Vous êtes sur la page 1sur 13

Review

The interaction of hepatic lipid and glucose metabolism in liver diseases


Lars P. Bechmann1, Rebekka A. Hannivoort2, Guido Gerken1, Gkhan S. Hotamisligil3, Michael Trauner4, Ali Canbay1,
1 Department of Gastroenterology and Hepatology, University Hospital Essen, University of Duisburg-Essen, Essen, Germany; 2Department of Gastroenterology and Hepatology, University Medical Center Groningen, University of Groningen, Groningen, The Netherlands; 3Departments of Genetics and Complex Diseases and Nutrition, Harvard School of Public Health, Boston, MA, USA; 4Division of Gastroenterology and Hepatology, Department of Internal Medicine III, Medical University of Vienna, Austria

Summary
It is widely known that the liver is a central organ in lipogenesis, gluconeogenesis and cholesterol metabolism. However, over the last decades, a variety of pathological conditions highlighted the importance of metabolic functions within the diseased liver. As observed in Western societies, an increase in the prevalence of obesity and the metabolic syndrome promotes pathophysiological changes that cause non-alcoholic fatty liver disease (NAFLD). NAFLD increases the susceptibility of the liver to acute liver injury and may lead to cirrhosis and hepatocellular cancer. Alterations in insulin response, b-oxidation, lipid storage and transport,

autophagy and an imbalance in chemokines and nuclear receptor signaling are held accountable for these changes. Furthermore, recent studies revealed a role for lipid accumulation in inammation and ER stress in the clinical context of liver regeneration and hepatic carcinogenesis. This review focuses on novel ndings related to nuclear receptor signaling including the vitamin D receptor and the liver receptor homolog 1 in hepatic lipid and glucose uptake, storage and metabolism in the clinical context of NAFLD, liver regeneration, and cancer. 2012 European Association for the Study of the Liver. Published by Elsevier B.V. All rights reserved.

Keywords: NAFLD; Fatty acid transporters; HCC; Liver regeneration; Nuclear receptors; ER stress. Received 13 May 2011; received in revised form 9 August 2011; accepted 10 August 2011 Corresponding author. Address: Department of Gastroenterology and Hepatology, University Hospital Essen, Hufelandstr 55, 45122 Essen, Germany. Tel.: +49 201 723 84713; fax: +49 201 723 5719. E-mail address: ali.canbay@uni-due.de (A. Canbay). Abbreviations: ACC, acetyl-CoA carboxylase; ATF6, activating transcription factor6; AMPK, AMP-activated protein kinase; apoB-48/100, apoprotein B-48/100; BA, bile acids; CD36/FAT, fatty acid translocase; ChREBP, carbohydrate responsive element binding protein; CPT-1, carnitine palmitoyltransferase-1; DAG, diacylglycerol; DGAT, diacylglycerol-acyltransferase; DNL, de novo lipogenesis; ER, endoplasmatic reticulum; FA, fatty acids; FABP, fatty acid binding protein; FAS, fatty acid synthase; FATPs, fatty acid transport proteins; FFA, free FA; FGF15/19, broblast growth factors 15 (mouse) and 19 (human); FoxO, forkhead box protein O; FXR, farnesoid X receptor; GCKR, glucokinase regulatory protein; GLP-1, glucagon like peptide-1; GLUT2, glucose transporter type 2; Got2 or mitochondrial aspartate aminotransferase [mAspAT], glutamate-oxaloacetate-transaminase 2; GPAT, glycerol-3-phosphate-acyltransferase; GS, glycogen synthase; G6Pase, glucose-6-phosphase; HCC, Hepatocellular carcinoma; HNF4a, hepatic nuclear factor-4-alpha; HNF6, hepatic nuclear factor 6; IGF-1, insulin-like-growth factor 1; IR, insulin resistance; LCFAs, long chain fatty acids; L-GCK, liver glucokinase; LRH1, liver receptor homolog 1; LXR, liver X receptor; MODY, maturity onset diabetes of the young; mTOR, mmmalian target of rapamycin; NAFLD, non-alcoholic fatty liver disease; NASH, non-alcoholic steatohepatitis; NRs, nuclear receptors; PEPCK, phosphoenolpyruvate carboxykinase; PGC1a, peroxisome proliferator-activated receptor c co-activator-1a; PHx, partial hepatectomy; PI3K, phosphoinositide-3-kinase; PKA, protein kinase A; PPARs, peroxisome proliferator-activated receptors; PYGL, glycogen phosphorylase; ROS, reactive oxygen species; SIRT1, sirtuin 1; SREBP-1c, sterol-response-binding-protein-1c; TAGs, triacylglycerols; TCA, tricarboxylic acid cycle; TRB3, tribbles-homologue 3; UPR, unfolded protein response; USF1, upstream stimulatory factor 1; VDR, vitamin D receptor; VLDL, very low-density lipoproteins.

Key Points 1

The prevalence of obesity and the metabolic syndrome is rising dramatically Obesity and the metabolic syndrome promote alterations in hepatic lipid and glucose metabolism and are linked to the pathophysiologies of NAFLD and liver cancer Lipid accumulation and fatty acid oxidation are important mechanisms in liver damage, repair and regeneration, partially due to induction of autophagy and ROS production Insulin and nuclear receptor (including PPAR, LXR, FXR, LRH1 and vitamin D receptor) signaling regulate hepatic lipid and glucose metabolism, and are closely interrelated Both metabolic pathways share common regulatory elements as well as metabolites and indistinguishably contribute to NAFLD, liver cancer, and liver regeneration

Introduction As the main detoxifying organ of the body, the liver also plays a central role in metabolic homeostasis and is a major site for synthesis, metabolism, storage and redistribution of carbohydrates,

Journal of Hepatology 2012 vol. 56 j 952964

JOURNAL OF HEPATOLOGY
proteins and lipids. The rapid increase in obesity worldwide is associated with an increase in the prevalence of non-alcoholic fatty liver disease (NAFLD), making NAFLD the most common liver disease in Western societies [1,2]. In order to understand the pathogenesis of NAFLD, we discuss the basic physiologic mechanisms of hepatic lipid and glucose metabolism. We also aimed at integrating recent clinical and mechanistic data and point out novel links between basic metabolic pathways and the pathophysiologies of NAFLD, liver regeneration, and carcinogenesis. NAFLD is characterized by lipid accumulation within hepatocytes and may progress to non-alcoholic steatohepatitis (NASH). Lipids derive from circulating fatty acids (FA) upon insulin resistance (IR)-induced dysregulation of peripheral lipolysis. FAs are translocated into the hepatocyte mainly by membrane bound transport proteins [3]. De novo lipogenesis (DNL) further contributes to hepatic steatosis [4]. Hepatocellular accumulation of lipotoxic intermediates such as diacylglycerol (DAG) and ceramids causes hepatic IR [5]. Hepatocytic lipid accumulation predisposes to overproduction of reactive oxygen species (ROS), endoplasmatic reticulum (ER) stress and lipotoxicity [68]. Recently, autophagy (especially in the form of macrolipophagy) has been identied to regulate intracellular lipid stores through degradation of lipid droplets and release of FAs into the cytosol as a rapid response to starvation [911]. Thus, disrupted autophagy might be essential to the pathogenesis of NASH via lipotoxicity-induced ER stress [12]. The individual steps in hepatic lipid metabolism are orchestrated by a delicate interplay of hormones, nuclear receptors, intracellular signaling pathways and transcription factors. Insulin signaling plays an important role in the regulation of FA metabolism, underscoring the close relation between lipid and glucose metabolism. Insulin affects DNL on multiple levels, via induction of lipogenic genes, activation of sterol-response-binding-protein1c (SREBP-1c) and Akt-regulated production of very low-density lipoproteins (VLDLs) [13]. Insulin, along with other mediators (e.g., calpain-1), further represses autophagy within the hepatocyte, and thus induces lipogenesis and represses lipid-degradation in the fed state as well [9,12]. On the other hand, glucagon-like peptide-1 (GLP-1) also alters hepatic lipid metabolism [14]. FA oxidation and the expression of fatty acid transport proteins (FATPs) are closely regulated by the nuclear receptor peroxisome proliferator-activated receptor (PPAR)a, and in the steatotic liver also by PPARc [15]. In the following paragraphs, we will discuss the function of hepatic lipid and glucose metabolism under normal conditions and their roles in NAFLD, acute liver injury and regeneration as well as hepatic carcinogenesis. This review focuses on the central role of nuclear receptor signaling in hepatic glucose and lipid metabolism including novel mechanisms like vitamin D receptor (VDR) and liver receptor homolog 1 (LRH1) signaling, autophagy and the interrelation of ER stress and metabolism Table 1. the bile duct to emulsify lipid droplets by their amphiphilic properties rendering them accessible to lipase hydrolyzation. Hydrolyzed lipids are then absorbed by enterocytes, where lipids are re-synthesized and packed into lipoprotein particles (i.e., nascent chylomicrons). Nascent chylomicrons are secreted into the lymphatic system, where they bypass the liver and enter the circulation within two hours after food intake [16]. During their journey through the vascular system, nascent chylomicrones lose two minor apoproteins (apoA-I and apoA-IV), which are replaced by apoE and apoC-II that are crucial for their further processing. ApoC-II activates adipocyte lipoprotein lipase (LPL), which facilitates the digestion of the chylomicron triacylglycerols (TAGs) into FAs and glycerol [17]. FAs are then partially taken up and stored in adipocytes, while chylomicron remnants re-enter the blood stream. ApoE is then recognized by the hepatocyte LDL receptor, the LDL receptor-related protein (LRP) and scavenger receptor B-1, which facilitate endocytotic uptake of the chylomicron remnants. As cholesteryl-ester enriched and triglyceride-depleted products of chylomicron metabolism, chylomicron remnants are nally processed by intracellular lysosomes and their glycerol, FA, cholesterol, amino acid and phosphate residues are metabolized and recycled into new VLDLs (see Fig. 1A for overview). Hepatic lipogenesis As mentioned above, hepatic FAs either derive from endogenous lipogenesis, are released from lysosomes by autophagy, or derive from the free FA (FFA) plasma pool via active uptake into the hepatocyte. Depending on the metabolic state, FAs are then either processed to TAGs and stored or rapidly metabolized. Indeed, b-oxidation is the predominant source of energy during the fasting state. Hepatic lipogenesis includes de novo synthesis of FAs from acetyl-CoA or malonyl-CoA and further processing to TAGs. In mammals, FA synthesis is catalyzed by acetyl-CoA carboxylase (ACC) and fatty acid synthase (FAS) an enzyme that is complexly regulated by various nuclear receptors (PPARa, PPARc and the bile acid receptor/farnesoid X receptor [FXR]) [1820]. FA elongation requires NADPH as a reducing reagent, which is provided by the pentose phosphate pathway (Fig. 2B). Remarkably, PPARa itself is activated by a phospholipid synthesized by FAS, indicating a feedback loop [21]. A close link between glucose and lipid metabolism is indicated by the fact that nuclear receptors (NRs) are also important mediators of insulin signaling and since DNL occurs under anabolic condition. The existence of such a link is further supported by the fact that insulin stimulates FAS expression via the phosphoinositide-3-kinase (PI3K) pathway [22]. On a transcriptional level, SREBP-1c and carbohydrate-responsive element binding protein (ChREBP), a glucose dependant transcription factor, synergistically induce expression of FAS and ACC [23]. As FAs and their metabolites are the major cause for lipotoxicity and promote the formation of ROS, FAs are stored for future use as TAGs, which are relatively inert and consist of three FAs esteried to a glycerol backbone. TAGs are then either stored in lipid droplets within the hepatocyte or processed to VLDL [7]. TAG synthesis is catalyzed by the enzymes mitochondrial glycerol-3-phosphate-acyltransferase (mtGPAT) and diacylglycerolacyltransferase (DGAT) [24]. TAGs are then packaged into VLDL particles, by conjugation to apoB-100 in a 5:1 TAG/cholesterol

Hepatic lipid metabolism Extrahepatic lipid metabolism Lipid metabolism starts with the intestinal absorption of dietary fats. In order to cross the intestinal lumen into the plasma, lipids are emulsied and hydrolyzed within the lumen. The healthy liver is crucial for intestinal lipid absorption via bile acids (BA) that are synthesized within the hepatocyte and secreted into

Journal of Hepatology 2012 vol. 56 j 952964

953

Review
Table 1. Overview of ligands and function of the most abundant nuclear receptors in hepatocytes.

Nuclear Natural ligands receptor PPAR Fatty acids

Chemical ligands/ agonists Fibrates

Function in lipid metabolism Regulates expression of FAS lipogenesis; CD36/FAT, FATPs FA-Uptake; Acetyl-CoA-synthetase, CPT-1 -oxidation

Function in glucose Role in NAFLD metabolism Regulates PEPCK, Fibrate treatment GSK3, improves IR Glycogen synthase glycogen metabolism; insulin sensitivity Activation; PPAR KO mice are protected from diet induced steatosis; Glitazones improve IR and TAG accumulation and increase adiponectin levels Agonist treatment improves hepatic steatosis in mice Induced; SHP is upregulated in NAFLD; FXR KO mice develop steatosis

Role in liver regeneration Delayed regeneration in PPAR KO mice

Role in HCC

PPAR activation is associated with liver carcinogenesis

PPAR

Prostaglandins

Glitazones

Regulates expression of Regulates GLUT-4 CD36/FAT FA-Uptake; expression SCD-1 FA-metabolism insulin sensitivity

Downregulated; glitazones inhibit hepatocyte proliferation

PPAR KO increases susceptibility to hepatic carcinogens; PPAR activation induces cell cycle arrest in hepatoma cells PPAR affects COX2 expression in hepatomas FXR KO mice develop hepatic tumors; interaction with Wnt/-catenin signaling Downstream target SHP reduced in human HCC Unknown

PPAR

Fatty acids

Glitazones; GW501516

Regulates SREBP-1c lipogenesis

Induces glycolysis and pentose phosphate pathway shunt

Agonist treatment improves liver regeneration in mice Mediates liver regeneration after partial hepatectomy; impaired regeneratory capacity in FXR KO mice Reduced activation after partial hepatectomy

FXR

Bile acids

Chenodeoxycholic acid (CDCA); GW4064

Regulates SREBP-1c lipogenesis;

Regulates PEPCK, glucose6-phosphatase Regulates VLDL formation gluconeogenesis via SHP

LXR

Hydroxysterols

T0901317; GW3965

Regulates SREBP1c, SCD-1, FAS lipogenesis; cholesterol metabolism

Regulates insulin receptor expression, GLUT-4 and IRS expression insulin sensitivity VDR represses PPAR signaling, direct effects unknown

Induced, LXR promotes hepatic lipogenesis

VDR

(1),25-Hydroxy- Calcitriolvitamin D3 derivates

VDR represses PPAR signaling, direct effects unknown

Downregulated; VDR KO mice develop steatosis; vitamin D protects from diet induced steatosis DLPC treatment improves hepatic TAG and FA accumulation

Vitamin D deficiency leads to impaired hepatic regeneration

VDR polymorphisms are associated with HCC

LRH1

Phospholipids

Dilauroyl Repression of SHP phosphatidylpotential effects on choline (DLPC) VLDL synthesis; DLPC treatment suppresses SREBP-1c, FAS, ACC-2 and SCD-1 expression in vivo

DLPC treatment reduces hepatic gluconeogenesis and improves insulin response in vivo

Unknown; Unknown; potential proliferative repression of SHP effects might promote tumor growth

ratio. These processes are controlled by SREBP-1c, the liver X receptor (LXR), FXR and ChREBP, which again links glucose and lipid metabolism [25]. Hepatic fatty acid uptake Another source for hepatic FAs is FFA recruitment from the plasma pool. FFAs are derived from lipolysis in adipocytes. This occurs usually in the fasting state, where it is promoted by catecholamines, natriuretic peptides and glucagon, while it is usually repressed by insulin [26]. However, the insulin-resistant state (obesity; metabolic syndrome) goes along with increased adipocyte lipolysis, leading to abundant FFAs in the plasma pool independently from the nutritional status [27]. FFAs are then 954

taken up by the hepatocytes in a facilitated fashion rather than by passive processes [28]. FATPs are thus in the focus of NAFLD research in which a variety of FATPs have been identied. While FATP1 is abundant in muscle and adipose tissue and is barely detectable in the liver [29], FATP2 and FATP5 are expressed in hepatocytes and most likely facilitate the major amount of FA uptake in the liver [30]. Other transport proteins include fatty acid binding protein (FABP), glutamateoxaloacetate-transaminase 2 (Got2; or mitochondrial aspartate aminotransferase [mAspAT], a membrane bound protein that mediates the endocytotic uptake of long-chain FAs), and caveolin-1 [3133]. Fatty acid translocase (CD36/FAT) is a membrane glycoprotein present on platelets, mononuclear phagocytes, adipocytes and hepatocytes with multiple functions, including thrombospondin-1 receptor

Journal of Hepatology 2012 vol. 56 j 952964

JOURNAL OF HEPATOLOGY
A
LRH1 BA LRP/ LDLR
Chylomicron remnants

B
Insulin, mTOR FoxO, Calpain1

Autophagy FXR BA Enterocytes FA uptake Dietary lipids Cholesterol CR LYS TAG FA Insulin, PPAR HDL FFA FA
Esterification

FGF-19

mtFA-oxidation Acetyl-CoA

Lym
NC

tic pha

trac

Insulin, PPAR, PPAR

ATP Malonyl-CoA De novo lipogenesis

Bloo
C LPL

rea d st

m
Insulin, LXR FXR, PPAR

Ketones TCA Acetyl-CoA (from PDC) LD

Adipocytokines Myocytes Adipocytes

VLDL

VLDL

TAG Cholesterol LDL/HDL

C
Mitochondrial -oxidation Acetyl-CoA

SCFA MCFA LCFA CPT1

Malonyl-CoA

D
Lipolysis mtFA-oxidation

TCA/ketones Peroximal -oxidation Acyl-CoA

LCFA VLCFA ROS


Hepatocyte stress

FFA

FA uptake

FA ER -oxidation

DNL

Lipotoxicity ROS

LCFA ER -oxidation Dicarboxylic acid

Apoptosis

Inflammation

Fibrosis

Fig. 1. Hepatic lipid metabolism in health and disease. (A) Dietary lipids are emulsied in the intestinal tract by bile acids (BAs). Hydrolyzed lipids are absorbed by enterocytes and packed into nascent chylomicrons (NCs). NCs enter the bloodstream via the thoracic duct where they receive important apoproteins (apoE; apoC-II) from HDL. These apoproteins are important for chylomicrons (C) to deliver TAGs and FAs to adipocytes and myocytes via lipoprotein lipase (LPL) degradation. Chylomicron remnants are taken up by hepatocytes via LDL-receptor (LDLR) and LDL receptor-related protein (LRP)-mediated endocytosis. BA synthesis is regulated by LRH1 and FXR, which activate BA export pumps. BA re-uptake by enterocytes stimulates FGF-19 release into the portal blood, which inhibits BA synthesis. (B) Free fatty acids (FFA) derive from lipolysis in adipose tissue and are actively taken up by various FA transporters under the control of insulin (Ins) and nuclear receptor signaling. Under physiologic conditions, the bulk of FAs is oxidized intramitochondrially and provides ATP and acetyl-CoA for the tricarboxylic acid cycle (TCA). Triglycerides (TAGs) derived from de novo lipogenesis are either stored in lipid droplets (LD) or packed into VLDL and exported into the blood stream. Acetyl-CoA for de novo lipogenesis is provided by the pyruvate dehydrogenase complex (PDC), which catalyzes oxidation of pyruvate, the end product of glycolysis. (C) Under physiologic conditions, b-oxidation of short-, medium- and long-chain FAs (SCFA, MCFA, LCFA) are degraded in mitochondria. Therefore, FAs are activated to acyl-CoA and shuttled across the mitochondrial membrane by carnitine palmitoyltransferase-1 (CPT1). Malonyl-CoA, an intermediate of lipogenesis, inhibits CPT1 and thus FA oxidation in the mitochondria. With FA abundance and in the insulin resistant state, LCFA and verylong-chain FAs (VLCFA) are oxidized in peroxisomes and the ER. This leads to an abundance of metabolites that induce formation of ROS and contribute to lipotoxicity. (D) In NASH, increased peripheral lipolysis, upregulation of FA transporters, an increase in DNL, and a switch from mitochondrial b-oxidation to peroxisomal and x-oxidation promote FA toxicity and the release of ROS. This leads to the induction of hepatocyte apoptosis, the invasion and activation of inammatory cells, as well as brogenesis.

Journal of Hepatology 2012 vol. 56 j 952964

955

Review
A
Insulin, SREBP-1c, HNF4, HNF6, USF1

B
GLY

GCK Glucose Glucose


GC KR

Insulin Akt/PI3K GSK3, Glu-6-P AMPK

Insulin, PKA, AMP

GLY Glycogen Glycogenosynthesis lysis

GLUT-2

Glucose

GLUT-2

Glucose

R GCK

ATP GCK ADP


P

GCK

G-6-Pase
P

GCK

Glucose Glucose GCKR GCK Pyruvate


Insulin Akt FoxO, PPAR, FXR, HNF4, CREBP, Acetyl-CoA Intermediates, AMP, PI3K CHREBP

Glycolysis

Gluconeo- Glycolysis genesis

Acetyl-CoA

Pyruvate Acetyl-CoA
Extracellular TCA cycle Glycogen synthesis

Pentose phosphate shunt NADPH DNL FA oxidation

Lactate
TCA cycle

Fig. 2. Regulation of hepatic glucose metabolism. (A) After intestinal absorption, glucose (Glu) reaches the hepatocyte via the portal vein. The insulin-independent glucose transporter 2 (GLUT2) shuttles Glu across the membrane. Abundance of glucose induces conformational changes of the glucokinase regulatory protein (GCKR), which binds to glucokinase (GCK) and keeps it in the nucleus in the fasting state. GCK is then released into the cytosol and phosphorylates Glu to glucose-6-phosphate (Glu6-P); depending on the nutritional state, it serves as a substrate for glycolysis or glycogen synthesis, respectively. GCK is transcriptionally regulated by insulin and nuclear receptor signaling. (B) Glu-6-P is a central intermediate in the hepatic glucose metabolism. It is degraded during glycolysis, which provides energy in the form of two ATP and two NADH molecules per glucose molecule. The product pyruvate is further decarboxylized to acetyl-CoA, which enters the intramitochondrial tricarboxylic acid cycle (TCA). Alternatively, Glu-6-P is degraded in the pentose-phosphate shunt, which provides NADPH, a co-substrate for DNL. Acetyl-CoA is an important product of the TCA, linking glucose and lipid metabolism, as it is the substrate for DNL (Fig. 1). Gluconeogenesis and glycogenolysis provide Glu-6-P as a substrate for glucose synthesis in the fasting state. Glucogenolysis is catalyzed by glycogen phosphorylase, activated by AMP, and repressed by insulin. The key enzyme in gluconeogenesis is PEPCK, which is repressed by insulin signaling via Akt-mediated FoxO phosphorylation and activated by PPARa.

activity, which has also been identied to facilitate FA uptake [34,35]. Besides the fact that the regulation of FATP activity is generally complex, the individual contribution of these FATPs to FA uptake has not been entirely claried yet. Nevertheless, signaling via PPARa again predominantly regulates the transcription of these transport proteins as combined with hormonal regulation via insulin and leptin [30,36]. Macroautophagy Autophagy has recently been implied to play a role in hepatic lipid homeostasis [37]. As a lysosomal pathway, it recycles dispensable cellular constituents into important energy sources during the fasting state [38]. Recent animal studies revealed that autophagy is a key process in hepatic lipolysis and lipid droplet degradation [10,39]. As mentioned above, lysosomes process chylomicron remnants as well as TAGs that accumulate during hepatic lipogen956

esis. During starvation, macroautophagy leads to the fusion of lysosomes and lipid droplets into autophagosomes, which are then degraded; FAs are thus released and can be catabolized via b-oxidation. Starvation leads to repression of the so-called mammalian target of rapamycin (mTOR), an insulin downstream target that inhibits autophagy. Interestingly, rapamycin treatment also downregulates SREBP-1c in primary hepatocytes, which suggests an effect of rapamycin, independent of mTOR, on the activity of forkhead box protein O (FoxO) [40]. Long-term repression of autophagy is accounted for by insulin action, and Akt mediated de-phosphorylation of FoxO, a transcriptional activator of autophagy related genes (ATGs). Indeed, FoxO simultaneously represses SREBP-1c activation and thus DNL. Thus, insulin receptor activation induces DNL and represses autophagy-mediated lipid droplet degradation, both short- and long-term, via two distinct mechanisms. However, in the insulin-resistant state and in obesity per se, hepatic mTOR is over-activated and calpain, a repressor of

Journal of Hepatology 2012 vol. 56 j 952964

JOURNAL OF HEPATOLOGY
ATGs, is induced [9]. Despite FoxO activation in IR, mTOR and calpain activation account for the repression of hepatic autophagy in obese individuals. Fatty acid oxidation Oxidation of FAs occurs within mitochondria, peroxisomes and the ER, and facilitates degradation of activated FAs to acetyl-CoA. It is a rapid and effective way of energy allocation since, for example, the oxidation of one molecule of palmitate produces up to 129 ATP equivalents [41]. FAs are activated by acyl-CoA-synthetase to acyl-CoA in the cytosol. This process is indispensable for enabling FAs to cross membranes and enter organelles. While short- and medium-chain FAs pass the mitochondrial membrane without activation, activated long-chain fatty acids (LCFAs) are shuttled across the membrane via carnitine palmitoyltransferase-1 (CPT1). Malonyl-CoA, an early intermediate of DNL that accumulates upon insulin receptor activation, is an allosteric inhibitor of CPT1 [42]. Thus, in the fed state, FA oxidation is inhibited and DNL-promoted, allowing for storage and distribution of lipids [43]. In general, short-, medium- and LCFAs are oxidized within mitochondria (b-oxidation), while toxic, very-long-chain FAs are oxidized within peroxisomes (Fig. 1C). In diabetes or FA overload, cytochrome P450 (a.k.a. CYP4A)-dependent x-oxidation of LCFAs occurs in the ER and induces ROS and lipid peroxidation [44]. During the process of b-oxidation, electrons are indirectly donated to the electron transport chain to drive ATP synthesis. Acetyl-CoA can be further processed via the tricarboxylic acid cycle (TCA) or, in the case of FA abundance, be converted into ketone bodies [45]. PPARa and insulin signaling are again involved in the regulation of FA oxidation and the formation of ketone bodies via transcriptional regulation of mitochondrial HMG-CoA synthase. Interestingly, succinyl-CoA, an intermediate of the TCA, inactivates mitochondrial HMG-CoA synthase by succinylation, which once more intercalates glucose- and lipid metabolism [46]. For a simplied overview of key metabolic pathways in hepatocellular lipid homeostasis, see Fig. 1B. In contrast to GLUT4, which is expressed by muscle and adipose tissue, the expression and activity of GLUT2 is independent of insulin signaling. In pancreatic islet cells, GLUT2 is thus also referred to as a glucose sensor [47]. Once taken up by the hepatocyte, glucose is phosphorylated to glucose-6-phosphate by liver glucokinase (L-GCK; Fig. 2), the rate limiting enzyme for hepatic glucose utilization [48]. In contrast to other hexokinases, GCK (syn.: hexokinase IV) is not inhibited by its product, which allows for postprandial glycogen storage within the hepatocyte. In the fasting state, L-GCK is inactive and bound to glucokinase regulatory protein (GCKR) within the nucleus. Post-prandial glucose abundance and insulin-action synergistically cause rapid dissociation of L-GCK from GCKR and translocation to the cytoplasm [49]. L-GCK is transcriptionally regulated by SREBP-1c, hepatic nuclear factor-4-alpha (HNF4a), hepatic nuclear factor 6 (HNF6), FoxO1, and upstream stimulatory factor 1 (USF1) (see Fig. 2A). Indeed, mutations in the GCK gene have been associated with IR and the pathogenesis of maturity-onset diabetes of the young (MODY) in several studies [50,51]. Glycolysis and glycogen synthesis Glucose-6-phosphate is either further processed in glycolysis or utilized for glycogen synthesis, depending on the systemic metabolic state. Glycolysis, a ten-step process, metabolizes glucose to pyruvate with a net gain of two ATP and two NADH molecules per glucose molecule. Glycolysis is regulated by L-GCK, which provides glucose-6-phosphate, phosphofructokinase, which is inhibited by its product fructose-1,6-bisphosphate, AMP and pyruvatekinase (PK), the nal step in glycolysis. PK is activated by its substrate and inhibited by abundance of ATP. Insulin, epinephrine, and glucagon also regulate PK via the PI3K pathway and ChREBP induces transcription of PK in the presence of glucose [23]. Pyruvate is further decarboxylized to acetyl-CoA and then processed in the TCA or utilized for DNL. The pentose phosphate pathway is an alternative way for degradation of glucose-6-phosphate in hepatocytes, which provides the cell with NADPH, an important antioxidant and co-substrate for DNL and cholesterol synthesis. In hepatocellular carcinoma (HCC), glycolytic activity is dramatically upregulated and associated with increased hexokinase 2 activity and expression of GLUT1, leading to altered glucose utilization, which has therapeutic and diagnostic implications (for the so-called Warburg effect, Fig. 3B and C) [52]. Glycogen synthesis is catalyzed by glycogen synthase (GS) after conversion of glucose-6-phosphate to UDP-glucose [53]. GS is regulated by the allosteric activator glucose-6-phosphate and is inactive in the phosphorylated state. Glycogen synthase kinase 3 (GSK3) phosphorylates GS and is a downstream target of Akt/PI3K and thus insulin signaling. GSK3 is a multifunctional kinase, involved in cell senescence, apoptosis and lipid metabolism via phosphorylation of SREBP-1c [54]. Other protein kinases that phosphorylate GS are AMP-activated protein kinase (AMPK) and protein kinase A (PKA). Insulin activates glycogen synthesis via repression of PKA. GS synthesizes the glycogen polymer, which is further branched by a branching enzyme. Glycogenolysis and gluconeogenesis In the fasting state, the liver supplies the body with energy by breaking down glycogen, and following prolonged fasting by gluconeogenesis [55]. Glycogen breakdown is catalyzed by glycogen 957

Key Points 2

Insulin activates DNL via PI3K-mediated FAS expression SREBP-1c and ChREBP transcriptionally activate FAS and ACC as well as VLDL assembly Insulin represses peripheral lipolysis and liberation of FFAs from adipocytes Insulin represses autophagy via activation of mTOR and repression of FoxO phosphorylation Insulin represses FA oxidation

Hepatic glucose metabolism Hepatocyte glucose uptake In the postprandial state, blood glucose is taken up by the hepatocyte via the glucose transporter type 2 (GLUT2) a membranebound transporter with high capacity and low afnity for glucose.

Journal of Hepatology 2012 vol. 56 j 952964

Review
A B
Hepatoma cell

Obesity

Healthy liver

FFA

Glucose

Insulin resistance

Steatosis

GLUT-2 Glucose

Tumor proliferation

TNF

GLUT-1

GCK

HK2

Inflammation

NASH

Glucose IL-6 IL-1 ER-stress ROS

Glycolysis Pentose phosphate shunt NADPH

Fibrosis

Cirrhosis

Pyruvate
TUMORIGENESIS

Acetyl-CoA Lactate
TCA cycle

JNK JAK/STAT NFB

DNL Acetyl-CoA

HCC

Fig. 3. Hepatic lipid and glucose metabolism in HCC. (A) Mediators of NAFLD progression also contribute to carcinogenesis in the liver. In general, obesity and diabetes have been identied as risk factors for cancer development as well as inammation. Cirrhosis is a precancerous condition and, in fact, most HCCs derive from cirrhotic livers. Important mediators of insulin resistance and lipotoxicity also induce dysplasia and carcinogenesis. This gure gives a brief overview of different cytokines and cell signaling pathways involved in HCC development. (B) Expression of GLUT1 in hepatoma cells leads to increased hepatic glucose utilization. Glucokinase is downregulated, but hexokinase 2 (HK2) is now expressed and phosphorylates glucose with a higher afnity. Aerobic glycolysis leads to a rapid, but rather ineffective energy supply to the proliferating cancer cell. However, this process, referred to as the Warburg effect, facilitates uptake and de novo synthesis of nutrients (nucleotides, amino acids, lipids), available for cell proliferation and tumor growth. (C) Clinically, this effect is utilized in PET diagnostics. An increase in cancer cell glucose uptake, as visualized by FDG PET/CT, acts as an important tool in HCC diagnostics.

958

Journal of Hepatology 2012 vol. 56 j 952964

JOURNAL OF HEPATOLOGY
phosphorylase (PYGL), which cleaves glucose from the glycogen polymer and produces glucose-1-phosphate, which is converted to glucose-6-phosphate by phosphoglucomutase. A debranching enzyme cleaves the last four glucose monomers. PYGL is regulated through allosteric activation by AMP and via phosphorylation by PKA, which is inhibited by insulin. Hepatic gluconeogenesis occurs during prolonged fasting episodes and begins intramitochondrially by the induction of pyruvate carboxylase in abundance of acetyl-CoA. Interestingly, inhibition of hepatic CPT-1 and thus mitochondrial fatty acid oxidation signicantly represses hepatic gluconeogenesis in mice [56]. Gluconeogenesis is further regulated via allosteric and

Key Points 3

NRs are highly conserved ligand-activated transcription factors. Specific ligands of most NRs are known, are either endogenous (bile acids; fatty acids; hormones; vitamins) or exogenous (drugs; xenobiotic toxins), and induce conformational changes affecting the transcriptional activation of downstream targets [64, 65]. Most NRs heterodimerize with the retinoid-X-receptor. Table 1 sketches the most abundant NRs in the liver, their ligands, and their roles in hepatic lipid and glucose homeostasis PPAR is the most abundant PPAR in the healthy liver. It regulates FA oxidation and uptake as well as gluconeogenesis (PEPCK) and glycogen synthesis (GSK3; GS) (see above). Moreover, PPAR activation has anti-inflammatory effects and (in mice) induces hepatocyte proliferation via induction of microRNA let7c-signaling cascades [66, 67]. In mouse models, but not in humans, activation of PPAR induces hepatic carcinogenesis, which is in part explained by ROS production [68, 69] and induction of microRNA let7c that degrades the c-myc oncogene. Initially an orphan NR, PPAR is known to be the receptor for fibrates in the therapy of hypertriglyceridemia and recently a diacylglycerol has been identified as an endogeneous ligand [21], produced by the enzyme FAS, which in turn is activated by PPAR and other NRs. Emphasizing the cross-talk between NRs, FXR and LXR are known to activate PPAR signaling while VDR appears to repress PPAR signaling [58, 70, 71]. Post-transcriptionally, PPAR protein expression is repressed by microRNA 10b, which is associated with metastasis formation in a variety of cancers [72] PPAR is mainly expressed in adipocytes, where it promotes lipid uptake and TAG storage by upregulation of the LDL receptor and CD36/FAT, induces stearoylCoA-desaturase-1 (SCD1), increases insulin sensitivity by induction of GLUT4, and decreases TNF levels and thus systemic IR [73]. In NAFLD, PPAR is upregulated in liver tissue, and liver-specific PPAR KO mice are protected from diet-induced steatosis [74, 75]. PPAR has anti-inflammatory features as it represses NFB signaling [76]. Recently, activation of PPAR has been shown to reduce hepatic lipogenesis by suppression of SREPB-1c [76, 77] in addition to increasing the hepatic glucose catabolism as well as muscular oxidation The PPAR co-activator-1 (PGC1) is a transcriptional co-activator that regulates mitochondrial biology and energy homeostasis [78]. PGC1 is induced by fasting and activates FA oxidation and gluconeogenesis via induction of PPAR, FoxO1 and HNF-4 [79]. Sirtuin 1 (SIRT1) deacetylates PGC1 and thus regulates its activity [80]

FXR, the bile acid receptor and its downstream target, small heterodimer partner (SHP), have recently been identified to play a central role in hepatic lipid metabolism as BA administration lowers blood lipids and the TAGs hepatic content. This effect is partially reversed in SHP KO mice and is mediated by transcriptional repression of SREBP-1c [81]. Thus, FXR represses hepatic DNL and VLDL export. FXR KO mice develop hepatic steatosis, and treatment with FXR agonist attenuates hepatic steatosis in mice when fed a methionine- and choline- deficient diet (MCD), a model that induces steatosis and oxidative stress, rather than insulin resistance [82, 83]. Other studies also revealed that FXR agonists improve IR [64, 84]. In the intestine, FXR activation induces secretion of fibroblast growth factors 15 (mouse) and 19 (human) (FGF15/19), which activates hepatic FA oxidation and insulin response via activation of the FGF receptor 4 [85, 86]. BAs have additional effects on lipid and glucose metabolisms via a G-protein-coupled receptor (TGR5/GBAR), which regulates energy expenditure in brown adipose tissue (and possibly skeletal muscle), and promotes intestinal GLP-1 secretion [87]. Thus, BAs may now be viewed as enterohepatic hormones that circulate with absorbed dietary lipids and fine-tune their metabolism Natural ligands of LXR are hydroxysterols. LXR is expressed primarily in hepatocytes, adipose tissue and macrophages, whereas LXR is expressed ubiquitously [88]. LXR regulates SREBP-1c, FAS and SCD-1 expression and LXR agonist treatment leads to hypertriglyceremia secondary to induction of hepatic lipogenesis [89]. LXR increases hepatic glucose uptake by induction of GLUT4 expression and has anti-inflammatory attributes as it modulates the innate immune response [90, 91] LRH1 has recently been implied as a novel NR in BA homeostasis by transcriptional activation of Cyp7A1 [92]. Interestingly, LRH1 activation by dilauroyl phosphatidylcholine improves the hepatic insulin response and hepatic lipid accumulation in high fat dietfed mice [93]. Accordingly, SREBP-1c, FAS and SCD-1 mRNA levels were downregulated in the treatment group The VDR is a novel target in the field of hepatology. Recent studies have shown that vitamin D supplementation protects mice from diet-induced steatosis and VDR KO mice spontaneously develop hepatic steatosis [94, 95]. Polymorphisms in VDR are associated with HCC development in patients with alcoholic cirrhosis, and VDR activation acts antiproliferative effects [96, 97]. In HCV patients, low vitamin D levels are associated with fibrosis progression, and the VDR regulates T-cell activation [97, 98]

Journal of Hepatology 2012 vol. 56 j 952964

959

Review
transcriptional activation of phosphoenolpyruvate carboxykinase (PEPCK), fructose-1,6-bisphosphatase and glucose-6-phosphatase (G6Pase). Overexpression of PEPCK promotes IR in mice [57]. Insulin action represses PEPCK expression via Akt-mediated FoxO1 phosphorylation. This pathway is repressed by activation of PPARa signaling-mediated induction of the pseudokinase tribbles-homologue 3 (TRB3), and TRB3 expression is associated with IR in patients [58,59]. Other NRs (PGC1a, FXR), glucagon and glucocorticoids also mediate expression of PEPCK and G6Pase [55,60]. FoxO1, a transcriptional activator of PEPCK and G6Pase, is both directly and indirectly activated by PGC1a, HNF4a, CREBP and PPARa [61]. PEPCK knock-out mice not only show decreased gluconeogenesis, but more importantly a decreased removal of TCA anions, which causes hepatic TAG accumulation and steatosis [62]. In mice, PEPCK overexpression in the striated muscle interestingly leads to longevity and an impressive increase in muscle strength and endurance; this phenotype is partially explained by an optimized mitochondrial oxidation of TAG [63]. These processes again demonstrate the close inter-relation between gluconeogenesis and hepatic lipid metabolism (Fig. 2B). stress and inhibits autophagy, a novel pathway in the biology of hepatic lipid droplets [79,80].

NAFLD, insulin resistance and lipotoxicity NAFLD represents the most prevalent liver disease in Western societies. It presents with a wide spectrum ranging from simple steatosis or non-alcoholic fatty liver (NAFL) to fully developed NASH with or without brosis. NASH can progress to brosis with an increased risk to develop end-stage liver disease or hepatocellular carcinoma (HCC) [1]. Hepatic steatosis is dened as an intrahepatic accumulation of TAGs. In parallel, abundant FAs cause lipotoxicity via the induction of ROS release, which causes inammation, apoptosis, and thus the progression to NASH and brogenesis [81]. As described above, most FAs derive from the circulation secondary to increased lipolysis in adipose tissue as well as DNL [3]. Obesity increases the TNFa production in adipocytes, which facilitates adipocyte IR and increases lipolysis rate [82]. Thus, the circulating FFA pool is increased in obese individuals and accounts for the majority of liver lipids in NAFLD [13]. As mentioned, uptake of FFAs into the hepatocyte is facilitated by a variety of FATPs; several studies found an upregulation of these transporters in NAFLD and NASH as well as a correlation with disease severity [65,83,84]. Fifteen percent of the lipid content within the steatotic liver derives from an increased dietary intake of lipids [3]. DNL may account for up to 30% of TAGs in steatotic livers, a mechanism that involves dysregulation in SREBP-1c- and FoxO-mediated hepatic insulin signaling [4,85,86]. Since autophagy-related genes are transcriptionally activated by FoxO and insulin action modulates autophagy, recent studies showed that macroautophagy is dysregulated in the metabolic syndrome [87]. Accordingly, in conditional atg7-knockout mice, Singh et al. observed an increase in hepatic lipid accumulation, and in genetic and dietary mouse models for obesity and hepatic steatosis, autophagy-related genes were downregulated [10,12]. The long-lasting paradigm claiming TAG accumulation to be the rst hit that predisposes to further liver damage in the pathogenesis of NASH has recently been replaced by a more complex model as emerging evidence points to FAs and their metabolites as the true lipotoxic agents [7]. Interestingly, lipid accumulation and altered composition of phospholipids within ER membranes further promotes ER stress and IR in obese mice [8]. Cytosolic TAGs are thus now considered to be inert and, in fact, lipid droplet accumulation has recently been found to be hepatoprotective [88]. Notably, genetic deletion of DGAT2 (responsible for TAG formation) increases hepatocellular injury in MCD-fed mice despite a reduction in the content of hepatocellular TAGs [89]. However, TAG accumulation and lipid droplet formation accompany and parallel pathophysiologic mechanisms in NASH. FFAs are thus now in the focus of basic research and FAs as well as acyl-CoA and acetyl-CoA have been identied as potential causes of lipotoxicity [90]. FAs have been found to activate Toll-like receptors and initiate the extrinsic apoptosis cascade [91,92]. FAs also interfere with NR signaling, which might additionally inuence the extent of hepatocyte damage and further promote IR and ER stress [93,94]. Accordingly, b-oxidation of LCFA within peroxisomes and x-oxidation within the ER are upregulated in NASH and contribute to lipotoxicity and ROS formation [95,96]. This might be secondary to inhibition of

Hepatic lipid and glucose metabolism in liver injury Serum FFA levels correlate with hepatocyte apoptosis and FAs were found to activate death receptor-mediated apoptosis [64,65]. On the other hand, high glucose concentrations induce apoptosis in hepatoma cell lines, and the HOMA score is associated with hepatocyte apoptosis in NAFLD patients [66,67]. These observations indicate that hepatocyte apoptosis due to an imbalance in cell metabolism has clinical implications for NAFLD, liver regeneration, as well as brogenesis and carcinogenesis. The effects of systemic IR on liver injury via induction of TNFa are discussed in detail below. Intrahepatic fat deposition and obesity decrease hepatic blood ow by direct compression and systemic hypercatecholemia and thus inhibit mitochondrial function and cause formation of ROS, which induces Kupffer cell activation, hepatic inammation, and hepatocyte apoptosis [68]. Interestingly, mitochondrial stress may be partially reversed by treatment with insulin-like-growth factor 1 (IGF-1) and PPARc agonists [69,70]. Alcohol-fed PPARa knock-out mice develop a phenotype that mimics alcoholic liver disease in humans, which is linked to ROS accumulation [71]. Overexpression studies identied that the expression of PEPCK links mitochondrial dysfunction with the ER stress response [72]. As mitochondria are the main organelles in energy combustion, the ER is the major site of protein folding and trafcking. Recently, activation of the unfolded protein response (UPR) within the ER has been implied as a key modulator of cellular inammation and is linked to IR, lipid and glucose metabolism [73]. Three membrane-bound proteins regulate the UPR within the ER; PKR-like eukaryotic initiation factor 2a kinase (PERK), inositol-requiring enzyme 1 (IRE1), and activating transcription factor-6 (ATF6). The protein kinase PERK affects transcriptional regulation of rRNA, which activates NFjB and ATF4 signaling. NFjB regulates inammatory signaling (IL-6, TNFa), and ATF4 regulates glucose metabolism [74,75]. SREBP-1c activation occurs during ER stress and thus affects the lipid metabolism and the ER possibly regulates the number, composition and quality of lipid droplets [7678]. Furthermore, induction of the mTOR pathway in obese individuals activates SREBP-1c, promotes ER

960

Journal of Hepatology 2012 vol. 56 j 952964

JOURNAL OF HEPATOLOGY
mitochondrial b-oxidation due to an accumulation of malonylCoA and the inhibition of CPT1 [3,97,98]. In fact, recent studies indicate that activation of mitochondrial FA oxidation protects from steatosis and IR [99,100]. As mentioned above, DAGs and ceramides might as well contribute to hepatocyte damage and IR in NAFLD [7]. These stressors induce a variety of intracellular and paracrine mechanisms that may promote hepatocellular damage. FAs induce the production of TNFa, and hepatic TNF receptor expression correlates with the disease severity in NAFLD [101]. TNF receptor activation increases expression of SREBP-1c, which induces hepatic lipogenesis and lipid accumulation [102]. As TNFa-mediated effects are antagonized by adiponectin, adiponectin receptors are actually downregulated in NASH [103]. TNFa activation is further paralleled by death-receptor expression, which facilitates activation of the extrinsic apoptosis cascade. Apoptosis indeed is the predominant form of hepatocellular injury in NASH [64,104]. In fact, apoptotic activity within the diseased liver correlates with disease severity and thus cleaved cytokeratin-18 fragments in the serum of NAFLD could effectively be utilized as surrogate markers for the progression of NAFLD [105]. As previously mentioned, FA accumulation also leads to induction of ER stress and ROS formation, which again promotes hepatic injury [1,106]. In summary, while hepatic TAG accumulation seems to be a benign symptom of hepatic steatosis, FA metabolites contribute to the progression of NAFLD to NASH. IR promotes the recruitment of FFAs from the serum pool as well as intrahepatic FA accumulation, which induces apoptosis and ROS formation. FAs themselves also promote hepatic IR via TNF receptor activation, hence indicating a vicious circle of lipid accumulation and IR as a crucial mechanism in the pathogenesis of NASH (Fig. 1D). Liver cancer Emerging data supports a role for lipid and glucose metabolism in hepatocellular carcinogenesis [115]. First of all, the increase in the prevalence of NAFLD is closely associated with the development of cirrhosis and NASH-related HCC [116]. Accordingly, NASH might account for the majority of HCC in the context of a cryptogenic cirrhosis [117]. Obesity and diabetes have been identied as independent risk factors for developing HCC (Fig. 3A) [118,119]. Several studies could recently dissect the effects of individual confounders of the metabolic syndrome and clearly demonstrate a signicant risk for NAFLD patients to develop HCC [120]. Intriguingly, HCC in NAFLD not only arises from cirrhotic tissue; this might imply a special role for lipid metabolism and lipotoxicity in the carcinogenesis within this cohort [121]. To date however, the vast majority of HCC cases arise from patients with cirrhosis secondary to chronic HBV or HCV infection [122]. Wu et al. recently identied various alterations in multiple genes involved in hepatic lipogenesis, FA oxidation, and TAG metabolism in tumors derived from HCV patients compared to HCV-positive control patients [123]. Mechanistically, mediators and signaling cascades involved in the hepatic IR and the regulation of hepatic lipid metabolism are in the focus of HCC research. Several studies identied the insulin receptor downstream targets PI3-kinase/Akt and IRS to be activated in HCC when compared to surrounding healthy tissue, which mediates GSK3b phosphorylation and mTOR activation [124]. These mediators not only interact with b-catenin signaling, but also with autophagy and glycogen synthesis. In fact, knock-out mice with specic deletions in autophagy-related genes show spontaneous development of hepatocellular dysplasia [125]. As pointed out before, TNFa induces hepatic IR and thus promotes lipid accumulation. TNF and IL-6 signaling is also involved in the activation of the JAK/STAT and ERK pathways; both mediators are crucial for the development of obesity induced-HCC in a rodent model [126]. On the other hand, JAK/STAT signaling is closely linked to, and partially modulated by, the adipocytokines leptin and adiponectin [127]. NRs play a crucial role in hepatic lipid and glucose homeostasis (see also above) and are important mediators of hepatic carcinogenesis. PPARa mediates hepatic FA-oxidation and transport as well as Akt phosphorylation via its downstream target TRB-3 [15]. PPARa agonists promote carcinogenesis in a rodent model for HCV infection [128]. Interestingly, PPARa agonists not only interfere with the binding of PPARa to SREBP-1c, but also with STAT3 in hepatoma cell lines, thus indicating a potential mechanism for its carcinogenic properties [129]. In contrast, PPARc agonists prevent NAFLD progression and, more interestingly, were in some studies described to act antineoplastically [130,131]. Notably, the majority of mechanistical data in hepatic carcinogenesis derives from hepatoma cell lines and murine models that are only of limited value for the elucidation of hepatic carcinogenesis in humans.

Acute liver injury and regeneration Injured livers produce cytokine signals that trigger adipose tissue to release FAs into the circulation. Indeed, in the acute response, hepatocytes initiate the transcription of lipogenic genes and accumulate TAGs in intracellular lipid droplets. In liver regeneration (e.g., after partial hepatectomy (PHx)), lipid droplet formation is essential to propagate the proliferative response by sufciently supplying the organs energy household. Droplet-stored lipids are also used for synthesizing new lipoproteins, bile acids, and entire membranes [107,108]. As previously mentioned, emerging data supports a close connection between BA and FA metabolism. In this context, FATP-5 has also been suggested to participate in BA metabolism as a bile acid-CoA ligase [109,110]. FATP-5 is expressed preferentially towards the space of Disse and closely follows the hepatic sinusoids [109] where FAs and BAs are absorbed from the enterohepatic circulation. Intriguingly, it has been recently shown that elevated BAs accelerate liver regeneration after PHx by bile acid receptor (FXR)-dependent signals [107,111]. Autophagy in acute liver injury might to a certain degree act hepatoprotectively as it rapidly supplies the hepatocyte with energy. However, hyperactivation of autophagy induces cell death, and the necrosis rate is actually a predictor of liver failure [9,112]. Furthermore, as we previously elucidated key processes in the interplay between adipose tissue and the liver, adiponectin was identied as an important mediator of STAT3 signaling in the regenerating liver [113,114].

Conclusions Hepatic lipid and glucose metabolism are closely interrelated with inammatory, proliferative and apoptotic signaling within the liver. In the liver, these catabolic and anabolic pathways can hardly be separated. They share intermediate metabolites and receptor signaling, and go hand in hand in the pathogenesis

Journal of Hepatology 2012 vol. 56 j 952964

961

Review
of the most common liver diseases. Intriguingly, the case that these metabolic pathways are also involved in cell proliferation, regeneration and carcinogenesis implies potent future therapeutic approaches for life-threatening diseases. The enhanced understanding of these basic mechanisms is thus imperative as we witness a rising prevalence of obesity and the metabolic syndrome.
[13] Savage DB, Semple RK. Recent insights into fatty liver, metabolic dyslipidaemia and their links to insulin resistance. Curr Opin Lipidol 2010;21:329336. [14] Gupta NA, Mells J, Dunham RM, Grakoui A, Handy J, Saxena NK, et al. Glucagon-like peptide-1 receptor is present on human hepatocytes and has a direct role in decreasing hepatic steatosis in vitro by modulating elements of the insulin signaling pathway. Hepatology (Baltimore, Md) 2010;51:15841592. [15] Pyper SR, Viswakarma N, Yu S, Reddy JK. PPARalpha: energy combustion, hypolipidemia, inammation and cancer. Nucl Recept Signal 2010;8:e002. [16] Timlin MT, Parks EJ. Temporal pattern of de novo lipogenesis in the postprandial state in healthy men. Am J Clin Nutr 2005;81:3542. [17] Merkel M, Eckel RH, Goldberg IJ. Lipoprotein lipase: genetics, lipid uptake, and regulation. J Lipid Res 2002;43:19972006. [18] Knight BL, Hebbachi A, Hauton D, Brown AM, Wiggins D, Patel DD, et al. A role for PPARalpha in the control of SREBP activity and lipid synthesis in the liver. Biochem J 2005;389:413421. [19] Schadinger SE, Bucher NL, Schreiber BM, Farmer SR. PPARgamma2 regulates lipogenesis and lipid accumulation in steatotic hepatocytes. Am J Physiol Endocrinol Metab 2005;288:E11951205. [20] Shen LL, Liu H, Peng J, Gan L, Lu L, Zhang Q, et al. Effects of farnesoid X receptor on the expression of the fatty acid synthetase and hepatic lipase. Mol Biol Rep 2011;38:553559. [21] Chakravarthy MV, Lodhi IJ, Yin L, Malapaka RR, Xu HE, Turk J, et al. Identication of a physiologically relevant endogenous ligand for PPARalpha in liver. Cell 2009;138:476488. [22] Sul HS, Latasa MJ, Moon Y, Kim KH. Regulation of the fatty acid synthase promoter by insulin. J Nutr 2000;130:315S320S. [23] Dentin R, Girard J, Postic C. Carbohydrate responsive element binding protein (ChREBP) and sterol regulatory element binding protein-1c (SREBP1c): two key regulators of glucose metabolism and lipid synthesis in liver. Biochimie 2005;87:8186. [24] Coleman RA, Lee DP. Enzymes of triacylglycerol synthesis and their regulation. Prog Lipid Res 2004;43:134176. [25] Postic C, Girard J. The role of the lipogenic pathway in the development of hepatic steatosis. Diabetes Metab 2008;34:643648. [26] Arner P. Human fat cell lipolysis: biochemistry, regulation and clinical role. Best Pract Res Clin Endocrinol Metab 2005;19:471482. [27] Delarue J, Magnan C. Free fatty acids and insulin resistance. Curr Opin Clin Nutr Metab Care 2007;10:142148. [28] Berk PD. Regulatable fatty acid transport mechanisms are central to the pathophysiology of obesity, fatty liver, and metabolic syndrome. Hepatology (Baltimore, Md) 2008;48:13621376. [29] Martin G, Nemoto M, Gelman L, Geffroy S, Najib J, Fruchart JC, et al. The human fatty acid transport protein-1 (SLC27A1; FATP-1) cDNA and gene: organization, chromosomal localization, and expression. Genomics 2000;66:296304. [30] Ge F, Zhou S, Hu C, Lobdell Ht, Berk PD. Insulin- and leptin-regulated fatty acid uptake plays a key causal role in hepatic steatosis in mice with intact leptin signaling but not in ob/ob or db/db mice. Am J Physiol Gastrointest Liver Physiol 2010;299:G855866. [31] Zhou SL, Stump D, Sorrentino D, Potter BJ, Berk PD. Adipocyte differentiation of 3T3L1 cells involves augmented expression of a 43-kDa plasma membrane fatty acid-binding protein. J Biol Chem 1992;267:1445614461. [32] Zhou SL, Stump D, Kiang CL, Isola LM, Berk PD. Mitochondrial aspartate aminotransferase expressed on the surface of 3T3L1 adipocytes mediates saturable fatty acid uptake. Proc Soc Exp Biol Med 1995;208:263270. [33] Trigatti BL, Anderson RG, Gerber GE. Identication of caveolin-1 as a fatty acid binding protein. Biochem Biophys Res Commun 1999;255:3439. [34] Abumrad NA, el-Maghrabi MR, Amri EZ, Lopez E, Grimaldi PA. Cloning of a rat adipocyte membrane protein implicated in binding or transport of longchain fatty acids that is induced during preadipocyte differentiation. Homology with human CD36. J Biol Chem 1993;268:1766517668. [35] Silverstein RL, Febbraio M. CD36, a scavenger receptor involved in immunity, metabolism, angiogenesis, and behavior. Sci Signal 2009;2:re3. [36] Wierzbicki M, Chabowski A, Zendzian-Piotrowska M, Gorski J. Differential effects of in vivo PPAR alpha and gamma activation on fatty acid transport proteins expression and lipid content in rat liver. J Physiol Pharmacol 2009;60:99106. [37] Czaja MJ. Autophagy in health and disease. 2. Regulation of lipid metabolism and storage by autophagy: pathophysiological implications. Am J Physiol Cell Physiol 2010;298:C973978. [38] Finn PF, Dice JF. Proteolytic and lipolytic responses to starvation. Nutrition 2006;22:830844.

Conict of interest The authors do not have a relationship with the manufacturers of the drugs involved either in the past or present and did not receive funding from the manufacturers to carry out their research. The authors received support from DFG, Wilhelm Laupitz Foundation, EASL, and IFORES.

Financial support AC is supported by the Deutsche Forschungsgemeinschaft (DFG, grant 267/6-1 and 267/8-1), and the Wilhelm Laupitz Foundation. LPB is supported by an EASL Sheila Sherlock short-term fellowship and the IFORES Program of the University of Duisburg-Essen. Acknowledgements The authors would like to thank Robert Gieseler von der Crone, Jan-Peter Sowa, Judith Ertle and Thomas Schreiter for instructive discussions and careful review of this manuscript. References
[1] Feldstein AE. Novel insights into the pathophysiology of nonalcoholic fatty liver disease. Semin Liver Dis 2010;30:391401. [2] Browning JD, Szczepaniak LS, Dobbins R, Nuremberg P, Horton JD, Cohen JC, et al. Prevalence of hepatic steatosis in an urban population in the United States: impact of ethnicity. Hepatology (Baltimore, Md) 2004;40: 13871395. [3] Donnelly KL, Smith CI, Schwarzenberg SJ, Jessurun J, Boldt MD, Parks EJ. Sources of fatty acids stored in liver and secreted via lipoproteins in patients with nonalcoholic fatty liver disease. J Clin Invest 2005;115:13431351. [4] Postic C, Girard J. Contribution of de novo fatty acid synthesis to hepatic steatosis and insulin resistance. lessons from genetically engineered mice. J Clin Invest 2008;118:829838. [5] Samuel VT, Petersen KF, Shulman GI. Lipid-induced insulin resistance. unravelling the mechanism. Lancet 2010;375:22672277. [6] Sanyal AJ, Campbell-Sargent C, Mirshahi F, Rizzo WB, Contos MJ, Sterling RK, et al. Nonalcoholic steatohepatitis: association of insulin resistance and mitochondrial abnormalities. Gastroenterology 2001;120:11831192. [7] Neuschwander-Tetri BA. Hepatic lipotoxicity and the pathogenesis of nonalcoholic steatohepatitis: the central role of nontriglyceride fatty acid metabolites. Hepatology (Baltimore, Md) 2010;52:774788. [8] Fu S, Yang L, Li P, Hofmann O, Dicker L, Hide W, et al. Aberrant lipid metabolism disrupts calcium homeostasis causing liver endoplasmic reticulum stress in obesity. Nature 2011. [9] Rautou PE, Mansouri A, Lebrec D, Durand F, Valla D, Moreau R. Autophagy in liver diseases. J. Hepatol. 2010;53:11231134. [10] Singh R, Kaushik S, Wang Y, Xiang Y, Novak I, Komatsu M, et al. Autophagy regulates lipid metabolism. Nature 2009;458:11311135. [11] Wang Y, Singh R, Xiang Y, Czaja MJ. Macroautophagy and chaperonemediated autophagy are required for hepatocyte resistance to oxidant stress. Hepatology (Baltimore, Md) 2010;52:266277. [12] Yang L, Li P, Fu S, Calay ES, Hotamisligil GS. Defective hepatic autophagy in obesity promotes ER stress and causes insulin resistance. Cell Metab 2010;11:467478.

962

Journal of Hepatology 2012 vol. 56 j 952964

JOURNAL OF HEPATOLOGY
[39] Shibata M, Yoshimura K, Furuya N, Koike M, Ueno T, Komatsu M, et al. The MAP1-LC3 conjugation system is involved in lipid droplet formation. Biochem Biophys Res Commun 2009;382:419423. [40] Li S, Brown MS, Goldstein JL. Bifurcation of insulin signaling pathway in rat liver: mTORC1 required for stimulation of lipogenesis, but not inhibition of gluconeogenesis. Proc Natl Acad Sci U S A 2010;107:34413446. [41] Lehninger AL, Nelson DL, Cox MM. Lehninger principles of biochemistry. 4th ed. New York: W.H. Freeman; 2005. [42] McGarry JD, Brown NF. The mitochondrial carnitine palmitoyltransferase system. From concept to molecular analysis. Eur J Biochem 1997;244:114. [43] Akkaoui M, Cohen I, Esnous C, Lenoir V, Sournac M, Girard J, et al. Modulation of the hepatic malonyl-CoA-carnitine palmitoyltransferase 1A partnership creates a metabolic switch allowing oxidation of de novo fatty acids. Biochem J 2009;420:429438. [44] Reddy JK. Nonalcoholic steatosis and steatohepatitis. III. Peroxisomal betaoxidation, PPAR alpha, and steatohepatitis. Am J Physiol Gastrointest Liver Physiol 2001;281:G13331339. [45] Nguyen P, Leray V, Diez M, Serisier S, Le Bloch J, Siliart B, et al. Liver lipid metabolism. J Anim Physiol Anim Nutr (Berl) 2008;92:272283. [46] Hegardt FG. Mitochondrial 3-hydroxy-3-methylglutaryl-CoA synthase: a control enzyme in ketogenesis. Biochem J 1999;338:569582. [47] Leturque A, Brot-Laroche E, Le Gall M. GLUT2 mutations, translocation, and receptor function in diet sugar managing. Am J Physiol Endocrinol Metab 2009;296:E985992. [48] Agius L. Glucokinase and molecular aspects of liver glycogen metabolism. Biochem J 2008;414:118. [49] Chu CA, Fujimoto Y, Igawa K, Grimsby J, Grippo JF, Magnuson MA, et al. Rapid translocation of hepatic glucokinase in response to intraduodenal glucose infusion and changes in plasma glucose and insulin in conscious rats. Am J Physiol Gastrointest Liver Physiol 2004;286:G627634. [50] Miller SP, Anand GR, Karschnia EJ, Bell GI, LaPorte DC, Lange AJ. Characterization of glucokinase mutations associated with maturity-onset diabetes of the young type 2 (MODY-2): different glucokinase defects lead to a common phenotype. Diabetes 1999;48:16451651. [51] Cuesta-Munoz AL, Tuomi T, Cobo-Vuilleumier N, Koskela H, Odili S, Stride A, et al. Clinical heterogeneity in monogenic diabetes caused by mutations in the glucokinase gene (GCK-MODY). Diabetes Care 2010; 33:290292. [52] Amann T, Hellerbrand C. GLUT1 as a therapeutic target in hepatocellular carcinoma. Expert Opin Ther Targets 2009;13:14111427. [53] Roach PJ. Glycogen and its metabolism. Curr Mol Med 2002;2:101120. [54] Kim YM, Seo YH, Park CB, Yoon SH, Yoon G. Roles of GSK3 in metabolic shift toward abnormal anabolism in cell senescence. Ann N Y Acad Sci 2010;1201:6571. [55] Raddatz D, Ramadori G. Carbohydrate metabolism and the liver: actual aspects from physiology and disease. Z Gastroenterol 2007;45:5162. [56] Conti R, Mannucci E, Pessotto P, Tassoni E, Carminati P, Giannessi F, et al. Selective reversible inhibition of liver carnitine palmitoyl-transferase 1 by teglicar reduces gluconeogenesis and improves glucose homeostasis. Diabetes 2011;60:644651. [57] Valera A, Pujol A, Pelegrin M, Bosch F. Transgenic mice overexpressing phosphoenolpyruvate carboxykinase develop non-insulin-dependent diabetes mellitus. Proc Natl Acad Sci U S A 1994;91:91519154. [58] Stayrook KR, Bramlett KS, Savkur RS, Ficorilli J, Cook T, Christe ME, et al. Regulation of carbohydrate metabolism by the farnesoid X receptor. Endocrinology 2005;146:984991. [59] Oberkoer H, Pfeifenberger A, Soyal S, Felder T, Hahne P, Miller K, et al. Aberrant hepatic TRIB3 gene expression in insulin-resistant obese humans. Diabetologia 2010;53:19711975. [60] Cao R, Cronk ZX, Zha W, Sun L, Wang X, Fang Y, et al. Bile acids regulate hepatic gluconeogenic genes and farnesoid X receptor via G(alpha)iprotein-coupled receptors and the AKT pathway. J Lipid Res 2010;51:22342244. [61] Puigserver P, Rhee J, Donovan J, Walkey CJ, Yoon JC, Oriente F, et al. Insulinregulated hepatic gluconeogenesis through FOXO1-PGC-1alpha interaction. Nature 2003;423:550555. [62] Hakimi P, Johnson MT, Yang J, Lepage DF, Conlon RA, Kalhan SC, et al. Phosphoenolpyruvate carboxykinase and the critical role of cataplerosis in the control of hepatic metabolism. Nutr Metab (Lond) 2005;2:33. [63] Hanson RW, Hakimi P. Born to run; the story of the PEPCK-Cmus mouse. Biochimie 2008;90:838842. [64] Feldstein AE, Canbay A, Guicciardi ME, Higuchi H, Bronk SF, Gores GJ. Diet associated hepatic steatosis sensitizes to Fas mediated liver injury in mice. J. Hepatol. 2003;39:978983. [65] Bechmann LP, Gieseler RK, Sowa JP, Kahraman A, Erhard J, Wedemeyer I, et al. Apoptosis is associated with CD36/fatty acid translocase upregulation in non-alcoholic steatohepatitis. Liver Int 2010;30:850859. [66] Chandrasekaran K, Swaminathan K, Chatterjee S, Dey A. Apoptosis in HepG2 cells exposed to high glucose. Toxicol In Vitro 2010;24:387396. [67] Civera M, Urios A, Garcia-Torres ML, Ortega J, Martinez-Valls J, Cassinello N, et al. Relationship between insulin resistance, inammation and liver cell apoptosis in patients with severe obesity. Diabetes Metab Res Rev 2010;26:187192. [68] Sato N. Central role of mitochondria in metabolic regulation of liver pathophysiology. J Gastroenterol Hepatol 2007;22:S16. [69] Puche JE, Garcia-Fernandez M, Muntane J, Rioja J, Gonzalez-Baron S, Castilla Cortazar I. Low doses of insulin-like growth factor-I induce mitochondrial protection in aging rats. Endocrinology 2008;149:26202627. [70] Wang X, Wang Z, Liu JZ, Hu JX, Chen HL, Li WL, et al. Double antioxidant activities of rosiglitazone against high glucose-induced oxidative stress in hepatocyte. Toxicol In Vitro 2011. [71] Okiyama W, Tanaka N, Nakajima T, Tanaka E, Kiyosawa K, Gonzalez FJ, et al. Polyenephosphatidylcholine prevents alcoholic liver disease in PPARalphanull mice through attenuation of increases in oxidative stress. J Hepatol 2009;50:12361246. [72] Lim JH, Lee HJ, Ho Jung M, Song J. Coupling mitochondrial dysfunction to endoplasmic reticulum stress response: a molecular mechanism leading to hepatic insulin resistance. Cell Signal 2009;21:169177. [73] Hotamisligil GS. Endoplasmic reticulum stress and the inammatory basis of metabolic disease. Cell 2010;140:900917. [74] Jiang HY, Wek SA, McGrath BC, Scheuner D, Kaufman RJ, Cavener DR, et al. Phosphorylation of the alpha subunit of eukaryotic initiation factor 2 is required for activation of NF-kappaB in response to diverse cellular stresses. Mol Cell Biol 2003;23:56515663. [75] Seo J, Fortuno 3rd ES, Suh JM, Stenesen D, Tang W, Parks EJ, et al. Atf4 regulates obesity, glucose homeostasis, and energy expenditure. Diabetes 2009;58:25652573. [76] Kammoun HL, Chabanon H, Hainault I, Luquet S, Magnan C, Koike T, et al. GRP78 expression inhibits insulin and ER stress-induced SREBP-1c activation and reduces hepatic steatosis in mice. J Clin Invest 2009;119: 12011215. [77] Rutkowski DT, Wu J, Back SH, Callaghan MU, Ferris SP, Iqbal J, et al. UPR pathways combine to prevent hepatic steatosis caused by ER stressmediated suppression of transcriptional master regulators. Dev Cell 2008;15:829840. [78] Gregor MG, Hotamisligil GS. Adipocyte stress: the endoplasmic reticulum and metabolic disease. J Lipid Res 2007. [79] Porstmann T, Santos CR, Grifths B, Cully M, Wu M, Leevers S, et al. SREBP activity is regulated by mTORC1 and contributes to Akt-dependent cell growth. Cell Metab 2008;8:224236. [80] Jung CH, Ro SH, Cao J, Otto NM, Kim DH. mTOR regulation of autophagy. FEBS Lett 2010;584:12871295. [81] Cheung O, Sanyal AJ. Abnormalities of lipid metabolism in nonalcoholic fatty liver disease. Semin Liver Dis 2008;28:351359. [82] Hotamisligil GS. Inammation and metabolic disorders. Nature 2006;444: 860867. [83] Bieghs V, Wouters K, van Gorp PJ, Gijbels MJ, de Winther MP, Binder CJ, et al. Role of scavenger receptor A and CD36 in diet-induced nonalcoholic steatohepatitis in hyperlipidemic mice. Gastroenterology 2010;138: 24772486 (2486 e2471-2473). [84] Berk PD, Zhou S, Bradbury MW. Increased hepatocellular uptake of long chain fatty acids occurs by different mechanisms in fatty livers due to obesity or excess ethanol use, contributing to development of steatohepatitis in both settings. Trans Am Clin Climatol Assoc 2005;116:335344 (discussion 345). [85] Sajan MP, Standaert ML, Rivas J, Miura A, Kanoh Y, Soto J, et al. Role of atypical protein kinase C in activation of sterol regulatory element binding protein-1c and nuclear factor kappa B (NFkappaB) in liver of rodents used as a model of diabetes, and relationships to hyperlipidaemia and insulin resistance. Diabetologia 2009;52:11971207. [86] Taniguchi CM, Kondo T, Sajan M, Luo J, Bronson R, Asano T, et al. Divergent regulation of hepatic glucose and lipid metabolism by phosphoinositide 3kinase via Akt and PKClambda/zeta. Cell Metab 2006;3:343353. [87] Liu HY, Han J, Cao SY, Hong T, Zhuo D, Shi J, et al. Hepatic autophagy is suppressed in the presence of insulin resistance and hyperinsulinemia: inhibition of FoxO1-dependent expression of key autophagy genes by insulin. J Biol Chem 2009;284:3148431492. [88] Ricchi M, Odoardi MR, Carulli L, Anzivino C, Ballestri S, Pinetti A, et al. Differential effect of oleic and palmitic acid on lipid accumulation and

Journal of Hepatology 2012 vol. 56 j 952964

963

Review
[89] apoptosis in cultured hepatocytes. J Gastroenterol Hepatol 2009;24: 830840. Yamaguchi K, Yang L, McCall S, Huang J, Yu XX, Pandey SK, et al. Inhibiting triglyceride synthesis improves hepatic steatosis but exacerbates liver damage and brosis in obese mice with nonalcoholic steatohepatitis. Hepatology (Baltimore, Md) 2007;45:13661374. Han MS, Park SY, Shinzawa K, Kim S, Chung KW, Lee JH, et al. Lysophosphatidylcholine as a death effector in the lipoapoptosis of hepatocytes. J Lipid Res 2008;49:8497. Fessler MB, Rudel LL, Brown JM. Toll-like receptor signaling links dietary fatty acids to the metabolic syndrome. Curr Opin Lipidol 2009;20:379385. Feldstein AE, Werneburg NW, Canbay A, Guicciardi ME, Bronk SF, Rydzewski R, et al. Free fatty acids promote hepatic lipotoxicity by stimulating TNF-alpha expression via a lysosomal pathway. Hepatology (Baltimore, Md) 2004;40:185194. Nolan CJ, Larter CZ. Lipotoxicity: why do saturated fatty acids cause and monounsaturates protect against it? J Gastroenterol Hepatol 2009;24: 703706. Erbay E, Babaev VR, Mayers JR, Makowski L, Charles KN, Snitow ME, et al. Reducing endoplasmic reticulum stress through a macrophage lipid chaperone alleviates atherosclerosis. Nat Med 2009;15:13831391. Robertson G, Leclercq I, Farrell GC. Nonalcoholic steatosis and steatohepatitis. II. Cytochrome P-450 enzymes and oxidative stress. Am J Physiol Gastrointest Liver Physiol 2001;281:G11351139. Kohjima M, Enjoji M, Higuchi N, Kato M, Kotoh K, Yoshimoto T, et al. Reevaluation of fatty acid metabolism-related gene expression in nonalcoholic fatty liver disease. Int J Mol Med 2007;20:351358. Dentin R, Benhamed F, Hainault I, Fauveau V, Foufelle F, Dyck JR, et al. Liver-specic inhibition of ChREBP improves hepatic steatosis and insulin resistance in ob/ob mice. Diabetes 2006;55:21592170. Tamura S, Shimomura I. Contribution of adipose tissue and de novo lipogenesis to nonalcoholic fatty liver disease. J Clin Invest 2005;115: 11391142. Orellana-Gavalda JM, Herrero L, Malandrino MI, Paneda A, Sol RodriguezPena M, Petry H, et al. Molecular therapy for obesity and diabetes based on a long-term increase in hepatic fatty-acid oxidation. Hepatology (Baltimore, Md) 2011;53:821832. Rosselli MS, Burgueno AL, Pirola CJ, Sookoian S. Cyclooxygenase inhibition Up-regulates liver carnitine palmitoyltransferase 1A expression and improves fatty liver. Hepatology (Baltimore, Md) 2011;53:21432144. Crespo J, Cayon A, Fernandez-Gil P, Hernandez-Guerra M, Mayorga M, Dominguez-Diez A, et al. Gene expression of tumor necrosis factor alpha and TNF-receptors, p55 and p75, in nonalcoholic steatohepatitis patients. Hepatology (Baltimore, Md) 2001;34:11581163. Endo M, Masaki T, Seike M, Yoshimatsu H. TNF-alpha induces hepatic steatosis in mice by enhancing gene expression of sterol regulatory element binding protein-1c (SREBP-1c). Exp Biol Med (Maywood) 2007;232:614621. Kaser S, Moschen A, Cayon A, Kaser A, Crespo J, Pons-Romero F, et al. Adiponectin and its receptors in non-alcoholic steatohepatitis. Gut 2005;54:117121. Feldstein AE, Canbay A, Angulo P, Taniai M, Burgart LJ, Lindor KD, et al. Hepatocyte apoptosis and fas expression are prominent features of human nonalcoholic steatohepatitis. Gastroenterology 2003;125:437443. Feldstein AE, Wieckowska A, Lopez AR, Liu YC, Zein NN, McCullough AJ. Cytokeratin-18 fragment levels as noninvasive biomarkers for nonalcoholic steatohepatitis: a multicenter validation study. Hepatology (Baltimore, Md) 2009;50:10721078. Horoz M, Bolukbas C, Bolukbas FF, Sabuncu T, Aslan M, Sarifakiogullari S, et al. Measurement of the total antioxidant response using a novel automated method in subjects with nonalcoholic steatohepatitis. BMC gastroenterology 2005;5:35. Huang W, Ma K, Zhang J, Qatanani M, Cuvillier J, Liu J, et al. Nuclear receptor-dependent bile acid signaling is required for normal liver regeneration. Science 2006;312:233236. Fernandez MA, Albor C, Ingelmo-Torres M, Nixon SJ, Ferguson C, Kurzchalia T, et al. Caveolin-1 is essential for liver regeneration. Science 2006;313:16281632. Doege H, Baillie RA, Ortegon AM, Tsang B, Wu Q, Punreddy S, et al. Targeted deletion of FATP5 reveals multiple functions in liver metabolism: alterations in hepatic lipid homeostasis. Gastroenterology 2006;130:12451258. [110] Hubbard B, Doege H, Punreddy S, Wu H, Huang X, Kaushik VK, et al. Mice deleted for fatty acid transport protein 5 have defective bile acid conjugation and are protected from obesity. Gastroenterology 2006;130: 12591269. [111] Geier A, Trautwein C. Bile acids are homeotrophic sensors of the functional hepatic capacity and regulate adaptive growth during liver regeneration. Hepatology 2007;45:251253. [112] Bechmann LP, Jochum C, Kocabayoglu P, Sowa JP, Kassalik M, Gieseler RK, et al. Cytokeratin 18-based modication of the MELD score improves prediction of spontaneous survival after acute liver injury. J. Hepatol. 2010;53:639647. [113] Shu RZ, Zhang F, Wang F, Feng DC, Li XH, Ren WH, et al. Adiponectin deciency impairs liver regeneration through attenuating STAT3 phosphorylation in mice. Lab Invest 2009;89:10431052. [114] Wree A, Kahraman A, Gerken G, Canbay A. Obesity affects the liver - the link between adipocytes and hepatocytes. Digestion 2011;83:124133. [115] Hirsch HA, Iliopoulos D, Joshi A, Zhang Y, Jaeger SA, Bulyk M, et al. A transcriptional signature and common gene networks link cancer with lipid metabolism and diverse human diseases. Cancer Cell 2010;17: 348361. [116] Petta S, Craxi A. Hepatocellular carcinoma and non-alcoholic fatty liver disease: from a clinical to a molecular association. Curr Pharm Des 2010;16:741752. [117] Bugianesi E. Non-alcoholic steatohepatitis and cancer. Clin Liver Dis 2007;11:191207 (x-xi). [118] Calle EE, Rodriguez C, Walker-Thurmond K, Thun MJ. Overweight, obesity and mortality from cancer in a prospectively studied cohort of U.S. adults. N Engl J Med 2003;348:16251638. [119] Davila JA, Morgan RO, Shaib Y, McGlynn KA, El-Serag HB. Diabetes increases the risk of hepatocellular carcinoma in the United States: a population based case control study. Gut 2005;54:533539. [120] Adams LA, Lymp JF, St Sauver J, Sanderson SO, Lindor KD, Feldstein A, et al. The natural history of nonalcoholic fatty liver disease: a population-based cohort study. Gastroenterology 2005;129:113121. [121] Ertle J, Dechene A, Sowa JP, Penndorf V, Herzer K, Kaiser G, et al. Nonalcoholic fatty liver disease progresses to HCC in the absence of apparent cirrhosis. Int J Cancer 2010. [122] Venook AP, Papandreou C, Furuse J, de Guevara LL. The incidence and epidemiology of hepatocellular carcinoma: a global and regional perspective. Oncologist 2010;15:513. [123] Wu JM, Skill NJ, Maluccio MA. Evidence of aberrant lipid metabolism in hepatitis C and hepatocellular carcinoma. HPB (Oxford) 2010;12: 625636. [124] Villanueva A, Chiang DY, Newell P, Peix J, Thung S, Alsinet C, et al. Pivotal role of mTOR signaling in hepatocellular carcinoma. Gastroenterology 2008;135:19721983 (1983 e1971-1911). [125] Qu X, Yu J, Bhagat G, Furuya N, Hibshoosh H, Troxel A, et al. Promotion of tumorigenesis by heterozygous disruption of the beclin 1 autophagy gene. J Clin Invest 2003;112:18091820. [126] Park EJ, Lee JH, Yu GY, He G, Ali SR, Holzer RG, et al. Dietary and genetic obesity promote liver inammation and tumorigenesis by enhancing IL-6 and TNF expression. Cell 2010;140:197208. [127] Handy JA, Saxena NK, Fu P, Lin S, Mells JE, Gupta NA, et al. Adiponectin activation of AMPK disrupts leptin-mediated hepatic brosis via suppressors of cytokine signaling (SOCS-3). J Cell Biochem 2010;110:11951207. [128] Tanaka N, Moriya K, Kiyosawa K, Koike K, Gonzalez FJ, Aoyama T. PPARalpha activation is essential for HCV core protein-induced hepatic steatosis and hepatocellular carcinoma in mice. J Clin Invest 2008;118:683694. [129] van der Meer DL, Degenhardt T, Vaisanen S, de Groot PJ, Heinaniemi M, de Vries SC, et al. Proling of promoter occupancy by PPARalpha in human hepatoma cells via ChIP-chip analysis. Nucleic Acids Res 2010;38: 28392850. [130] Ratziu V, Giral P, Jacqueminet S, Charlotte F, Hartemann-Heurtier A, Serfaty L, et al. Rosiglitazone for nonalcoholic steatohepatitis: one-year results of the randomized placebo-controlled Fatty Liver Improvement with Rosiglitazone Therapy (FLIRT) Trial. Gastroenterology 2008;135: 100110. [131] Yu J, Qiao L, Zimmermann L, Ebert MP, Zhang H, Lin W, et al. Troglitazone inhibits tumor growth in hepatocellular carcinoma in vitro and in vivo. Hepatology (Baltimore, Md) 2006;43:134143.

[90]

[91] [92]

[93]

[94]

[95]

[96]

[97]

[98]

[99]

[100]

[101]

[102]

[103]

[104]

[105]

[106]

[107]

[108]

[109]

964

Journal of Hepatology 2012 vol. 56 j 952964

Vous aimerez peut-être aussi