You are on page 1of 10

TEM-980; No.

of Pages 10

Review

Role of metabolic lipases and lipolytic


metabolites in the pathogenesis of
NAFLD
Claudia D. Fuchs, Thierry Claudel, and Michael Trauner
Hans Popper Laboratory of Molecular Hepatology, Division of Gastroenterology and Hepatology, Department of Internal Medicine
III, Medical University of Vienna, Vienna, Austria

Non-alcoholic fatty liver disease (NAFLD) is the most About 60% of FAs for hepatic TG accumulating in NAFLD
frequent chronic liver disease in Western countries, rang- patients originate from WAT-derived FAs [2], whereas 25%
ing from simple steatosis to steatohepatitis, cirrhosis, and result from increased de novo lipogenesis controlled via the
hepatocellular cancer. Although the mechanisms under- transcription factors sterol regulatory element-binding pro-
lying disease progression are incompletely understood, tein-1c (SREBP-1c), carbohydrate response element-bind-
lipotoxic events in the liver resulting in inflammation and ing protein (ChREBP), and peroxisome proliferator-
fibrosis appear to be central. Free fatty acids and their activated receptor g (PPARg). The remaining 15% are
metabolites are potentially lipotoxic mediators triggering obtained from the diet [2].
liver injury, suggesting a central role for metabolic lipases. Although IR may initiate hepatic steatosis as first hit,
These enzymes are major players in lipid partitioning additional (second) hits such as drugs, ischemia, or endo-
between tissues and within cells, and provide ligands toxin result in inflammation with disease progression to
for nuclear receptors (NRs). We discuss the potential role nonalcoholic steatohepatitis (NASH) and beyond [2]. How-
of intracellular lipases and their lipolytic products in ever, this ‘second hit hypothesis’ is increasingly replaced by
NAFLD. Because tissue-specific modulation of lipases is the concept that benign steatosis (which may be non-pro-
currently impossible, targeting NRs with ligands may gressive in the majority of patients) and NASH could be two
open novel therapeutic perspectives.
Glossary
Non-alcoholic fatty liver disease (NAFLD; see Glossary)
G-protein-coupled receptors (GPCRs): GPCRs are proteins located at the cell
comprises a disease spectrum ranging from ‘simple’ hepatic
membrane. They bind to various extracellular substances (such as fatty acids,
triglyceride (TG) accumulation (steatosis), to steatosis with bile acids) and therefore can transduce extracellular signals to intracellular
inflammation (steatohepatitis), more advanced fibrosis, cir- proteins.
Lipases: enzymes responsible for the hydrolysis of neutral lipids in free FAs
rhosis, and liver cancer [1,2]. NAFLD is considered as the and glycerol. Triglycerides and phospholipids are common lipase substrates.
hepatic manifestation of metabolic syndrome and is strongly Intracellular/metabolic lipases are expressed specifically in various cells types
associated with type 2 diabetes mellitus (T2DM) and obesi- found in different organs such as adipose tissue (macrophages and white
adipocytes), liver (hepatocytes, Kupffer cells, stellate cells), intestine (enter-
ty. Obesity and T2DM are well known consequences of ocytes), etc.
modern lifestyle characterized by increased dietary calorie Non-alcoholic fatty liver disease (NAFLD): NAFLD comprises a wide spectrum
intake of saturated and trans-unsaturated fatty acids (FAs), of liver injury ranging from simple, relatively benign steatosis to steatohepa-
titis (NASH), fibrosis, cirrhosis, and liver cancer. Estimated prevalence is in the
as well as fructose [3], and a sedentary lifestyle with reduced range 40–50% of the general population depending on country and region.
physical activity [3]. This contributes to an increasing prev- NAFLD diagnosis is based on minimal or absent alcohol consumption (<20–
alence for NAFLD that now affects up to 40–50% of the total 30 g/day), steatosis on imaging, and exclusion of other liver diseases. Firm
diagnosis of NASH still relies on histological criteria (ballooning and
population in the EU and USA, with even higher preva- inflammation in addition to steatosis) revealed by liver biopsy. Simple
lences (75% and more) in risk groups with T2DM and steatosis usually has a benign liver prognosis; however, patients displaying
obesity. Although the precise pathogenesis and progression histological inflammatory changes and/or fibrosis on liver biopsy have a
considerable risk to progress to end-stage liver disease and cancer. Treatment
of NAFLD are still poorly understood [2], insulin resistance of NAFLD is focused on lifestyle modification (diet and exercise), antioxidant
(IR) in muscle, white adipose tissue (WAT), and liver strategies (e.g., vitamin E), and treatment of associated metabolic conditions
such as obesity, insulin resistance (insulin sensitizers), and hyperlipidemia
appears to play a central role [4,5]. Impaired insulin signal-
(statins).
ing in WAT results in increased lipolysis, generating a flux of Nuclear receptors (NR): NRs belong to a family of zinc-finger proteins generally
FA from WAT to the liver, thus causing substrate overload binding DNA and cofactors modulating polymerase II activity therefore
regulating positively or negatively specific gene expression. Some but not all
and further promoting hepatic IR, with subsequently en- NRs are activated after ligand binding (e.g., fatty acids for PPARs, bile acids for
hanced de novo lipogenesis and TG accumulation [6,7]. FXR).
Type 2 diabetes (T2DM): T2DM is characterized by hyperglycemia and
disturbances in carbohydrate and lipid metabolism resulting from defects in
Corresponding author: Trauner, M. (michael.trauner@meduniwien.ac.at). insulin signaling termed insulin resistance (IR). Therefore the metabolic
complications of T2DM derive from deficient insulin action on target tissues.
1043-2760/ Patients suffering from T2DM are at higher risk of cardiovascular, peripheral
ß 2014 Published by Elsevier Ltd. http://dx.doi.org/10.1016/j.tem.2014.08.001 vascular and cerebrovascular disease.

Trends in Endocrinology and Metabolism xx (2014) 1–10 1


TEM-980; No. of Pages 10

Review Trends in Endocrinology and Metabolism xxx xxxx, Vol. xxx, No. x

different entities with divergent pathogenic pictures [8], and Box 1. Lipolysis and metabolic lipases
that ‘multiple hits’ may occur in parallel rather than in strict Physiologically under fasting conditions, and pathologically as a
sequence. In this context, lipid accumulation may be con- result of IR, lipolysis of TGs in WAT by key lipases generates FAs
sidered as a ‘bystander phenomenon’ [8]. Constant ectopic and glycerol, which are secreted into blood and transported to other
exposure of hepatocytes to potentially toxic lipid metabolites tissues such as liver and muscle.
such as FAs, phosphatidic acid, lysophosphatidic acid, cer- Intracellular lipolysis consists of three steps, each releasing FAs,
and involves at least three different enzymes: adipose triglyceride
amides, and diacylglycerols (DAG) may result in ‘lipotoxic’ lipase (ATGL/PNPLA2), a member of the patatin-like phospholipase
insults [2], characterized by endoplasmic reticulum (ER) domain-containing protein family, catalyzes the first step; convert-
stress, inflammation, apoptosis, necrosis, ballooning, and ing TG into DAG. ATGL has a coactivator (comparative gene
Mallory–Denk body formation in hepatocytes, all of which identification 58, CGI-58) that considerably augments ATGL activity
are histopathological hallmarks of NASH [9]. Moreover, [100]. Hormone sensitive lipase (HSL) is the enzyme responsible for
the second step and converts DAGs into monoacylglycerols (MAGs).
these observations strengthen the assumption that not Finally, monoacylglycerol lipase (MGL) hydrolyzes MGs as last step
TG but instead TG-derived metabolites may be toxic to in lipolysis (Figure 1) [54]. Furthermore, MGL also plays an
the liver, and TG breakdown via metabolic lipases contrib- important role in endocannabinoid signaling by creating arachido-
utes to development of NASH [9]. Conversely, specific FAs nic acid (a prostaglandin precursor) through hydrolysis of 2-
arachidonoylglycerol, a major endocannabinoid [101], therefore
are also known to have non-toxic properties such as ligand
playing a potential role in inflammation. ATGL and HSL account for
function NRs and, in addition to signaling function, they act 90% of TG hydrolysis in adipose tissue [100]. Hormonal (e.g.,
also as energy substrates. NRs belong to a superfamily of catecholamines) or neuronal [102] stimulation of lipolysis increases
ligand-activated transcription factors [10]. Several key NRs FA release from WAT more than 100-fold. Although the enzymes
[e.g., PPARs, farnesoid X receptor (FXR) and liver X receptor mentioned above are also expressed in non-adipose tissue such as
muscle and liver, much less is known about lipolytic mechanisms in
(LXR)] are expressed in the liver and activated by ligands
these organs [37]. Therefore, additional lipases may be necessary
such as FAs, bile acids, and oxysterols, respectively [11]. In for efficient lipolysis in non-adipose tissue such as liver [54].
addition, synthetic drugs designed to activate these recep- A closely related enzyme of ATGL/PNPLA2, PNPLA3 (also known
tors such as fibrates (PPARa), glitazones (PPARg), and as adiponutrin), might be such candidate. PNPLA3 is suggested to
obeticholic acid (FXR) are available [12]. Activation of these have both lipase and lysophosphatidic acid acyltransferase (LPAAT)
functions, the latter contributing to TG formation rather than to
NRs promotes FA catabolism (PPARa and FXR), storage breakdown [15,103,104]. Interestingly, PNPLA3 expression in adi-
(PPARg), or endogenous synthesis (LXR) (see below). Be- pose tissue and liver [15,105–107] decreases with fasting and
cause disturbances in these metabolic pathways favor increases in response to insulin and glucose [15], which is more in
NAFLD development, targeting NRs with synthetic drugs line with a suggested LPAAT function and subsequent role in
might counteract and/or reverse fatty liver development. lipogenesis [15]. Additional studies to determine the role of PNPLA3
in hepatic lipid metabolism will be necessary to decipher whether
(or how) possible polymorphisms (see below) affect its lipase and/or
Potential role of lipases in the pathogenesis and LPAAT function and alter the accumulation of key signaling lipid
progression of NAFLD metabolites such as DAG or ceramides in liver or elsewhere [15].
Because the pathogenesis of NAFLD is characterized by
hepatic lipid accumulation [13], metabolic lipases (Box 1)
are in the center of interest of NAFLD/NASH pathogen- Because free FAs are metabolized via several different
esis. Free FAs derived from adipose tissue are taken up pathways, it may be difficult to distinguish whether FAs
via cell specific transporters such as cluster of differen- or their metabolites are the toxic component [9].
tiation 36 (CD36) and fatty acid transport protein
(FATP) or via passive diffusion (Figure 1) [14]. Within PNPLA3 polymorphisms
hepatocytes, FAs are esterified for storage in TGs. How- A genome-wide association study (GWAS) screen within
ever, it appears that not TGs but FAs or other metab- the Dallas Heart study has uncovered an association of
olites such as DAGs and ceramides are cytotoxic. NAFLD with a single-nucleotide polymorphism (SNP)
Because non-adipose tissue has a limited capacity to (rs738409; C>G) in the PNPLA3 gene (Box 1) [15,16],
store neutral lipids in the form of lipid droplets, excess which was subsequently confirmed by several studies
free FA spillover from the periphery or from intracellular from other centers [17–19]. The rs738409 SNP encodes
lipolysis results in hepatic lipid accumulation and lipo- the I148 M substitution, which leads to loss of lipolytic
toxicity. Studies using mice overexpressing diacylgly- function. Although most SNP carriers have increased
cerol acyltransferase (DGAT) 2 (the enzyme catalyzing hepatic fat content, several studies show that it is inde-
the last step in TG formation) in the liver demonstrated pendent of IR [20–22]. However, it has to be mentioned
increased hepatic TG accumulation and hepatic phos- that most of these studies were performed in relatively
phorylation of protein kinase R-like ER kinase (PERK) insulin-resistant obese patients. Therefore, to identify
[9], without inflammation or IR. Conversely, liver-specif- definitively whether PNPLA3 confers a risk independent
ic knockdown of DGAT2 prevented TG accumulation in of IR, insulin sensitivity needs to be assessed in lean
the liver but resulted in lipotoxic injury caused by excess NAFLD patients who are carriers of the wild type
of free FAs [9]. Based on these findings, it can be postu- PNPLA3 versus lean NAFLD patients carrying the
lated that lipid droplet/TG formation may protect I148 M mutant [15].
against toxic lipid-mediated liver injury [9]. However, Overexpression of the mutant I148 M PNPLA3 gene in
the anti-lipotoxic properties of TG may be limited be- mice resulted in increased lipid droplet size and TG accu-
cause FAs can be released from TGs by lipases, thus mulation, which is in line with the human data [15,23].
generating a new source of toxic lipid intermediates. This observation strengthens the assumption that
2
TEM-980; No. of Pages 10

Review Trends in Endocrinology and Metabolism xxx xxxx, Vol. xxx, No. x

TAG Acetyl-CoA
ATGL PPARγ TAG
synthesis
DAG Fatp5 FA-CoA TAG
ATGL
HSL FFA FFA CD36
HSL
MAG MGL
PPAR
FFA
MGL α
RXR
Gene
Glycerol expression
Nucleus
Adipocyte
Hepatocyte
TRENDS in Endocrinology & Metabolism

Figure 1. Lipid partitioning and lipid nuclear receptor (NR) signaling via PPAR. Lipolysis of triglycerides (TGs) in white adipose tissue (WAT) via metabolic lipases adipose
triglyceride lipase (ATGL), hormone-sensitive lipase (HSL), and monoacylglycerol lipase (MGL), initiates free fatty acid (FA) flux from WAT to the liver. In the liver, FAs
deriving from the periphery (or from endogenous sources) need to be activated with acyl-CoA, and are subsequently incorporated into TGs. Intrahepatic hydrolysis of TGs
via the metabolic lipases ATGL, HSL, and MGL create potential ligands (or ligand precursors) for nuclear receptors such as PPARa. In addition, PPARg in WAT is also
activated via free FAs, which stimulate TG hydrolysis.

PNPLA3 acts as a lipase. However, overexpression of wild ATGL and NAFLD


type PNPLA3 did not change hepatic lipid content in mice Animal studies have shown that absence of ATGL (Box 1)
[15,23,24]. Moreover, Pnpla3 knockout (KO) mice did not in mice, in addition to TG accumulation in multiple tissues,
develop hepatic steatosis or disturbances in glucose ho- also leads to increased susceptibility to inflammation be-
meostasis, even when challenged with various diets cause of disturbances in PPARa signaling [31], emphasiz-
[25,26]. Therefore, the function of PNPLA3 in vivo is still ing the important function of lipases for NAFLD. ATGL
controversial and needs to be understood better. Addition- deficiency in humans has been linked to neutral lipid
al studies using transgenic mice expressing the human storage disease (NLSD) characterized by lipid deposition
form of PNPLA3 and its variant I148 M under the control in liver (and other organs such as skin, muscle, and leu-
of their natural promoter at the liver cell specific level may cocytes) [32]. Two phenotypes can be distinguished, name-
help to clarify the exact role/function of PNPLA3 in ly NLSD with ichtiosis (NLSDI, also known as Chanarin–
NAFLD development and progression. Dorfman syndrome) and NLSD with myopathy (NLSDM).
Interestingly, not only steatosis but also disease pro- Whereas NLSDM is caused by mutations in ATGL, NLSDI
gression towards steatohepatitis and fibrosis might be is based on mutations in the ATGL cofactor CGI 58 [32].
initiated via disturbed lipid metabolism. The PNPLA3 NLSDI is characterized by hepatomegaly, liver steatosis
polymorphism may reduce TG hydrolysis, which could and neurological disorders, and NLSDM by skeletal my-
make the cells more susceptible to lipid peroxidation opathy [32]. Interestingly, these patients do not develop
and subsequent oxidative stress and inflammation [27]. impaired glucose tolerance or IR, emphasizing that ectopic
The release of inflammatory cytokines could activate Kupf- fat accumulation does not necessarily cause IR. Apart from
fer cells as well as hepatic stellate cells (HSCs), subse- these rare mutations, the role of polymorphisms of these
quently initiating fibrosis, possibly explaining the link genes in NAFLD is still to be explored. To clarify the role of
between genetic variants and fibrosis progression. More- ATGL in liver physiology and disease, this enzyme was
over, it was observed that PNPLA3 is highly expressed in knocked down either genetically [33] or via adenovirus
human HSC and that it exerts retinyl palmitate lipase injection [34]. In comparison to wild type (WT) mice,
activity and therefore plays an essential role in HSC liver-specific Atgl KO mice developed hepatic steatosis
activation and subsequent fibrosis development. The but the inflammatory response were slightly lower in liver
I148 M PNPLA3 variant was identified as a loss of function Atgl KO than in WT mice [33]. Interestingly, under ER
mutation and therefore reduces retinol release from HSC stress conditions it was also observed that (global) Atgl KO
[28]. Notably, PNPLA3 variants have also been linked to mice develop hepatic steatosis but no inflammation [35],
fibrosis and are considered as marker of disease progres- suggesting that loss of hepatic ATGL activity may protect
sion in several other liver diseases such as alcoholic liver from the disease progression. Of note, in contrast to global
disease [29] and chronic hepatitis C [30]. Therefore further Atgl KO mice, liver-specific Atgl knockdown did not show
studies need to clarify the role of I148 M PNPLA3 variant increased insulin sensitivity [33]. Knockdown of Cgi58 at
in HSCs in the susceptibility toward hepatic fibrogenesis the hepatocellular level results in a much more severe
and carcinogenesis [28] NAFLD phenotype than hepatic Atgl knockdown [33,36].
3
TEM-980; No. of Pages 10

Review Trends in Endocrinology and Metabolism xxx xxxx, Vol. xxx, No. x

In contrast to Atgl liver-specific KO mice, Cgi58 liver- SFA


specific KO mice not only develop steatosis but also severe
inflammation and fibrosis [36], emphasizing an important
role of CGI 58 in NAFLD development.
TLR2 or TLR4
Cell membrane or FATP5
HSL deficiency
In humans, loss of HSL is accompanied by a smaller size of SFA

adipocytes accumulating DGs [37]. To date only two indi-


viduals with HSL deficiency have been identified. These BIM/PUMA (pro-apoptoc proteins)
patients showed IR, low high-density lipoprotein (HDL)
cholesterol, hypertriglyceridemia, and ectopic fat accumu- BCL2 family (an-apoptoc proteins)
lation in the liver [38]. These findings are in line with the
phenotype of Hsl KO mice [39], where impaired PPARg BAX/BAX Caspase 3/6/7
activity causes decreased de novo lipogenesis and subse-
quent reduction in lipid storage [40]. Interestingly, aged Hsl
Apoptosis
KO mice show a decrease in adipose tissue mass, which was
ascribed to disturbances in PPARg signaling [40] as a result TRENDS in Endocrinology & Metabolism

of reduced free FAs, which are PPARg ligands [41]. An Figure 2. Saturated fatty acids (SFAs) induce lipoapoptosis. SFAs activate pro-
additional factor contributing to reduction of WAT in Hsl apoptotic proteins (as such BIM and PUMA), leading to inhibition of the anti-
apoptotic proteins of the BCL 2 family and subsequent activation of pro-apoptotic
KO mice may be cell death caused by increased inflamma-
proteins BAK and BAX. Once activated, BAK and BAX cause mitochondrial
tion, macrophage infiltration, and the formation of ‘crown- dysfunction, which leads to caspase activation and apoptotic cell death.
like structures’ [37] characterized by macrophages sur-
rounding hypertrophic adipocytes [42]. HSL expression is
very low in the liver [39] and Hsl KO mice do not develop
steatosis. Therefore, from the hepatic point of view loss of (Figure 2). Genetic deletion of Puma leads to resistance to
HSL does not appear to be a major factor in NAFLD devel- SFA induced apoptosis in vitro. Free FA-mediated activa-
opment. However, when HSL is overexpressed in livers of tion of BIM and PUMA is accompanied by induction of the
ob/ob mice fed a high-fat diet (HFD) hepatic TG content is multi-domain proapoptotic member of the BCL2 family
reduced by about 50%, suggesting that HSL (despite its BAX. This initiates a mitochondrial pathway of cell death,
minor role in fatty liver development) might be a potential including activation of caspases 3, 6, and 7, resulting in
drug target in treatment of NAFLD [43] apoptosis. Of note, BAX expression is increased in a murine
Because hepatic TG hydrolysis provides substrates for NASH model, and is related to structural and functional
very low density lipoprotein (VLDL)–TG synthesis (ATGL mitochondrial abnormalities found in patients with NASH
as well as HSL levels are very low, and neither overexpres- [45]. Interestingly, the effects of SFAs are counteracted by
sion nor knockdown affects hepatic VLDL secretion), it is unsaturated FAs [44,46]. As mentioned above, unsaturated
tempting to speculate that other lipases may take over this FAs have the potential to initiate TG formation and there-
important role in the liver [44]. It is suggested that trigly- fore have anti-lipotoxic properties. Furthermore, SFAs such
cerol hydrolase (TGH) could be the important component in as palmitic acid (PA) have been shown to act as proinflam-
VLDL assembling. In mice overexpressing human TGH in matory lipid mediators and activate inflammatory path-
the liver, increased secretion of ApoB was observed and, in ways in different cell types such as macrophages,
line, global deletion of TGH resulted in reduced plasma TG adipocytes, myocytes, and hepatocytes [47], which leads
as well as ApoB levels [44]. Interestingly, no concomitant to activation of NF-kB and JNK1/AP1 pathways and cyto-
liver phenotype was observed in the Tgh KO mouse. Lack of kine release [47]. Although it has been reported that the
steatosis in this mouse model was linked to increased FA proinflammatory effects of SFAs are mediated via Toll-like
oxidation and low FA flux from adipose tissue to the liver receptor 4 (TLR4) and TLR2 [47], recent studies demon-
[44]. strated that SFAs indirectly activate TLRs by interacting
with TLR coreceptors [47] such as CD36 or LDL receptor
FAs as lipotoxic mediators [48,49], and/or stimulate TLR4 dimerization [50]. Emerging
Hepatic lipotoxicity is known to induce hepatocyte death, evidence suggests that fetuin A could act as an endogenous
referred to as lipoapoptosis [45]. Free FAs [especially TLR4 ligand and subsequent connector through which FAs
saturated FAs (SFAs)] have hepatotoxic properties. The may induce TLR4 driven insulin resistance [51] (Figure 3).
difference in toxicity between SFAs and unsaturated FAs is An additional possibility could be that SFAs release a tissue
related to the possibility that unsaturated FAs are easily damage signal which may function via TLRs [47]. Never-
esterified and incorporated into neutral lipids [45]. Hepa- theless, SFAs are also able to induce inflammation indepen-
tocytes can undergo apoptosis either via an extrinsic or an dently of TLR pathways and via activation of stress kinases
intrinsic pathway. The extrinsic pathway is activated via and subsequent reactive oxygen species (ROS) production.
death ligands, whereas the intrinsic pathway is initiated by Increased ROS levels deriving from SFA stimulation may
intracellular stress of organelles such as ER, lysosomes, and account for the activation of nucleotide binding domain
mitochondria [45]. Hepatocytes treated with SFAs dis- leucine-rich repeats containing family, pyrin domain-con-
played increased expression of the proapoptotic proteins taining 3/apoptotic speck protein containing a caspase re-
BIM and p53 upregulated modulator of apoptosis (PUMA) cruitment domain (NLRP3/ASC) inflammasome complex,
4
TEM-980; No. of Pages 10

Review Trends in Endocrinology and Metabolism xxx xxxx, Vol. xxx, No. x

ATGL HSL MGL


TAG DAG MAG Glycerol
Adipocyte FFA FFA FFA

(B) SFA
SFA
(A) FetA
ω3-FAs Insulin
Hepatocyte
TLR2 or TLR4
GPR120 FATP5

Cell IR
SFA IRS
membrane PKCε Ser-P
ω3-FAs
DAG
TAK
PI3K
β arr2
TAB1
ATGL
TAG FFA PPARa NF-κB JNK Ceramides PKCε AKT

DAG
FA oxidaon
FOXO1
Inflammaon (gluconeogenesis)

(cytokine producon)
mTOR GSK-3
(protein synthesis) (glycogen synthesis)
TRENDS in Endocrinology & Metabolism

Figure 3. Impact of fatty acid (FA) signaling on inflammation and insulin resistance. (A) Omega-3 fatty acids (v3-FA) are able to bind the G-protein-coupled receptor GPR120,
leading to receptor internalization and binding of b arrestin 2 (b arr2). Furthermore, TAK1 binding protein (TAB) will be complexed by b arr2. Complex formation with TAK1
(TGFb-activated kinase 1) and TAB is inhibited, resulting in TAK1 inhibition and repression of downstream inflammatory signaling. (B) Saturated FAs (SFAs) activate the
inflammatory cascade via TLR4- or TLR2-dependent or -independent pathways. Moreover, SFAs are precursors for diacylglycerol (DAG) as well as ceramide biosynthesis.

thereby increasing interleukin 1b (IL1b) release [52], insulin receptor substrate (IRS) 1 and 2, resulting in
which was shown to impair insulin signaling, suggesting repressed activation of AKT. Reduced AKT function is
an additional SFA-mediated mechanism leading to IR. In accompanied with attenuated glycogen synthesis and de-
line, Nlrp3, Asc, and caspase1 KO mice are protected from creased suppression of gluconeogenesis, which in turn
HFD-induced inflammation and IR [52]. Paradoxically, leads to increased glucose release [59] (Figure 3). Of note,
Asc KO mice developed a more severe NASH phenotype it has also been shown that intracellular DAG concentra-
with a methionine- and choline-deficient (MCD) diet owing tion can increase independently of impaired insulin sig-
to abnormal microbiota composition [53]. Furthermore, naling [60]. Moreover, localization and intracellular
SFAs lead to increased production of ROS which activate compartmentation of DAGs may play a major role in
redox-sensitive serine kinases that subsequently inhibit determining its lipotoxic properties because it has been
insulin signaling [54]. Although SFAs appear to be lipo- demonstrated that internalization of DAGs into lipid dro-
toxic because of proinflammatory and IR properties, other plets prevents PKCe activation at the plasma membrane
classes of FAs such as long chain polyunsaturated omega-3 and subsequently impairs insulin receptor signaling [61].
FAs may have protective functions [47] (Figure 3). Omega- In addition to localization, the structure of DAGs is also of
3 FAs have been suggested as ligands for the G-protein- particular importance for lipotoxicity. It has been demon-
coupled receptor (GPCR) 120, as well as for PPARa and strated that only the 1,2-DAG stereoisomer is able to
PPARg [55], and therefore may exert anti-inflammatory activate PKCe [62], whereas both saturated as well as
effects in macrophages [56]. Importantly, SFAs are also unsaturated acyl side chains had the same effect on kinase
precursors for other lipid products such as DAGs activation [63].
(Figure 3). Intracellular concentrations of DAG is In addition to potential lipotoxic TG degradation pro-
increase during IR [57] leading to increased protein kinase ducts, cholesterol may also be an important trigger for
C (PKC) e activation and impairment of insulin signaling NAFLD development. Patients suffering from NAFLD dis-
[58] with subsequent development of NAFLD [59]. DAG- played increased cholesterol synthesis [64]. Interestingly,
dependent PKCe activation leads to insulin receptor ki- cholesterol can activate LXRa and subsequent de novo
nase inhibition and reduced tyrosine phosphorylation of lipogenesis via increased oxysterol levels (see below). This

5
TEM-980; No. of Pages 10

Review Trends in Endocrinology and Metabolism xxx xxxx, Vol. xxx, No. x

observation indicates that enhanced cholesterol intake can Interestingly, a recently identified variant of transmem-
cause steatosis development [65]. Furthermore, increased brane 6 superfamily member 2 (TM6SF2) that confers
dietary cholesterol uptake leads to hepatic inflammation susceptibility to NAFLD was shown to lower hepatic VLDL
and oxidative stress in mice [66], suggesting that increased secretion and increase liver TG [79]. Thus, it is tempting to
cholesterol does not only act as disease initiator but can also speculate that FXR may impact on regulation of TM6SF2.
promote its progression beyond the more severe stages. Free Furthermore, FXR is also known to inhibit ChREBP [80],
cholesterol was found to accumulate in HSC, where it another lipogenic transcription factor potentially involved
upregulates TLR4 protein levels and subsequently in fatty liver development [60]. Finally, human FXR also
represses the inhibitory TGF-pseudoreceptor BMP and acti- upregulates PPARa expression [81], suggesting that some
vin membrane-bound inhibitor homolog (BAMBI). The con- FXR effects on lipid homeostasis might be mediated via
sequential increase in TGF signaling promotes HSC PPARa-induced FA oxidation [12].
activation and liver fibrosis [67].
LXR signaling
FAs and NR signaling LXRs (LXRa and LXRb) are sterol-activated transcription
PPAR signaling factors which are regulated via oxysterols and other cho-
FAs, especially polyunsaturated FAs (PUFAs) [68], are lesterol derivatives [82]. In addition to their role in reverse
ligands for the NRs PPARa (NR1C1), PPARg (NR1C3), cholesterol transport and bile acid synthesis, LXRs also
and PPARd (NR1C2), and can also activate GPCRs. PPARa affect glucose and lipid metabolism. The lipogenic activity
is highly expressed in oxidative tissues such as liver, cardiac of LXR results from the activation of SREBP1c (master
and skeletal muscle, whereas PPARg is mainly found in regulator of de novo lipogenesis) as well as direct regula-
adipocytes, enterocytes, and macrophages, and PPARd is tion of some SREBP1c downstream targets such as fatty
ubiquitously expressed [11,69]. PPARa increases FA oxida- acid synthase (FAS) and stearyl-CoA desaturase (SCD1)
tion and reverse cholesterol transport, PPARg promotes [82]. In line, feeding mice with an LXR agonist led to
lipid storage, adipocyte differentiation, and normalizes glu- increased hepatic TG formation and VLDL secretion
cose levels [9], whereas PPARd increases FA and glucose [82]. Ob/ob mice lacking LXRa and LXRb displayed re-
catabolism but also plays a role in fertility and cancer [70]. duced hepatic steatosis and improved insulin sensitivity in
Interestingly, free FAs originating from adipose tissue comparison to ob/ob mice. However, loss of hepatic de novo
lipolysis are not able to activate hepatic PPARa directly lipogenesis in ob/ob mice deficient for LXRs was accompa-
but are ligands for PPARd in the liver [68]. However, nied by increased adipose-tissue lipid storage, which was
activation of cardiac and hepatic gene transcription via ascribed to tissue-selective action of LXR [82]. Therefore
PPARa largely depends on lipolytic breakdown of intracel- the mice remained obese despite improved insulin sensi-
lular TG depots [71] by ATGL, HSL, and MGL (Figure 1) tivity [82].
[31] in these tissues. Therefore, FA-induced activation of
PPARa drives a feed-forward loop protecting against FA- Therapeutic implications
derived lipotoxicity by promoting their oxidation in heart Metabolic lipases control the balance between FA storage
and liver [68]. This observation suggests that lipid droplets and FA release, thus regulating FA flux from periphery to
do not only serve as energy stores but also as a reservoir for the liver. To protect the liver from potentially toxic FA
lipid mediators that can act as PPARa ligands [71]. There- spillover and development of NAFLD, targeted suppres-
fore, FA-activated PPARs play a protective role against sion of lipase activity in adipose tissue (especially visceral
hepatic lipotoxicity [72]. Metabolic lipases such as ATGL fat) might represent a therapeutic aim because general
release the intracellular ligands, providing a potential suppression of lipase activity has many risks including
mechanistic link and drawing attention to lipases as po- impaired PPAR signaling. The lipolysis inhibitor nicotinic
tential ‘ligand providers’ for anti-lipotoxic pathways. In acid (niacin) reduces free FA levels in the serum by acti-
addition, it was also shown that not only FAs derived from vating the high-affinity receptor for niacin, GPR109A, in
intracellular lipolysis but also 1-palmitoyl-2-oleoyl-sn- WAT [83]. Owing to multiple side effects such as flushing
glycerol-3-phosphocholine (POPC) is an endogenous ligand and lack of efficiency (AIM-HIGH study [84]), nicotinic acid
for PPARa [73]. In addition to their endogenous ligands, is not suitable as therapy of NAFLD/NASH. Orlistat, an
synthetic drugs are also able to activate PPARs, including inhibitor of extracellular but not intracellular lipases, was
fibrates for PPARa or glitazones for PPARg [74] (see tested in some small NAFLD trials, but the effects on liver
therapeutic implications). histology and body weight were not superior to placebo
[85–87].
FXR signaling NRs such as PPARs and FXR, which have metabolic,
In addition to PPARs, FXR (NR1H4) is activated by PUFAs anti-inflammatory, and anti-fibrotic effects, may be valid
[75]. FXR, which is mainly known for its regulatory func- targets. PPARa agonists such as fibrates are used to
tion in bile acid metabolism, is also a key regulator of treat dyslipidemia and hypertriglyceridemia [88,89].
hepatic lipid metabolism [12,76]. Moreover FXR and the However, studies with PPARa agonists in NAFLD have
FXR target gene small heterodimer partner (SHP) play key generated divergent results. Fenofibrate treatment did
roles in lipid metabolism by downregulating SREBP1c not show beneficial effects on steatosis, assessed by MR
after inhibition of LXRa signaling [77] and repress spectroscopy (MRS), in a study with only 15 patients
microsomal triglyceride transfer protein (MTTP) induction [90], whereas another small (n = 16) study reported a
mediated by liver receptor homolog 1 (LRH-1) [78]. reduction in steatosis by fenofibrate in combination with
6
TEM-980; No. of Pages 10

Review Trends in Endocrinology and Metabolism xxx xxxx, Vol. xxx, No. x

dietary intervention [88]. Unfortunately, no placebo further improve liver histology, suggesting a plateau effect
group was included and histological data were not of histological improvement when the maximal insulin-
obtained. Another small (n = 16) biopsy-controlled study sensitizing effect is reached [93]. Furthermore, glitazone-
revealed a reduction in hepatic alanine transaminase based improvement of ALT and hepatic IR returned to
(ALT) levels, but no histological improvement of key baseline after 3 months of treatment cessation, and in few
features such as hepatocellular ballooning, steatosis, patients steatohepatitis recurred a 1 year follow-up liver
inflammation, or fibrosis was observed [88]. Finally, biopsy [96]. The main restriction for the use of glitazones
short-term (4 weeks) treatment with gemfibrozil resulted came from their rather poor safety profile, and persistent
in reduction of serum transaminases [88]. In light of weight gain that arises mainly from increased peripheral
these data, there is currently very limited evidence to fat mass [97]. Indeed, glitazones promote preadipocyte
support a role for fibrates in the treatment of NAFLD/ differentiation in subcutaneous fat depots, while inhibiting
NASH. adipocyte maturation in visceral fat, thus shifting free FA
However, treatment with the PPARad dual agonist storage from liver to more protective fat depots such as
GFT505 has lead to promising results [91,92]. After 4– subcutaneous fat [98]. However, weight gain as a cosmetic
12 weeks of treatment, serum transaminases, as well as (not metabolic) side effect is not desirable. In addition,
alkaline phosphatase and g-glutamyl transpeptidase based on the increase in cardiovascular events, congestive
(gGT), were reduced in a small Phase II trial with patients heart failure, and bone fractures in women, as well as
displaying dyslipidemia, diabetes, and IR. A biopsy-con- bladder cancer risk, glitazones were withdrawn from the
trolled Phase II study evaluating GFT505 in NASH is market in France and Germany [93].
currently ongoing (http://clinical.trials.gov, identifier PPARd, which is also expressed in HSC and macro-
NCT01694849) [91]. phages, was shown to have anti-fibrotic features in vivo.
The therapeutic effects of the PPARg ligands (glita- However, human data on NAFLD treatment are lacking
zones) in NAFLD can be explained by their actions in [88].
adipose tissue and muscle. PPARg was shown to activate FXR ligands also improved fatty liver and related IR in
directly ATGL activity in adipose tissue at the transcrip- various rodent models of obesity and NAFLD (reviewed in
tional level in vitro and in vivo [93], a finding that was [12]). Moreover, different FXR agonists such as PX-104
verified by increased ATGL mRNA and protein expression (non-steroidal) and INT747, a chenodeoxycholic acid deriv-
in adipose tissue of mice treated with the PPARg agonist ative also known as obeticholic acid, have entered clinical
rosiglitazone [93]. Therefore, as a lipogenic master regula- trials for NAFLD treatment. A double-blind, placebo-con-
tor, PPARg may regulate lipid partitioning via ATGL trolled proof-of-concept study evaluated the effects of
regulation in WAT (Figure 3). Moreover, PPARg activation INT747 in diabetic patients with presumed NAFLD [99].
increases insulin sensitivity and reduces FA flux to the Administration of a low (25 mg) and a high (50 mg) INT747
liver. In liver, PPARg is expressed in HSCs (maintaining for 6 weeks improved insulin sensitivity as well as liver
them in a quiescent state, and showing anti-fibrotic prop- inflammation and fibrosis [99]. INT747 was also tested in a
erties in preclinical studies), and macrophages (where it biopsy-controlled NIH/NDDK-sponsored Phase IIb trial
displays anti-inflammatory effects) [88]. Another impor- investigating its effects on histologically proven NASH
tant mechanism is the capacity of glitazones to induce the [FXR ligand NASH Treatment (FLINT) trial; http://clin-
production of adiponectin, an insulin-sensitizing and anti- icaltrials.gov/ct2/show/NCT01265498] which was prema-
steatogenic adipokine, which increases FA oxidation in turely stopped because of positive effects on liver histology.
liver and muscle [93]. Studies with PPARg agonists in
diabetic and non-diabetic patients with biopsy-proven Concluding remarks and future perspectives
NASH showed improvements in IR, liver enzyme levels, Metabolic lipases are the enzymes responsible for TG
and in hepatic fat accumulation (assessed by MR), some- breakdown in WAT and liver. Lipolysis in WAT creates
times with improvement of some histological NASH fea- a free FA flux to the liver, which promotes steatosis devel-
tures such as steatosis and inflammation, but not opment in IR conditions. Saturated FAs induce hepatocyte
ballooning and fibrosis [88]. A randomized double-blind death and inflammation, which are key players for NASH
placebo-controlled study of 55 patients with impaired glu- development. However, PUFAs exert anti-inflammatory
cose tolerance or T2DM and biopsy-proven NASH, testing and anti-steatotic effects. Increased hepatic ATGL activity
the effects of caloric restriction in combination with 6 generates free FAs, which are PPARa ligands. Pharmaco-
month pioglitazone versus placebo treatment, uncovered logical activation of FXR and PPARa results in increased
a reduction in steatosis, inflammation, and ballooning by FA catabolism in the liver, whereas PPARg stimulation
pioglitazone [94]. Importantly, fibrosis did not significantly favors TG storage, therefore limiting hepatic lipotoxicity.
improve compared to placebo. Another pioglitazone study Indeed, clinical trials with FXR and PPARagd ligands
in non-diabetic patients with biopsy-proven NASH have resulted in promising results for NAFLD/NASH
revealed reduced liver enzyme levels, hepatic steatosis, treatment. In contrast to NRs, lipases are not yet suitable
and inflammation, again without improvement of fibrosis as direct pharmacological targets. However, the develop-
[95]. Although the glitazone trials are very heterogeneous ment of tissue-specific lipase modulators (i.e., inhibiting
in outcome, the two most robust and reproducible findings ATGL and HSL in WAT), while increasing ATGL activity
are the reduction of serum ALT levels and the grade of in the liver, may be desirable. Analysis of tissue-specific
steatosis [93]. However, the optimal duration of therapy is Atgl KO mice might validate this concept (Box 2). In
not yet established. A prolonged 3 year study did not addition, small peptides targeting PNPLA3 variants
7
TEM-980; No. of Pages 10

Review Trends in Endocrinology and Metabolism xxx xxxx, Vol. xxx, No. x

Box 2. Outstanding questions 19 Sookoian, S. and Pirola, C.J. (2011) Meta-analysis of the influence of
I148 M variant of patatin-like phospholipase domain containing 3
 Which non-invasive markers can stratify the risk for progression in gene (PNPLA3) on the susceptibility and histological severity of
NAFLD/NASH? nonalcoholic fatty liver disease. Hepatology 53, 1883–1894
 Does PNPLA3 function as a lipase or an acyltransferase? 20 Kantartzis, K. et al. (2009) Dissociation between fatty liver and
 How does PNPLA3 promote fibrosis development in progression insulin resistance in humans carrying a variant of the patatin-like
of liver disease? phospholipase 3 gene. Diabetes 58, 2616–2623
 What is the role of genetic variants of metabolic lipases (other 21 Kotronen, A. et al. (2009) A common variant in PNPLA3, which
than PNPLA3) in NAFLD/NASH pathogenesis and progression? encodes adiponutrin, is associated with liver fat content in
 Which as yet unknown/metabolic lipases account for the remain- humans. Diabetologia 52, 1056–1060
ing lipolytic activity in liver (not accounted by ATGL, HSL, and 22 Speliotes, E.K. et al. (2010) PNPLA3 variants specifically confer
PNPLA3)? increased risk for histologic nonalcoholic fatty liver disease but not
 How can we target genetic variants of metabolic lipases metabolic disease. Hepatology 52, 904–912
pharmacologically? 23 He, S. et al. (2010) A sequence variation (I148 M) in PNPLA3
associated with nonalcoholic fatty liver disease disrupts
triglyceride hydrolysis. J. Biol. Chem. 285, 6706–6715
24 Qiao, A. et al. (2011) Mouse patatin-like phospholipase domain-
associated with NAFLD/NASH progression reverting the containing 3 influences systemic lipid and glucose homeostasis.
enzyme activity to normal could provide an innovative Hepatology 54, 509–5021
approach to treat NAFLD. 25 Basantani, M.K. et al. (2011) Pnpla3/Adiponutrin deficiency in mice
does not contribute to fatty liver disease or metabolic syndrome. J.
Lipid Res. 52, 318–329
Acknowledgments
26 Chen, W. et al. (2010) Patatin-like phospholipase domain-containing
This work was supported by grants SFB 30 and SFB 35 (F3008-B19 and
3/adiponutrin deficiency in mice is not associated with fatty liver
F3517-B20) from the Austrian Science Foundation (to M.T.).
disease. Hepatology 52, 1134–1142
27 Valenti, L. et al. (2010) Homozygosity for the patatin-like
References phospholipase-3/adiponutrin I148 M polymorphism influences liver
1 Cohen, J.C. et al. (2011) Human fatty liver disease: old questions and fibrosis in patients with nonalcoholic fatty liver disease. Hepatology
new insights. Science 332, 1519–1523 51, 1209–1217
2 Trauner, M. et al. (2010) Fatty liver and lipotoxicity. Biochim. 28 Pirazzi, C. et al. (2014) PNPLA3 has retinyl-palmitate lipase activity
Biophys. Acta 1801, 299–310 in human hepatic stellate cells. Hum. Mol. Genet. 23, 4077–4085
3 Stefan, N. et al. (2011) Dissociation between fatty liver and insulin 29 Tian, C. et al. (2010) Variant in PNPLA3 is associated with alcoholic
resistance: the role of adipose triacylglycerol lipase. Diabetologia 54, liver disease. Nat. Genet. 42, 21–23
7–9 30 Valenti, L. et al. (2011) Modulation of the effect of PNPLA3 I148 M
4 Utzschneider, K.M. and Kahn, S.E. (2006) The role of insulin mutation on steatosis and liver damage by alcohol intake in patients
resistance in nonalcoholic fatty liver disease. J. Clin. Endocrinol. with chronic hepatitis C. J. Hepatol. 55, 1470–1471
Metab. 91, 4753–4761 31 Jha, P. et al. (2014) Role of adipose triglyceride lipase (PNPLA2) in
5 Bertolani, C. and Marra, F. (2008) The role of adipokines in liver protection from hepatic inflammation in mouse models of
fibrosis. Pathophysiology 15, 91–101 steatohepatitis and endotoxemia. Hepatology 59, 858–869
6 Anderson, N. and Borlak, J. (2008) Molecular mechanisms and 32 Schweiger, M. et al. (2009) Neutral lipid storage disease: genetic
therapeutic targets in steatosis and steatohepatitis. Pharmacol. disorders caused by mutations in adipose triglyceride lipase/
Rev. 60, 311–357 PNPLA2 or CGI-58/ABHD5. Am. J. Physiol. Endocrinol. Metab.
7 Postic, C. and Girard, J. (2008) Contribution of de novo fatty acid 297, E289–E296
synthesis to hepatic steatosis and insulin resistance: lessons from 33 Wu, J.W. et al. (2011) Deficiency of liver adipose triglyceride lipase in
genetically engineered mice. J. Clin. Invest. 118, 829–838 mice causes progressive hepatic steatosis. Hepatology 54, 122–132
8 Tilg, H. and Moschen, A.R. (2010) Evolution of inflammation in 34 Ong, K.T. et al. (2013) Hepatic ATGL knockdown uncouples glucose
nonalcoholic fatty liver disease: the multiple parallel hits intolerance from liver TAG accumulation. FASEB J. 27, 313–321
hypothesis. Hepatology 52, 1836–1846 35 Fuchs, C.D. et al. (2012) Absence of adipose triglyceride lipase protects
9 Neuschwander-Tetri, B.A. (2010) Hepatic lipotoxicity and the from hepatic endoplasmic reticulum stress in mice. Hepatology 56,
pathogenesis of nonalcoholic steatohepatitis: the central role of 270–280
nontriglyceride fatty acid metabolites. Hepatology 52, 774–788 36 Guo, F. et al. (2013) Deficiency of liver comparative gene
10 Evans, R.M. and Mangelsdorf, D.J. (2014) Nuclear receptors, RXR, identification-58 causes steatohepatitis and fibrosis in mice. J.
and the big bang. Cell 157, 255–266 Lipid Res. 54, 2109–2120
11 Bookout, A.L. et al. (2006) Anatomical profiling of nuclear receptor 37 Young, S.G. and Zechner, R. (2013) Biochemistry and pathophysiology
expression reveals a hierarchical transcriptional network. Cell 126, of intravascular and intracellular lipolysis. Genes Dev. 27, 459–484
789–799 38 Albert, J.S. et al. (2014) Null mutation in hormone-sensitive lipase
12 Fuchs, C. et al. (2013) Bile acid-mediated control of liver triglycerides. gene and risk of type 2 diabetes. N. Engl. J. Med. 370, 2307–2315
Semin. Liver Dis. 33, 330–432 39 Haemmerle, G. et al. (2002) Hormone-sensitive lipase deficiency in
13 Browning, J.D. and Horton, J.D. (2004) Molecular mediators of mice causes diglyceride accumulation in adipose tissue, muscle, and
hepatic steatosis and liver injury. J. Clin. Invest. 114, 147–152 testis. J. Biol. Chem. 277, 4806–4815
14 Snel, M. et al. (2012) Ectopic fat and insulin resistance: 40 Zimmermann, R. et al. (2003) Decreased fatty acid esterification
pathophysiology and effect of diet and lifestyle interventions. Int. compensates for the reduced lipolytic activity in hormone-sensitive
J. Endocrinol. 2012, 983814 lipase-deficient white adipose tissue. J. Lipid Res. 44, 2089–2099
15 Samuel, V.T. and Shulman, G.I. (2012) Mechanisms for insulin 41 Shen, W.J. et al. (2011) Hormone-sensitive lipase modulates adipose
resistance: common threads and missing links. Cell 148, 852–871 metabolism through PPARgamma. Biochim. Biophys. Acta 1811, 9–
16 Romeo, S. et al. (2008) Genetic variation in PNPLA3 confers 16
susceptibility to nonalcoholic fatty liver disease. Nat. Genet. 40, 42 Cinti, S. et al. (2005) Adipocyte death defines macrophage localization
1461–1465 and function in adipose tissue of obese mice and humans. J. Lipid Res.
17 Sevastianova, K. et al. (2011) Genetic variation in PNPLA3 46, 2347–2355
(adiponutrin) confers sensitivity to weight loss-induced decrease in 43 Reid, B.N. et al. (2008) Hepatic overexpression of hormone-sensitive
liver fat in humans. Am. J. Clin. Nutr. 94, 104–111 lipase and adipose triglyceride lipase promotes fatty acid oxidation,
18 Hyysalo, J. et al. (2012) Genetic variation in PNPLA3 but not APOC3 stimulates direct release of free fatty acids, and ameliorates steatosis.
influences liver fat in non-alcoholic fatty liver disease. J. J. Biol. Chem. 283, 13087–13099
Gastroenterol. Hepatol. 27, 951–956

8
TEM-980; No. of Pages 10

Review Trends in Endocrinology and Metabolism xxx xxxx, Vol. xxx, No. x

44 Quiroga, A.D. and Lehner, R. (2012) Liver triacylglycerol lipases. 71 Haemmerle, G. et al. (2011) ATGL-mediated fat catabolism regulates
Biochim. Biophys. Acta 1821, 762–769 cardiac mitochondrial function via PPAR-alpha and PGC-1. Nat. Med.
45 Ibrahim, S.H. et al. (2011) Mechanisms of lipotoxicity in NAFLD and 17, 1076–1085
clinical implications. J. Pediatr. Gastroenterol. Nutr. 53, 131–140 72 Nolan, C.J. and Larter, C.Z. (2009) Lipotoxicity: why do saturated
46 Listenberger, L.L. et al. (2003) Triglyceride accumulation protects fatty acids cause and monounsaturates protect against it? J.
against fatty acid-induced lipotoxicity. Proc. Natl. Acad. Sci. U.S.A. Gastroenterol. Hepatol. 24, 703–706
100, 3077–3082 73 Chakravarthy, M.V. et al. (2009) Identification of a physiologically
47 Glass, C.K. and Olefsky, J.M. (2012) Inflammation and lipid signaling relevant endogenous ligand for PPARalpha in liver. Cell 138, 476–488
in the etiology of insulin resistance. Cell Metab. 15, 635–645 74 Pirat, C. et al. (2012) Targeting peroxisome proliferator-activated
48 Seimon, T.A. et al. (2010) Atherogenic lipids and lipoproteins trigger receptors (PPARs): development of modulators. J. Med. Chem. 55,
CD36–TLR2-dependent apoptosis in macrophages undergoing 4027–4061
endoplasmic reticulum stress. Cell Metab. 12, 467–482 75 Zhao, A. et al. (2004) Polyunsaturated fatty acids are FXR ligands and
49 Stewart, C.R. et al. (2010) CD36 ligands promote sterile inflammation differentially regulate expression of FXR targets. DNA Cell Biol. 23,
through assembly of a Toll-like receptor 4 and 6 heterodimer. Nat. 519–526
Immunol. 11, 155–161 76 Thomas, C. et al. (2008) Targeting bile-acid signalling for metabolic
50 Wong, S.W. et al. (2009) Fatty acids modulate Toll-like receptor 4 diseases. Nat. Rev. Drug Discov. 7, 678–693
activation through regulation of receptor dimerization and 77 Watanabe, M. et al. (2004) Bile acids lower triglyceride levels via a
recruitment into lipid rafts in a reactive oxygen species-dependent pathway involving FXR, SHP, and SREBP-1c. J. Clin. Invest. 113,
manner. J. Biol. Chem. 284, 27384–27392 1408–1418
51 Pal, D. et al. (2012) Fetuin-A acts as an endogenous ligand of TLR4 to 78 Huang, J. et al. (2007) Molecular characterization of the role of orphan
promote lipid-induced insulin resistance. Nat. Med. 18, 1279–1285 receptor small heterodimer partner in development of fatty liver.
52 Wen, H. et al. (2011) Fatty acid-induced NLRP3–ASC inflammasome Hepatology 46, 147–157
activation interferes with insulin signaling. Nat. Immunol. 12, 408– 79 Kozlitina, J. et al. (2014) Exome-wide association study identifies a
4015 TM6SF2 variant that confers susceptibility to nonalcoholic fatty liver
53 Henao-Mejia, J. et al. (2012) Inflammasome-mediated dysbiosis disease. Nat. Genet. 46, 352–356
regulates progression of NAFLD and obesity. Nature 482, 179–185 80 Caron, S. et al. (2013) Farnesoid X receptor inhibits the
54 Zechner, R. et al. (2012) Fat signals – lipases and lipolysis in lipid transcriptional activity of carbohydrate response element binding
metabolism and signaling. Cell Metab. 15, 279–291 protein in human hepatocytes. Mol. Cell. Biol. 33, 2202–2211
55 Deckelbaum, R.J. et al. (2006) n-3 fatty acids and gene expression. Am. 81 Pineda Torra, I. et al. (2003) Bile acids induce the expression of the
J. Clin. Nutr. 83 (Suppl.), 1520S–1525S human peroxisome proliferator-activated receptor alpha gene via
56 Oh, D.Y. et al. (2010) GPR120 is an omega-3 fatty acid receptor activation of the farnesoid X receptor. Mol. Endocrinol. 17, 259–272
mediating potent anti-inflammatory and insulin-sensitizing effects. 82 Hong, C. and Tontonoz, P. (2014) Liver X receptors in lipid
Cell 142, 687–698 metabolism: opportunities for drug discovery. Nat. Rev. Drug
57 Kumashiro, N. et al. (2011) Cellular mechanism of insulin resistance Discov. 13, 433–444
in nonalcoholic fatty liver disease. Proc. Natl. Acad. Sci. U.S.A. 108, 83 Ren, N. et al. (2009) Phenolic acids suppress adipocyte lipolysis via
16381–16385 activation of the nicotinic acid receptor GPR109A (HM74a/PUMA-G).
58 Samuel, V.T. et al. (2010) Lipid-induced insulin resistance: J. Lipid Res. 50, 908–914
unravelling the mechanism. Lancet 375, 2267–2277 84 Boden, W.E. et al. (2011) Niacin in patients with low HDL cholesterol
59 Jornayvaz, F.R. and Shulman, G.I. (2012) Diacylglycerol activation of levels receiving intensive statin therapy. N. Engl. J. Med. 365, 2255–
protein kinase Cepsilon and hepatic insulin resistance. Cell Metab. 15, 2267
574–584 85 Dowman, J.K. et al. (2011) Current therapeutic strategies in non-
60 Benhamed, F. et al. (2012) The lipogenic transcription factor ChREBP alcoholic fatty liver disease. Diabetes Obes. Metab. 13, 692–702
dissociates hepatic steatosis from insulin resistance in mice and 86 Zelber-Sagi, S. et al. (2006) A double-blind randomized placebo-
humans. J. Clin. Invest. 122, 2176–2194 controlled trial of orlistat for the treatment of nonalcoholic fatty
61 Cantley, J.L. et al. (2013) CGI-58 knockdown sequesters liver disease. Clin. Gastroenterol. Hepatol. 4, 639–644
diacylglycerols in lipid droplets/ER-preventing diacylglycerol- 87 Hussein, O. et al. (2007) Orlistat reverse fatty infiltration and
mediated hepatic insulin resistance. Proc. Natl. Acad. Sci. U.S.A. improves hepatic fibrosis in obese patients with nonalcoholic
110, 1869–1874 steatohepatitis (NASH). Dig. Dis. Sci. 52, 2512–2519
62 Ganong, B.R. et al. (1986) Specificity and mechanism of protein kinase 88 Tailleux, A. et al. (2012) Roles of PPARs in NAFLD: potential
C activation by sn-1,2-diacylglycerols. Proc. Natl. Acad. Sci. U.S.A. 83, therapeutic targets. Biochim. Biophys. Acta 1821, 809–818
1184–2118 89 Peters, J.M. et al. (2012) The role of peroxisome proliferator-activated
63 Ebeling, J.G. et al. (1985) Diacylglycerols mimic phorbol diester receptors in carcinogenesis and chemoprevention. Nat. Rev. Cancer
induction of leukemic cell differentiation. Proc. Natl. Acad. Sci. 12, 181–195
U.S.A. 82, 815–819 90 Bajaj, M. et al. (2007) Effects of peroxisome proliferator-activated
64 Simonen, P. et al. (2011) Cholesterol synthesis is increased and receptor (PPAR)-alpha and PPAR-gamma agonists on glucose and
absorption decreased in non-alcoholic fatty liver disease lipid metabolism in patients with type 2 diabetes mellitus.
independent of obesity. J. Hepatol. 54, 153–159 Diabetologia 50, 1723–1731
65 Enjoji, M. et al. (2012) Nutrition and nonalcoholic fatty liver disease: 91 Staels, B. et al. (2013) Hepatoprotective effects of the dual peroxisome
the significance of cholesterol. Int. J. Hepatol. 00, 925807 proliferator-activated receptor alpha/delta agonist, GFT505, in rodent
66 Subramanian, S. et al. (2011) Dietary cholesterol exacerbates hepatic models of nonalcoholic fatty liver disease/nonalcoholic
steatosis and inflammation in obese LDL receptor-deficient mice. J. steatohepatitis. Hepatology 58, 1941–1952
Lipid Res. 52, 1626–1635 92 Cariou, B. et al. (2013) Dual peroxisome proliferator-activated
67 Schwabe, R.F. and Maher, J.J. (2012) Lipids in liver disease: looking receptor alpha/delta agonist GFT505 improves hepatic and
beyond steatosis. Gastroenterology 142, 8–11 peripheral insulin sensitivity in abdominally obese subjects.
68 Georgiadi, A. and Kersten, S. (2012) Mechanisms of gene regulation Diabetes Care 36, 2923–2930
by fatty acids. Adv. Nutr. Res. 3, 127–134 93 Ratziu, V. (2013) Pharmacological agents for NASH. Nat. Rev.
69 Escher, P. et al. (2001) Rat PPARs: quantitative analysis in adult rat Gastroenterol. Hepatol. 10, 676–685
tissues and regulation in fasting and refeeding. Endocrinology 142, 94 Belfort, R. et al. (2006) A placebo-controlled trial of pioglitazone in
4195–4202 subjects with nonalcoholic steatohepatitis. N. Engl. J. Med. 355,
70 Nakamura, M.T. et al. (2014) Regulation of energy metabolism by 2297–2307
long-chain fatty acids. Prog. Lipid Res. 53, 124–144 95 Sanyal, A.J. et al. (2010) Pioglitazone, vitamin E, or placebo for
nonalcoholic steatohepatitis. N. Engl. J. Med. 362, 1675–1785

9
TEM-980; No. of Pages 10

Review Trends in Endocrinology and Metabolism xxx xxxx, Vol. xxx, No. x

96 Lutchman, G. et al. (2007) The effects of discontinuing pioglitazone in 102 Scherer, T. et al. (2011) Brain insulin controls adipose tissue lipolysis
patients with nonalcoholic steatohepatitis. Hepatology 46, 424–429 and lipogenesis. Cell Metab. 13, 183–194
97 Balas, B. et al. (2007) Pioglitazone treatment increases whole body fat 103 Jenkins, C.M. et al. (2004) Identification, cloning, expression, and
but not total body water in patients with non-alcoholic purification of three novel human calcium-independent
steatohepatitis. J. Hepatol. 47, 565–570 phospholipase A2 family members possessing triacylglycerol lipase
98 Adams, M. et al. (1997) Activators of peroxisome proliferator- and acylglycerol transacylase activities. J. Biol. Chem. 279, 48968–
activated receptor gamma have depot-specific effects on human 48975
preadipocyte differentiation. J. Clin. Invest. 100, 3149–3153 104 Kumari, M. et al. (2012) Adiponutrin functions as a nutritionally
99 Mudaliar, S. et al. (2013) Efficacy and safety of the farnesoid X regulated lysophosphatidic acid acyltransferase. Cell Metab. 15, 691–
receptor agonist obeticholic acid in patients with type 2 diabetes 702
and nonalcoholic fatty liver disease. Gastroenterology 145, 105 Caimari, A. et al. (2007) Regulation of adiponutrin expression by
574–582 feeding conditions in rats is altered in the obese state. Obesity 15,
100 Schweiger, M. et al. (2006) Adipose triglyceride lipase and hormone- 591–599
sensitive lipase are the major enzymes in adipose tissue 106 Kershaw, E.E. et al. (2006) Adipose triglyceride lipase: function,
triacylglycerol catabolism. J. Biol. Chem. 281, 40236–40241 regulation by insulin, and comparison with adiponutrin. Diabetes
101 Cao, Z. et al. (2013) Monoacylglycerol lipase controls endocannabinoid 55, 148–157
and eicosanoid signaling and hepatic injury in mice. Gastroenterology 107 Huang, Y. et al. (2010) A feed-forward loop amplifies nutritional
144, 808–817 regulation of PNPLA3. Proc. Natl. Acad. Sci. U.S.A. 107, 7892–78987

10

You might also like