argue for a currently poorly understood role of auto-
phagy in the biogenesis of LDs. Interestingly, while adipose-
specific Atg7-deficient mice display reduced WAT mass, BAT
mass increases (Baerga et al., 2009; Singh et al., 2009b). Singh
et al. argued that inhibition of white adipocyte differentiation
may lead to a defect in lipogenesis or that blocked autophagy
may promote WAT to BAT transdifferentiation.
Cell Metabolism 15, March 7, 2012 ª2012 Elsevier Inc. 283
Cell Metabolism
Review
In summary, autophagy may have pleiotropic roles in lipid
metabolism depending on the cell- or tissue-type. In the liver,
autophagy may contribute to lipolysis under a high-fat diet or pro-
longed fasting. Alternatively, autophagy may promote lipid accu-
mulation under normal fasting conditions. In WAT, autophagy
appears to be involved in adipocyte differentiation and lipogen-
esis but not in lipolysis. The role of autophagy in the breakdown
of fat in other lipolytically active tissues, such as muscle, macro-
phages, and steroidogenic cells remains to be determined.
Lipolysis: A Role in Lipid-Mediated Signaling?
Textbook knowledge tells us that TGs are the most efficient way
to store large amounts of FAs as an energy reserve. Consistent
with this view, cellular LDs were long seen as a relatively inert
storage depot for fat in adipose tissue that is mobilized at times
of increased energy demand. This view changed when it was
realized that 1) LDs are present in essentially all cell types,
including those for which mere energy storage does not seem
to be the main purpose; 2) LDs, particularly in nonadipose
tissues, undergo very dynamic changes of formation and degra-
dation; and 3) LDs represent a reservoir of bioactive lipids and
lipid-derived hormones in adipose and nonadipose tissues.
A role for neutral lipid metabolism in signaling gained substan-
tial interest when it was noted that increased cellular TG concen-
trations are strongly associated with insulin resistance in skeletal
muscle and liver (Cohen et al., 2011; Kelley et al., 2002). The rela-
tively inert nature of TGs makes it unlikely that they interfere
directly with insulin signaling. The concept that TGs themselves
are not the culprit was also supported by the so-called athletes’
paradox. Endurance athletes accumulate more TGs in LDs of
skeletal myocytes than do untrained individuals, yet their muscle
is highly insulin-sensitive. Similarly, mice that lack ATGL accu-
mulate large amounts of fat in numerous tissues (including skel-
etal muscle, cardiac muscle, liver, kidneys, and macrophages)
but exhibit increased insulin sensitivity (Haemmerle et al.,
2006; Kienesberger et al., 2009a). Increased insulin sensitivity
is observed despite the fact that ATGL-deficiency leads to an
insulin-secretion defect in pancreatic islets (Peyot et al., 2009).
In humans, ATGL-deficiency leads to neutral lipid storage
disease with myopathy (NLSDM), with a similar lipid phenotype
as observed in ATGL-deficient mice (Fischer et al., 2007).
Patients lacking the ATGL coactivator CGI-58 not only develop
neutral lipid storage disease but also exhibit a severe skin defect
(neutral lipid storage disease with ichthyosis, NLSDI) (Lefe
`vre
et al., 2001). To date, defective pancreatic insulin production
and alterations in insulin sensitivity have not been reported in
patients with NLSDM or NLSDI. In conclusion, cellular TG
content can be a marker of insulin resistance under certain phys-
iological conditions but is not a regulator of insulin signaling.
Lipolysis-Derived FAs and PPAR Signaling
Besides their powerful role as energy substrates and precursors
of other lipids, FAs are directly involved in cellular signaling path-
ways and regulation of gene transcription. FAs or FA derivatives
can bind to and activate members of the nuclear receptor family
of transcription factors that control the expression of genes
involved in lipid and energy homeostasis and inflammation.
The best-studied FA-activated nuclear receptors are the PPARs.
The PPAR family consists of four members: PPARa, PPARg-1
and -2, and PPARd(also designated PPARb). PPARaand
PPARdare highly expressed in oxidative tissues and regulate
genes involved in substrate delivery, substrate oxidation, and
oxidative phosphorylation (OXPHOS). In contrast, PPARgis
more important in lipogenesis and lipid synthesis, with highest
expression levels in WAT. The full transcriptional activity of
PPARs requires the binding of cognate lipid ligands, heterodime-
rization with another nuclear receptor (retinoid-X receptor, RXR),
and interaction with a number of transcriptional coactivators,
including PPARgcoactivator-1 (PGC-1). In addition to FAs, other
lipid ligands also have been described to activate PPARs, such
as acyl-CoAs, glycerol-phospholipids, and eicosanoids.
FAs involved in signaling originate from import of exogenous
FAs (from circulating FA-albumin complexes or from LPL-medi-
ated hydrolysis of plasma VLDL and chylomicrons) or from
endogenous de novo synthesis. Recently it was shown that
neither source of FA can generate PPAR ligands directly; rather,
a cycle of FA esterification and rehydrolysis is required (Haem-
merle et al., 2011)(Figure 2). As a consequence, lipolysis-
impaired ATGL-deficient mice exhibit a severe defect in PPARa
signaling in oxidative tissues such as liver (Ong et al., 2011),
macrophages (Chandak et al., 2010), and BAT (Ahmadian
et al., 2011). The most dramatic phenotype is observed in
cardiac muscle (Haemmerle et al., 2011). The reduced expres-
sion of PPARatarget genes in ATGL knockout animals causes
severe mitochondrial dysfunction, decreased rates of substrate
oxidation and OXPHOS, massive cardiac lipid accumulation,
and lethal cardiomyopathy within a few months after birth. HSL
deficiency is also associated with moderately decreased PPARa
target-gene expression but does not generate a comparable
cardiac phenotype, indicating the specific importance of ATGL
activity in the generation of PPARaligands or ligand precursors.
Figure 2. ATGL-Mediated Lipolysis Is Required for PPAR Signaling
and OXPHOS
Fatty acids from exogenous or endogenous sources are activated to acyl-
CoAs, which are subject to mitochondrial oxidation or TG formation. ATGL-
mediated lipolysis of TG generates lipolytic products (FA and DG), which may
act directly (e.g., FA) or after conversion (e.g., DG to phospholipids) as ligands
for nuclear receptors (for details see text). Activation of nuclear receptor
PPARavia lipolytic cleavage of TGs is required for normal mitochondrial
function and OXPHOS. In ATGL-deficient mice, defective PPARaactivation
and OXPHOS can be restored by treatment with PPARaagonists. ATGL,
adipose triglyceride lipase; CD36, cluster of differentiation 36; DG, diac-
ylglycerol; FA, fatty acid; FATP, fatty acid transport protein; LPL, lipoprotein
lipase; OXPHOS, oxidative phosphorylation; PPARa/d, peroxisome pro-
liferator-activated receptor alpha/delta; RA, retinoic acid; RXR, retinoid X
receptor; TG, triacylglycerol.
284 Cell Metabolism 15, March 7, 2012 ª2012 Elsevier Inc.
Cell Metabolism
Review
NLSDM or NLSDI in humans due to the deficiency of ATGL or
CGI-58, respectively, may also lead to reduced PPARasignaling
and defective OXPHOS. Although not proven experimentally,
this assumption stems from the finding that NLSDM patients
also develop systemic TG accumulation and cardiomyopathy.
In many patients, this condition is lethal if they do not undergo
heart transplantation (Hirano et al., 2008). Importantly, at least
in mice, the mitochondrial defect can be prevented by applica-
tion of PPARaactivators, such as Wy16453 or fenofibrate
(Haemmerle et al., 2011). Treatment of ATGL-deficient mice
leads to increased PPARasignaling, disappearance of cardiac
steatosis, and improved mitochondrial function and OXPHOS,
as well as prolonged survival. Whether pharmacological PPARa
activation will also be beneficial for patients with NLSDM is
currently not known but remains a promising possibility.
In addition to the instrumental role of ATGL in PPARasignaling,
the enzyme may also affect PPARgfunction. Festuccia et al.
(Festuccia et al., 2006) showed that rosiglitazone-mediated
PPARgactivation and lipid accumulation are associated with
increased lipolysis in WAT of rats and that increased lipolysis
was due to induction of ATGL and MGL. Although it seems
counterintuitive that lipolysis is induced during increased TG
synthesis, it is conceivable that this step is required to promote
PPARg-activated expression of lipogenic genes. Moreover, the
fact that HSL-deficiency also leads to downregulation of PPARg
target-gene expression in WAT (Shen et al., 2011; Zimmermann
et al., 2003) suggests that lipolysis is involved in PPARgsignaling.
Lipolysis-Derived FAs and Insulin Signaling
It is well established that both plasma and cellular FA concen-
trations correlate positively with increased insulin resistance
(Boden and Shulman, 2002). Several mechanisms are discussed.
Increased cellular concentrations of nonesterifiedFAs, particularly
palmitate,can drive the synthesisof lipotoxic lipids suchas ceram-
ides, which interfere with functional insulin signaling (Summers,
2010). Additionally, FAs can directly or indirectly—via increased
production of reactive oxygen species—activate redox-sensitive
serine kinases, which, in turn, inactivate the insulin response
(Vallerie and Hotamisligil, 2010). Yet, not all FA species seem to
have the same inhibitory effect on insulin signaling. Whereas
palmitate consistently decreases the insulin response, palmito-
leate actually enhances the insulin signal in liver and muscle
(Cao et al., 2008). Palmitoleate is mostly generated by de novo
synthesis in adipose tissue. Its effects on insulin signaling in liver
and muscle suggest that this FA is a lipokine with endocrine func-
tion. Additionally, it is conceivable that hepatic palmitoleate also
contributes to insulin sensitization in a paracrine fashion. Whether
lipolysis contributes to the generation of palmitoleate or other FAs
with a downstream effect on insulin signaling requires additional
studies. ATGL-deficient mice have low plasma FA concentrations
and are highly insulin-sensitive, arguing for a protective role of
reduced lipolysis and low FA levels. Whether low plasma FAs in
HSL-deficient mice also affect insulin sensitivity is controversial
(Mulder et al., 2003; Park et al., 2005; Voshol et al., 2003).
Lipolysis-Derived DGs Are Unlikely to Affect
PKC Signaling
The potential of DGs to act as second messengers was discov-
ered approximately 50 years ago when researchers realized that
DGs affect many metabolic and mitogenic activities via activa-
tion of protein kinase-C (PKC). Metabolic regulation involves
suppression of insulin signaling via phosphorylation of insulin
receptor substrate-1, leading to insulin resistance in muscle
and liver (Samuel et al., 2010). Only one DG stereoisomer,
1,2-diacyl-sn-glycerols (1,2-DGs), is able to activate PKCs,
whereas the others, 1,3-diacyl-sn-glycerols (1,3-DGs) and
2,3-diacyl-sn-glycerols (2,3-DGs), lack this bioactivity (Boni
and Rando, 1985). 1,2-DGs activate conventional and novel
PKC isoforms after recruitment of the enzymes to the plasma
membrane by receptor of activated C kinase (RACK) proteins
(Turban and Hajduch, 2011). Accordingly, both stereo-specific
and location-specific preconditions are required for DGs to acti-
vate PKCs. Three potential sources exist for the generation of
1,2-DGs (Figure 3). Classical signaling 1,2-DGs derive from
phospholipase C (PLC)-mediated hydrolysis of phosphatidylino-
sitol-4,5-phosphate in the plasma membrane. Vertebrates have
13 isoforms of PLC grouped within 6 isotypes. Both products of
Figure 3. Lipolysis and Lipid Signaling
Lipid intermediates involved in cellular signaling are generated by anabolic and
catabolic reactions in distinct cellular compartments. 1,2-DGs, the ligands of
conventional and novel PKCs, are formed at the plasma membrane by
PLC-mediated degradation of PIP2. This reaction also generates IP3,
a signaling molecule, which leads to Ca
2+
efflux from the ER. De novo
synthesis of 1,2-DGs at the ER may also contribute to PKC activation. FAs
are ligands for nuclear receptors. They are generated by de novo synthesis or
hydrolysis of neutral lipids or phospholipids. 2-AG is an important MG involved
in endocannabinoid signaling. It originates from membrane-associated
phospholipid hydolysis by PLCs and the subsequent hydrolysis of DGs by
DAGLs. The contribution of TG hydrolysis by ATGL and HSL to cellular 2-AG
concentrations is not known. The 2-AG signal is inactivated by MGL. AGPAT,
acyl-CoA acylglycerol-3-phosphate acyltransferase; 2-AG, 2-arachidonoyl-
glycerol; ATGL, adipose triglyceride lipase; DAGL, diacylglycerol lipase; DG,
diacylglycerol; 11, 2-DG, diacyl-sn1,2-glycerol, DGAT, acyl-CoA: diac-
ylglycerol acyltransferase; FA, fatty acid; G, glycerol; G3P, glycerol-3-phos-
phate; GPAT, glycerol-3-phosphate acyltransferase; HSL, hormone-sensitive
lipase; IP3, inositol-1,4,5-trisphosphate; LPA, lysophosphatidic acid; MG,
monoacylglycerol; MGL, monoglyceride lipase; PA, phosphatidic acid;
PAPase, PA phosphohydrolase; PIP2, phosphatidylinositol 4,5-bisphosphate;
PKC, protein kinase C; PLC, phospholipase C; TG, triacylglycerol.
Cell Metabolism 15, March 7, 2012 ª2012 Elsevier Inc. 285
Cell Metabolism
Review
the enzymatic reaction, 1,2-DGs and inositol-1,4,5-phosphate
(IP
3
), are potent second messengers. Whereas 1,2-DGs remain
plasma membrane-associated, IP
3
dissociates and induces
Ca
2+
release from the ER, which is required in addition to
1,2-DGs for activation of conventional PKCs.
Although PLCs generate the ‘‘correct’’ stereoisomer (1,2-DGs)
in the ‘‘correct’’ cellular location for PKC activation, it remains
controversial whether other cellular sources also contribute
to the production of signaling 1,2-DGs. De novo synthesis via
dual acylation of sn-glycerol-3-phosphate by acyl-CoA glyc-
erol-3-phosphate acyltransferases (GPATs) and acyl-CoA acyl-
glycerol-3-phosphate acyltransferases (AGPATs), and subse-
quent dephosphorylation of phosphatidic acid by phosphatidic
acid phosphohydrolases (PAPases) also leads to the formation
of 1,2-DGs (Figure 3). However, this pathway of DG synthesis
is restricted to ER membranes. Accordingly, a model proposing
activation of PKC by ER-associated 1,2-DGs needs to include
PKC localization to the ER or contact sites between the plasma
membrane and the ER membrane.
The third potential source of DGs derives from the lipolysis of
LD-associated TGs by ATGL (Figure 3). The stereospecificity of
ATGL has not been reported. However, our unpublished obser-
vations showed that the enzyme preferentially hydrolyzes sn-1
and sn-2 ester bonds but not sn-3 esters. This indicates that
ATGL generates 1,3-DGs and 2,3-DGs but not 1,2-DGs. In
accordance with the subsequent hydrolysis of 1,3-DGs and
2,3-DGs by HSL, this enzyme has a stereo-preference for the
hydrolysis of FAs in the sn-3 position of DGs (Rodriguez et al.,
2010). In TGs, HSL preferably hydrolyses sn-1(3) ester bonds
(Fredrikson and Belfrage, 1983). Therefore, neither the ATGL
nor the HSL reaction generate 1,2-DGs on LDs. Additionally, it
is questionable whether LD-associated DGs would dissociate
from LDs to participate in the recruitment and activation of
PKC at the plasma membrane. From all this, it seems unlikely
that lipolytically generated DGs act as signaling mediators.
In a recent review, Shulman and colleagues (Samuel et al.,
2010) summarized numerous animal and human studies pro-
viding evidence that cellular DG concentrations account for the
development of lipid-induced insulin resistance in type 2 dia-
betes, lipodystrophy, and other conditions. Consistent with this
hypothesis, mice lacking DG-kinase-d, the major enzyme that
inactivates the DG signal, have increased DG levels and
increased insulin resistance (Chibalin et al., 2008). However,
considering the structural complexity of DG species and their
localization in different cellular compartments, a general correla-
tion between total cellular DG concentrations and insulin re-
sistance seems unlikely. This is supported by studies where
increased total cellular DG concentrations in mutant mouse
models were not associated with insulin resistance. For ex-
ample, HSL-deficient mice accumulate large amounts of DGs
in adipose and many nonadipose tissues due to defective DG
catabolism. Yet, most studies agree that this does not lead to
PKC hyperactivation or a severely defective insulin response
(Mulder et al., 2003; Park et al., 2005; Voshol et al., 2003).
Similarly, CGI-58 silencing in the liver leads to increased DG
levels, but normal (chow diet) or increased (high-fat diet) glucose
tolerance and insulin sensitivity (Brown et al., 2010).
Taken together, determination of the specific 1,2-DG concen-
trations in the plasma membrane may provide a more reliable
predictor for lipid-induced, PKC-mediated insulin resistance
than total cellular DG concentrations.
Monoacylglycerol Signaling and Lipolysis
The signaling potential of MGs was recognized when it
was found that the phospholipid-derived MG 2-AG activates
cannabinoid receptors (CBR), thereby regulating food intake,
lipid metabolism, and energy homeostasis. The endocannabi-
noid system (ECS) refers to a group of neuromodulatory lipids
(endocannabinoids, ECs), two G protein-coupled receptors
(CBR1 and CBR2), and enzymes involved in the synthesis and
degradation of ECs (Di Marzo, 2009). The best-characterized
ECs are N-arachidonoyl ethanolamine (AEA, anandamide) and
2-AG. Their biological effect is mimicked by D
9
-tetrahydrocan-
nabinol (THC), the major psychoactive component of marijuana.
The ECS regulates a diverse spectrum of physiological pro-
cesses, including motor function, pain, appetite, cognition,
emotional behavior, and immunity. In the nervous system,
2-AG acts as a retrograde messenger, inhibiting presynaptic
neurotransmitter release (Alger and Kim, 2011). It is produced
postsynaptically and traverses the synaptic cleft to stimulate
presynaptic CBR1. Subsequently, 2-AG is internalized into the
presynaptic terminal and inactivated by the MGL reaction,
forming glycerol and arachidonic acid. The ECS is active in
neurons and nonneuronal cells such as immune cells, hepato-
cytes, and adipocytes. Treatment of obese patients with the
CBR1-antagonist rimonabant (Christopoulou and Kiortsis,
2011) and studies with animal models lacking CBR1 revealed
that blockade of the ECS reduces food intake, decreases lipo-
genesis, and increases energy consumption. Conversely, an
overactive ECS has a central orexigenic effect and reduces
energy expenditure, promoting lipid deposition in peripheral
tissues like the liver and WAT (Cota, 2008). Because of these
biological effects, the ECS has been linked to the pathogenesis
of metabolic diseases. Obese patients may have an overactive
ECS, which stimulates appetite and promotes lipid deposition
(Perkins and Davis, 2008).
It is generally assumed that signaling 2-AG originates from
the degradation of glycerophospholipids containing arachidonic
acid in the sn-2 position (Figure 3). Various isoforms of PLC
generate 1,2-DGs (see above), which are subsequently hydro-
lyzed by DG lipase (DAGL) to 2-AG. Whether HSL also partici-
pates in the hydrolysis of plasma membrane-associated
1,2-DGs to generate 2-AG is not known. Current evidence
suggests that at least two isoforms of DAGL (DAGLaand DAGLb)
exist in the brain and liver (Bisogno et al., 2003). Mice lacking
DAGLaexhibit a substantial decrease in brain and spinal cord
2-AG levels, whereas DAGLbappears to be more important in
peripheral tissues, such as the liver (Gao et al., 2010). Whether
the catabolism of arachidonic acid-containing TGs in LDs by
ATGL and HSL also contributes to the cellular 2-AG and arach-
idonic acid pool is not known.
Recent studies using an MGL-specific small-molecule inhib-
itor (JZL184) and MGL knockout mice provided compelling
evidence that MGL is the major enzyme in the degradation of
2-AG and other MGs esterified with long-chain FAs (Chanda
et al., 2010; Schlosburg et al., 2010; Taschler et al., 2011).
Animals lacking MGL or mice treated with JZL184 show abnor-
mally high amounts of various MG species in the brain and
286 Cell Metabolism 15, March 7, 2012 ª2012 Elsevier Inc.
Cell Metabolism
Review
peripheral tissues. In the brain, lack of MGL activity leads to
a more than 20-fold increase in 2-AG, suggesting that MGL-defi-
ciency could lead to hyperactivation of the ECS. Indeed, JZL184
treatment of mice provoked cannabimimetic effects in mice,
including analgesia, hypothermia, and hypomotility (Long et al.,
2009). However, genetic ablation of MGL in mice did not result
in a hyperactive ECS or any obvious phenotype. This surprising
observation was explained by desensitization of brain CBR1
leading to functional antagonism, and highlights the important
role of 2-AG as a retrograde neurotransmitter (Chanda et al.,
2010; Schlosburg et al., 2010). Obviously, increased brain
2-AG concentrations provoke counter-regulatory mechanisms
similar to those observed when animals are chronically fed
CBR agonists (Lichtman and Martin, 2005).
Although MGL-deficiency in mice does not produce cannabi-
mimetic effects, the lack of MGL activity substantially affects
lipolysis and metabolism in adipose tissue and nonadipose
tissues (Taschler et al., 2011). MGL deficiency results in the
accumulation of MGs and a reduction of circulating TG and glyc-
erol levels in fasted animals. Unexpectedly, and in contrast to the
proposed role of the ECS in obesity-related metabolic diseases,
MGL knockout mice exhibit improved insulin sensitivity and
glucose tolerance when fed a high-fat diet. The cause for this
finding may be complex, considering that MGL-deficiency is
associated with desensitized CBRs. Investigation of mice lack-
ing MGL, specifically in the brain or peripheral tissues, should
help to unravel the question of whether central or peripheral
effects cause attenuation of insulin resistance.
Lipolysis in the Cell Cycle, Cancer, and Cachexia:
A Question of Lipid Signaling?
The first observation to indicate that lipolysis is linked to efficient
cell-cycle progression was reported in the yeast S. cerevisiae.
Yeast expresses three TG lipases of the patatin domain-contain-
ing family termed Tgl3 to 5 (Czabany et al., 2007). Tgl4 is a func-
tional ortholog of mammalian ATGL (Kurat et al., 2006). Deletion
of Tgl3 and Tgl4 abolishes virtually all cellular TG lipase activity
and causes a marked delay of entry into the cell division cycle
of starved cells upon refeeding (Kurat et al., 2009). Tgl4 is
activated via phosphorylation by the cyclin-dependent kinase
Cdk1/Cdc28 (ortholog of mammalian Cdc2). Concomitantly,
Cdk1/Cdc28 inhibits lipogenesis by phosphorylation of phos-
phatidic acid phosphohydrolase (Pah1). This suggests that TG
levels oscillate during the cell-division cycle to either deposit
de novo synthesized FAs in TGs or, conversely, to provide FAs
during phases of increased demand. Tgl4 phosphorylation and
activation occur at the G1/S transition of the cell-division cycle,
which coincides with bud emergence and requires increased
amounts of membrane lipids. Pah1 phosphorylation and inacti-
vation occur at the G2/M transition of the cell cycle, indicating
that a window exists during the cell cycle in which both the initial
step of lipogenesis and lipolysis may operate in parallel. This is
feasible because both activities are confined to different organ-
elles, namely the ER and LDs, respectively. The specific check-
point proteins that regulate cell-cycle progression in response
to lipolysis are currently unknown. Recent evidence shows that
the synthesis of phosphatidylinositol (PI), a precursor for signal-
ing molecules in cell-cycle regulation in yeast (e.g., IP
3
and
inositol-containing ceramides), strongly depends on intact lipol-
ysis (Gaspar et al., 2011). Thus, cell cycle-regulated TG lipolysis
may provide critical precursors for signaling molecules for cell
division.
A highly interesting study recently demonstrated that MGL
promotes the oncogenic properties of mammalian cancer cells
(Nomura et al., 2010). The authors demonstrated that overex-
pression or disruption of MGL activity increased or decreased,
respectively, the proliferation of cancer cells, and that MGL is
highly expressed in aggressive tumor cell lines or primary
tumors. The study also provided evidence that MGL influences
tumor proliferation by metabolic effects rather than by CBR-
dependent mechanisms. The lack of MGL reduced cellular
nonesterified FA concentrations. Growth of tumor cells in
response to MGL-knockdown was not impaired after addition
of exogenous nonesterified FAs or in mice fed a high-fat diet.
This led to the conclusion that MGL affects the concentration
of FA-derived tumorigenic lipid metabolites, such as LPA and
PGE
2
. Although the involved lipid signal(s) requires identification,
these studies clearly designate MGL as an interesting target for
cancer therapy.
Lipolytic signaling may also be causally involved in the patho-
genesis of cancer-associated cachexia (CAC). In a recent study,
Das et al. (Das et al., 2011) demonstrated that ATGL-deficient
mice are protected against tumor-induced loss of adipose tissue
and skeletal muscle. HSL-deficient mice were also protected,
although to a lesser degree. ATGL-deficient mice maintained
body weight and composition despite increased circulating
factors that induce lipolysis, muscle proteolysis, and apoptosis
(e.g., TNFa, interleukin-6, and zinc-a-glycoprotein 1). This sug-
gests that lipolysis is integrated in a signal transduction network
that eventually leads to the loss of adipose tissue and muscle.
This view is also consistent with the observation that the activity
of lipolytic enzymes and release of FAs and glycerol are
increased in adipose tissue of cancer patients with cachexia
(Agustsson et al., 2007; Das et al., 2011; Ryde
´n et al., 2008).
The nature of the lipolytic signal involved is currently unknown.
It is also unclear whether the signal originates from lipolysis in
one tissue (such as adipose tissue) and promotes wasting in
an endocrine manner or whether wasting requires autonomous
lipolysis in all tissues that are affected by wasting. Although the
underlying mechanism is not yet defined, this study suggests
that inhibition of lipolysis may help to prevent cachexia in
patients with cancer or other chronic diseases.
Conclusion
Recent discoveries of enzymes and regulatory factors have led
to a revision of our perception of lipolysis. The complexity of
the process and its regulation are still only partially understood.
Additionally, we have just begun to address the role of lipolytic
products and intermediates in cellular signaling. Important
topics for future investigations include: 1) better understanding
of the biochemical factors and processes that coordinately
regulate the lipolytic machinery in response to hormonal activa-
tors and inhibitors, 2) the physiological function of lipolysis in
numerous nonadipose tissues and the tissue-specific differ-
ences in lipolytic mechanisms, and 3) the characterization
of lipolytic signals and the molecular mechanisms of their
effects on gene transcription, the cell cycle, and cell growth.
The recent examples of lipases affecting tumor proliferation or
Cell Metabolism 15, March 7, 2012 ª2012 Elsevier Inc. 287
Cell Metabolism
Review
cancer-associated cachexia emphasize the potential impor-
tance of lipolysis in human disease.
ACKNOWLEDGMENTS
This work was supported by the grants P20602, P18434, P21296, F30 SFB
LIPOTOX, the Doktoratskollegs W901 and W1226, and the Wittgenstein Award
Z136, which are funded by the Austrian Science Foundation. Funding was also
provided by the grant ‘‘GOLD: Genomics Of Lipid-associated Disorders,’
which is part of the Austrian Genome Project ‘‘GEN-AU: Genome Research
in Austria’’ funded by the Austrian Ministry of Science and Research and the
FFG. Additional support was obtained from the European commission grant
agreements no. 202272 (LipidomicNet), the City of Graz and the Province of
Styria. We thank Dr. Ellen Zechner, Mag. Caroline Schober-Trummler, and
Mag. Dr. Tarek Mustafa for careful and critical reading of the manuscript.
REFERENCES
Agustsson, T., Ryde
´n, M., Hoffstedt, J., van Harmelen, V., Dicker, A., Lauren-
cikiene, J., Isaksson, B., Permert, J., and Arner, P. (2007). Mechanism of
increased lipolysis in cancer cachexia. Cancer Res. 67, 5531–5537.
Ahmadian, M., Abbott, M.J., Tang, T., Hudak, C.S., Kim, Y., Bruss, M., Heller-
stein, M.K., Lee, H.Y., Samuel, V.T., Shulman, G.I., et al. (2011). Desnutrin/
ATGL is regulated by AMPK and is required for a brown adipose phenotype.
Cell Metab. 13, 739–748.
Alger, B.E., and Kim, J. (2011). Supply and demand for endocannabinoids.
Trends Neurosci. 34, 304–315.
Anthonsen, M.W., Ro
¨nnstrand, L., Wernstedt, C., Degerman, E., and Holm, C.
(1998). Identification of novel phosphorylation sites in hormone-sensitive
lipase that are phosphorylated in response to isoproterenol and govern activa-
tion properties in vitro. J. Biol. Chem. 273, 215–221.
Baerga, R., Zhang, Y., Chen, P.H., Goldman, S., and Jin, S. (2009). Targeted
deletion of autophagy-related 5 (atg5) impairs adipogenesis in a cellular model
and in mice. Autophagy 5, 1118–1130.
Balser, W.A. (1882). Ueber Fettnekrose, eine zuweilen to
¨dtliche Krankheit des
Menschen. Virchows Arch. 90, 520–535.
Bartz, R., Zehmer, J.K., Zhu, M., Chen, Y., Serrero, G., Zhao, Y., and Liu, P.
(2007). Dynamic activity of lipid droplets: protein phosphorylation and GTP-
mediated protein translocation. J. Proteome Res. 6, 3256–3265.
Basantani, M.K., Sitnick, M.T., Cai, L., Brenner, D.S., Gardner, N.P., Li, J.Z.,
Schoiswohl, G., Yang, K., Kumari, M., Gross, R.W., et al. (2011). Pnpla3/Adipo-
nutrin deficiency in mice does not contribute to fatty liver disease or metabolic
syndrome. J. Lipid Res. 52, 318–329.
Baulande, S., Lasnier, F., Lucas, M., and Pairault, J. (2001). Adiponutrin,
a transmembrane protein corresponding to a novel dietary- and obesity-linked
mRNA specifically expressed in the adipose lineage. J. Biol. Chem. 276,
33336–33344.
Bell, M., Wang, H., Chen, H., McLenithan, J.C., Gong, D.W., Yang, R.Z., Yu, D.,
Fried, S.K., Quon, M.J., Londos, C., and Sztalryd, C. (2008). Consequences of
lipid droplet coat protein downregulation in liver cells: abnormal lipid droplet
metabolism and induction of insulin resistance. Diabetes 57, 2037–2045.
Beller, M., Sztalryd, C., Southall, N., Bell, M., Ja
¨ckle, H., Auld, D.S., and Oliver,
B. (2008). COPI complex is a regulator of lipid homeostasis. PLoS Biol. 6,e292.
Bernard, C. (1856). Me
´moire sur le pancre
´as et sur la role du suc pancre
´atique
dans les phe
´nome
`mes digestifs particulie
`rement dans la digestion des
matie
`res grasses neutres (Paris: Academie des Sciences).
Bertrand, T., Auge
´, F., Houtmann, J., Rak, A., Valle
´e, F., Mikol, V., Berne, P.F.,
Michot, N., Cheuret, D., Hoornaert, C., and Mathieu, M. (2010). Structural basis
for human monoglyceride lipase inhibition. J. Mol. Biol. 396, 663–673.
Bezaire, V., Mairal, A., Ribet, C., Lefort, C., Girousse, A., Jocken, J., Laurenci-
kiene, J., Anesia, R., Rodriguez, A.M., Ryden, M., et al. (2009). Contribution of
adipose triglyceride lipase and hormone-sensitive lipase to lipolysis in hMADS
adipocytes. J. Biol. Chem. 284,18282–18291.
Bisogno, T., Howell, F., Williams, G., Minassi, A., Cascio, M.G., Ligresti, A.,
Matias, I., Schiano-Moriello, A., Paul, P., Williams, E.J., et al. (2003). Cloning
of the first sn1-DAG lipases points to the spatial and temporal regulation of
endocannabinoid signaling in the brain. J. Cell Biol. 163, 463–468.
Boden, G., and Shulman, G.I. (2002). Free fatty acids in obesity and type 2 dia-
betes: defining their role in the development of insulin resistance and beta-cell
dysfunction. Eur. J. Clin. Invest. 32 (Suppl 3), 14–23.
Boni, L.T., and Rando, R.R. (1985). The nature of protein kinase C activation by
physically defined phospholipid vesicles and diacylglycerols. J. Biol. Chem.
260, 10819–10825.
Borg, M.L., Andrews, Z.B., Duh, E.J., Zechner, R., Meikle, P.J., and Watt, M.J.
(2011). Pigment epithelium-derived factor regulates lipid metabolism via
adipose triglyceride lipase. Diabetes 60, 1458–1466.
Brown, J.M., Betters, J.L., Lord, C., Ma, Y., Han, X., Yang, K., Alger, H.M.,
Melchior, J., Sawyer, J., Shah, R., et al. (2010). CGI-58 knockdown in mice
causes hepatic steatosis but prevents diet-induced obesity and glucose
intolerance. J. Lipid Res. 51, 3306–3315.
Cao, H., Gerhold, K., Mayers, J.R., Wiest, M.M., Watkins, S.M., and Hotamisli-
gil, G.S. (2008). Identification of a lipokine, a lipid hormone linking adipose
tissue to systemic metabolism. Cell 134, 933–944.
Chakrabarti, P., English, T., Shi, J., Smas, C.M., and Kandror, K.V. (2010).
Mammalian target of rapamycin complex 1 suppresses lipolysis, stimulates
lipogenesis, and promotes fat storage. Diabetes 59, 775–781.
Chakrabarti, P., English, T., Karki, S., Qiang, L., Tao, R., Kim, J., Luo, Z.,
Farmer, S.R., and Kandror, K.V. (2011). SIRT1 controls lipolysis in adipocytes
via FOXO1-mediated expression of ATGL. J. Lipid Res. 52, 1693–1701.
Chanda, P.K., Gao, Y., Mark, L., Btesh, J., Strassle, B.W., Lu, P., Piesla, M.J.,
Zhang, M.Y., Bingham, B., Uveges, A., et al. (2010). Monoacylglycerol lipase
activity is a critical modulator of the tone and integrity of the endocannabinoid
system. Mol. Pharmacol. 78, 996–1003.
Chandak, P.G., Radovic, B., Aflaki, E., Kolb, D., Buchebner, M., Fro
¨hlich, E.,
Magnes, C., Sinner, F., Haemmerle, G., Zechner, R., et al. (2010). Efficient
phagocytosis requires triacylglycerol hydrolysis by adipose triglyceride lipase.
J. Biol. Chem. 285, 20192–20201.
Chen, W., Chang, B., Li, L., and Chan, L. (2010). Patatin-like phospholipase
domain-containing 3/adiponutrin deficiency in mice is not associated with fatty
liver disease. Hepatology 52, 1134–1142.
Chibalin, A.V., Leng, Y., Vieira, E., Krook, A., Bjo
¨rnholm, M., Long, Y.C.,
Kotova, O., Zhong, Z., Sakane, F., Steiler, T., et al. (2008). Downregulation of
diacylglycerol kinase delta contributes to hyperglycemia-induced insulin
resistance. Cell 132, 375–386.
Christopoulou, F.D., and Kiortsis, D.N. (2011). An overview of the metabolic
effects of rimonabant in randomized controlled trials: potential for other canna-
binoid 1 receptor blockers in obesity. J. Clin. Pharm. Ther. 36, 10–18.
Chung, C., Doll, J.A., Gattu, A.K., Shugrue, C., Cornwell, M., Fitchev, P., and
Crawford, S.E. (2008). Anti-angiogenic pigment epithelium-derived factor
regulates hepatocyte triglyceride content through adipose triglyceride lipase
(ATGL). J. Hepatol. 48, 471–478.
Cohen, J.C., Horton, J.D., and Hobbs, H.H. (2011). Human fatty liver disease:
old questions and new insights. Science 332, 1519–1523.
Cota, D. (2008). Role of the endocannabinoid system in energy balance regu-
lation and obesity. Front. Horm. Res. 36, 135–145.
Czabany, T., Athenstaedt, K., and Daum, G. (2007). Synthesis, storage and
degradation of neutral lipids in yeast. Biochim. Biophys. Acta 1771, 299–309.
Das, S.K., Eder, S., Schauer, S., Diwoky, C., Temmel, H., Guertl, B., Gorkie-
wicz, G., Tamilarasan, K.P., Kumari, P., Trauner, M., et al. (2011). Adipose
triglyceride lipase contributes to cancer-associated cachexia. Science 333,
233–238.
Daval, M., Diot-Dupuy, F., Bazin, R., Hainault, I., Viollet, B., Vaulont, S.,
Hajduch, E., Ferre
´,P., and Foufelle, F. (2005). Anti-lipolytic action of AMP-
activated protein kinase in rodent adipocytes. J. Biol. Chem. 280, 25250–
25257.
Di Marzo, V. (2009). The endocannabinoid system: its general strategy of
action, tools for its pharmacological manipulation and potential therapeutic
exploitation. Pharmacol. Res. 60, 77–84.
288 Cell Metabolism 15, March 7, 2012 ª2012 Elsevier Inc.
Cell Metabolism
Review
Dubuquoy, C., Robichon, C., Lasnier, F., Langlois, C., Dugail, I., Foufelle, F.,
Girard, J., Burnol, A.F., Postic, C., and Moldes, M. (2011). Distinct regulation
of adiponutrin/PNPLA3 gene expression by the transcription factors ChREBP
and SREBP1c in mouse and human hepatocytes. J. Hepatol. 55, 145–153.
Duncan, R.E., Wang, Y., Ahmadian, M., Lu, J., Sarkadi-Nagy, E., and Sul, H.S.
(2010). Characterization of desnutrin functional domains: critical residues for
triacylglycerol hydrolysis in cultured cells. J. Lipid Res. 51, 309–317.
Ellong, E.N., Soni, K.G., Bui, Q.T., Sougrat, R., Golinelli-Cohen, M.P., and
Jackson, C.L. (2011). Interaction between the triglyceride lipase ATGL and
the Arf1 activator GBF1. PLoS ONE 6, e21889.
Enoksson, S., Degerman, E., Hagstro
¨m-Toft, E., Large, V., and Arner, P.
(1998). Various phosphodiesterase subtypes mediate the in vivo antilipolytic
effect of insulin on adipose tissue and skeletal muscle in man. Diabetologia
41, 560–568.
Festuccia, W.T., Laplante, M., Berthiaume, M., Ge
´linas, Y., and Deshaies, Y.
(2006). PPARgamma agonism increases rat adipose tissue lipolysis, expres-
sion of glyceride lipases, and the response of lipolysis to hormonal control.
Diabetologia 49, 2427–2436.
Filleur, S., Nelius, T., de Riese, W., and Kennedy, R.C. (2009). Characterization
of PEDF: a multi-functional serpin family protein. J. Cell. Biochem. 106,
769–775.
Fischer, J., Lefe
`vre, C., Morava, E., Mussini, J.M., Lafore
ˆt, P., Negre-Salvayre,
A., Lathrop, M., and Salvayre, R. (2007). The gene encoding adipose triglyc-
eride lipase (PNPLA2) is mutated in neutral lipid storage disease with myop-
athy. Nat. Genet. 39, 28–30.
Flexner, S. (1897). On the Occurrence of the Fat-Splitting Ferment in Peritoneal
Fat Necroses and the Histology of These Lesions. J. Exp. Med. 2, 413–425.
Fredrikson, G., and Belfrage, P. (1983). Positional specificity of hormone-
sensitive lipase from rat adipose tissue. J. Biol. Chem. 258, 14253–14256.
Fredrikson, G., Tornqvist, H., and Belfrage, P. (1986). Hormone-sensitive
lipase and monoacylglycerol lipase are both required for complete degrada-
tion of adipocyte triacylglycerol. Biochim. Biophys. Acta 876, 288–293.
Gaidhu, M.P., Fediuc, S., Anthony, N.M., So, M., Mirpourian, M., Perry, R.L.,
and Ceddia, R.B. (2009). Prolonged AICAR-induced AMP-kinase activation
promotes energy dissipation in white adipocytes: novel mechanisms inte-
grating HSL and ATGL. J. Lipid Res. 50, 704–715.
Gandotra, S., Lim, K., Girousse, A., Saudek, V., O’Rahilly, S., and Savage, D.B.
(2011). Human frameshift mutations affecting the carboxyl terminus of perilipin
increase lipolysis by failing to sequester the adipose triglyceride lipase (ATGL)
coactivator, AB-hydrolase containing 5 (ABHD5). J Biol Chem. 286, 34998–
35006.
Gao, Y., Vasilyev, D.V., Goncalves, M.B., Howell, F.V., Hobbs, C., Reisenberg,
M., Shen, R., Zhang, M.Y., Strassle, B.W., Lu, P., et al. (2010). Loss of retro-
grade endocannabinoid signaling and reduced adult neurogenesis in diacyl-
glycerol lipase knock-out mice. J. Neurosci. 30, 2017–2024.
Gaspar, M.L., Hofbauer, H.F., Kohlwein, S.D., and Henry, S.A. (2011). Coordi-
nation of storage lipid synthesis and membrane biogenesis: evidence for
cross-talk between triacylglycerol metabolism and phosphatidylinositol
synthesis. J. Biol. Chem. 286, 1696–1708.
Gauthier, M.S., Miyoshi, H., Souza, S.C., Cacicedo, J.M., Saha, A.K., Green-
berg, A.S., and Ruderman, N.B. (2008). AMP-activated protein kinase is acti-
vated as a consequence of lipolysis in the adipocyte: potential mechanism and
physiological relevance. J. Biol. Chem. 283, 16514–16524.
Ghosh, A.K., Ramakrishnan, G., and Rajasekharan, R. (2008). YLR099C (ICT1)
encodes a soluble Acyl-CoA-dependent lysophosphatidic acid acyltransfer-
ase responsible for enhanced phospholipid synthesis on organic solvent
stress in Saccharomyces cerevisiae. J. Biol. Chem. 283, 9768–9775.
Granneman, J.G., Moore, H.P., Krishnamoorthy, R., and Rathod, M. (2009).
Perilipin controls lipolysis by regulating the interactions of AB-hydrolase con-
taining 5 (Abhd5) and adipose triglyceride lipase (Atgl). J. Biol. Chem. 284,
34538–34544.
Granneman, J.G., Moore, H.P., Mottillo, E.P., Zhu, Z., and Zhou, L. (2011).
Interactions of perilipin-5 (Plin5) with adipose triglyceride lipase. J. Biol.
Chem. 286,5126–5135.
Guo, Y., Walther, T.C., Rao, M., Stuurman, N., Goshima, G., Terayama, K.,
Wong, J.S., Vale, R.D., Walter, P., and Farese, R.V. (2008). Functional genomic
screen reveals genes involved in lipid-droplet formation and utilization. Nature
453, 657–661.
Haemmerle, G., Zimmermann, R., Hayn, M., Theussl, C., Waeg, G., Wagner,
E., Sattler, W., Magin, T.M., Wagner, E.F., and Zechner, R. (2002). Hormone-
sensitive lipase deficiency in mice causes diglyceride accumulation in adipose
tissue, muscle, and testis. J. Biol. Chem. 277, 4806–4815.
Haemmerle, G., Lass, A., Zimmermann, R., Gorkiewicz, G., Meyer, C.,
Rozman, J., Heldmaier, G., Maier, R., Theussl, C., Eder, S., et al. (2006).
Defective lipolysis and altered energy metabolism in mice lacking adipose
triglyceride lipase. Science 312, 734–737.
Haemmerle, G., Moustafa, T., Woelkart, G., Buttner, S., Schmidt, A., van de
Weijer, T., Hesselink, M., Jaeger, D., Kienesberger, P.C., Zierler, K., et al.
(2011). ATGL-mediated fat catabolism regulates cardiac mitochondrial func-
tion via PPAR-alpha and PGC-1. Nat Med. 17, 1076–1085.
He, S., McPhaul, C., Li, J.Z., Garuti, R., Kinch, L., Grishin, N.V., Cohen, J.C.,
and Hobbs, H.H. (2010). A sequence variation (I148M) in PNPLA3 associated
with nonalcoholic fatty liver disease disrupts triglyceride hydrolysis. J. Biol.
Chem. 285, 6706–6715.
Hirano, K., Ikeda, Y., Zaima, N., Sakata, Y., and Matsumiya, G. (2008). Triglyc-
eride deposit cardiomyovasculopathy. N. Engl. J. Med. 359, 2396–2398.
Ho, P.C., Chuang, Y.S., Hung, C.H., and Wei, L.N. (2011). Cytoplasmic
receptor-interacting protein 140 (RIP140) interacts with perilipin to regulate
lipolysis. Cell. Signal. 23, 1396–1403.
Holm, C., Osterlund, T., Laurell, H., and Contreras, J.A. (2000). Molecular
mechanisms regulating hormone-sensitive lipase and lipolysis. Annu. Rev.
Nutr. 20, 365–393.
Holst, L.S., Langin, D., Mulder, H., Laurell, H., Grober, J., Bergh, A., Mohren-
weiser, H.W., Edgren, G., and Holm, C. (1996). Molecular cloning, genomic
organization, and expression of a testicular isoform of hormone-sensitive
lipase. Genomics 35, 441–447.
Huang, Y., Cohen, J.C., and Hobbs, H.H. (2011). Expression and characteriza-
tion of a PNPLA3 isoform (I148M) associated with nonalcoholic fatty liver
disease. J. Biol Chem. 286, 37085–37093.
Jenkins, C.M., Mancuso, D.J., Yan, W., Sims, H.F., Gibson, B., and Gross,
R.W. (2004). Identification, cloning, expression, and purification of three novel
human calcium-independent phospholipase A2 family members possessing
triacylglycerol lipase and acylglycerol transacylase activities. J. Biol. Chem.
279, 48968–48975.
Kelley, D.E., Goodpaster, B.H., and Storlien, L. (2002). Muscle triglyceride and
insulin resistance. Annu. Rev. Nutr. 22, 325–346.
Kershaw, E.E., Hamm, J.K., Verhagen, L.A., Peroni, O., Katic, M., and Flier,
J.S. (2006). Adipose triglyceride lipase: function, regulation by insulin, and
comparison with adiponutrin. Diabetes 55, 148–157.
Kienesberger, P.C., Lee, D., Pulinilkunnil, T., Brenner, D.S., Cai, L., Magnes,
C., Koefeler, H.C., Streith, I.E., Rechberger, G.N., Haemmerle, G., et al.
(2009a). Adipose triglyceride lipase deficiency causes tissue-specific changes
in insulin signaling. J. Biol. Chem. 284, 30218–30229.
Kienesberger, P.C., Oberer, M., Lass, A., and Zechner, R. (2009b). Mammalian
patatin domain containing proteins: a family with diverse lipolytic activities
involved in multiple biological functions. J. Lipid Res. Suppl. 50, S63–S68.
Kralisch, S., Klein, J., Lossner, U., Bluher, M., Paschke, R., Stumvoll, M., and
Fasshauer, M. (2005). Isoproterenol, TNFalpha, and insulin downregulate
adipose triglyceride lipase in 3T3-L1 adipocytes. Mol. Cell. Endocrinol. 240,
43–49.
Krawczyk, M., Gru
¨nhage, F., Zimmer, V., and Lammert, F. (2011). Variant adi-
ponutrin (PNPLA3) represents a common fibrosis risk gene: non-invasive elas-
tography-based study in chronic liver disease. J. Hepatol. 55, 299–306.
Kurat, C.F., Natter, K., Petschnigg, J., Wolinski, H., Scheuringer, K., Scholz, H.,
Zimmermann, R., Leber, R., Zechner, R., and Kohlwein, S.D. (2006). Obese
yeast: triglyceride lipolysis is functionally conserved from mammals to yeast.
J. Biol. Chem. 281, 491–500.
Kurat, C.F., Wolinski, H., Petschnigg, J., Kaluarachchi, S., Andrews, B., Natter,
K., and Kohlwein, S.D. (2009). Cdk1/Cdc28-dependent activation of the major
Cell Metabolism 15, March 7, 2012 ª2012 Elsevier Inc. 289
Cell Metabolism
Review
triacylglycerol lipase Tgl4 in yeast links lipolysis to cell-cycle progression. Mol.
Cell 33, 53–63.
Labar, G., Bauvois, C., Borel, F., Ferrer, J.L., Wouters, J., and Lambert, D.M.
(2010). Crystal structure of the human monoacylglycerol lipase, a key actor
in endocannabinoid signaling. ChemBioChem 11, 218–227.
Lake, A.C., Sun, Y., Li, J.L., Kim, J.E., Johnson, J.W., Li, D., Revett, T., Shih,
H.H., Liu, W., Paulsen, J.E., and Gimeno, R.E. (2005). Expression, regulation,
and triglyceride hydrolase activity of Adiponutrin family members. J. Lipid Res.
46, 2477–2487.
Langerhans, R. (1890). U
¨ber multiple Fettgewebsnekrose. Virchows Arch. 122,
252–270.
Langin, D., Laurell, H., Holst, L.S., Belfrage, P., and Holm, C. (1993). Gene
organization and primary structure of human hormone-sensitive lipase:
possible significance of a sequence homology with a lipase of Moraxella
TA144, an antarctic bacterium. Proc. Natl. Acad. Sci. USA 90, 4897–4901.
Lass, A., Zimmermann, R., Haemmerle, G., Riederer, M., Schoiswohl, G.,
Schweiger, M., Kienesberger, P., Strauss, J.G., Gorkiewicz, G., and Zechner,
R. (2006). Adipose triglyceride lipase-mediated lipolysis of cellular fat stores is
activated by CGI-58 and defective in Chanarin-Dorfman Syndrome. Cell
Metab. 3, 309–319.
Lass, A., Zimmermann, R., Oberer, M., and Zechner, R. (2011). Lipolysis -
a highly regulated multi-enzyme complex mediates the catabolism of cellular
fat stores. Prog. Lipid Res. 50, 14–27.
Lefe
`vre, C., Jobard, F., Caux, F., Bouadjar, B., Karaduman, A., Heilig, R.,
Lakhdar, H., Wollenberg, A., Verret, J.L., Weissenbach, J., et al. (2001). Muta-
tions in CGI-58, the gene encoding a new protein of the esterase/lipase/
thioesterase subfamily, in Chanarin-Dorfman syndrome. Am. J. Hum. Genet.
69, 1002–1012.
Levine, B., and Kroemer, G. (2008). Autophagy in the pathogenesis of disease.
Cell 132, 27–42.
Lichtman, A.H., and Martin, B.R. (2005). Cannabinoid tolerance and depen-
dence. Handb Exp Pharmacol. 168, 691–717.
Listenberger, L.L., Ostermeyer-Fay, A.G., Goldberg, E.B., Brown, W.J., and
Brown, D.A. (2007). Adipocyte differentiation-related protein reduces the lipid
droplet association of adipose triglyceride lipase and slows triacylglycerol
turnover. J. Lipid Res. 48, 2751–2761.
Long, J.Z., Li, W., Booker, L., Burston, J.J., Kinsey, S.G., Schlosburg, J.E.,
Pavo
´n, F.J., Serrano, A.M., Selley, D.E., Parsons, L.H., et al. (2009). Selective
blockade of 2-arachidonoylglycerol hydrolysis produces cannabinoid behav-
ioral effects. Nat. Chem. Biol. 5, 37–44.
Lu, X., Yang, X., and Liu, J. (2010). Differential control of ATGL-mediated lipid
droplet degradation by CGI-58 and G0S2. Cell Cycle 9, 2719–2725.
Miyoshi, H., Perfield, J.W., 2nd, Souza, S.C., Shen, W.J., Zhang, H.H.,
Stancheva, Z.S., Kraemer, F.B., Obin, M.S., and Greenberg, A.S. (2007).
Control of adipose triglyceride lipase action by serine 517 of perilipin A globally
regulates protein kinase A-stimulated lipolysis in adipocytes. J. Biol. Chem.
282, 996–1002.
Montero-Moran, G., Caviglia, J.M., McMahon, D., Rothenberg, A., Subrama-
nian, V., Xu, Z., Lara-Gonzalez, S., Storch, J., Carman, G.M., and Brasaemle,
D.L. (2010). CGI-58/ABHD5 is a coenzyme A-dependent lysophosphatidic
acid acyltransferase. J. Lipid Res. 51, 709–719.
Mulder, H., So
¨rhede-Winzell, M., Contreras, J.A., Fex, M., Stro
¨m, K., Ploug, T.,
Galbo, H., Arner, P., Lundberg, C., Sundler, F., et al. (2003). Hormone-sensitive
lipase null mice exhibit signs of impaired insulin sensitivity whereas insulin
secretion is intact. J. Biol. Chem. 278, 36380–36388.
Narbonne, P., and Roy, R. (2009). Caenorhabditis elegans dauers need
LKB1/AMPK to ration lipid reserves and ensure long-term survival. Nature
457, 210–214.
Nomura, D.K., Long, J.Z., Niessen, S., Hoover, H.S., Ng, S.W., and Cravatt,
B.F. (2010). Monoacylglycerol lipase regulates a fatty acid network that
promotes cancer pathogenesis. Cell 140, 49–61.
Notari, L., Baladron, V., Aroca-Aguilar, J.D., Balko, N., Heredia, R., Meyer, C.,
Notario, P.M., Saravanamuthu, S., Nueda, M.L., Sanchez-Sanchez, F., et al.
(2006). Identification of a lipase-linked cell membrane receptor for pigment
epithelium-derived factor. J. Biol. Chem. 281,38022–38037.
Ong, K.T., Mashek, M.T., Bu, S.Y., Greenberg, A.S., and Mashek, D.G. (2011).
Adipose triglyceride lipase is a major hepatic lipase that regulates triacylgly-
cerol turnover and fatty acid signaling and partitioning. Hepatology 53,
116–126.
Osuga, J., Ishibashi, S., Oka, T., Yagyu, H., Tozawa, R., Fujimoto, A., Shionoiri,
F., Yahagi, N., Kraemer, F.B., Tsutsumi, O., and Yamada, N. (2000). Targeted
disruption of hormone-sensitive lipase results in male sterility and adipocyte
hypertrophy, but not in obesity. Proc. Natl. Acad. Sci. USA 97, 787–792.
Park, S.Y., Kim, H.J., Wang, S., Higashimori, T., Dong, J., Kim, Y.J., Cline, G.,
Li, H., Prentki, M., Shulman, G.I., et al. (2005). Hormone-sensitive lipase
knockout mice have increased hepatic insulin sensitivity and are protected
from short-term diet-induced insulin resistance in skeletal muscle and heart.
Am. J. Physiol. Endocrinol. Metab. 289, E30–E39.
Perkins, J.M., and Davis, S.N. (2008). Endocannabinoid system overactivity
and the metabolic syndrome: prospects for treatment. Curr. Diab. Rep. 8,
12–19.
Peyot, M.L., Guay, C., Latour, M.G., Lamontagne, J., Lussier, R., Pineda, M.,
Ruderman, N.B., Haemmerle, G., Zechner, R., Joly, E., et al. (2009). Adipose
triglyceride lipase is implicated in fuel- and non-fuel-stimulated insulin secre-
tion. J. Biol. Chem. 284, 16848–16859.
Radner, F.P., Streith, I.E., Schoiswohl, G., Schweiger, M., Kumari, M.,
Eichmann, T.O., Rechberger, G., Koefeler, H.C., Eder, S., Schauer, S., et al.
(2010). Growth retardation, impaired triacylglycerol catabolism, hepatic stea-
tosis, and lethal skin barrier defect in mice lacking comparative gene identifi-
cation-58 (CGI-58). J. Biol. Chem. 285, 7300–7311.
Reid, B.N., Ables, G.P., Otlivanchik, O.A., Schoiswohl, G., Zechner, R., Blaner,
W.S., Goldberg, I.J., Schwabe, R.F., Chua, S.C., Jr., and Huang, L.S. (2008).
Hepatic overexpression of hormone-sensitive lipase and adipose triglyceride
lipase promotes fatty acid oxidation, stimulates direct release of free fatty
acids, and ameliorates steatosis. J. Biol. Chem. 283, 13087–13099.
Rodriguez, J.A., Ben Ali, Y., Abdelkafi, S., Mendoza, L.D., Leclaire, J., Fotiadu,
F., Buono, G., Carrie
`re, F., and Abousalham, A. (2010). In vitro stereoselective
hydrolysis of diacylglycerols by hormone-sensitive lipase. Biochim. Biophys.
Acta 1801, 77–83.
Romeo, S., Kozlitina, J., Xing, C., Pertsemlidis, A., Cox, D., Pennacchio, L.A.,
Boerwinkle, E., Cohen, J.C., and Hobbs, H.H. (2008). Genetic variation in
PNPLA3 confers susceptibility to nonalcoholic fatty liver disease. Nat. Genet.
40, 1461–1465.
Rydel, T.J., Williams, J.M., Krieger, E., Moshiri, F., Stallings, W.C., Brown,
S.M., Pershing, J.C., Purcell, J.P., and Alibhai, M.F. (2003). The crystal struc-
ture, mutagenesis, and activity studies reveal that patatin is a lipid acyl hydro-
lase with a Ser-Asp catalytic dyad. Biochemistry 42, 6696–6708.
Ryde
´n, M., Jocken, J., van Harmelen, V., Dicker, A., Hoffstedt, J., Wire
´n, M.,
Blomqvist, L., Mairal, A., Langin, D., Blaak, E., and Arner, P. (2007). Compar-
ative studies of the role of hormone-sensitive lipase and adipose triglyceride
lipase in human fat cell lipolysis. Am. J. Physiol. Endocrinol. Metab. 292,
E1847–E1855.
Ryde
´n, M., Agustsson, T., Laurencikiene, J., Britton, T., Sjo
¨lin, E., Isaksson, B.,
Permert, J., and Arner, P. (2008). Lipolysis—not inflammation, cell death, or
lipogenesis—is involved in adipose tissue loss in cancer cachexia. Cancer
113, 1695–1704.
Sakurada, T., and Noma, A. (1981). Subcellular localization and some proper-
ties of monoacylglycerol lipase in rat adipocytes. J. Biochem. 90, 1413–1419.
Samuel, V.T., Petersen, K.F., and Shulman, G.I. (2010). Lipid-induced insulin
resistance: unravelling the mechanism. Lancet 375, 2267–2277.
Scherer, T., O’Hare, J., Diggs-Andrews, K., Schweiger, M., Cheng, B., Lindt-
ner, C., Zielinski, E., Vempati, P., Su, K., Dighe, S., et al. (2011). Brain insulin
controls adipose tissue lipolysis and lipogenesis. Cell Metab. 13, 183–194.
Schlosburg, J.E., Blankman, J.L., Long, J.Z., Nomura, D.K., Pan, B., Kinsey,
S.G., Nguyen, P.T., Ramesh, D., Booker, L., Burston, J.J., et al. (2010). Chronic
monoacylglycerol lipase blockade causes functional antagonism of the endo-
cannabinoid system. Nat. Neurosci. 13, 1113–1119.
Schweiger, M., Schreiber, R., Haemmerle, G., Lass, A., Fledelius, C., Jacob-
sen, P., Tornqvist, H., Zechner, R., and Zimmermann, R. (2006). Adipose
triglyceride lipase and hormone-sensitive lipase are the major enzymes in
adipose tissue triacylglycerol catabolism. J. Biol. Chem. 281, 40236–40241.
290 Cell Metabolism 15, March 7, 2012 ª2012 Elsevier Inc.
Cell Metabolism
Review
Schweiger, M., Schoiswohl, G., Lass, A., Radner, F.P., Haemmerle, G., Malli,
R., Graier, W., Cornaciu, I., Oberer, M., Salvayre, R., et al. (2008). The
C-terminal region of human adipose triglyceride lipase affects enzyme activity
and lipid droplet binding. J. Biol. Chem. 283, 17211–17220.
Shen, W.J., Patel, S., Miyoshi, H., Greenberg, A.S., and Kraemer, F.B. (2009).
Functional interaction of hormone-sensitive lipase and perilipin in lipolysis.
J. Lipid Res. 50, 2306–2313.
Shen, W.J., Yu, Z., Patel, S., Jue, D., Liu, L.F., and Kraemer, F.B. (2011).
Hormone-sensitive lipase modulates adipose metabolism through PPARg.
Biochim. Biophys. Acta 1811, 9–16.
Shibata, M., Yoshimura, K., Furuya, N., Koike, M., Ueno, T., Komatsu, M., Arai,
H., Tanaka, K., Kominami, E., and Uchiyama, Y. (2009). The MAP1-LC3 conju-
gation system is involved in lipid droplet formation. Biochem. Biophys. Res.
Commun. 382, 419–423.
Shibata, M., Yoshimura, K., Tamura, H., Ueno, T., Nishimura, T., Inoue, T.,
Sasaki, M., Koike, M., Arai, H., Kominami, E., and Uchiyama, Y. (2010). LC3,
a microtubule-associated protein1A/B light chain3, is involved in cytoplasmic
lipid droplet formation. Biochem. Biophys. Res. Commun. 393, 274–279.
Singh, R., Kaushik, S., Wang, Y., Xiang, Y., Novak, I., Komatsu, M., Tanaka, K.,
Cuervo, A.M., and Czaja, M.J. (2009a). Autophagy regulates lipid metabolism.
Nature 458, 1131–1135.
Singh, R., Xiang, Y., Wang, Y., Baikati, K., Cuervo, A.M., Luu, Y.K., Tang, Y.,
Pessin, J.E., Schwartz, G.J., and Czaja, M.J. (2009b). Autophagy regulates
adipose mass and differentiation in mice. J. Clin. Invest. 119, 3329–3339.
Soni, K.G., Mardones, G.A., Sougrat, R., Smirnova, E., Jackson, C.L., and
Bonifacino, J.S. (2009). Coatomer-dependent protein delivery to lipid droplets.
J. Cell Sci. 122, 1834–1841.
Stra
˚lfors, P., and Belfrage, P. (1983). Phosphorylation of hormone-sensitive
lipase by cyclic AMP-dependent protein kinase. J. Biol. Chem. 258, 15146–
15152.
Summers, S.A. (2010). Sphingolipids and insulin resistance: the five Ws. Curr.
Opin. Lipidol. 21, 128–135.
Tang, Y., Chen, Y., Jiang, H., and Nie, D. (2011). Short-chain fatty acids
induced autophagy serves as an adaptive strategy for retarding mitochon-
dria-mediated apoptotic cell death. Cell Death Differ. 18, 602–618.
Taschler, U., Radner, F.P., Heier, C., Schreiber, R., Schweiger, M., Schois-
wohl, G., Preiss-Landl, K., Jaeger, D., Reiter, B., Koefeler, H.C., et al. (2011).
Monoglyceride lipase deficiency in mice impairs lipolysis and attenuates
diet-induced insulin resistance. J. Biol. Chem. 286, 17467–17477.
Tian, C., Stokowski, R.P., Kershenobich, D., Ballinger, D.G., and Hinds, D.A.
(2010). Variant in PNPLA3 is associated with alcoholic liver disease. Nat.
Genet. 42, 21–23.
Turban, S., and Hajduch, E. (2011). Protein kinase C isoforms: mediators of
reactive lipid metabolites in the development of insulin resistance. FEBS
Lett. 585, 269–274.
Unger, R.H., Clark, G.O., Scherer, P.E., and Orci, L. (2010). Lipid homeostasis,
lipotoxicity and the metabolic syndrome. Biochim. Biophys. Acta 1801, 209–
214.
Vallerie, S.N., and Hotamisligil, G.S. (2010). The role of JNK proteins in metab-
olism. Sci. Transl. Med. 2, 60rv65.
Vaughan, M., Berger, J.E., and Steinberg, D. (1964). Hormone-Sensitive
Lipase and Monoglyceride Lipase Activities in Adipose Tissue. J. Biol.
Chem. 239, 401–409.
Villena, J.A., Roy, S., Sarkadi-Nagy, E., Kim, K.H., and Sul, H.S. (2004). Desnu-
trin, an adipocyte gene encoding a novel patatin domain-containing protein, is
induced by fasting and glucocorticoids: ectopic expression of desnutrin
increases triglyceride hydrolysis. J. Biol. Chem. 279, 47066–47075.
Voshol, P.J., Haemmerle, G., Ouwens, D.M., Zimmermann, R., Zechner, R.,
Teusink, B., Maassen, J.A., Havekes, L.M., and Romijn, J.A. (2003). Increased
hepatic insulin sensitivity together with decreased hepatic triglyceride stores in
hormone-sensitive lipase-deficient mice. Endocrinology 144, 3456–3462.
Wang, H., Hu, L., Dalen, K., Dorward, H., Marcinkiewicz, A., Russell, D., Gong,
D., Londos, C., Yamaguchi, T., Holm, C., et al. (2009). Activation of hormone-
sensitive lipase requires two steps, protein phosphorylation and binding to the
PAT-1 domain of lipid droplet coat proteins. J. Biol. Chem. 284, 32116–32125.
Wang, H., Bell, M., Sreenevasan, U., Hu, H., Liu, J., Dalen, K., Londos, C.,
Yamaguchi, T., Rizzo, M.A., Coleman, R., et al. (2011). Unique regulation of
adipose triglyceride lipase (ATGL) by perilipin 5, a lipid droplet-associated
protein. J. Biol. Chem. 286, 15707–15715.
Watt, M.J., Holmes, A.G., Pinnamaneni, S.K., Garnham, A.P., Steinberg, G.R.,
Kemp, B.E., and Febbraio, M.A. (2006). Regulation of HSL serine phosphory-
lation in skeletal muscle and adipose tissue. Am. J. Physiol. Endocrinol. Metab.
290, E500–E508.
Wei, E., Ben Ali, Y., Lyon, J., Wang, H., Nelson, R., Dolinsky, V.W., Dyck, J.R.,
Mitchell, G., Korbutt, G.S., and Lehner, R. (2010). Loss of TGH/Ces3 in mice
decreases blood lipids, improves glucose tolerance, and increases energy
expenditure. Cell Metab. 11, 183–193.
Welch, C., Santra, M.K., El-Assaad, W., Zhu, X., Huber, W.E., Keys, R.A.,
Teodoro, J.G., and Green, M.R. (2009). Identification of a protein, G0S2, that
lacks Bcl-2 homology domains and interacts with and antagonizes Bcl-2.
Cancer Res. 69, 6782–6789.
Whitehead, R.H. (1909). A note on the absorption of fat. Am. J. Physiol. 24,
294–296.
Wilson, P.A., Gardner, S.D., Lambie, N.M., Commans, S.A., and Crowther,
D.J. (2006). Characterization of the human patatin-like phospholipase family.
J. Lipid Res. 47, 1940–1949.
Wu, J.W., Wang, S.P., Alvarez, F., Casavant, S., Gauthier, N., Abed, L., Soni,
K.G., Yang, G., and Mitchell, G.A. (2011). Deficiency of liver adipose triglyc-
eride lipase in mice causes progressive hepatic steatosis. Hepatology 54,
122–132.
Yang, L., Li, P., Fu, S., Calay, E.S., and Hotamisligil, G.S. (2010a). Defective
hepatic autophagy in obesity promotes ER stress and causes insulin resis-
tance. Cell Metab. 11, 467–478.
Yang, X., Lu, X., Lombe
`s, M., Rha, G.B., Chi, Y.I., Guerin, T.M., Smart, E.J., and
Liu, J. (2010b). The G(0)/G(1) switch gene 2 regulates adipose lipolysis through
association with adipose triglyceride lipase. Cell Metab. 11, 194–205.
Yin, W., Mu, J., and Birnbaum, M.J. (2003). Role of AMP-activated protein
kinase in cyclic AMP-dependent lipolysis In 3T3-L1 adipocytes. J. Biol.
Chem. 278, 43074–43080.
Yuan, X., Waterworth, D., Perry, J.R., Lim, N., Song, K., Chambers, J.C.,
Zhang, W., Vollenweider, P., Stirnadel, H., Johnson, T., et al. (2008). Popula-
tion-based genome-wide association studies reveal six loci influencing plasma
levels of liver enzymes. Am. J. Hum. Genet. 83, 520–528.
Zandbergen, F., Mandard, S., Escher, P., Tan, N.S., Patsouris, D., Jatkoe, T.,
Rojas-Caro, S., Madore, S., Wahli, W., Tafuri, S., et al. (2005). The G0/G1
switch gene 2 is a novel PPAR target gene. Biochem. J. 392, 313–324.
Zechner, R., and Madeo, F. (2009). Cell biology: Another way to get rid of fat.
Nature 458, 1118–1119.
Zimmermann, R., Haemmerle, G., Wagner, E.M., Strauss, J.G., Kratky, D., and
Zechner, R. (2003). Decreased fatty acid esterification compensates for the
reduced lipolytic activity in hormone-sensitive lipase-deficient white adipose
tissue. J. Lipid Res. 44, 2089–2099.
Zimmermann, R., Strauss, J.G., Haemmerle, G., Schoiswohl, G., Birner-
Gruenberger, R., Riederer, M., Lass, A., Neuberger, G., Eisenhaber, F.,
Hermetter, A., and Zechner, R. (2004). Fat mobilization in adipose tissue is
promoted by adipose triglyceride lipase. Science 306, 1383–1386.
Cell Metabolism 15, March 7, 2012 ª2012 Elsevier Inc. 291
Cell Metabolism
Review
... Obesity, characterized as excess adipose tissue accumulation, has become a serious global health issue because of its induced risk of metabolic diseases, immune dysfunctions, cardiovascular diseases, and cancers (1)(2)(3). Proverbially, adipose tissues can be divided into two types: white adipose tissue (WAT) mainly for the storage of fat, and brown adipose tissue (BAT) for the generation of energy by nonshivering thermogenesis (4)(5)(6). Increasing evidence also indicates that white adipocytes can be differentiate into beige adipocytes containing a small number of mitochondria (7). ...
... Adipose triglyceride lipase (ATGL), hormonesensitive lipase (HSL), and monoglyceride lipase (MG) work together to hydrolyze triacylglycerol in adipose tissue, which is then released into the bloodstream. Part of the fatty acids are oxidized by beta oxidation and then the products go into the tricarboxylic acid cycle to produce ATP (5,37). ...
Article
Full-text available
Interactions between macrophages and adipocytes in adipose tissue are critical for the regulation of energy metabolism and obesity. Macrophage polarization induced by cold or other stimulations can drive metabolic reprogramming of adipocytes, browning, and thermogenesis. Accordingly, investigating the roles of macrophages and adipocytes in the maintenance of energy homeostasis is critical for the development of novel therapeutic approaches specifically targeting macrophages in metabolic disorders such as obesity. Current review outlines macrophage polarization not only regulates the release of central nervous system and inflammatory factors, but controls mitochondrial function, and other factor that induce metabolic reprogramming of adipocytes and maintain energy homeostasis. We also emphasized on how the adipocytes conversely motivate the polarization of macrophage. Exploring the interactions between adipocytes and macrophages may provide new therapeutic strategies for the management of obesity-related metabolic diseases.
... Kilka lat później Balser, Langerhans i Flexner scharakteryzowali "ferment" trzustkowy, który rozszczepiał tłuszcz. Z kolei w 1909 roku Whitehead udowodnił, że tłuszcz nie jest w stanie dostać się do komórki w postaci niezhydrolizowanej [2]. W następstwie opisanych wyżej odkryć świat naukowy potwierdził powszechne występowanie lipaz w mikroorganizmach, organizmach roślinnych i zwierzęcych. ...
Article
Full-text available
Lipases are enzymes commonly found in microorganisms, fungi, plants and animals. Their main function in cell metabolism is the hydrolysis (lipolysis) of ester bonds between fatty acids and glycerol in mono-, di- and triacylglycerols. In plants, lipases play an important role in ontogeny, participating in both vegetative development and generative stages. These enzymes may also be a component of plant responses to biotic and abiotic stresses. Based on the similarity of the amino acid sequence and vacuolar localization of some plant lipases to yeast Atg15, we present a hypothesis about the participation of lipases in autophagy (precisely, in the degradation of the autophagic body) in plants. Despite the narrow substrate specificity and the type of reactions catalysed in cells, lipases find numerous biotechnological applications. The physicochemical features of lipases, which determine, for example, wide substrate specificity in vitro or high stability in a wide range of pH and temperature, make these enzymes the subject of applied research, and plant lipases show an increasing potential in this area of science and industry.
... The sequential action of three major hydrolases, namely adipose triglyceride lipase (ATGL), hormone-sensitive lipase (HSL), and monoacylglycerol lipase (MGL), catalyzes the neutral hydrolysis of TGs. Lipolysis (Fig. 2B) consists of ATGL initiating the process by catalyzing TGs to diacylglycerols (DGs) and mediating triglyceride hydrolysis during basal lipolysis; HSL is mainly responsible for the hydrolysis of DGs to monoacylglycerols (MGs) and MGL hydrolyzes MGs (Zechner et al., 2012;Zimmermann et al., 2004). Therefore, enhancing lipolysis enzyme activity in adipose tissue is a metabolic pathway to be investigated for pharmacological strategies against obesity. ...
Article
Full-text available
Obesity, a major public health problem, causes numerous complications that threaten human health and increase the socioeconomic burden. The pathophysiology of obesity is primarily attributed to lipid metabolism disorders. Conventional anti-obesity medications have a high abuse potential and frequently deliver insufficient efficacy and have negative side-effects. Hence, functional foods are regarded as effective alternatives to address obesity. Coffee, tea, and cocoa, three widely consumed beverages, have long been considered to have the potential to prevent obesity, and several studies have focused on their intrinsic molecular mechanisms in past few years. Therefore, in this review, we discuss the mechanisms by which the bioactive ingredients in these three beverages counteract obesity from the aspects of adipogenesis, lipolysis, and energy expenditure (thermogenesis). The future prospects and challenges for coffee, tea, and cocoa as functional products for the treatment of obesity are also discussed, which can be pursued for future drug development and prevention strategies against obesity.
... The primary cause of obesity stems from an imbalance between energy intake expenditure, accumulating excess energy in the form of triglycerides. This results in an excessive buildup of lipid and adipose tissue enlargement [6]. The proper remodeling adipose tissue hinges on a delicate equilibrium between de novo adipogenesis, where mesenchymal stem cells differentiate into new adipocytes, and the enlargement of existing adipocytes (hypertrophy) [7][8][9][10]. ...
Article
Full-text available
Obesity stands as a significant risk factor for type 2 diabetes, hyperlipidemia, and cardiovascular diseases, intertwining increased inflammation and decreased adipogenesis with metabolic disorders. Studies have highlighted the correlation between Caspase-1 and inflammation in obesity, elucidating its essential role in the biological functions of adipose tissue. However, the impact of Caspase-1 on adipogenesis and the underlying mechanisms remain largely elusive. In our study, we observed a positive correlation between Caspase-1 expression and obesity and its association with adipogenesis. In vivo experiments revealed that, under normal diet conditions, Caspase-1 deficiency improved glucose homeostasis, stimulated subcutaneous adipose tissue expansion, and enhanced adipogenesis. Furthermore, our findings indicate that Caspase-1 deficiency promotes the expression of autophagy-related proteins and inhibits autophagy with 3-MA or CQ blocked Caspase-1 deficiency-induced adipogenesis in vitro. Notably, Caspase-1 deficiency promotes adipogenesis via Atg7-mediated autophagy activation. In addition, Caspase-1 deficiency resisted against high-fat diet-induced obesity and glucose intolerance. Our study proposes the downregulation of Caspase-1 as a promising strategy for mitigating obesity and its associated metabolic disorders.
... Excess dietary fatty acids are typically converted into neutral triglycerides, which are stored in adipose tissue. Upon receiving catabolic stimuli such as starvation and exercise, triglycerides undergo sequential hydrolysis by adipose triglyceride lipase (ATGL), hormone-sensitive lipase (HSL) and monoglyceride lipase (MGL) to liberate free fatty acids into the circulation [97]. Fatty acids are transported into myocytes through various mechanisms, including muscle lipoprotein lipase (mLPL), cluster of differentiation 36 (CD36), fatty acid banding proteins (FABPs) and fatty acid transport proteins (FATPs) [98]. ...
... Inhibition of SREBPs with inhibitor fotostatin increases apoptosis and inhibits cell growth in AR-negative prostate cancer [161]. Similarly, pharmacological or genetic inhibition of the SREBP pathway decreases the progression of hepatocellular carcinoma [162]. Treatment with an agonist of PPARγ-suppressed lipogenic pathway in CSCs leads to a reduction in the expression of lipogenic enzymes, including ACLY, FASN, NR1D1, and MIG12 [163]. ...
Article
Full-text available
Lipids are the key component of all membranes composed of a variety of molecules that transduce intracellular signaling and provide energy to the cells in the absence of nutrients. Alteration in lipid metabolism is a major factor for cancer heterogeneity and a newly identified cancer hallmark. Reprogramming of lipid metabolism affects the diverse cancer phenotypes, especially epithelial–mesenchymal transition (EMT). EMT activation is considered to be an essential step for tumor metastasis, which exhibits a crucial role in the biological processes including development, wound healing, and stem cell maintenance, and has been widely reported to contribute pathologically to cancer progression. Altered lipid metabolism triggers EMT and activates multiple EMT-associated oncogenic pathways. Although the role of lipid metabolism-induced EMT in tumorigenesis is an attractive field of research, there are still significant gaps in understanding the underlying mechanisms and the precise contributions of this interplay. Further study is needed to clarify the specific molecular mechanisms driving the crosstalk between lipid metabolism and EMT, as well as to determine the potential therapeutic implications. The increased dependency of tumor cells on lipid metabolism represents a novel therapeutic target, and targeting altered lipid metabolism holds promise as a strategy to suppress EMT and ultimately inhibit metastasis.
Article
Full-text available
Hormone-sensitive lipase, detergent-solubilized and purified from rat adipose tissue, was phosphorylated with the catalytic subunit of cyclic AMP-dependent protein kinase from the same tissue. Maximally 1.05 +/- 0.05 (mean +/- S.E. (n = 3) ) mol of phosphate/mol of hormone-sensitive lipase Mr = 84,000 subunit was incorporated. Phosphoserine was the only phosphorylated amino acid residue. A single phosphorylation site was demonstrated by digestion with Staphylococcus aureus V8 protease and trypsin that produced a single acidic phosphopeptide of about 10 amino acid residues length, which was isolated by two-dimensional electrophoresis-thin layer chromatography. Enzyme activity was enhanced, 2.5-fold against trioleoylglycerol, concomitant with phosphorylation, with half-maximal effect within 30 sec, a rate of phosphorylation of the enzyme comparable to that obtained in vivo (Nilsson, N. O., Strålfors, P., Fredrikson, G., and Belfrage, P. (1980) FEBS Lett. 111, 125-130). The initial rate of phosphorylation was approximately half that with phosphorylase kinase as substrate. The effects of modifications of hormone-sensitive lipase and of various additions and variation in pH were examined.
Article
Full-text available
A genetic variant of PNPLA3 (patatin-like phospholipase domain-containing 3; PNPLA3-I148M), a serine protease of unknown function, is associated with accumulation of triacylglycerol (TAG) in the liver. To determine the biological substrates of PNPLA3 and the effect of the I148M substitution on enzymatic activity and substrate specificity, we purified and characterized recombinant human PNPLA3 and PNPLA3-I148M. Maximal hydrolytic activity of PNPLA3 was observed against the three major glycerolipids, TAG, diacylglycerol, and monoacylglycerol, with a strong preference for oleic acid as the acyl moiety. Substitution of methionine for isoleucine at position 148 markedly decreased the V(max) of the enzyme for glycerolipids but had only a modest effect on the K(m). Purified PNPLA3 also catalyzed the hydrolysis of oleoyl-CoA, but the V(max) was 100-fold lower for oleoyl-CoA than for triolein. The thioesterase activity required the catalytic serine but was only modestly decreased by the I148M substitution. The enzyme had little or no hydrolytic activity against the other lipid substrates tested, including phospholipids, cholesteryl ester, and retinyl esters. Neither the wild-type nor mutant enzyme catalyzed transfer of oleic acid from oleoyl-CoA to glycerophosphate, lysophosphatidic acid, or diacylglycerol, suggesting that the enzyme does not promote de novo TAG synthesis. Taken together, our results are consistent with the notion that PNPLA3 plays a role in the hydrolysis of glycerolipids and that the I148M substitution causes a loss of function, although we cannot exclude the possibility that the enzyme has additional substrates or activities.
Article
Full-text available
Peroxisome proliferator-activated receptors (PPARs) are nuclear hormone receptors that regulate genes involved in energy metabolism and inflammation. For biological activity, PPARs require cognate lipid ligands, heterodimerization with retinoic X receptors, and coactivation by PPAR-γ coactivator-1α or PPAR-γ coactivator-1β (PGC-1α or PGC-1β, encoded by Ppargc1a and Ppargc1b, respectively). Here we show that lipolysis of cellular triglycerides by adipose triglyceride lipase (patatin-like phospholipase domain containing protein 2, encoded by Pnpla2; hereafter referred to as Atgl) generates essential mediator(s) involved in the generation of lipid ligands for PPAR activation. Atgl deficiency in mice decreases mRNA levels of PPAR-α and PPAR-δ target genes. In the heart, this leads to decreased PGC-1α and PGC-1β expression and severely disrupted mitochondrial substrate oxidation and respiration; this is followed by excessive lipid accumulation, cardiac insufficiency and lethal cardiomyopathy. Reconstituting normal PPAR target gene expression by pharmacological treatment of Atgl-deficient mice with PPAR-α agonists completely reverses the mitochondrial defects, restores normal heart function and prevents premature death. These findings reveal a potential treatment for the excessive cardiac lipid accumulation and often-lethal cardiomyopathy in people with neutral lipid storage disease, a disease marked by reduced or absent ATGL activity.
Article
Plasma liver-enzyme tests are widely used in the clinic for the diagnosis of liver diseases and for monitoring the response to drug treat- ment. There is considerable evidence that human genetic variation influences plasma levels of liver enzymes. However, such genetic variation has not been systematically assessed. In the present study, we performed a genome-wide association study of plasma liver-en- zyme levels in three populations (total n ¼ 7715) with replication in three additional cohorts (total n ¼ 4704). We identified two loci influencing plasma levels of alanine-aminotransferase (ALT) (CPN1-ERLIN1-CHUK on chromosome 10 and PNPLA3-SAMM50 on chro- mosome 22), one locus influencing gamma-glutamyl transferase (GGT) levels (HNF1A on chromosome 12), and three loci for alkaline phosphatase (ALP) levels (ALPL on chromosome 1, GPLD1 on chromosome 6, and JMJD1C-REEP3 on chromosome 10). In addition, we confirmed the associations between the GGT1 locus and GGT levels and between the ABO locus and ALP levels. None of the ALP-asso- ciated SNPs were associated with other liver tests, suggesting intestine and/or bone specificity. The mechanisms underlying the associ- ations may involve cis- or trans-transcriptional effects (some of the identified variants were associated with mRNA transcription in hu- man liver or lymphoblastoid cells), dysfunction of the encoded proteins (caused by missense variations at the functional domains), or other unknown pathways. These findings may help in the interpretation of liver-enzyme tests and provide candidate genes for liver dis- eases of viral, metabolic, autoimmune, or toxic origin. The specific associations with ALP levels may point to genes for bone or intestinal diseases. Plasma liver-enzyme tests are widely used in the clinic to identify patients with liver diseases, to monitor the course and severity of these diseases and the effect of therapies, and to detect drug-induced liver injury.1,2 These tests also have substantial epidemiologic significance that extends beyond the liver, given that they have been shown to be prospective risk factors for type 2 diabetes, cardiovascular disease, and all-cause mortality in multiple large stud- ies.3-6 Therefore, it is of interest to identify genes or loci af- fecting these markers in order to establish whether such loci are also associated with these clinical endpoints. Plasma liver-enzyme levels are influenced by environ- mental and genetic factors. The estimated heritabilities range from 33% for alanine-aminotransferase (ALT) to 61% for gamma-glutamyl transferase (GGT).7,8 So far,
Article
The endogenous cannabinoid system (ECS) is a neuromodulatory system recently recognized to have a role in the regulation of various aspects of eating behavior and energy balance through central and peripheral mechanisms. In the central nervous system, cannabinoid type 1 receptors and their endogenous ligands, the endocannabinoids, are involved in modulating food intake and motivation to consume palatable food. Moreover, the ECS is present in peripheral organs, such as liver, white adipose tissue, muscle, and pancreas, where it seems to be involved in the regulation of lipid and glucose homeostasis. Dysregulation of the ECS has been associated with the development of obesity and its sequelae, such as dyslipidemia and diabetes. Conversely, recent clinical trials have shown that cannabinoid type 1 receptor blockade may ameliorate these metabolic abnormalities. Although further investigation is needed to better define the actual mechanisms of action, pharmacologic approaches targeting the ECS may provide a novel, effective option for the management of obesity, type 2 diabetes and cardiovascular disease.
Article
By catalyzing the rate-limiting step in adipose tissue lipolysis, hormone-sensitive lipase (HSL) is an important regulator of energy homeostasis. The role and importance of HSL in tissues other than adipose are poorly understood. We report here the cloning and expression of a testicular isoform, designated HSLtes. Due to an addition of amino acids at the NH2-termini, rat and human HSLtesconsist of 1068 and 1076 amino acids, respectively, compared to the 768 and 775 amino acids, respectively, of the adipocyte isoform (HSLadi). A novel exon of 1.2 kb, encoding the human testis-specific amino acids, was isolated and mapped to the HSL gene, 16 kb upstream of the exons encoding HSLadi. The transcribed mRNA of 3.9 kb was specifically expressed in testis. No significant similarity with other known proteins was found for the testis-specific sequence. The amino acid composition differs from the HSLadisequence, with a notable hydrophilic character and a high content of prolines and glutamines. COS cells, transfected by the 3.9-kb human testis cDNA, expressed a protein of the expected molecular mass (Mr≈ 120,000) that exhibited catalytic activity similar to that of HSLadi. Immunocytochemistry localized HSL to elongating spermatids and spermatozoa; HSL was not detected in interstitial cells.