3 Triglyceride Metabolism in The Liver

Download as pdf or txt
Download as pdf or txt
You are on page 1of 44

HHS Public Access

Author manuscript
Compr Physiol. Author manuscript; available in PMC 2019 February 15.
Author Manuscript

Published in final edited form as:


Compr Physiol. ; 8(1): 1–8. doi:10.1002/cphy.c170012.

Triglyceride metabolism in the liver


Michele Alves-Bezerra and David E. Cohen
Joan & Sanford I. Weill Department of Medicine, Weill Cornell Medical College, New York, NY,
USA

Abstract
Triglyceride molecules represent the major form of storage and transport of fatty acids within cells
Author Manuscript

and in the plasma. The liver is the central organ for fatty acid metabolism. Fatty acids accrue in
liver by hepatocellular uptake from the plasma and by de novo biosynthesis. Fatty acids are
eliminated by oxidation within the cell or by secretion into the plasma within triglyceride-rich very
low density lipoproteins. Notwithstanding high fluxes through these pathways, under normal
circumstances the liver stores only small amounts of fatty acids as triglycerides. In the setting of
overnutrition and obesity, hepatic fatty acid metabolism is altered, commonly leading the
accumulation of triglycerides within hepatocytes, and to a clinical condition known as non-
alcoholic fatty liver disease (NAFLD). In this review, we describe the current understanding of
fatty acid and triglyceride metabolism in the liver and its regulation in health and disease,
identifying potential directions for future research. Advances in understanding the molecular
mechanisms underlying the hepatic fat accumulation are critical to the development of targeted
therapies for NAFLD.
Author Manuscript

Introduction
The liver is the central organ that controls lipid homeostasis by means of complex, but
precisely regulated biochemical, signaling and cellular pathways. Hepatocytes are the main
liver parenchymal cells, which control hepatic biochemical and metabolic functions in the
liver, including triglyceride metabolism. Additional cell types within the liver include
cholangiocytes, Kupffer cells, stellate cells, and endothelial cells, each of which has
specialized functions in hepatic pathobiology (261).

This review will focus on the hepatic metabolism of fatty acids (FA) and their neutral
storage form, triglycerides (TG), which occurs primarily in hepatocytes. Under normal
Author Manuscript

circumstances on a daily basis, the liver processes large quantities of FA, but stores only
small amounts in the form of TG, with steady state TG contents of less than 5% (32). This is
because the rates of acquisition of FA by uptake from the plasma and from de novo synthesis
within the liver are balanced by rates of FA oxidation and secretion into plasma as TG-

Cross-References
Energy metabolism in the liver
Metabolism of lipids in chylomicrons and very low density lipoproteins (legacy)
Obesity
Steatosis in the liver
Triglyceride metabolism (legacy)
Alves-Bezerra and Cohen Page 2

enriched very low-density lipoprotein (VLDL-TG). The relatively small quantities of TG


Author Manuscript

stored within the liver are localized in cytoplasmic lipid droplets.

NAFLD is the most common chronic liver disease and is characterized by excess TG
accumulation within the liver. It is associated with obesity, type 2 diabetes and dyslipidemia,
and commonly occurs in the setting of insulin resistance (22). NAFLD encompasses a
spectrum of liver histopathologies from simple hepatic steatosis, often referred to as non-
alcoholic fatty liver (NAFL), to non-alcoholic steatohepatitis (NASH) which hepatic
steatosis is accompanied by inflammation, cell death and fibrosis (5). Hepatic complications
of NASH include cirrhosis and hepatocellular carcinoma. Currently, there are no effective
therapies for NAFLD apart from changes in lifestyle that lead to improved physical fitness
and weight loss. In order to develop effective pharmacologic therapies, detailed knowledge
of the cellular and molecular mechanisms leading to altered hepatic lipid metabolism is
essential. In addition to describing the major pathways of hepatic FA and TG metabolism
Author Manuscript

and their regulation, we will discuss recent advances in our understanding of how they are
altered in the setting of NAFLD.

Hepatic fatty acid metabolism


FA within the liver originate from either dietary or endogenous sources. Dietary TG are
emulsified by bile acids within the intestinal lumen following their hydrolysis primarily by
pancreatic lipase (166), which yields sn-2-monoacylglycerols and free FA as products.
Following emulsification, these lipid molecules are taken by enterocytes and resynthesized
into TG. TG are packaged into chylomicrons, secreted into the lymphatic system and
ultimately reach the plasma (112, 122). Much of the chylomicron TG are taken up by muscle
and adipose tissue due to the activity of lipoprotein lipase, which is expressed on the luminal
surfaces of capillary endothelial cells of these tissues (Fig. 1). TG remaining within the
Author Manuscript

chylomicron remnants are delivered to the liver when these particles are taken up by receptor
mediated endocytosis, and FA are released during lysosomal processing of the particles (54).

When carbohydrates are abundant, the liver converts glucose into FA, a process referred to
as de novo lipogenesis (DNL) (120) (Fig.1). The control of DNL is primarily transcriptional.
Plasma insulin activates the endoplasmic reticulum membrane-bound transcription factor
sterol regulatory element binding protein 1C (SREBP1c), the N-terminus of which
translocates to the nucleus and upregulates all genes in the FA biosynthetic pathway (116).
The hepatic uptake of excess plasma glucose promotes the nuclear translocation of
carbohydrate response element binding protein (ChREBP), a transcription factor that also
upregulates transcription of the majority of FA biosynthetic genes plus pyruvate kinase (89,
205), which increases the availability of citrate for FA synthesis.
Author Manuscript

Another important source of FA is direct uptake from the plasma. In human subjects, FA
from plasma constitute the main source of hepatic triglycerides during fasting (72). Under
fasting conditions when plasma insulin concentrations are low, a lipolytic program is
initiated in white adipose tissue, which increases the plasma FA pool that is available for
uptake by the liver (12, 214) (Fig. 1). FA are largely albumin-bound within the circulation.
Several steps are involved in hepatic FA uptake. These include dissociation of FAs from
albumin, transport across the hepatocyte plasma membrane, binding to intracellular proteins,

Compr Physiol. Author manuscript; available in PMC 2019 February 15.


Alves-Bezerra and Cohen Page 3

and esterification to coenzyme A (CoA). Plasma FA are also the principal source of VLDL-
Author Manuscript

TG in both fasting and fed states (72).

Within the hepatocyte, FA are esterified to glycerol-3-phosphate (G3P) and to cholesterol in


order to generate TG or cholesteryl esters, respectively. These neutral lipids can be either
stored in cytoplasmic lipid droplets (LDs), or secreted into bloodstream as VLDL particles
(135) (Fig. 1). FA within the liver may also be used for the synthesis of other complex lipids,
including phospholipids (PL). During fasting, FA are used both as local energy supply as
well as substrate for ketone bodies production. Overall, hepatic FA metabolism is tightly
regulated by multiple interrelated transcriptional and signaling pathways that remain the
subject of intensive investigation and discovery.

Fatty acid uptake


Notwithstanding their ability to diffuse across a lipid bilayer (232), FA are taken up by
Author Manuscript

hepatocytes via plasma membrane-associated proteins. A variety of proteins have been


associated with long-chain FA transport, including plasma membrane FA-binding protein
(FABPpm), FA translocase (FAT)/CD36, caveolin-1, and very long-chain acyl-CoA
synthetases (ACSVL/FA transport proteins, also named FATP/solute carrier family 27A1–6,
SLC27A1–6) (27, 253) (Fig. 2).

The roles of CD36 and FABPpm proteins in FA uptake are established for heart and skeletal
muscle, but their function in hepatocytes remains unresolved. Although expression levels are
low in the liver, CD36 mRNA levels are positively correlated with hepatic TG contents in a
rat model of hepatic steatosis (35) and its protein expression is increased in patients with
NAFLD (181). However, the pathophysiological relevance of this protein to the liver is
unclear. Whereas CD36 liver-specific deletion in mice leads to reduced hepatic lipid
Author Manuscript

contents and decreased FA uptake in the setting of high fat feeding, suggesting the direct
contribution of CD36 to hepatic FA metabolism under these conditions (279), whole body
knockout of CD36 in mice significantly impairs FA uptake by heart, adipose tissue, and
muscle, but normal uptake is observed in the liver (52). There also seems to be a role for
CD36 in hepatic lipogenesis. CD36 is a target gene for the aryl hydrocarbon receptor (AhR),
and is required for the TG accumulation observed in the liver in response to its activation
(154). In addition, the pregnane X receptor (PXR) induces hepatic CD36 mRNA expression,
along with increased FA uptake and TG accumulation in the liver (294). Finally, the CD36 is
a transcriptional target of peroxisome proliferator activated receptor (PPAR) γ (259), which
is upregulated by PXR (294).

Functional evidence for protein-mediated fatty acid transport in the rat hepatocyte plasma
Author Manuscript

membrane led to the identification of FABPpm as a protein that mediates this activity (247,
249). FABPpm was subsequently identified as mitochondrial aspartate aminotransferase (20,
28). The mRNA encoding this protein is increased in hepatocytes of genetically obese
mouse models (180), and the binding of oleate to rat hepatocyte plasma membranes is
reduced after treatment with specific antiserum (246). In cultured hepatocytes, ethanol
treatment increases mRNA and plasma membrane protein contents, and these changes
correlate with increased FA uptake (295). However, the relative contribution of FABPpm to
FA uptake by the liver remains unknown.

Compr Physiol. Author manuscript; available in PMC 2019 February 15.


Alves-Bezerra and Cohen Page 4

FATP proteins (1–6) are FA-transporting proteins that are situated in the plasma membrane,
Author Manuscript

as well as intracellular membranes. Because of their capacity to also activate long- or very-
long-chain FA (69), these proteins have been re-classified in the very-long-chain acyl-CoA
synthetase (ACSVL) family as ACSVL1–6 (106). FATPs couple FA uptake with intracellular
FA esterification into acyl-CoAs, a metabolic trapping mechanism that has been referred to
as vectorial acylation (77, 258). Although well characterized, neither FATP1 nor FATP4 are
normally expressed at substantial levels in the liver (239). Expression of FATP4, an ER
membrane-associated isoform (226, 273), is increased in the setting of hepatic steatosis and
may play a role in palmitate-mediated lipoapoptosis (226). FATP2 is highly expressed in
liver and kidney (85). In HepG2 cells, FATP2 localizes to the ER, but its overexpression
increases uptake of long-chain FA in cell culture (148). FATP5 is expressed in the liver, and
studies in knockout mice and cell culture systems support a role in FA uptake (70).

Caveolins (types 1–3) are integral membrane proteins that are essential components of
Author Manuscript

caveolae, vesicular membrane microdomains that are rich in cholesterol and sphingolipids
(6). Caveolins were initially characterized with respect to cholesterol transport (204).
However, their roles in FA transport have been appreciated in experiments using HepG2
hepatoma cells (203), wherein Caveolin-1, CD36, FABPpm, and calcium-independent
membrane phospholipase A2 (iPLA2) form a heterotetrameric protein complex within
plasma membrane microdomains (248). These complexes promote FA accumulation within
caveolae-derived vesicular structures for transport into the cell (247). In support of this
mechanism, mice deficient in caveolin-1 show reduced hepatic TG content and impaired LD
accumulation (87).

Although multiple factors lead to hepatic steatosis, circulating FA are the main source of
hepatic lipids in NAFLD (72). The overflow of FA derived from excessive lipolysis in the
Author Manuscript

adipose tissue contributes to the pathogenesis of insulin resistance in mice and humans
(159), leading to both type 2 diabetes (213) and NAFLD (27, 171). Although FA transporters
may exert a central role in the control of FA flux into the liver, additional studies are needed
to clarify specific functions and physiological contributions of each FA transporter, along
with their differential regulation in the context of NAFLD.

De novo lipogenesis
Under normal circumstances in humans, pathways of hepatic DNL appear to be utilized
sparingly (114). However, in addition to the enhanced FA uptake, increased DNL can
contribute substantially to hepatic steatosis (72). The first step in DNL pathway is catalyzed
by ATP-citrate lyase (ACLY), which converts citrate to acetyl-CoA, which is then
carboxylated to malonyl-CoA by acetyl-CoA carboxylase (ACC). ACC is followed by a
Author Manuscript

series of reactions that convert malonyl-CoA into palmitate. Fatty acid synthase (FAS) is the
key rate-limiting enzyme in palmitate synthesis, and its activity is regulated by multiple
mechanisms (125, 143, 218, 228) (Fig. 2). Palmitate can be modified by elongases and
desaturases to generate a variety of FA species.

There are two ACC isoforms, cytosolic ACC1 and mitochondrial ACC2, which are encoded
by separate genes. The presence of distinct ACC enzymes enables the formation of two
distinct pools of malonyl-CoA. ACC1-generated malonyl-CoA is utilized as a substrate by

Compr Physiol. Author manuscript; available in PMC 2019 February 15.


Alves-Bezerra and Cohen Page 5

FAS, whereas the malonyl-CoA produced by ACC2 plays a key role in the negative
regulation of β-oxidation by inhibiting carnitine palmitoyltransferase 1 (CPT1) (266). ACC1
Author Manuscript

is highly expressed in lipogenic organs, including the liver, wherein the gene expression and
enzyme activity are induced by carbohydrate-rich, low-fat diet, and are suppressed in the
setting of starvation or diabetes (139, 149). Whole body disruption of the ACC1 gene results
in embryonic lethality, but liver-specific knockout mice are viable and exhibit 70–75%
reductions in malonyl-CoA levels and 40–70% reductions in hepatic TG contents (170).
Despite similar malonyl-CoA levels as wild type animals, livers of ACC2 knockout mice are
characterized by reduced rates of β-oxidation and TG contents. These mice resist to diet-
induced obesity and diabetes (2, 3). In keeping with these observations, antisense-mediated
knockdown of both ACC1 and ACC2 led to improved hepatic insulin responsiveness in rats
with experimental NAFLD (217).

FAS expression is ubiquitous, but evidence of elevated transcriptional and enzymatic


Author Manuscript

activities are observed in the liver of murine models of obesity, such as Zucker (fa/fa) rats
(227). The transcription of both ACC and FAS genes is enhanced in livers of NAFLD
subjects (146). Similar to ACC knockout mice, whole body disruption of FAS leads to
embryonic lethality (51). Surprisingly, liver-specific knockout of FAS results in
hypoglycemia and hepatic steatosis in mice fed a zero-fat diet (45). Because this phenotype
is reversed by treatment with a PPARα agonist, it has been suggested that newly synthesized
FA and their derivatives, including phospholipids (44), constitute a distinct pool that
provides endogenous ligands for PPARα. This in turn stimulates hepatic gluconeogenesis
and FA oxidation (45).

In leptin-deficient obese (ob/ob) mice, increased transcription of Acc1 and Fas is linked to
the increased activation of ChREBP (64). Additionally, SREBP-1c and ChREBP can
Author Manuscript

synergistically induce the mRNA expression of Acc and Fas in the liver (65). Expression of
both transcriptional factors are upregulated in livers of mouse models of NAFLD (19),
however only SREBP-1c is upregulated in patients with NAFLD (115).

Among TG found in VLDL in normal subjects, less than 5% contain FA derived from DNL
(68). However, inhibition of DNL impairs production and secretion of VLDL-TG in rats
(95). The rate of incorporation of DNL-synthesized FA into VLDL-TG is increased in
normal subjects in the setting of high-carbohydrate diet (223, 224), alcohol consumption and
infectious states (88), and is positively correlated with increased plasma VLDL-TG content
(223). DNL is also increased in obese (67, 244) and NAFLD subjects (68, 88). These
observations suggest that hypertriglyceridemia is at least in part attributable to increased
availability of DNL-synthesized FA for TG synthesis and VLDL production (223). Ongoing
Author Manuscript

research is needed to clarify mechanisms of altered DNL in a variety of pathogenic


conditions, including NAFLD.

Fatty acid activation


Activation by thioesterification to CoA to form fatty acyl-CoA molecules is an obligatory
step in the metabolism of long-chain FA. This reaction is catalyzed by members of the long-
chain acyl-CoA synthetase (ACSL) family of enzymes (Fig. 2). The five ACSL isoforms
found in mammalians have specific tissue distributions, subcellular locations and substrate

Compr Physiol. Author manuscript; available in PMC 2019 February 15.


Alves-Bezerra and Cohen Page 6

preferences (106). Additionally, distinct ACSL isoforms play key roles in partitioning fatty
Author Manuscript

acyl-CoAs into different metabolic pathways (57, 106).

ACSL1, the best characterized isoform, is abundantly expressed in a variety of tissues.


Because Acsl1 is a target gene of PPARα in rat liver, this enzyme has been associated with
FA catabolism (220). ACSL1 interacts physically with CPT1 in the outer mitochondrial
membrane of rat hepatocytes, possibly functioning to channel fatty acyl-CoA products into
mitochondria for β-oxidation (155). However, this process is not fully understood because
liver-specific ACSL1 knockout mice exhibit only modestly decreased rates of β-oxidation
and reduced fatty acyl-chain incorporation into newly-synthesized TG, although the total
hepatic TG contents remain unchanged (161).

ACSL3 is expressed on lipid droplets in HuH7 hepatoma cells (94), and its expression is up-
regulated in livers of ob/ob mice and in mice fed a high-carbohydrate diet (34). Knockdown
Author Manuscript

of ACSL3 in rat hepatocytes reduces DNL through diminished activation of lipogenic


transcriptional factors (i.e. PPARγ, ChREBP, SREBP-1c and LXRα) possibly due to the
reduced availability of FA species, acyl-CoA or their lipid intermediates (34).

Suggestive of a role for ACSL5 in TG synthesis, mRNA levels are increased by insulin and
by SREBP-1c activation (4). In support of this lipogenic role, overexpression of ACSL5 in
rat hepatoma cells increases uptake of exogenous FA, as well as its partitioning into TG, but
not into PL (174). Knockdown of ACSL5 in rat primary hepatocytes strongly influences the
partitioning of fatty acyl-CoAs between lipogenic and lipolytic pathways. In the absence of
this enzyme, the incorporation of exogenous and endogenous FA into complex lipids is
reduced, the formation of LD is decreased and the oxidation of FA is stimulated (33).

The reaction catalyzed by ACSL isoforms is reversed by the activities of acyl-CoA


Author Manuscript

thioesterases (ACOTs) (Fig. 2). These enzymes catalyze the hydrolysis of acyl-CoA
molecules into FA and CoA. Despite their relatively well-investigated enzymatic activities,
the biological functions of the fifteen ACOT family members remain less well understood
(53, 142, 256). Acots 1–6, which constitute the type I enzymes, are upregulated in the livers
of fasted and high fat-fed mice under the transcriptional control of PPARα (79). These
observations suggest a role of type I ACOTs in lipid catabolism. Indeed, overexpression of
ACOT2 in the liver leads to increased fatty acid oxidation and ketogenesis, although the
mechanism underlying this effect remains unclear (182).

ACOTS 7–15 comprise the type II enzymes, and some of these contribute to hepatic FA
metabolism. ACOT11 (synonym, Thioesterase superfamily member 1, Them1) is highly
expressed in brown adipose tissue, where it functions to suppress energy expenditure (292)
Author Manuscript

by decreasing rates of fat oxidation (196). Acot11 gene expression is also strongly induced
in the liver of obese mice (79). Whereas deficiency of this enzyme results in decreased
hepatic steatosis, improved glucose homeostasis, and resistance to diet-induced diabetes in
high fat fed mice (292), it is unclear whether these phenotypes are indirect effects of
increased energy expenditure.

ACOT13 (a.k.a. Them2) is a mitochondria-associated enzyme that is highly expressed in the


liver and other oxidative tissues. Them2 knockout mice are resistant to diet-induced hepatic

Compr Physiol. Author manuscript; available in PMC 2019 February 15.


Alves-Bezerra and Cohen Page 7

steatosis and exhibit improved glucose metabolism (130). Moreover, Them2-deficiency


Author Manuscript

results in diminished expression of PPARα and HNF4α, possibly due to the role of this
enzyme in the control of FA ligands of these nuclear receptors (53) and the reduced FA
oxidation (135). Them2 also controls lipid and glucose metabolism by interacting with
phosphatidylcholine transfer protein (PC-TP) to suppress hepatic insulin signaling by
reducing the activation of both the insulin receptor substrate 2 (IRS2) and the mammalian
target of rapamycin (mTOR) (80).

ACOT15 (a.k.a. Them5) appears to play a key role in maintaining the integrity of
mitochondrial membranes. Mice lacking ACOT15 experience deficiencies in cardiolipin
remodeling in the inner mitochondrial membrane of hepatocytes. This phenotype leads to
mitochondrial dysfunction, including altered morphology, decreased mitochondrial FA
contents, reduced FA oxidation, and the development of liver steatosis (296).
Author Manuscript

It is plausible that ACSL and ACOT enzymes coordinate the control of the intracellular
balance of fatty acyl-CoAs and FA, intracellular and intra-organelle concentrations of CoA,
and the availability of lipid substrates for a variety of metabolic pathways. Both types of
enzymes exhibit specific transcriptional patterns of expression that vary depending on cell
type and physiological state, and the coordinated regulation of ACSL and ACOT enzymes
has been recently described (79). Although elevated hepatic concentrations of long-chain
acyl-CoAs have been associated with increased plasma insulin levels and hepatic insulin
resistance (48), decreased hepatic fatty acyl-CoA concentrations in the setting of liver-
specific deletion of ACSL1 do not prevent the development of insulin resistance in high-fat
fed mice (161). As illustrated by these apparently contradictory observations, mechanisms
whereby ACSL and ACOT enzymes control FA metabolism within hepatocytes are
incompletely understood, and it remains uncertain whether their altered regulation
Author Manuscript

contributes to insulin resistance and related metabolic abnormalities.

Intracellular fatty acid transport


Long-chain FAs and their acyl-CoA derivatives act not only as substrates for lipid synthesis
and oxidation, but are also involved in a range of diverse intracellular processes, such as
protein palmitoylation, intracellular signaling, and activation of transcription factors (83,
106, 222). Considering this multiplicity of cellular functions, their relative insolubility and
potential cytotoxicity (102, 103), it is logical that the intracellular concentrations and
localization of free FA and fatty acyl-CoAs are tightly regulated. This is accomplished at
least in part due to the activities of lipid binding proteins, several of which have been
implicated in the control of the intracellular concentration and the partitioning of long-chain
FA and acyl-CoA within hepatocytes. These include the liver fatty acid binding protein (L-
Author Manuscript

FABP, synonym FABP1), acyl-CoA binding protein (ACBP; also named as acyl-CoA-
binding domain containing protein-1, ACBD1, or diazepam binding inhibitor, DBI), and
sterol carrier protein-2 (SCP2) (8, 103) (Fig. 2).

The FABP protein family comprises 10 members (237). FABP1 is abundant in the cytosol of
hepatocytes, comprising 2–5% of the cytosolic protein content (177) and reversibly binds
long-chain FA, fatty acyl-CoAs, lysophospholipids, cholesterol and other lipids (14, 36, 254,
255). Fabp1−/− male mice fed standard chow diet exhibit normal gain weight and liver

Compr Physiol. Author manuscript; available in PMC 2019 February 15.


Alves-Bezerra and Cohen Page 8

weights (172, 192). However, these animals have reduced total hepatic TG contents as well
Author Manuscript

as decreased rates of FA esterification into TG, with preserved or increased esterification


into PL (172, 192). Because the overexpression of FABP1 in L-cell fibroblasts enhances the
intracellular traffic of short-, medium-, and long-chain FA from the cytosol to the nucleus
(119), it has been conjectured that FABP1 must enter nuclei to promote the entry of FA
(222). Indeed, FABP1 directly interacts with PPARα within nuclei of primary hepatocytes
(117). This interaction may represent a mechanism for the delivery of FA ligands that
activate this key transcription factor (222). FABP1 ablation is protective against diet-induced
obesity and hepatic steatosis and improves glucose metabolism in mice fed a Western (high
saturated fat, high cholesterol) diet (193). Taken together, these findings suggest a role for
FABP1 in FA partitioning into specific metabolic pathways.

ACBP binds medium- and long-chain acyl-CoA with high affinity (84, 207, 211). Levels of
ACBP mRNA and protein are highest in the liver (26). In the mouse, ACBP expression is
Author Manuscript

modulated by nutritional status. Fasting results in decreased hepatic ACBP expression levels,
whereas high fat feeding increases it (21). Overexpression of ACBP in rat hepatoma cells
increases TG content (283). In addition, hepatic TG and PL concentrations are increased in
transgenic mice that overexpress ACBP in the liver (118). Under these conditions,
microsome-associated glycerol-3-phosphate acyltransferase (GPAT) activity is induced,
suggesting that ACBP has a role in the formation and partitioning of an acyl-CoA pool
towards glycerolipid synthesis (118).

Expression of SCP-2 is high in liver and intestine (15). This protein has high affinity for
binding both long-chain FA and acyl-CoAs, with Kd values in the same range as reported for
FABP and ACBP (103, 144, 177). Knockout mice exhibit reduced plasma FA
concentrations, decreased hepatic cholesteryl ester and TG contents, as well as increased
Author Manuscript

rates of β-oxidation (225). SCP-2 ablation also results in alterations in the composition and
physical properties of lipid rafts from mice primary hepatocytes, including the enrichment in
PL contents (9). Moreover, hepatic SCP-2 and FABP1 act to facilitate hydrolysis of
cholesteryl ester molecules associated with high-density lipoprotein (HDL) particles and
increasing the biliary secretion of bile acids (270) and cholesterol (92).

Because the metabolic fates of FA and acyl-CoA molecules are linked to these intracellular
lipid biding proteins, it is tempting to speculate that they could be leveraged to manipulate
the flux of lipids between anabolic and catabolic pathways in order to defend against the FA-
mediated lipotoxicity, that contributes to the pathogenesis of NAFLD.

Triglyceride metabolism
Author Manuscript

The assembly of TG molecules constitutes the principal means by which the liver stores and
exports FA. Under normal conditions, the liver stores little TG, but exports considerable
amounts in the form of VLDL particles that deliver FA to muscle and fat tissue, depending
on nutritional status. Whereas previously it was considered that excess TG stores in the
setting of NAFLD contributed to lipotoxicity (61), emerging concepts suggest that increased
TG storage and VLDL secretion are instead protective against FA-mediated hepatotoxicity.

Compr Physiol. Author manuscript; available in PMC 2019 February 15.


Alves-Bezerra and Cohen Page 9

Triglyceride synthesis
Author Manuscript

In most mammalian cell types, the G3P pathway is the principal route for the synthesis of
TG, contributing over 90% of total TG synthesis (56, 62). The first and rate-limiting step of
this pathway is the esterification of long-chain acyl-CoA to G3P, which is catalyzed by
mitochondrial and microsomal G3P acyltransferase (GPAT) enzymes. Lysophosphatidic acid
(LPA) molecules produced in this reaction are then acylated to form phosphatidic acid (PA)
by the acylglycerol-3-phosphate acyltransferases (AGPAT) present in the ER membrane. PA
can be converted into cytidine diphosphate diacylglycerol (CDP-DG), which is a substrate
for the synthesis of certain glycerolphospholipids and cardiolipins (113, 231) or can be
dephosphorylated by phosphatidate phosphohydrolase (PAP, synonym Lipin) to form DG,
which serve as precursor molecules for the synthesis of TG, as well as phosphatidylcholine
(PC) and phosphatidylethanolamine (PE) (56, 73). DG acyltransferase (DGAT) catalyzes the
acylation of DG, constituting the final step of TG synthesis (55). Newly synthesized TG
Author Manuscript

molecules are then directed from ER lipid bilayer to form cytosolic LDs (99, 278) (Fig. 3).

Four isoforms of GPAT are encoded by distinct genes (276). In the liver, the mitochondria-
associated GPAT isoforms (GPAT1 and 2) contribute 30–50% of total GPAT activity (56).
GPAT1 is highly expressed in the liver (157). It is induced after fasting/refeeding and in
response to insulin, and downregulated via cAMP-dependent signaling pathway during
nutrient deprivation (76, 186, 230). GPAT1-deficient mice exhibit increased rates of FA
oxidation and concentrations of plasma β-hydroxybutyrate. Moreover, hepatic TG contents
are decreased by 60%, suggesting that GPAT1 functions in the partitioning of FA towards
TG synthesis and away from oxidation in the liver (110). GPAT1 deficiency also leads to
reduced plasma TG levels, reduced VLDL secretion rates and decreased total body weights
(110, 191). Accordingly, the overexpression of GPAT1 in rat hepatocytes leads to increased
esterification of oleate into LPA, DG and TG, and to increased intracellular TG contents
Author Manuscript

(158, 165, 188). Although highly expressed in the liver, the metabolic role(s) of GPAT2 in
the lipid metabolism remain unclear. This isoform is not regulated by either insulin or
fasting/refeeding (271), and mouse models with GPAT2 loss or gain of function have yet to
be reported.

GPAT3 and GPAT4 account for microsomal GPAT activity. These isoforms are modulated by
nutritional status and hormonal variations. Although initially considered to mainly mediate
PL synthesis (55), more recent studies have revealed their contributions to TG metabolism in
adipose tissue. Specific roles for GPAT3 and GPAT4 in hepatic lipid metabolism remain
largely undefined. The overexpression of GPAT4 in HepG2 cells leads to 20% increase in
intracellular TG content (189), and its knockout in mice results in 40–50% reduction in
hepatic TG (37, 262). It is noteworthy that hepatocytes from mice lacking GPAT4 are able to
Author Manuscript

incorporate de novo synthesized FA into TG (275), suggesting that this isoform could be
relevant for the metabolism of exogenous FA.

The second step in TG synthesis is mediated by AGPAT enzymes. Although ten AGPAT
enzymes have been identified on the basis of sequence homology, the enzymatic activity for
LPA acylation has been confirmed for only a few (252). Certain isoforms initially described
as AGPAT were later reclassified into different acyltransferase groups. For example,
AGPAT6 and AGPAT8 are currently designated as GPAT4 and GPAT3, respectively (38, 49,

Compr Physiol. Author manuscript; available in PMC 2019 February 15.


Alves-Bezerra and Cohen Page 10

189). AGPAT1 and AGPAT2 are localized in the ER (74, 97, 150, 277) and are highly
Author Manuscript

expressed in the liver (75, 240, 252, 277). However, the contribution of these enzymes to the
hepatic metabolism remains unknown.

Lipins-1–3 constitute the PAP enzyme family. Under lipogenic stimuli, these proteins
translocate from the cytosol to ER membrane in order to access their substrate PA and
contribute the TG biosynthetic pathway (31, 173). Additionally, lipin-1 can localize in the
nucleus of hepatocytes, where it acts as a transcriptional co-activator (234). Fasting and
glucocorticoid administration result in increased lipin-1 expression in mouse livers (90, 169,
195). This effect is mediated by PPARγ coactivator 1α (PGC-1α) and is required to sustain
the fasting-induced expression of PPARα and its downstream targets. Lipin-1 physically
interacts with PGC-1α and PPARα in the nucleus, leading to upregulation of oxidative fatty
acid metabolism by mitochondria (90). Accordingly, mice lacking lipin-1 exhibit hepatic
steatosis (90), enhanced stearoyl-CoA desaturase-1 (Scd1) gene expression, increased rates
Author Manuscript

of VLDL secretion, and increased plasma concentrations of TG without changes in TG


synthetic rates (50). Reduced lipin-1 expression is observed in livers of obese diabetic mice,
and the overexpression of lipin-1 in this animal inhibits VLDL secretion and improves
hepatic insulin responsiveness (50). The effects of lipin-1 on VLDL secretion are
independent of PAP enzymatic activity and instead linked to its nuclear receptor interactions
(50). In addition to lipin-1, high levels of lipin-2 expression are also observed in the liver
(71) and hepatic lipin-3 transcription is increased in lipin-1-deficient mice (71). Moreover,
lipin-2 mRNA levels are induced in lipin-1-deficient mice, supporting the hypothesis that
lipin isoforms could compensate in order to preserve PAP activity and restore the levels of
TG synthesis (107).

In mouse liver, DGAT activity is detected in two microsomal membrane locations: luminal
Author Manuscript

activity (DGAT1, measured as latent – luminal – DGAT activity) and cytosol-accessible


activity (predominantly DGAT2, measured as overt DGAT activity) (200). The former
contributes to the synthesis of TG that are packaged into VLDL. The latter contributes to the
synthesis of a TG pool to be stored in the cytoplasmic LDs (1, 200, 272). Interestingly, the
DGAT proteins are genetically unrelated. DGAT1 is part of the membrane bound O-acyl
transferase (MBOAT) protein family, which comprises a diverse group of acyltransferases,
including acyl-CoA:cholesterol acyltransferase (ACAT) and protein-cysteine N-
palmitoyltransferase (286). DGAT2 is a member of diacylglycerol acyltransferase (DAGAT)
family that also includes the monoacylglycerol acyltransferase (MGAT) and acyl-CoA wax
alcohol acyltransferases (AWAT) (286).

Although DGAT1 and DGAT2 are able to catalyze the same reaction, these enzymes do not
Author Manuscript

appear to be redundant in their respective functions. DGAT1 mRNA is modestly expressed


in the liver in comparison to other organs, such as skeletal muscle and small intestine (41).
This enzyme is able to catalyze acyltransferase reactions using substrates other than DG. For
example, DGAT1 acylates MG, wax esters and retinol in vitro (285). DGAT1 predicted
topology combined with results obtained from loss of function studies supports the
involvement of this enzyme in both the overt and latent DGAT activities (286, 289).
Accordingly, overexpression of DGAT1 in rat hepatoma cells leads to increased intracellular
TG contents and also increased rates of VLDL-TG secretion (164). Dgat1 knockout mice

Compr Physiol. Author manuscript; available in PMC 2019 February 15.


Alves-Bezerra and Cohen Page 11

have slightly reductions in hepatic TG concentrations when fed a low fat diet, with more
Author Manuscript

prominent decreases being observed in response to high fat feeding (47, 238). High fat fed
Dgat1 knockout high fat fed mice are also resistant to obesity and have improved insulin
responsiveness (238). Moreover, DGAT1 expression is increased in NAFLD subjects and
may contribute to hypertriglyceridemia (146).

By contrast, Dgat2 transcripts are abundant in the liver, and are decreased upon fasting and
increased after refeeding (179). DGAT2 exhibits in vitro specificity for the use of DG
molecules as substrates (42). The predicted topology of this protein suggests its contribution
to overt activity, and this was confirmed in HepG2 cells by studies using DGAT2 inhibitors
(98). Accordingly, DGAT2 is found in close association to cytosolic LDs and mitochondria-
associated membranes, where enzymes involved in TG synthesis and storage are also
localized (178, 242). Dgat2 knockout mice exhibit severe lipopenia and die in the first day
after birth (243). The liver-specific overexpression of DGAT2 in chow-fed mouse leads to
Author Manuscript

the development of hepatic steatosis but no abnormalities in plasma glucose and insulin
signaling were observed, suggesting that hepatic TG accumulation is not a cause of insulin
resistance (183). In agreement with this observation, human DGAT2 genetic variants reduce
hepatic steatosis. However, the decrease in fat content does not lead to improved insulin
sensitivity (131).

In keeping with specialized cellular functions of the enzymes, gain and loss of function
studies in mice have revealed different phenotypes for DGAT1 and DGAT2, without
evidence of compensation (243). In fact, a more recent study revealed a unique role of
hepatic DGAT2 in its use of de novo-synthesized acyl-CoA and DG as substrates (281). The
TG formed by this reaction are stored in cytosolic LDs and ultimately undergoes lipase-
mediated hydrolysis. The resulting lipid intermediates (i.e., MG and DG) are available for
Author Manuscript

re-esterification by overt DGAT1 activity present in the cytosolic compartment, followed by


TG transport into the ER. These intermediates can also undergo translocation to the ER
lumen followed by acylation catalyzed by latent DGAT1 activity (Fig. 3A). Overt and latent
activity of DGAT1 contribute to lipoprotein lipidation during VLDL assembly and utilize
pre-formed, but not de novo-synthesized acyl-CoA as preferred substrate (Fig. 3B).
Therefore, DGAT2 and DGAT1 use independent pools of acyl-CoA for TG synthesis, and
act in a sequential and integrated way for TG packaging into nascent VLDL particles (281).

Although hepatic TG accumulation is the major determinant of NAFLD (61), lipid and
lipidomic analyses have revealed changes in the compositions of a variety of lipids,
including FA, DG, LPA, PA, PL and ceramides (78, 127, 168, 216, 260). The specific
contributions of hepatic lipid species to NAFLD remain the subject of active investigation
Author Manuscript

(136).

VLDL assembly and secretion


TG-rich VLDL particles represent the mechanism by which FA are exported from the liver
and delivered to muscle for oxidation and adipose tissue for storage, respectively. VLDL
assembly is a two-step process that begins in the ER lumen. In the first step, microsomal
triglyceride transfer protein (MTP) acts to incorporate a small amount of TG in apoB100 as
it is being translated by ribosomes and translocated across the ER membrane (101). In the

Compr Physiol. Author manuscript; available in PMC 2019 February 15.


Alves-Bezerra and Cohen Page 12

second step, additional TG is packaged into the nascent apoB100-containing particles as


Author Manuscript

they are traverse from the ER to the Golgi apparatus to form VLDL particles (Fig. 4) (54).

The link between increased VLDL secretion and metabolic diseases, such as insulin
resistance and diabetes, is well stablished (54, 251, 297). The development of insulin
resistance leads to hepatic TG accumulation owing to both enhanced FA uptake into the liver
and increased DNL. The greater availability of TG together with higher MTP activity
promote the overproduction of VLDL particles (10) and greater plasma TG concentrations
(54). A positive correlation between plasma TG concentrations and the hepatic steatosis is
not always observed (63). Obese leptin-deficient ob/ob mice exhibit decreased VLDL
production rates despite marked hepatic steatosis (163). In nondiabetic obese NAFLD
subjects, VLDL-TG secretion rates are elevated up to 2-times in comparison to normal
subjects (82). In this setting, higher rates of TG accumulation fail to increase VLDL-TG
proportionately, suggesting a limited capacity for increasing VLDL production and secretion
Author Manuscript

(81).

Reductions in VLDL secretion can also lead to hepatic steatosis. This occurs in the setting of
genetic defects in apoB100 and MTP, which lead to hypobetalipoproteinemia and
abetalipoproteinemia, respectively (274). Mipomersen and lomitapide are drugs that
respectively target apoB100 and MTP in the interest of reducing plasma LDL cholesterol
concentrations by blocking VLDL secretion (58, 263). A mechanism-based side effect of
these medications is hepatic steatosis. Finally, a polymorphism in Transmembrane 6
superfamily member 2 (TM6SF2) has been identified by human genome-wide association
studies to be associated with hepatic steatosis (147, 236). The polymorphism results in
reduced VLDL secretion rates, because of impaired lipidation of nascent VLDL particles
(236). Whether the control of VLDL secretion can be leveraged in the management of
Author Manuscript

NAFLD remains unclear.

Lipid droplet biology


LDs are dynamic cellular structures that transiently store lipids. Second to white adipose
tissue, the liver has the next greatest capacity to store TG in LDs, such that overnight fasting
leads to hepatic LD formation that accommodates FA derived from adipose tissue lipolysis
(267). The neutral lipid composition of LDs differs among different hepatic cell types.
Whereas hepatocytes harbor TG-enriched LD cores, stellate cell LDs store predominantly
vitamin A as retinyl esters (23). The TG-rich cores of LDs in hepatocytes are surrounded by
a coat comprised of PL and proteins. The PL are primarily PC and form a monolayer around
the neutral lipid core (13). CTP-phosphocholine cytidylyltransferase (CCT) is the rate-
limiting enzyme in the Kennedy pathway for PC biosynthesis. Indicative of its role in LD
Author Manuscript

biogenesis, CCT is recruited from the cytosol to expanding LDs in a variety of cell types in
order to provide PC for LD surface expansion (278). However, the precise role of CCT on
LD formation within hepatocytes is uncertain.

In addition to preserving the LD structure, LD-associated proteins regulate the formation,


expansion and contraction of LDs. These proteins include selected enzymes of lipid
synthesis, as ACSL3 and DGAT2 (Fig. 4). Changes in the composition and activity of the
protein components of the LD have been implicated as contributing to metabolic disease (40,

Compr Physiol. Author manuscript; available in PMC 2019 February 15.


Alves-Bezerra and Cohen Page 13

99, 100, 105). The predominant hepatocellular LD proteins are the members of perilipin
Author Manuscript

(PLIN; former perilipin/ADRP/TIP47, PAT) protein superfamily (PLIN1–5). These proteins


in adipocytes are involved both in the stabilization of LD structure as well as in the control
of availability of substrates for several LD-associated enzymes (141). Perilipin 1 (PLIN1;
synonym Perilipin A) is normally only expressed in LDs from adipocytes, whereas PLIN2
(synonym ADRP) and PLIN3 (synonym TIP47) are associated with LDs in hepatocytes
(245). Plin1 mRNA is expressed in livers of subjects with NAFLD (104), and PLIN1 protein
is found in LDs of steatotic hepatocytes (245). A key role for PLIN2 in hepatocytes is
supported by the observation that it is the most abundant LD protein in the HuH7 hepatocyte
cell line (94), and its levels are also increased in steatotic livers of both humans and mice
(185) in response to PPARγ activation (197). Overexpression of PLIN2 in rat hepatic stellate
cells results in lipid accumulation within LDs (96), while PLIN2-deficient mice exhibit a
60% reduction in hepatic TG content and are resistant to diet-induced hepatic steatosis (46).
Author Manuscript

Liver-specific ablation of PLIN2 in mice results in reduced LD size and is protective against
diet-induced NASH (190). PLIN1 is frequently associated with larger LDs in the liver of
NAFLD subjects while PLIN2 is predominantly found in smaller LDs (93). The specific
targeting of these proteins raises the suggestion of distinct PLIN1 and PLIN2 functions.

Plin3 expression is also induced in livers of high fat-fed mice and its knockdown reduces
hepatic TG content, attenuates steatosis, and improves insulin sensitivity and glucose
tolerance (39). In hepatocytes, the degradation of PLIN2 and PLIN3 are steps required for
lipolysis during nutrient deprivation. The removal of PLIN proteins from LD occurs via
chaperone-mediated autophagy (CMA) and facilitates the LD association of components of
the lipolytic machinery, including ATGL (134). Moreover, the combined knockdown of
Plin2 and Plin3 results in increased insulin resistance in mouse AML12 hepatocytes
stimulated with oleate (17). PLIN5 is highly expressed in oxidative tissues, including liver.
Author Manuscript

Hepatic Plin5 mRNA levels are increased after fasting in response to activation of PPARα
and/or PPARβ/δ (140). PLIN5 expression is also enhanced in steatotic livers of humans and
obese mice, and the knockout of this protein results in decreased hepatic TG content and
smaller-sized LDs (268).

In hepatocytes, LDs are tightly associated with ER membrane cisternae, allowing the
physical interaction between LD components and microsomal proteins, such as apoB100
(194) (Fig. 4). The fat-inducing transmembrane proteins (FIT1 and 2) are localized in the
ER and have the ability to bind DG and TG, with FIT2 showing higher affinity (108). The
overexpression on FIT2, the most abundant hepatic FIT isoform, in mouse liver results in
increased TG content and LD accumulation (128). These data suggest that FIT2 functions in
the binding of neutral lipids synthesized within ER and their channeling to nascent LDs (99).
Author Manuscript

The DFF45-like effector (CIDE) proteins, Cidea, Cideb, and Fsp27, are associated both to
the ER and LDs and have been linked to LD metabolism (100). Recent studies have
demonstrated that these proteins act to promote lipid exchange and fusion among LDs (282).
However, they are differentially distributed among hepatocytes based on sizes of LD that
they harbor. Cideb localizes both to hepatocytes with small and with large LDs, whereas
Cidea and Fsp27 specifically localize to hepatocytes containing large LDs (282).

Compr Physiol. Author manuscript; available in PMC 2019 February 15.


Alves-Bezerra and Cohen Page 14

Cidea expression is increased in livers of diabetic mice (137) and a coding region
Author Manuscript

polymorphism is associated with obesity in human subjects (60, 290). Moreover, expression
of both Cidea and Fsp27 are upregulated in mice with hepatic steatosis induced by PPARγ
activation (176, 264, 288). Fsp27 is a target gene of PPARγ and its transcription is decreased
in ob/ob mice with liver-specific knockdown of PPARγ, accompanied by decreased TG
content in the liver (175, 176). Cideb is highly expressed in the liver and Cideb knockout
mice are resistant to obesity and liver steatosis induced by high-fat diet (160). Moreover,
these animals have decreased levels of circulating TG and FA, and exhibit improved hepatic
insulin sensitivity. These effects are mediated by the hepatic downregulation of SREBP-1c,
resulting in decreased lipogenesis and increased FA oxidation (160). Cideb also interacts
with apoB100, and mice lacking Cideb exhibit reduced VLDL secretion and increased
hepatic steatosis (284). In addition to its role in VLDL assembly, Cideb has been associated
in VLDL export from ER to the Golgi. This occurs through the interaction with components
Author Manuscript

of coat complex II (COPII), namely Sar1 and Sec24, which are present in VLDL transport
vesicles (257).

Because NAFLD is characterized by excessive deposition of TG within cytosolic LDs and


because proteomic studies have revealed that numerous LD-associated proteins are
modulated in human and mice steatotic livers (138, 250), detailed knowledge of the biology
hepatocyte LDs could reveal novel targets for therapeutic intervention.

Lipolysis
The LD-associated adipose triglyceride lipase (ATGL, also known as PNPLA2) is the rate-
limiting step in TG lipolysis within adipocytes. The resulting DG molecules are then
hydrolyzed by the hormone sensitive lipase (HSL) to release monoglycerides (MG). In the
final step, the monoacylglycerol lipase (MGL) cleaves MG into glycerol and FA (151) (Fig.
Author Manuscript

5). Gain and loss of function studies have revealed that hepatic ATGL is required for the
lipolysis of TG stored in LDs, controls substrate availability for FA oxidation, and modulates
the progression of hepatosteatosis (198, 208, 280). Full activation of ATGL depends on co-
activation by the comparative gene identification-58 (CGI-58) (152). Liver-specific ablation
of CGI-58 results in NAFLD phenotypes in mice that include hepatic steatosis and fibrosis
(109). In the setting of hepatic steatosis, lipolysis mediated by hepatic ATGL is decreased
when PLIN5 binds competitively to CGI-58, displacing ATGL (268, 269). Insulin-resistant
NAFLD patients who exhibit higher degrees of liver steatosis when compared to NAFLD
subjects without insulin resistance, also exhibit decreased CGI-58 mRNA levels. These
observations suggest that insulin-resistance could induce hepatic TG accumulation through
CGI-58-mediated reductions in ATGL-dependent lipolysis (132). Less is known concerning
MGL and HSL activities in hepatocytes, although Hsl expression is down-regulated in livers
Author Manuscript

of NAFLD patients (146).

In the setting of insulin resistance, enhanced rates of lipolysis within white adipose tissue are
a major contributor to the increased FFA plasma levels and hepatic steatosis (159). However,
the contribution of TG hydrolysis within hepatocytes to aberrant lipid accumulation in the
liver is less clear. Several lipases are present in close association with LDs in adipocytes (30,
66) (Fig. 5), but their contributions to LD TG hydrolysis remain topics of discovery.

Compr Physiol. Author manuscript; available in PMC 2019 February 15.


Alves-Bezerra and Cohen Page 15

The patatin-like phospholipase domain-containing 3 (PNPLA3/adiponutrin) is a TG


Author Manuscript

hydrolase/transacylase that localizes to LD, ER and cytoplasm (124, 212). A single-


nucleotide variant of the Pnpla3 gene (I148M) confers increased risk of NAFLD in humans
(210). In this connection, livers from transgenic mice expressing this allele show
accumulation of inactive PNPLA3 in lipid droplets and increased steatosis when fed a high-
sucrose diet (235). Although the genetic association of PNPLA3 with NALFD is robust, key
details are still lacking regarding the metabolic activity of this protein.

Autophagy
Autophagy is a lysosome-dependent process that selectively targets, transports and regulates
the storage of essential components, including lipids, proteins and carbohydrates. Three
types of autophagy have been described in hepatocytes (59). In macroautophagy, LDs or
other organelles are engulfed to form autophagosomes, which fuses to lysosome to generate
Author Manuscript

an autolysosome. After hydrolysis by lysosomal enzymes, lipid products are released to the
cytosol and recycled by other cellular processes (167). The autophagosome formation relies
on several autophagy receptors in addition to the autophagy-related proteins (ATGs) (241).
In microautophagy, small vesicles originated from lysosomal membrane invagination
mediate the engulfment of cytosolic components (167). In chaperone-mediated autophagy
(CMA), lysosomal degradation of specific proteins is regulated upon recognition by CMA
receptor (133).

Macroautophagy of LDs (also named lipophagy) is stimulated in hepatocytes during


starvation, leading to the release of FA in the cytosol after TG breakdown (233).
Additionally, products of lipid anabolism have inhibitory effects in lipophagy (167).
Although the mechanisms controlling lipophagy are not well understood, this process is
driven by the small GTPase Rab7 protein, which recognizes LDs and mediates their
Author Manuscript

engulfment by autophagosomes (221).

Decreased autophagy has been reported in obesity-related disorders, including


hepatosteatosis (219). Altered membrane lipid composition due to high-fat feeding reduces
the activity of lysosomal proteins and decreases autophagic activity in mouse liver (145,
209). Impaired CMA leads to hepatosteatosis through at least three mechanisms. First,
reduced CMA activity decreases mitochondrial function and β-oxidation. Second, CMA-
deficient mice exhibit increased expression of lipogenic enzymes, including DGAT2 (219).
Finally, CMA is required for the lipolysis of hepatic TG through hsp70-mediated
degradation of PLIN2 and PLIN3 (134). Hence, impaired autophagy in the setting of
steatosis exacerbates lipid accumulation within the liver.
Author Manuscript

Fatty acid oxidation


FA derived from hydrolysis of hepatic TG stores, circulating lipids or DNL, can be oxidized
by multiple pathways. Mitochondrial β-oxidation is the primary route for the oxidation of
the majority of FA found in hepatocytes, including short- (<C4), medium- (C4-C12), and
long-chain (C12-C20) FA. β-oxidation of very long- (C20-C26) and branched-chain FA
begins in the peroxisomes. Additional pathways for FA oxidation include α-oxidation and
ω-oxidation within the ER and are mediated by cytochrome P450 4A family members (153,

Compr Physiol. Author manuscript; available in PMC 2019 February 15.


Alves-Bezerra and Cohen Page 16

187) (Fig. 5). The expression of genes involved in mitochondrial and extramitochondrial FA
Author Manuscript

oxidation is regulated largely by PPARα activity. PPARα is expressed at high levels in liver,
and both whole-body and hepatocyte-specific ablation of PPARα in mice lead to reduced
transcription of hepatic genes related to mitochondrial β-oxidation, such as very long chain
acyl-CoA dehydrogenase (VLCAD), long chain acyl-CoA dehydrogenase (LCAD), and
ACSL1 (7), and genes involved in peroxisomal β-oxidation, including peroxisomal fatty
acyl-CoA oxidase (AOX) and cytochrome P450 4A (156, 184). These animals also
experience hypoketonemia and increased hepatic steatosis (111, 184).

It is noteworthy that AOX ablation leads to sustained activation of PPARα, likely due to
accumulation of PPARα ligands and induction of cytochrome P450 4A (86). Impaired
peroxisomal FA oxidation results in the accumulation of dicarboxylic acids within the liver
and leads to mitochondrial damage and microvesicular steatosis (153). Alterations in
dicarboxylic acid levels are also observed in conditions of FA overload and have been
Author Manuscript

correlated with increased risk of NAFLD development, although the cellular mechanism
remains unknown (215, 287). PGC1α and BAF60a form a complex that regulates PPARα
transcriptional activity in the liver. Overexpression of BAF60a increases fatty acid oxidation
and reduces hepatic steatosis in obese mice (162). Similar phenotypes have been observed in
livers of mice overexpressing the PPARα coactivator PGC1β (18).

Notwithstanding the relationship between PPARα activity and hepatic lipid metabolism, the
correlation between PPARα expression levels and hepatic steatosis is inconsistent, with
reports demonstrating unchanged, increased and decreased expression (16). This is also the
case for rates of hepatic mitochondrial β-oxidation. The transcription of CPT1a, is down-
regulated in NAFLD (146). Although previous studies revealed decreased hepatic
mitochondrial FA oxidation in the setting of hepatic steatosis (123, 202, 206), recent work
Author Manuscript

has demonstrated increased oxidation rates in livers of subjects with hepatic steatosis,
metabolic syndrome and NAFLD (16). In livers of ob/ob mice, mitochondrial and
peroxisomal oxidative capacities are elevated (29). Mechanisms underlying increased FA
oxidation in NAFLD have been reviewed (16). Hypotheses include enhanced activation of
PPARα due to increased hepatic FA uptake or biosynthesis.

ER-mediated control of lipid homeostasis


Among its many functions, the ER regulates the synthesis of lipids and proteins, as well as
intracellular calcium storage (24). Under physiological conditions, unfolded and misfolded
proteins accumulate in the ER lumen. In response to this ER stress, the cell initiates multiple
lines of defense. One is the activation of ER-associated degradation (ERAD), a ubiquitin/
proteasome protein degradation pathway that relies on E3 ligase complex activation (11).
Author Manuscript

Second, the unfolded protein response (UPR) is initiated (229). This comprises three main
signaling pathways mediated by microsomal transmembrane proteins: PKR-like ER kinase
(PERK; also known as eukaryotic translation initiation factor 2-α kinase 3, eIF2αk3),
inositol requiring enzyme 1 (IRE1), and activating transcription factor 6 (ATF6) (229). UPR
activation alleviates the accumulation of misfolded proteins into the ER through the
inhibition of protein translation and the induction of chaperones to increase protein folding
capacity, as well as expanding the ER membrane (11).

Compr Physiol. Author manuscript; available in PMC 2019 February 15.


Alves-Bezerra and Cohen Page 17

Lipid accumulation is also associated with chronic ER stress in hepatocytes (293). Lipid-
Author Manuscript

mediated induction of ER stress can be triggered by multiple mechanisms. Obese mice


exhibit altered microsomal membrane composition in the liver, with increased PC:PE ratios
and variations in PL acyl-chain saturation degree (91, 121). Hence, changes in membrane
fluidity leads to altered activity of membrane-associated proteins. For instance, sarco-
endoplasmic reticulum Ca2+-ATPase (SERCA) activity is reduced in steatotic mouse livers.
In this model, decreased ER Ca2+ content impairs intraluminal Ca2+-dependent chaperones
function and is a cause for ER stress (91).

UPR-related proteins directly sense membrane lipid composition in non-hepatic cells. In


mouse fibroblasts, PERK and IRE1 are responsive to increased lipid saturation even in the
absence of their ER-spanning transmembrane domain (265). However, it is uncertain
whether direct lipid sensing mechanisms modulate hepatic ER stress. Because ER stress in
turn promotes lipid accumulation in hepatocytes, chronic ER stress may directly contribute
Author Manuscript

to the pathogenesis of NAFLD (291). In cell culture systems, prolonged exposure to oleate
or palmitate induces ER stress when these saturated FA are incorporated to ER membrane
PL (25). A concomitant reduction in VLDL-TG secretion is attributable to enhanced
apoB100 degradation by proteasomal and non-proteasomal mechanisms (43, 199). In this
connection, treatment with 4-phenyl butyric acid (PBA), a chemical chaperone, improves ER
homeostasis and VLDL secretion after FA exposure (43). Although reduced VLDL secretion
due to chronic ER stress results in TG accumulation in McA-RH7777, its potential
contribution for the progression of hepatic steatosis was not demonstrated in animal models
(199).

ER stress also promotes lipid accumulation in hepatocytes that is attributable to increased


lipogenesis (11). In obese mice, hepatic induction of UPR pathway reduces Insig-1 levels
Author Manuscript

and leads to the activation of SREBP, increased expression of lipogenic enzymes and
steatosis (129). UPR-mediated activation of PERK may also contribute to hepatic steatosis
by increasing expression of VLDL receptors in hepatocytes (126). Moreover, ER stress
reduces autophagy through the cross-talk with components of ERAD and UPR in different
cell types, exacerbating the hepatic steatosis (11, 201)

In summary, excess lipid load potentially initiates a vicious cycle of hepatic lipid
accumulation due to abnormal activity of ER stress-activated pathways, leading to altered
lipid homeostasis and TG deposition in the liver (11, 291, 293).

Conclusion
The liver is a main organ for the metabolism of fatty acids and triglycerides and altered
Author Manuscript

activity of metabolic pathways in the setting of overnutrition leads to a common chronic


liver disease known as NAFLD. Despite this biomedical relevance, many important
questions regarding basic aspects of hepatic lipid metabolism remain unanswered. These
include: (a) Which FA transporters are responsible for the increased intracellular
concentration of FA in the setting of overnutrition? (b) What is the precise mechanism by
which DNL is increased in obese and NAFLD subjects? (c) What is the precise molecular
mechanism underlying the partitioning of FA and acyl-CoA among metabolic pathways,

Compr Physiol. Author manuscript; available in PMC 2019 February 15.


Alves-Bezerra and Cohen Page 18

which presumably involve ACOT, ACSL and lipid-binding proteins? (d) How are rates of FA
Author Manuscript

oxidation regulated in the setting of overnutrition? In summary, a deeper understanding is


needed regarding the precise mechanisms whereby fatty acid uptake and synthesis are tightly
balance against oxidation and secretion within the liver. This would be expected to lead to a
better understanding of how triglycerides accumulate in many, but not all, individuals who
consume excess calories and to new therapeutic strategies for the management of NAFLD.

Acknowledgements
This work was supported by NIH grants R37DK048873, R01DK056626 and R01DK103046 to D.E.C. M.A.B. is
the recipient of NASH Fatty Liver Disease Postdoctoral Research Fellowship from the American Liver Foundation.

References
1. Abo-Hashema KAH, Cake MH, Power GW, and Clarke D. Evidence for triacylglycerol synthesis in
Author Manuscript

the lumen of microsomes via a lipolysis-esterification pathway involving carnitine acyltransferases.


J Biol Chem 274: 35577–35582, 1999. [PubMed: 10585433]
2. Abu-Elheiga L, Matzuk MM, Abo-Hashema KA, and Wakil SJ. Continuous fatty acid oxidation and
reduced fat storage in mice lacking acetyl-CoA carboxylase 2. Science 291: 2613–2616, 2001.
[PubMed: 11283375]
3. Abu-Elheiga L, Oh WK, Kordari P, and Wakil SJ. Acetyl-CoA carboxylase 2 mutant mice are
protected against obesity and diabetes induced by high-fat/high-carbohydrate diets. Proc Natl Acad
Sci U S A 100: 10207–10212, 2003 10.1073/pnas.1733877100 [PubMed: 12920182]
4. Achouri Y, Hegarty BD, Allanic D, Becard D, Hainault I, Ferre P, and Foufelle F. Long chain fatty
acyl-CoA synthetase 5 expression is induced by insulin and glucose: Involvement of sterol
regulatory element-binding protein-1c. Biochimie 87: 1149–1155, 2005. [PubMed: 16198472]
5. Ahmed A, Wong RJ, and Harrison SA. Nonalcoholic Fatty Liver Disease Review: Diagnosis,
Treatment, and Outcomes. Clin Gastroenterol Hepatol 13: 2062–2070, 2015 10.1016/j.cgh.
2015.07.029 [PubMed: 26226097]
Author Manuscript

6. Anderson RGW. The caveolae membrane system. Annu Rev Biochem 67: 1998.
7. Aoyama T, Peters JM, Iritani N, Nakajima T, Furihata K, Hashimoto T, and Gonzalez FJ. Altered
constitutive expression of fatty acid-metabolizing enzymes in mice lacking the peroxisome
proliferator-activated receptor alpha (PPAR alpha). J Biol Chem 273: 5678–5684, 1998. [PubMed:
9488698]
8. Atshaves BP, Martin GG, Hostetler HA, McIntosh AL, Kier AB, and Schroeder F. Liver fatty acid-
binding protein and obesity. J Nutr Biochem 21: 1015–1032, 2010. [PubMed: 20537520]
9. Atshaves BP, McIntosh AL, Payne HR, Gallegos AM, Landrock K, Maeda N, Kier AB, and
Schroeder F. SCP-2/SCP-x gene ablation alters lipid raft domains in primary cultured mouse
hepatocytes. J Lipid Res 48: 2193–21211, 2007. [PubMed: 17609524]
10. Avramoglu RK, Basciano H, and Adeli K. Lipid and lipoprotein dysregulation in insulin resistant
states. Clin Chim Acta 368: 1–19, 2006. [PubMed: 16480697]
11. Baiceanu A, Mesdom P, Lagouge M, and Foufelle F. Endoplasmic reticulum proteostasis in hepatic
steatosis. Nat Rev Endocrinol 12: 710–722, 2016 10.1038/nrendo.2016.124 [PubMed: 27516341]
12. Barrows BR, and Parks EJ. Contributions of different fatty acid sources to very low-density
Author Manuscript

lipoprotein-triacylglycerol in the fasted and fed states. J Clin Endocrinol Metab 91: 1446–1452,
2006. [PubMed: 16449340]
13. Bartz R, Li WH, Venables B, Zehmer JK, Roth MR, Welti R, Anderson RGW, Liu PS, and
Chapman KD. Lipidomics reveals that adiposomes store ether lipids and mediate phospholipid
traffic. J Lipid Res 48: 837–847, 2007. [PubMed: 17210984]
14. Bass NM. Function and regulation of hepatic and intestinal fatty acid binding proteins. Chem Phys
Lipids 38: 95–114, 1985. [PubMed: 4064226]
15. Baum CL, Kansal S, and Davidson NO. Regulation of sterol carrier protein-2 gene expression in
rat liver and small intestine. J Lipid Res 34: 729–739, 1993. [PubMed: 8509712]

Compr Physiol. Author manuscript; available in PMC 2019 February 15.


Alves-Bezerra and Cohen Page 19

16. Begriche K, Massart J, Robin MA, Bonnet F, and Fromenty B. Mitochondrial adaptations and
dysfunctions in nonalcoholic fatty liver disease. Hepatology 58: 1497–1507, 2013. [PubMed:
Author Manuscript

23299992]
17. Bell M, Wang H, Chen H, McLenithan JC, Gong DW, Yang RZ, Yu D, Fried SK, Quon MJ,
Londos C, and Sztalryd C. Consequences of lipid droplet coat protein downregulation in liver
cells: abnormal lipid droplet metabolism and induction of insulin resistance. Diabetes 57: 2037–
2045, 2008. [PubMed: 18487449]
18. Bellafante E, Murzilli S, Salvatore L, Latorre D, Villani G, and Moschetta A. Hepatic-specific
activation of peroxisome proliferator-activated receptor γ coactivator-1β protects against
steatohepatitis. Hepatology 57: 1343–1356, 2013. [PubMed: 23299802]
19. Benhamed F, Denechaud PD, Lemoine M, Robichon C, Moldes M, Bertrand-Michel J, Ratziu V,
Serfaty L, Housset C, Capeau J, Girard J, Guillou H, and Postic C. The lipogenic transcription
factor ChREBP dissociates hepatic steatosis from insulin resistance in mice and humans. J Clin
Invest 122: 2176–2194, 2012 10.1172/JCI41636 [PubMed: 22546860]
20. Berk PD, Wada H, Horio Y, Potter BJ, Sorrentino D, Zhou SL, Isola LM, Stump D, Kiang CL, and
Thung S. Plasma membrane fatty acid-binding protein and mitochondrial glutamic-oxaloacetic
Author Manuscript

transaminase of rat liver are related. Proc Natl Acad Sci U S A 87: 3484–3488, 1990. [PubMed:
2185471]
21. Bhuiyan J, Pritchard PH, Pande SV, and Seccombe DW. Effects of high-fat diet and fasting on
levels of acyl-CoenzymeA binding-protein in liver, kidney, and heart of rat. Metabolism-Clinical
and Experimental 44: 1185–1189, 1995. [PubMed: 7666793]
22. Birkenfeld AL, and Shulman GI. Nonalcoholic fatty liver disease, hepatic insulin resistance, and
type 2 diabetes. Hepatology 59: 713–723, 2014 10.1002/hep.26672 [PubMed: 23929732]
23. Blaner WS, O’Byrne SM, Wongsiriroj N, Kluwe J, D’Ambrosio DM, Jiang H, Schwabe RF,
Hillman EM, Piantedosi R, and Libien J. Hepatic stellate cell lipid droplets: a specialized lipid
droplet for retinoid storage. Biochim Biophys Acta 1791: 467–473, 2009. [PubMed: 19071229]
24. Borgese N, Francolini M, and Snapp E. Endoplasmic reticulum architecture: structures in flux.
Curr Opin Cell Biol 18: 358–364, 2006 10.1016/j.ceb.2006.06.008 [PubMed: 16806883]
25. Borradaile NM, Han X, Harp JD, Gale SE, Ory DS, and Schaffer JE. Disruption of endoplasmic
reticulum structure and integrity in lipotoxic cell death. J Lipid Res 47: 2726–2737, 2006 10.1194/
jlr.M600299-JLR200 [PubMed: 16960261]
Author Manuscript

26. Bovolin P, Schlichting J, Miyata M, Ferrarese C, Guidotti A, and Alho H. Distribution and
characterization of diazepam binding inhibitor (DBI) in peripheral tissues of rat. Regul Pept 29:
267–281, 1990. [PubMed: 2171047]
27. Bradbury MW. Lipid metabolism and liver inflammation. I. Hepatic fatty acid uptake: possible role
in steatosis. American journal of physiology Gastrointestinal and liver physiology 290: G194–
G198, 2006. [PubMed: 16407588]
28. Bradbury MW, Stump D, Guarnieri F, and Berk PD. Molecular modeling and functional
confirmation of a predicted fatty acid binding site of mitochondrial aspartate aminotransferase. J
Mol Biol 412: 412–422, 2011. [PubMed: 21803047]
29. Brady LJ, Brady PS, Romsos DR, and Hoppel CL. Elevated hepatic mitochondrial and
peroxisomal oxidative capacities in fed and starved adult obese (ob/ob) mice. Biochem J 231: 439–
444, 1985. [PubMed: 4062906]
30. Brasaemle DL, Dolios G, Shapiro L, and Wang R. Proteomic analysis of proteins associated with
lipid droplets of basal and lipolytically stimulated 3T3-L1 adipocytes. J Biol Chem 279: 46835–
Author Manuscript

46842, 2004. [PubMed: 15337753]


31. Brindley DN. Intracellular translocation of phosphatidate phosphohydrolase and its possible role in
the control of glycerolipid synthesis. Prog Lipid Res 23: 115–133, 1984. [PubMed: 6100459]
32. Browning JD, Szczepaniak LS, Dobbins R, Nuremberg P, Horton JD, Cohen JC, Grundy SM, and
Hobbs HH. Prevalence of hepatic steatosis in an urban population in the United States: impact of
ethnicity. Hepatology 40: 1387–1395, 2004 10.1002/hep.20466 [PubMed: 15565570]
33. Bu SY, and Mashek DG. Hepatic long-chain acyl-CoA synthetase 5 mediates fatty acid channeling
between anabolic and catabolic pathways. J Lipid Res 51: 3270–3280, 2010. [PubMed: 20798351]

Compr Physiol. Author manuscript; available in PMC 2019 February 15.


Alves-Bezerra and Cohen Page 20

34. Bu SY, Mashek MT, and Mashek DG. Suppression of long chain acyl-CoA synthetase 3 decreases
hepatic de novo fatty acid synthesis through decreased transcriptional activity. J Biol Chem 284:
Author Manuscript

30474–30483, 2009. [PubMed: 19737935]


35. Buqué X, Martínez MJ, Cano A, Miquilena-Colina ME, García-Monzón C, Aspichueta P, and
Ochoa B. A subset of dysregulated metabolic and survival genes is associated with severity of
hepatic steatosis in obese Zucker rats. J Lipid Res 51: 500–513, 2010. [PubMed: 19783528]
36. Burrier RE, Manson CR, and Brecher P. Binding of acyl-CoA to liver fatty acid binding protein:
effect on acyl-CoA synthesis. Biochim Biophys Acta 919: 221–230, 1987. [PubMed: 3593745]
37. Cao G, Konrad RJ, Li SD, and Hammond C. Glycerolipid acyltransferases in triglyceride
metabolism and energy homeostasis-potential as drug targets. Endocrine, metabolic & immune
disorders drug targets 12: 197–206, 2012.
38. Cao JS, Li JL, Li DM, Tobin JF, and Gimeno RE. Molecular identification of microsomal acyl-
CoA : glycerol-3-phosphate acyltransferase, a key enzyme in de novo triacylglycerol synthesis.
Proc Natl Acad Sci U S A 103: 19695–19700, 2006. [PubMed: 17170135]
39. Carr RM, Patel RT, Rao V, Dhir R, Graham MJ, Crooke RM, and Ahima RS. Reduction of TIP47
improves hepatic steatosis and glucose homeostasis in mice. American Journal of Physiology
Author Manuscript

Regulatory, integrative and comparative physiology 302: R996–R1003, 2012.


40. Carr RM, and Ahima RS. Pathophysiology of lipid droplet proteins in liver diseases. Exp Cell Res
340: 187–192, 2016. [PubMed: 26515554]
41. Cases S, Smith SJ, Zheng YW, Myers HM, Lear SR, Sande E, Novak S, Collins C, Welch CB,
Lusis AJ, Erickson SK, and Farese RV. Identification of a gene encoding an acyl CoA :
diacylglycerol acyltransferase, a key enzyme in triacylglycerol synthesis. Proc Natl Acad Sci U S
A 95: 13018–13023, 1998. [PubMed: 9789033]
42. Cases S, Stone SJ, Zhou P, Yen E, Tow B, Lardizabal KD, Voelker T, and Farese RV. Cloning of
DGAT2, a second mammalian diacylglycerol acyltransferase, and related family members. J Biol
Chem 276: 38870–38876, 2001. [PubMed: 11481335]
43. Caviglia JM, Gayet C, Ota T, Hernandez-Ono A, Conlon DM, Jiang H, Fisher EA, and Ginsberg
HN. Different fatty acids inhibit apoB100 secretion by different pathways: unique roles for ER
stress, ceramide, and autophagy. J Lipid Res 52: 1636–1651, 2011 10.1194/jlr.M016931 [PubMed:
21719579]
44. Chakravarthy MV, Lodhi IJ, Yin L, Malapaka RR, Xu HE, Turk J, and Semenkovich CF.
Author Manuscript

Identification of a physiologically relevant endogenous ligand for PPARalpha in liver. Cell 138:
476–488, 2009 10.1016/j.cell.2009.05.036 [PubMed: 19646743]
45. Chakravarthy MV, Pan Z, Zhu Y, Tordjman K, Schneider JG, Coleman T, Turk J, and Semenkovich
CF. “New” hepatic fat activates PPARalpha to maintain glucose, lipid, and cholesterol
homeostasis. Cell Metabolism 1: 309–322, 2005. [PubMed: 16054078]
46. Chang BH, Li L, Paul A, Taniguchi S, Nannegari V, Heird WC, and Chan L. Protection against
fatty liver but normal adipogenesis in mice lacking adipose differentiation-related protein. Mol
Cell Biol 26: 1063–1076, 2006. [PubMed: 16428458]
47. Chen HC, Smith SJ, Ladha Z, Jensen DR, Ferreira LD, Pulawa LK, McGuire JG, Pitas RE, Eckel
RH, and Farese RV. Increased insulin and leptin sensitivity in mice lacking acyl CoA :
diacylglycerol acyltransferase 1. J Clin Investig 109: 1049–1055, 2002. [PubMed: 11956242]
48. Chen MT, Kaufman LN, Spennetta T, and Shrago E. Effects of high fat-feeding to rats on the
interrelationship of body weight, plasma insulin, and fatty acyl-coenzyme A esters in liver and
skeletal muscle. Metabolism-Clinical and Experimental 41: 564–569, 1992. [PubMed: 1588840]
Author Manuscript

49. Chen YQ, Kuo MS, Li SY, Bui HH, Peake DA, Sanders PE, Thibodeaux SJ, Chu SY, Qian YW,
Zhao Y, Bredt DS, Moller DE, Konrad RJ, Beigneux AP, Young SG, and Cao GQ. AGPAT6 is a
novel microsomal glycerol-3-phosphate acyltransferase. J Biol Chem 283: 10048–10057, 2008.
[PubMed: 18238778]
50. Chen ZJ, Gropler MC, Norris J, Lawrence JC, Harris TE, and Finck BN. Alterations in hepatic
metabolism in fld mice reveal a role for lipin 1 in regulating VLDL-triacylglyceride secretion.
Arterioscler Thromb Vasc Biol 28: 1738–1744, 2008. [PubMed: 18669885]
51. Chirala SS, Chang H, Matzuk M, Abu-Elheiga L, Mao J, Mahon K, Finegold M, and Wakil SJ.
Fatty acid synthesis is essential in embryonic development: fatty acid synthase null mutants and

Compr Physiol. Author manuscript; available in PMC 2019 February 15.


Alves-Bezerra and Cohen Page 21

most of the heterozygotes die in utero. Proc Natl Acad Sci U S A 100: 6358–6363, 2003.
[PubMed: 12738878]
Author Manuscript

52. Coburn CT, Knapp FF, Febbraio M, Beets AL, Silverstein RL, and Abumrad NA. Defective uptake
and utilization of long chain fatty acids in muscle and adipose tissues of CD36 knockout mice. J
Biol Chem 275: 32523–32529, 2000. [PubMed: 10913136]
53. Cohen DE. New players on the metabolic stage: How do you like Them Acots? Adipocyte 2: 3–6,
2013. [PubMed: 23700546]
54. Cohen DE, and Fisher EA. Lipoprotein metabolism, dyslipidemia, and nonalcoholic fatty liver
disease. Semin Liver Dis 33: 380–388, 2013 10.1055/s-0033-1358519 [PubMed: 24222095]
55. Coleman RA, and Lee DP. Enzymes of triacylglycerol synthesis and their regulation. Prog Lipid
Res 43: 134–176, 2004. [PubMed: 14654091]
56. Coleman RA, Lewin TM, and Muoio DM. Physiological and nutritional regulation of enzymes of
triacylglycerol synthesis. Annu Rev Nutr 20: 77–103, 2000. [PubMed: 10940327]
57. Cooper DE, Young PA, Klett EL, and Coleman RA. Physiological consequences of
compartmentalized acyl-CoA metabolism. J Biol Chem 290: 20023–20031, 2015. [PubMed:
26124277]
Author Manuscript

58. Cuchel M, and Rader DJ. Microsomal transfer protein inhibition in humans. Curr Opin Lipidol 24:
246–250, 2013 10.1097/MOL.0b013e32836139df [PubMed: 23594709]
59. Czaja MJ, Ding WX, Donohue TM, Jr, Friedman SL, Kim JS, Komatsu M, Lemasters JJ, Lemoine
A, Lin JD, Ou JH, Perlmutter DH, Randall G, Ray RB, Tsung A, and Yin XM. Functions of
autophagy in normal and diseased liver. Autophagy 9: 1131–1158, 2013 10.4161/auto.25063
[PubMed: 23774882]
60. Dahlman I, Kaaman M, Jiao H, Kere J, Laakso M, and Arner P. The CIDEA gene V115F
polymorphism is associated with obesity in Swedish subjects. Diabetes 54: 3032–3034, 2005.
[PubMed: 16186410]
61. Day CP, and James OF. Steatohepatitis: a tale of two ““hits””? Gastroenterology 114: 842–845,
1998. [PubMed: 9547102]
62. Declercq PE, Haagsman HP, Van Veldhoven P, Debeer LJ, Van Golde LM, and Mannaerts GP. Rat
liver dihydroxyacetone-phosphate acyltransferases and their contribution to glycerolipid synthesis.
J Biol Chem 259: 9064–9075, 1984. [PubMed: 6746639]
Author Manuscript

63. den Boer M, Voshol PJ, Kuipers F, Havekes LM, and Romijn JA. Hepatic steatosis: a mediator of
the metabolic syndrome. Lessons from animal models. Arterioscler Thromb Vasc Biol 24: 644–
649, 2004. [PubMed: 14715643]
64. Denechaud PD, Dentin R, Girard J, and Postic C. Role of ChREBP in hepatic steatosis and insulin
resistance. FEBS Journal 582: 68–73, 2008.
65. Dentin R, Girard J, and Postic C. Carbohydrate responsive element binding protein (ChREBP) and
sterol regulatory element binding protein-1c (SREBP-1c): two key regulators of glucose
metabolism and lipid synthesis in liver. Biochimie 87: 81–86, 2005. [PubMed: 15733741]
66. Ding Y, Wu Y, Zeng R, and Liao K. Proteomic profiling of lipid droplet-associated proteins in
primary adipocytes of normal and obese mouse. Acta Biochimica et Biophysica Sinica 44: 394–
406, 2012. [PubMed: 22343379]
67. Diraison F, Dusserre E, Vidal H, Sothier M, and Beylot M. Increased hepatic lipogenesis but
decreased expression of lipogenic gene in adipose tissue in human obesity. American journal of
physiology Endocrinology and metabolism 282: E46–E51, 2002. [PubMed: 11739082]
68. Diraison F, Moulin P, and Beylot M. Contribution of hepatic de novo lipogenesis and
Author Manuscript

reesterification of plasma non-esterified fatty acids to plasma triglyceride synthesis during non-
alcoholic fatty liver disease. Diabetes Metab 29: 478–485, 2003. [PubMed: 14631324]
69. DiRusso CC, Li H, Darwis D, Watkins PA, Berger J, and Black PN. Comparative biochemical
studies of the murine fatty acid transport proteins (FATP) expressed in yeast. J Biol Chem 280:
16829–16837, 2005. [PubMed: 15699031]
70. Doege H, Baillie RA, Ortegon AM, Tsang B, Wu QW, Punreddy S, Hirsch D, Watson N, Gimeno
RE, and Stahl A. Targeted deletion of FATP5 reveals multiple functions in liver metabolism:
Alterations in hepatic lipid Homeostasis. Gastroenterology 130: 1245–1258, 2006. [PubMed:
16618416]

Compr Physiol. Author manuscript; available in PMC 2019 February 15.


Alves-Bezerra and Cohen Page 22

71. Donkor J, Sariahmetoglu M, Dewald J, Brindley DN, and Reue K. Three mammalian lipins act as
phosphatidate phosphatases with distinct tissue expression patterns. J Biol Chem 282: 3450–3457,
Author Manuscript

2007. [PubMed: 17158099]


72. Donnelly KL, Smith CI, Schwarzenberg SJ, Jessurun J, Boldt MD, and Parks EJ. Sources of fatty
acids stored in liver and secreted via lipoproteins in patients with nonalcoholic fatty liver disease. J
Clin Investig 115: 1343–1351, 2005. [PubMed: 15864352]
73. Dowhan W Molecular basis for membrane phospholipid diversity: why are there so many lipids?
Annu Rev Biochem 66: 199–232, 1997. [PubMed: 9242906]
74. Eberhardt C, Gray PW, and Tjoelker LW. cDNA cloning, expression and chromosomal localization
of two human lysophosphatidic acid acyltransferases In: Eicosanoids and Other Bioactive Lipids in
Cancer, Inflammation, and Radiation Injury, 4 New York, NY: Springer US, 1999, p. 351–356.
75. Eberhardt C, Gray PW, and Tjoelker LW. Human lysophosphatidic acid acyltransferase - cDNA
cloning, expression, and localization to chromosome 9q34.3. J Biol Chem 272: 20299–20305,
1997. [PubMed: 9242711]
76. Eberlé D, Hegarty B, Bossard P, Ferré P, and Foufelle F. SREBP transcription factors: master
regulators of lipid homeostasis. Biochimie 86: 839–848, 2004. [PubMed: 15589694]
Author Manuscript

77. Ehehalt R, Fullekrug J, Pohl J, Ring A, Herrmann T, and Stremmel W. Translocation of long chain
fatty acids across the plasma membrane - lipid rafts and fatty acid transport proteins. Mol Cell
Biochem 284: 135–140, 2006. [PubMed: 16477381]
78. Eisinger K, Liebisch G, Schmitz G, Aslanidis C, Krautbauer S, and Buechler C. Lipidomic analysis
of serum from high fat diet induced obese mice. Int J Mol Sci 15: 2991–3002, 2014 10.3390/
ijms15022991 [PubMed: 24562328]
79. Ellis JM, Bowman CE, and Wolfgang MJ. Metabolic and tissue-specific regulation of acyl-CoA
metabolism. PLos One 10: e0116587, 2015. [PubMed: 25760036]
80. Ersoy BA, Tarun A, D’Aquino K, Hancer NJ, Ukomadu C, White MF, Michel T, Manning BD, and
Cohen DE. Phosphatidylcholine transfer protein interacts with thioesterase superfamily member 2
to attenuate insulin signaling. Science Signaling 6: ra64, 2013. [PubMed: 23901139]
81. Fabbrini E, Mohammed BS, Magkos F, Korenblat KM, Patterson BW, and Klein S. Alterations in
adipose tissue and hepatic lipid kinetics in obese men and women with nonalcoholic fatty liver
disease. Gastroenterology 134: 424–431, 2008. [PubMed: 18242210]
Author Manuscript

82. Fabbrini E, Sullivan S, and Klein S. Obesity and nonalcoholic fatty liver disease: biochemical,
metabolic, and clinical implications. Hepatology 51: 679–689, 2010 10.1002/hep.23280 [PubMed:
20041406]
83. Faergeman NJ, and Knudsen J. Role of long-chain fatty acyl-CoA esters in the regulation of
metabolism and in cell signalling. Biochem J 323: 1–12, 1997. [PubMed: 9173866]
84. Faergeman NJ, Sigurskjold BW, Kragelund BB, Andersen KV, and Knudsen J. Thermodynamics of
ligand binding to acyl-coenzyme a binding protein studied by titration calorimetry. Biochemistry
35: 14118–14126, 1996. [PubMed: 8916897]
85. Falcon A, Doege H, Fluitt A, Tsang B, Watson N, Kay MA, and Stahl A. FATP2 is a hepatic fatty
acid transporter and peroxisomal very long-chain acyl-CoA synthetase. American Journal of
Physiology Endocrinology and Metabolism 299: E384–E393, 2010. [PubMed: 20530735]
86. Fan CY, Pan J, Usuda N, Yeldandi AV, Rao MS, and Reddy JK. Steatohepatitis, spontaneous
peroxisome proliferation and liver tumors in mice lacking peroxisomal fatty acyl-CoA oxidase.
Implications for peroxisome proliferator-activated receptor alpha natural ligand metabolism. J Biol
Chem 273: 15639–15645, 1998. [PubMed: 9624157]
Author Manuscript

87. Fernández MA, Albor C, Ingelmo-Torres M, Nixon SJ, Ferguson C, Kurzchalia T, Tebar F, Enrich
C, Parton RG, and Pol A. Caveolin-1 is essential for liver regeneration. Science 313: 1628–1132,
2006. [PubMed: 16973879]
88. Ferré P, and Foufelle F. Hepatic steatosis: a role for de novo lipogenesis and the transcription factor
SREBP-1c. Diabetes Obesity & Metabolism 12: 83–92, 2010.
89. Filhoulaud G, Guilmeau S, Dentin R, Girard J, and Postic C. Novel insights into ChREBP
regulation and function. Trends Endocrinol Metab 24: 257–268, 2013 10.1016/j.tem.2013.01.003
[PubMed: 23597489]

Compr Physiol. Author manuscript; available in PMC 2019 February 15.


Alves-Bezerra and Cohen Page 23

90. Finck BN, Gropler MC, Chen Z, Leone TC, Croce MA, Harris TE, Lawrence JCJ, and Kelly DP.
Lipin 1 is an inducible amplifier of the hepatic PGC-1alpha/PPARalpha regulatory pathway. Cell
Author Manuscript

Metabolism 4: 199–210, 2006. [PubMed: 16950137]


91. Fu S, Yang L, Li P, Hofmann O, Dicker L, Hide W, Lin X, Watkins SM, Ivanov AR, and
Hotamisligil GS. Aberrant lipid metabolism disrupts calcium homeostasis causing liver
endoplasmic reticulum stress in obesity. Nature 473: 528–531, 2011. [PubMed: 21532591]
92. Fuchs M, Lammert F, Wang DQ, Paigen B, Carey MC, and Cohen DE. Sterol carrier protein 2
participates in hypersecretion of biliary cholesterol during gallstone formation in genetically
gallstone-susceptible mice. Biochem J 336 ( Pt 1): 33–37, 1998. [PubMed: 9806881]
93. Fujii H, Ikura Y, Arimoto J, Sugioka K, Iezzoni JC, Park SH, Naruko T, Itabe H, Kawada N,
Caldwell SH, and Ueda M. Expression of perilipin and adipophilin in nonalcoholic fatty liver
disease; relevance to oxidative injury and hepatocyte ballooning. Journal of Atherosclerosis and
Thrombosis 16: 893–901, 2009. [PubMed: 20032580]
94. Fujimoto Y, Itabe H, Sakai J, Makita M, Noda J, Mori M, Higashi Y, Kojima S, and Takano T.
Identification of major proteins in the lipid droplet-enriched fraction isolated from the human
hepatocyte cell line HuH7. Biochimica Et Biophysica Acta-Molecular Cell Research 1644: 47–59,
Author Manuscript

2004.
95. Fukuda N, and Ontko JA. Interactions between fatty acid synthesis, oxidation, and esterification in
the production of triglyceride-rich lipoproteins by the liver. J Lipid Res 25: 831–842, 1984.
[PubMed: 6491528]
96. Fukushima M, Enjoji M, Kohjima M, Sugimoto R, Ohta S, Kotoh K, Kuniyoshi M, Kobayashi K,
Imamura M, Inoguchi T, Nakamuta M, and Nawata H. Adipose differentiation related protein
induces lipid accumulation and lipid droplet formation in hepatic stellate cells. In vitro cellular and
developmental biology Animal 41: 321–324, 2005. [PubMed: 16448220]
97. Gale SE, Frolov A, Han X, Bickel PE, Cao L, Bowcock A, Schaffer JE, and Ory DS. A regulatory
role for 1-acylglycerol-3-phosphate-O-acyltransferase 2 in adipocyte differentiation. J Biol Chem
281: 11082–11089, 2006. [PubMed: 16495223]
98. Ganji SH, Tavintharan S, Zhu DM, Xing YD, Kamanna VS, and Kashyap ML. Niacin
noncompetitively inhibits DGAT2 but not DGAT1 activity in HepG2 cells. J Lipid Res 45: 1835–
1845, 2004. [PubMed: 15258194]
99. Goh VJ, and Silver DL. The lipid droplet as a potential therapeutic target in NAFLD. Semin Liver
Author Manuscript

Dis 33: 312–320, 2013. [PubMed: 24222089]


100. Gong J, Sun Z, and Li P. CIDE proteins and metabolic disorders. Curr Opin Lipidol 20: 121–126,
2009. [PubMed: 19276890]
101. Gordon DA, Wetterau JR, and Gregg RE. Microsomal triglyceride transfer protein: a protein
complex required for the assembly of lipoprotein particles. Trends Cell Biol 5: 317–321, 1995.
[PubMed: 14732096]
102. Gordon GB. Saturated free fatty acid toxicity. II. Lipid accumulation, ultrastructural alterations,
and toxicity in mammalian cells in culture. . Exp Mol Pathol 27: 1977.
103. Gossett RE, Frolov AA, Roths JB, Behnke WD, Kier AB, and Schroeder F. Acyl-CoA binding
proteins: Multiplicity and function. Lipids 31: 895–918, 1996. [PubMed: 8882970]
104. Greco D, Kotronen A, Westerbacka J, Puig O, Arkkila P, Kiviluoto T, Laitinen S, Kolak M, Fisher
RM, Hamsten A, Auvinen P, and Yki-Järvinen H. Gene expression in human NAFLD. American
journal of physiology Gastrointestinal and liver physiology 294: G1281–G1287, 2008. [PubMed:
18388185]
Author Manuscript

105. Greenberg AS, Coleman RA, Kraemer FB, McManaman JL, Obin MS, Puri V, Yan QW, Miyoshi
H, and Mashek DG. The role of lipid droplets in metabolic disease in rodents and humans. J Clin
Investig 121: 2102–2110, 2011. [PubMed: 21633178]
106. Grevengoed TJ, Klett EL, and Coleman RA. Acyl-CoA metabolism and partitioning. Annu Rev
Nutr 34: 1–30, 2014. [PubMed: 24819326]
107. Gropler MC, Harris TE, Hall AM, Wolins NE, Gross RW, Han XL, Chen ZJ, and Finck BN. Lipin
2 Is a liver-enriched phosphatidate phosphohydrolase enzyme that is dynamically regulated by
fasting and obesity in Mice. J Biol Chem 284: 6763–6772, 2009. [PubMed: 19136718]

Compr Physiol. Author manuscript; available in PMC 2019 February 15.


Alves-Bezerra and Cohen Page 24

108. Gross DA, Zhan C, and Silver DL. Direct binding of triglyceride to fat storage-inducing
transmembrane proteins 1 and 2 is important for lipid droplet formation. Proc Natl Acad Sci U S
Author Manuscript

A 108: 19581–19586, 2011 10.1073/pnas.1110817108 [PubMed: 22106267]


109. Guo F, Ma Y, Kadegowda AK, Betters JL, Xie P, Liu G, Liu X, Miao H, Ou J, Su X, Zheng Z,
Xue B, Shi H, and Yu L. Deficiency of liver comparative gene identification-58 causes
steatohepatitis and fibrosis in mice. J Lipid Res 54: 2109–2120, 2013. [PubMed: 23733885]
110. Hammond LE, Neschen S, Romanelli AJ, Cline GW, Ilkayeva OR, Shulman GI, Muoio DM, and
Coleman RA. Mitochondrial glycerol-3-phosphate acyltransferase-1 is essential in liver for the
metabolism of excess acyl-CoAs. J Biol Chem 280: 25629–25636, 2005 10.1074/
jbc.M503181200 [PubMed: 15878874]
111. Hashimoto T, Cook WS, Qi C, Yeldandi AV, Reddy JK, and Rao MS. Defect in peroxisome
proliferator-activated receptor alpha-inducible fatty acid oxidation determines the severity of
hepatic steatosis in response to fasting. J Biol Chem 275: 28918–28928, 2000. [PubMed:
10844002]
112. Havel RJ. Postprandial hyperlipidemia and remnant lipoproteins. Curr Opin Lipidol 5: 102–109,
1994. [PubMed: 8044412]
Author Manuscript

113. Heacock AM, and Agranoff BW. CDP-diacylglycerol synthase from mammalian tissues. Biochim
Biophys Acta 1348: 166–172, 1997. [PubMed: 9370329]
114. Hellerstein MK, Schwarz JM, and Neese RA. Regulation of hepatic de novo lipogenesis in
humans. Annu Rev Nutr 16: 523–557, 1996. [PubMed: 8839937]
115. Higuchi N, Kato M, Shundo Y, Tajiri H, Tanaka M, Yamashita N, Kohjima M, Kotoh K,
Nakamuta M, Takayanagi R, and Enjoji M. Liver X receptor in cooperation with SREBP-1c is a
major lipid synthesis regulator in nonalcoholic fatty liver disease. Hepatol Res 38: 1122–1129,
2008 10.1111/j.1872-034X.2008.00382.x [PubMed: 18684130]
116. Horton JD, Goldstein JL, and Brown MS. SREBPs: activators of the complete program of
cholesterol and fatty acid synthesis in the liver. J Clin Investig 109: 1125–1131, 2002. [PubMed:
11994399]
117. Hostetler HA, McIntosh AL, Atshaves BP, Storey SM, Payne HR, Kier AB, and Schroeder F. L-
FABP directly interacts with PPAR alpha in cultured primary hepatocytes. J Lipid Res 50: 1663–
1675, 2009. [PubMed: 19289416]
118. Huang H, Atshaves BP, Frolov A, Kier AB, and Schroeder F. Acyl-coenzyme A binding protein
Author Manuscript

expression alters liver fatty acyl-coenzyme A metabolism. Biochemistry 44: 10282–10297, 2005.
[PubMed: 16042405]
119. Huang H, Starodub O, McIntosh A, Kier AB, Schroeder F, and 9;277(32):29139–51. JBCA. Liver
fatty acid-binding protein targets fatty acids to the nucleus. Real time confocal and multiphoton
fluorescence imaging in living cells. J Biol Chem 277: 29139–29151, 2002. [PubMed:
12023965]
120. Hudgins LC, Hellerstein M, Seidman C, Neese R, Diakun J, and Hirsch J. Human fatty acid
synthesis is stimulated by a eucaloric low fat, high carbohydrate diet. J Clin Investig 97: 2081–
2091, 1996. [PubMed: 8621798]
121. Hyslop PA, York DA, and Corina DL. Changes in the composition and fluidity of membranes in
obese (ob/ob) mice: a study of hepatic microsomal NADPH-cytochrome P450 oxidoreductase
activity. Int J Obes 6: 279–289, 1982. [PubMed: 6811450]
122. Iqbal J, and Hussain MM. Intestinal lipid absorption. American Journal of Physiology
Endocrinology and Metabolism 296: E1183–1194, 2009. [PubMed: 19158321]
Author Manuscript

123. James O, and Day C. Non-alcoholic steatohepatitis: another disease of affluence. Lancet 353:
1634–1636, 1999. [PubMed: 10335777]
124. Jenkins CM, Mancuso DJ, Yan W, Sims HF, Gibson B, and Gross RW. Identification, cloning,
expression, and purification of three novel human calcium-independent phospholipase A2 family
members possessing triacylglycerol lipase and acylglycerol transacylase activities. J Biol Chem
279: 48968–48975, 2004. [PubMed: 15364929]
125. Jensen-Urstad AP, and Semenkovich CF. Fatty acid synthase and liver triglyceride metabolism:
housekeeper or messenger? Biochim Biophys Acta 1821: 747–753, 2012 10.1016/j.bbalip.
2011.09.017 [PubMed: 22009142]

Compr Physiol. Author manuscript; available in PMC 2019 February 15.


Alves-Bezerra and Cohen Page 25

126. Jo H, Choe SS, Shin KC, Jang H, Lee JH, Seong JK, Back SH, and Kim JB. Endoplasmic
reticulum stress induces hepatic steatosis via increased expression of the hepatic very low-density
Author Manuscript

lipoprotein receptor. Hepatology 57: 1366–1377, 2013 10.1002/hep.26126 [PubMed: 23152128]


127. Jornayvaz FR, and Shulman GI. Diacylglycerol activation of protein kinase Cepsilon and hepatic
insulin resistance. Cell Metab 15: 574–584, 2012 10.1016/j.cmet.2012.03.005 [PubMed:
22560210]
128. Kadereit B, Kumar P, Wang WJ, Miranda D, Snapp EL, Severina N, Torregroza I, Evans T, and
Silver DL. Evolutionarily conserved gene family important for fat storage. Proc Natl Acad Sci U
S A 105: 94–99, 2008 10.1073/pnas.0708579105 [PubMed: 18160536]
129. Kammoun HL, Chabanon H, Hainault I, Luquet S, Magnan C, Koike T, Ferre P, and Foufelle F.
GRP78 expression inhibits insulin and ER stress-induced SREBP-1c activation and reduces
hepatic steatosis in mice. J Clin Investig 119: 1201–1215, 2009 10.1172/JCI37007 [PubMed:
19363290]
130. Kang HW, Niepel MW, Han S, Kawano Y, and Cohen DE. Thioesterase superfamily member 2/
acyl-CoA thioesterase 13 (Them2/Acot13) regulates hepatic lipid and glucose metabolism.
FASEB J 26: 2209–2221, 2012. [PubMed: 22345407]
Author Manuscript

131. Kantartzis K, Machicao F, Machann J, Schick F, Fritsche A, Haring HU, and Stefan N. The
DGAT2 gene is a candidate for the dissociation between fatty liver and insulin resistance in
humans. Clinical Science 116: 531–537, 2009. [PubMed: 18980578]
132. Kato M, Higuchi N, and Enjoji M. Reduced hepatic expression of adipose tissue triglyceride
lipase and CGI-58 may contribute to the development of non-alcoholic fatty liver disease in
patients with insulin resistance. Scand J Gastroenterol 43: 1018–1019, 2008. [PubMed:
19086170]
133. Kaushik S, and Cuervo AM. Chaperone-mediated autophagy: a unique way to enter the lysosome
world. Trends Cell Biol 22: 407–417, 2012 10.1016/j.tcb.2012.05.006 [PubMed: 22748206]
134. Kaushik S, and Cuervo AM. Degradation of lipid droplet-associated proteins by chaperone-
mediated autophagy facilitates lipolysis. Nat Cell Biol 17: 759–770, 2015 10.1038/ncb3166
[PubMed: 25961502]
135. Kawano Y, and Cohen DE. Mechanisms of hepatic triglyceride accumulation in non-alcoholic
fatty liver disease. J Gastroenterol 48: 434–441, 2013. [PubMed: 23397118]
136. Kawano Y, Nishiumi S, Saito M, Yano Y, Azuma T, and Yoshida M. Identification of lipid species
Author Manuscript

linked to the progression of non-alcoholic fatty liver disease. Curr Drug Targets 16: 1293–1300,
2015. [PubMed: 25850622]
137. Kelder B, Boyce K, Kriete A, Clark R, Berryman DE, Nagatomi S, List EO, Braughler M, and
Kopchick JJ. CIDE-A is expressed in liver of old mice and in type 2 diabetic mouse liver
exhibiting steatosis. Comparative Hepatology 6: 4, 2007. [PubMed: 17472743]
138. Khan SA, Wollaston-Hayden EE, Markowski TW, Higgins L, and Mashek DG. Quantitative
analysis of the murine lipid droplet-associated proteome during diet-induced hepatic steatosis. J
Lipid Res 56: 2260–2272, 2015 10.1194/jlr.M056812 [PubMed: 26416795]
139. Kim KH. Regulation of mammalian acetyl-coenzyme A carboxylase. Annu Rev Nutr 17: 77–99,
1997. [PubMed: 9240920]
140. Kimmel AR, and Sztalryd C. Perilipin 5, a lipid droplet protein adapted to mitochondrial energy
utilization. Curr Opin Lipidol 25: 110–117, 2014. [PubMed: 24535284]
141. Kimmel AR, and Sztalryd C. The Perilipins: Major Cytosolic Lipid Droplet-Associated Proteins
and Their Roles in Cellular Lipid Storage, Mobilization, and Systemic Homeostasis. Annu Rev
Author Manuscript

Nutr 36: 471–509, 2016. [PubMed: 27431369]


142. Kirkby B, Roman N, Kobe B, Kellie S, and Forwood JK. Functional and structural properties of
mammalian acyl-coenzyme A thioesterases. Prog Lipid Res 49: 366–377, 2010. [PubMed:
20470824]
143. Knight BL, Hebbachi A, Hauton D, Brown AM, Wiggins D, Patel DD, and Gibbons GF. A role
for PPARalpha in the control of SREBP activity and lipid synthesis in the liver. Biochem J 389:
413–421, 2005. [PubMed: 15777286]
144. Knudsen J Acyl-Coa-Binding Protein (ACBP) and Its Relation to Fatty Acid-Binding Protein
(FABP) - an Overview. Mol Cell Biochem 98: 217–223, 1990. [PubMed: 2266962]

Compr Physiol. Author manuscript; available in PMC 2019 February 15.


Alves-Bezerra and Cohen Page 26

145. Koga H, Kaushik S, and Cuervo AM. Altered lipid content inhibits autophagic vesicular fusion.
FASEB J 24: 3052–3065, 2010 10.1096/fj.09-144519 [PubMed: 20375270]
Author Manuscript

146. Kohjima M, Enjoji M, Higuchi N, Kato M, Kotoh K, Yoshimo T, Fujino T, Yada M, Yada R,
Harada N, Takayanagi R, and Nakamuta M. Re-evaluation of fatty acid metabolism-related gene
expression in nonalcoholic fatty liver disease. Int J Mol Med 20: 351–358, 2007. [PubMed:
17671740]
147. Kozlitina J, Smagris E, Stender S, Nordestgaard BG, Zhou HH, Tybjærg-Hansen A, Vogt TF,
Hobbs HH, and Cohen JC. Exome-wide association study identifies a TM6SF2 variant that
confers susceptibility to nonalcoholic fatty liver disease. Nat Genet 46: 352–356, 2014. [PubMed:
24531328]
148. Krammer J, Digel M, Ehehalt F, Stremmel W, Füllekrug J, and Ehehalt R. Overexpression of
CD36 and acyl-CoA synthetases FATP2, FATP4 and ACSL1 increases fatty acid uptake in human
hepatoma cells. International Journal of Medical Sciences 8: 599–614, 2011. [PubMed:
22022213]
149. Kreuz S, Schoelch C, Thomas L, Rist W, Rippmann JF, and Neubauer H. Acetyl-CoA
carboxylases 1 and 2 show distinct expression patterns in rats and humans and alterations in
Author Manuscript

obesity and diabetes. Diabetes Metab Res Rev 25: 577–586, 2009 10.1002/dmrr.997 [PubMed:
19618481]
150. Kume K, and Shimizu T. cDNA cloning and expression of murine 1-acyl-sn-glycerol-3-phosphate
acyltransferase. Biochem Biophys Res Commun 237: 663–666, 1997. [PubMed: 9299423]
151. Lafontan M, and Langin D. Lipolysis and lipid mobilization in human adipose tissue. Prog Lipid
Res 48: 275–297, 2009. [PubMed: 19464318]
152. Lass A, Zimmermann R, Haemmerle G, Riederer M, Schoiswohl G, Schweiger M, Kienesberger
P, Strauss JG, Gorkiewicz G, and Zechner R. Adipose triglyceride lipase-mediated lipolysis of
cellular fat stores is activated by CGI-58 and defective in Chanarin-Dorfman syndrome. Cell
Metabolism 3: 309–319, 2006. [PubMed: 16679289]
153. Lavoie JM, and Gauthier MS. Regulation of fat metabolism in the liver: link to non-alcoholic
hepatic steatosis and impact of physical exercise. Cellular and Molecular Life Sciences 63: 1393–
1409, 2006. [PubMed: 16649140]
154. Lee JH, Wada T, Febbraio M, He J, Matsubara T, Lee MJ, Gonzalez FJ, and Xie W. A novel role
for the dioxin receptor in fatty acid metabolism and hepatic steatosis. Gastroenterology 139: 653–
Author Manuscript

663, 2010. [PubMed: 20303349]


155. Lee K, Kerner J, and Hoppel CL. Mitochondrial carnitine palmitoyltransferase 1a (CPT1a) is part
of an outer membrane fatty acid transfer complex. J Biol Chem 286: 25655–25662, 2011.
[PubMed: 21622568]
156. Lee SS, Pineau T, Drago J, Lee EJ, Owens JW, Kroetz DL, Fernandez-Salguero PM, Westphal H,
and Gonzalez FJ. Targeted disruption of the alpha isoform of the peroxisome proliferator-
activated receptor gene in mice results in abolishment of the pleiotropic effects of peroxisome
proliferators. Mol Cell Biol 15: 3012–3022, 1995. [PubMed: 7539101]
157. Lewin TM, Granger DA, Kim JH, and Coleman RA. Regulation of mitochondrial sn-glycerol-3-
phosphate acyltransferase activity: Response to feeding status is unique in various rat tissues and
is discordant with protein expression. Arch Biochem Biophys 396: 119–127, 2001. [PubMed:
11716470]
158. Lewin TM, Wang SL, Nagle CA, Van Horn CG, and Coleman RA. Mitochondrial glycerol-3-
phosphate acyltransferase-1 directs the metabolic fate of exogenous fatty acids in hepatocytes.
Author Manuscript

Am J Physiol Endocrinol Metab 288: E835–E844, 2005. [PubMed: 15598672]


159. Lewis GF, Carpentier A, Adeli K, and Giacca A. Disordered fat storage and mobilization in the
pathogenesis of insulin resistance and type 2 diabetes. Endocr Rev 23: 201–229, 2002. [PubMed:
11943743]
160. Li JZ, Ye J, Xue B, Qi J, Zhang J, Zhou Z, Li Q, Wen Z, and Li P. Cideb regulates diet-induced
obesity, liver steatosis, and insulin sensitivity by controlling lipogenesis and fatty acid oxidation.
Diabetes 56: 2523–2532, 2007. [PubMed: 17646209]
161. Li LO, Ellis JM, Paich HA, Wang SL, Gong N, Altshuller G, Thresher RJ, Koves TR, Watkins
SM, Muoio DM, Cline GW, Shulman GI, and Coleman RA. Liver-specific loss of long chain

Compr Physiol. Author manuscript; available in PMC 2019 February 15.


Alves-Bezerra and Cohen Page 27

acyl-CoA synthetase-1 decreases triacylglycerol synthesis and beta-oxidation and alters


phospholipid fatty acid composition. J Biol Chem 284: 27816–27826, 2009. [PubMed:
Author Manuscript

19648649]
162. Li S, Liu C, Li N, Hao T, Han T, Hill DE, Vidal M, and Lin JD. Genome-wide coactivation
analysis of PGC-1α identifies BAF60a as a regulator of hepatic lipid metabolism. Cell
Metabolism 8: 105–117, 2008. [PubMed: 18680712]
163. Li X, Grundy SM, and Patel SB. Obesity in db and ob animals leads to impaired hepatic very low
density lipoprotein secretion and differential secretion of apolipoprotein B-48 and B-100. J Lipid
Res 38: 1277–1288, 1997. [PubMed: 9254055]
164. Liang JJ, Oelkers P, Guo CY, Chu PC, Dixon JL, Ginsberg HN, and Sturley SL. Overexpression
of human diacylglycerol acyltransferase 1, Acyl-CoA : cholesterol acyltransferase 1, or Acyl-
CoA : cholesterol acyltransferase 2 stimulates secretion of apolipoprotein B-containing
lipoproteins in McA-RH7777 cells. J Biol Chem 279: 44938–44944, 2004. [PubMed: 15308631]
165. Linden D, William-Olsson L, Rhedin M, Asztely AK, Clapham JC, and Schreyer S.
Overexpression of mitochondrial GPAT in rat hepatocytes leads to decreased fatty acid oxidation
and increased glycerolipid biosynthesis. J Lipid Res 45: 1279–1288, 2004. [PubMed: 15102885]
Author Manuscript

166. Lowe ME. The triglyceride lipases of the pancreas. J Lipid Res 43: 2007–2016, 2002. [PubMed:
12454260]
167. Madrigal-Matute J, and Cuervo AM. Regulation of Liver Metabolism by Autophagy.
Gastroenterology 150: 328–339, 2016 10.1053/j.gastro.2015.09.042 [PubMed: 26453774]
168. Magkos F, Su X, Bradley D, Fabbrini E, Conte C, Eagon JC, Varela JE, Brunt EM, Patterson BW,
and Klein S. Intrahepatic diacylglycerol content is associated with hepatic insulin resistance in
obese subjects. Gastroenterology 142: 1444–1446 e1442, 2012 10.1053/j.gastro.2012.03.003
[PubMed: 22425588]
169. Manmontri B, Sariahmetoglu M, Donkor J, Khalil MB, Sundaram M, Yao Z, Reue K, Lehner R,
and Brindley DN. Glucocorticoids and cyclic AMP selectively increase hepatic lipin-1
expression, and insulin acts antagonistically. J Lipid Res 49: 1056–1067, 2008. [PubMed:
18245816]
170. Mao J, DeMayo FJ, Li H, Abu-Elheiga L, Gu Z, Shaikenov TE, Kordari P, Chirala SS, Heird WC,
and Wakil SJ. Liver-specific deletion of acetyl-CoA carboxylase 1 reduces hepatic triglyceride
accumulation without affecting glucose homeostasis. Proc Natl Acad Sci U S A 103: 8552–8557,
Author Manuscript

2006. [PubMed: 16717184]


171. Marchesini G, Brizi M, Bianchi G, Tomassetti S, Bugianesi E, Lenzi M, McCullough AJ, Natale
S, Forlani G, and Melchionda N. Nonalcoholic fatty liver disease: a feature of the metabolic
syndrome. Diabetes 150: 2001.
172. Martin GG, Huang H, Atshaves BP, Binas B, and Schroeder F. Ablation of the liver fatty acid
binding protein gene decreases fatty acyl CoA binding capacity and alters fatty acyl CoA pool
distribution in mouse liver. Biochemistry 42: 11520–11532, 2003. [PubMed: 14516204]
173. Martin-Sanz P, Hopewell R, and Brindley DN. Long-chain fatty acids and their acyl-CoA esters
cause the translocation of phosphatidate phosphohydrolase from the cytosolic to the microsomal
fraction of rat liver. FEBS Lett 175: 284–288, 1984. [PubMed: 6090213]
174. Mashek DG, McKenzie MA, Van Horn CG, and Coleman RA. Rat long chain acyl-CoA
synthetase 5 increases fatty acid uptake and partitioning to cellular triacylglycerol in McArdle-
RH7777 cells. J Biol Chem 281: 945–950, 2006. [PubMed: 16263710]
175. Matsusue K, Haluzik M, Lambert G, Yim SH, Gavrilova O, Ward JM, Brewer BJ, Reitman ML,
Author Manuscript

and Gonzalez FJ. Liver-specific disruption of PPARgamma in leptin-deficient mice improves


fatty liver but aggravates diabetic phenotypes. J Clin Investig 111: 737–747, 2003. [PubMed:
12618528]
176. Matsusue K, Kusakabe T, Noguchi T, Takiguchi S, Suzuki T, Yamano S, and Gonzalez FJ.
Hepatic steatosis in leptin-deficient mice is promoted by the PPARgamma target gene Fsp27.
Cell Metabolism 7: 302–311, 2008. [PubMed: 18396136]
177. McArthur MJ, Atshaves BP, Frolov A, Foxworth WD, Kier AB, and Schroeder F. Cellular uptake
and intracellular trafficking of long chain fatty acids. J Lipid Res 40: 1371–1383, 1999.
[PubMed: 10428973]

Compr Physiol. Author manuscript; available in PMC 2019 February 15.


Alves-Bezerra and Cohen Page 28

178. McFie PJ, Banman SL, Kary S, and Stone SJ. Murine diacylglycerol acyltransferase-2 (DGAT2)
can catalyse triacylglycerol synthesis and promote lipid droplet formation independent of its
Author Manuscript

localization to the endoplasmic reticulum. J Biol Chem 286: 28235–28246, 2011. [PubMed:
21680734]
179. Meegalla RL, Billheimer JT, and Cheng D. Concerted elevation of acyl-coenzyme A :
diacylglycerol acyltransferase (DGAT) activity through independent stimulation of mRNA
expression of DGAT1 and DGAT2 by carbohydrate and insulin. Biochem Biophys Res Commun
298: 317–323, 2002. [PubMed: 12413942]
180. Memon RA, Fuller J, Moser AH, Smith PJ, Grunfeld C, and Feingold KR. Regulation of putative
fatty acid transporters and acyl-CoA synthetase in liver and adipose tissue in ob/ob mice.
Diabetes 48: 121–127, 1999. [PubMed: 9892232]
181. Miquilena-Colina ME, Lima-Cabello E, Sánchez-Campos S, García-Mediavilla MV, Fernández-
Bermejo M, Lozano-Rodríguez T, Vargas-Castrillón J, Buqué X, Ochoa B, Aspichueta P,
González-Gallego J, and García-Monzón C. Hepatic fatty acid translocase CD36 upregulation is
associated with insulin resistance, hyperinsulinaemia and increased steatosis in non-alcoholic
steatohepatitis and chronic hepatitis C. Gut 60: 1394–1402, 2011. [PubMed: 21270117]
Author Manuscript

182. Moffat C, Bhatia L, Nguyen T, Lynch P, Wang M, Wang D, Ilkayeva OR, Han X, Hirschey MD,
Claypool SM, and Seifert EL. Acyl-CoA thioesterase-2 facilitates mitochondrial fatty acid
oxidation in the liver. J Lipid Res 55: 2458–2470, 2014. [PubMed: 25114170]
183. Monetti M, Levin MC, Watt MJ, Sajan MP, Marmor S, Hubbard BK, Stevens RD, Bain JR,
Newgard CB, Farese RV, Hevener AL, and Farese RV. Dissociation of hepatic steatosis and
insulin resistance in mice overexpressing DGAT in the liver. Cell Metabolism 6: 69–78, 2007.
[PubMed: 17618857]
184. Montagner A, Polizzi A, Fouche E, Ducheix S, Lippi Y, Lasserre F, Barquissau V, Regnier M,
Lukowicz C, Benhamed F, Iroz A, Bertrand-Michel J, Al Saati T, Cano P, Mselli-Lakhal L,
Mithieux G, Rajas F, Lagarrigue S, Pineau T, Loiseau N, Postic C, Langin D, Wahli W, and
Guillou H. Liver PPARalpha is crucial for whole-body fatty acid homeostasis and is protective
against NAFLD. Gut 65: 1202–1214, 2016 10.1136/gutjnl-2015-310798 [PubMed: 26838599]
185. Motomura W, Inoue M, Ohtake T, Takahashi N, Nagamine M, Tanno S, Kohgo Y, and T. O. Up-
regulation of ADRP in fatty liver in human and liver steatosis in mice fed with high fat diet.
Biochem Biophys Res Commun 340: 1111–1118, 2006. [PubMed: 16403437]
Author Manuscript

186. Muoio DM, Seefeld K, Witters LA, and Coleman RA. AMP-activated kinase reciprocally
regulates triacylglycerol synthesis and fatty acid oxidation in liver and muscle: evidence that sn-
glycerol-3-phosphate acyltransferase is a novel target. Biochem J 338: 783–791, 1999. [PubMed:
10051453]
187. Musso G, Gambino R, and Cassader M. Recent insights into hepatic lipid metabolism in non-
alcoholic fatty liver disease (NAFLD). Prog Lipid Res 48: 1–26, 2009 [PubMed: 18824034]
188. Nagle CA, An J, Shiota M, Torres TP, Cline GW, Liu ZX, Wang S, Catlin RL, Shulman GI,
Newgard CB, and Coleman RA. Hepatic overexpression of glycerol-sn-3-phosphate
acyltransferase-1 causes insulin resistance. FASEB J 21: A699–A699, 2007.
189. Nagle CA, Vergnes L, Dejong H, Wang SL, Lewin TM, Reue K, and Coleman RA. Identification
of a novel sn-glycerol-3-phosphate acyltransferase isoform, GPAT4, as the enzyme deficient in
Agpat6(−/−) mice. J Lipid Res 49: 823–831, 2008. [PubMed: 18192653]
190. Najt CP, Senthivinayagam S, Aljazi MB, Fader KA, Olenic SD, Brock JR, Lydic TA, Jones AD,
and Atshaves BP. Liver-specific loss of Perilipin 2 alleviates diet-induced hepatic steatosis,
inflammation, and fibrosis. Am J Physiol Gastrointest Liver Physiol 310: G726–738, 2016
Author Manuscript

10.1152/ajpgi.00436.2015 [PubMed: 26968211]


191. Neschen S, Morino K, Hammond LE, Zhang DY, Liu ZX, Romanelli AJ, Cline GW, Pongratz RL,
Zhang XM, Choi CS, Coleman RA, and Shulman GI. Prevention of hepatic steatosis and hepatic
insulin resistance in mitochondrial acyl-CoA : glycerol-sn-3-phosphate acyltransferase 1
knockout mice. Cell Metabolism 2: 55–65, 2005. [PubMed: 16054099]
192. Newberry EP, Xie Y, Kennedy S, Han X, Buhman KK, Luo J, Gross RW, and Davidson NO.
Decreased hepatic triglyceride accumulation and altered fatty acid uptake in mice with deletion
of the liver fatty acid-binding protein gene. J Biol Chem 278: 51664–51672, 2003. [PubMed:
14534295]

Compr Physiol. Author manuscript; available in PMC 2019 February 15.


Alves-Bezerra and Cohen Page 29

193. Newberry EP, Xie Y, Kennedy SM, Luo J, and Davidson NO. Protection against Western diet-
induced obesity and hepatic steatosis in liver fatty acid-binding protein knockout mice.
Author Manuscript

Hepatology 44: 1191–1205, 2006. [PubMed: 17058218]


194. Ohsaki Y, Cheng J, Suzuki M, Fujita A, and Fujimoto T. Lipid droplets are arrested in the ER
membrane by tight binding of lipidated apolipoprotein B-100. J Cell Sci 121: 2415–2422, 2008.
[PubMed: 18577578]
195. Oishi K, Amagai N, Shirai H, Kadota K, Ohkura N, and Ishida N. Genome-wide expression
analysis reveals 100 adrenal gland-dependent circadian genes in the mouse liver. DNA Research
12: 191–202, 2005. [PubMed: 16303750]
196. Okada K, LeClair KB, Zhang Y, Li Y, Ozdemir C, Krisko TI, Hagen SJ, Betensky RA, Banks AS,
and Cohen DE. Thioesterase superfamily member 1 suppresses cold thermogenesis by limiting
the oxidation of lipid droplet-derived fatty acids in brown adipose tissue. Mol Metab 5: 340–351,
2016 10.1016/j.molmet.2016.02.002 [PubMed: 27110486]
197. Okumura T Role of lipid droplet proteins in liver steatosis. J Physiol Biochem 67: 629–636, 2011.
[PubMed: 21847662]
198. Ong KT, Mashek MT, Bu SY, Greenberg AS, and Mashek DG. Adipose triglyceride lipase is a
Author Manuscript

major hepatic lipase that regulates triacylglycerol turnover and fatty acid signaling and
partitioning. Hepatology 53: 116–126, 2011. [PubMed: 20967758]
199. Ota T, Gayet C, and Ginsberg HN. Inhibition of apolipoprotein B100 secretion by lipid-induced
hepatic endoplasmic reticulum stress in rodents. J Clin Investig 118: 316–332, 2008 10.1172/
JCI32752 [PubMed: 18060040]
200. Owen MR, Corstorphine CC, and Zammit VA. Overt and latent activities of diacylglycerol
acyltransferase in rat liver microsomes: Possible roles in very-low-density lipoprotein
triacylglycerol secretion. Biochem J 323: 17–21, 1997. [PubMed: 9173878]
201. Park C, and Cuervo AM. Selective autophagy: talking with the UPS. Cell Biochem Biophys 67:
3–13, 2013 10.1007/s12013-013-9623-7 [PubMed: 23709310]
202. Pessayre B, and Fromenty B. NASH: a mitochondrial disease. J Hepatol 42: 928–940, 2005.
[PubMed: 15885365]
203. Pohl J, Ring A, and Stremmel W. Uptake of long-chain fatty acids in HepG2 cells involves
caveolae: analysis of a novel pathway. J Lipid Res 43: 1390–1399, 2002. [PubMed: 12235170]
Author Manuscript

204. Pol A, Luetterforst R, Lindsay M, Heino S, Ikonen E, and Parton RG. A caveolin dominant
negative mutant associates with lipid bodies and induces intracellular cholesterol imbalance. J
Cell Biol 152: 1057–1070, 2001. [PubMed: 11238460]
205. Postic C, Dentin R, Denechaud PD, and Girard J. ChREBP, a transcriptional regulator of glucose
and lipid metabolism. Annu Rev Nutr 27: 179–192, 2007 10.1146/annurev.nutr.
27.061406.093618 [PubMed: 17428181]
206. Rao MS, and Reddy JK. Peroxisomal β-oxidation and steatohepatitis. Semin Liver Dis 21: 43–55,
2001. [PubMed: 11296696]
207. Rasmussen JT, Faergeman NJ, Kristiansen K, and Knudsen J. Acyl-CoA-binding protein (ACBP)
can mediate intermembrane acyl-CoA transport and donate acyl-CoA for b-oxidation and
glycerolipid synthesis. Biochem J 299: 165–170, 1994. [PubMed: 8166635]
208. Reid BN, Ables GP, Otlivanchik OA, Schoiswohl G, Zechner R, Blaner WS, Goldberg IJ,
Schwabe RF, Chua SC, Jr., and Huang LS. Hepatic overexpression of hormone-sensitive lipase
and adipose triglyceride lipase promotes fatty acid oxidation, stimulates direct release of free
fatty acids, and ameliorates steatosis. J Biol Chem 283: 13087–13099, 2008 10.1074/
Author Manuscript

jbc.M800533200 [PubMed: 18337240]


209. Rodriguez-Navarro JA, Kaushik S, Koga H, Dall’Armi C, Shui G, Wenk MR, Di Paolo G, and
Cuervo AM. Inhibitory effect of dietary lipids on chaperone-mediated autophagy. Proc Natl Acad
Sci U S A 109: E705–714, 2012 10.1073/pnas.1113036109 [PubMed: 22331875]
210. Romeo S, Kozlitina J, Xing C, Pertsemlidis A, Cox D, Pennacchio LA, Boerwinkle E, Cohen JC,
and Hobbs HH. Genetic variation in PNPLA3 confers susceptibility to nonalcoholic fatty liver
disease. Nat Genet 40: 1461–1465, 2008. [PubMed: 18820647]
211. Rosendal J, Ertbjerg P, and Knudsen J. Characterization of ligand binding to acyl-CoA-binding
protein. Biochem J 290: 321–326, 1993. [PubMed: 7680855]

Compr Physiol. Author manuscript; available in PMC 2019 February 15.


Alves-Bezerra and Cohen Page 30

212. Ruhanen H, Perttilä J, Hölttä-Vuori M, Zhou Y, Yki-Järvinen H, Ikonen E, Käkelä R, and


Olkkonen VM. PNPLA3 mediates hepatocyte triacylglycerol remodeling. J Lipid Res 55: 739–
Author Manuscript

746, 2014. [PubMed: 24511104]


213. Saltiel AR, and Kahn CR. Insulin signalling and the regulation of glucose and lipid metabolism.
Nature 414: 799–806, 2001. [PubMed: 11742412]
214. Samuel VT, and Shulman GI. The pathogenesis of insulin resistance: integrating signaling
pathways and substrate flux. J Clin Invest 126: 12–22, 2016 10.1172/JCI77812 [PubMed:
26727229]
215. Sanyal AJ, Campbell-Sargent C, Mirshahi F, Rizzo WB, Contos MJ, Sterling RK, Luketic VA,
Shiffman ML, and Clore JN. Nonalcoholic steatohepatitis: association of insulin resistance and
mitochondrial abnormalities. Gastroenterology 120: 1183–1192, 2001. [PubMed: 11266382]
216. Sanyal AJ, and Pacana T. A Lipidomic Readout of Disease Progression in A Diet-Induced Mouse
Model of Nonalcoholic Fatty Liver Disease. Trans Am Clin Climatol Assoc 126: 271–288, 2015.
[PubMed: 26330688]
217. Savage DB, Choi CS, Samuel VT, Liu ZX, Zhang D, Wang A, Zhang XM, Cline GW, Yu XX,
Geisler JG, Bhanot S, Monia BP, and Shulman GI. Reversal of diet-induced hepatic steatosis and
Author Manuscript

hepatic insulin resistance by antisense oligonucleotide inhibitors of acetyl-CoA carboxylases 1


and 2. J Clin Investig 116: 817–824, 2006. [PubMed: 16485039]
218. Schadinger SE, Bucher NL, Schreiber BM, and Farmer SR. PPARgamma2 regulates lipogenesis
and lipid accumulation in steatotic hepatocytes. American journal of physiology Endocrinology
and metabolism 288: E1195–E1205, 2005. [PubMed: 15644454]
219. Schneider JL, Suh Y, and Cuervo AM. Deficient chaperone-mediated autophagy in liver leads to
metabolic dysregulation. Cell Metab 20: 417–432, 2014 10.1016/j.cmet.2014.06.009 [PubMed:
25043815]
220. Schoonjans K, Watanabe M, Suzuki H, Mahfoudi A, Krey G, Wahli W, Grimaldi P, Staels B,
Yamamoto T, and Auwerx J. Induction of the acyl-Coenzyme-A synthetase gene by fibrates and
fatty-acids is mediated by a peroxisome proliferator response element in the C-promoter. J Biol
Chem 270: 19269–19276, 1995. [PubMed: 7642600]
221. Schroeder B, Schulze RJ, Weller SG, Sletten AC, Casey CA, and McNiven MA. The small
GTPase Rab7 as a central regulator of hepatocellular lipophagy. Hepatology 61: 1896–1907,
2015 10.1002/hep.27667 [PubMed: 25565581]
Author Manuscript

222. Schroeder F, Petrescu AD, Huang H, Atshaves BP, McIntosh AL, Martin GG, Hostetler HA,
Vespa A, Landrock D, Landrock KK, Payne HR, and Kier AB. Role of fatty acid binding proteins
and long chain fatty acids in modulating nuclear receptors and gene transcription. Lipids 43: 1–
17, 2008. [PubMed: 17882463]
223. Schwarz JM, Linfoot P, Dare D, and Aghajanian K. Hepatic de novo lipogenesis in
normoinsulinemic and hyperinsulinemic subjects consuming high-fat, low-carbohydrate and low-
fat, high-carbohydrate isoenergetic diets. Am J Clin Nutr 77: 43–50, 2003. [PubMed: 12499321]
224. Schwarz JM, Neese RA, Turner S, Dare D, and Hellerstein MK. Short-term alterations in
carbohydrate energy intake in humans. Striking effects on hepatic glucose production, de novo
lipogenesis, lipolysis, and whole-body fuel selection. J Clin Investig 96: 2735–2743, 1995.
[PubMed: 8675642]
225. Seedorf U, Raabe M, Ellinghaus P, Kannenberg F, Fobker M, Engel T, Denis S, Wouters F, Wirtz
KWA, Wanders RJ, Maeda N, and Assmann G. Defective peroxisomal catabolism of branched
fatty acyl coenzyme A in mice lacking the sterol carrier protein-2/sterol carrier protein-x gene
Author Manuscript

function. Genes Dev 12: 1189–1201, 1998. [PubMed: 9553048]


226. Seessle J, Liebisch G, Schmitz G, Stremmel W, and Chamulitrat W. Palmitate activation by fatty
acid transport protein 4 as a model system for hepatocellular apoptosis and steatosis. Biochim
Biophys Acta 1851: 549–565, 2015 10.1016/j.bbalip.2015.01.004 [PubMed: 25603556]
227. Semenkovich CF. Regulation of fatty acid synthase (FAS). Prog Lipid Res 36: 43–53, 1997.
[PubMed: 9373620]
228. Shen LL, Liu H, Peng J, Gan L, Lu L, Zhang Q, Li L, He F, and Jiang Y. Effects of farnesoid X
receptor on the expression of the fatty acid synthetase and hepatic lipase. Mol Biol Rep 38: 553–
559, 2011. [PubMed: 20373033]

Compr Physiol. Author manuscript; available in PMC 2019 February 15.


Alves-Bezerra and Cohen Page 31

229. Shen X, Zhang K, and Kaufman RJ. The unfolded protein response--a stress signaling pathway of
the endoplasmic reticulum. J Chem Neuroanat 28: 79–92, 2004 10.1016/j.jchemneu.2004.02.006
Author Manuscript

[PubMed: 15363493]
230. Shin DH, Paulauskis JD, Moustaïd N, and Sul HS. Transcriptional regulation of p90 with
sequence homology to Escherichia coli glycerol-3-phosphate acyltransferase. J Biol Chem 266:
23834–23839, 1991. [PubMed: 1721057]
231. Shindou H, Hishikawa D, Harayama T, Yuki K, and Shimizu T. Recent progress on acyl CoA:
lysophospholipid acyltransferase research. J Lipid Res 50: S46–S51, 2009. [PubMed: 18931347]
232. Simard JR, Kamp F, and Hamilton JA. Measuring the adsorption of Fatty acids to phospholipid
vesicles by multiple fluorescence probes. Biophys J 94: 4493–4503, 2008. [PubMed: 18296488]
233. Singh R, Kaushik S, Wang Y, Xiang Y, Novak I, Komatsu M, Tanaka K, Cuervo AM, and Czaja
MJ. Autophagy regulates lipid metabolism. Nature 458: 1131–1135, 2009 10.1038/nature07976
[PubMed: 19339967]
234. Siniossoglou S Phospholipid metabolism and nuclear function: roles of the lipin family of
phosphatidic acid phosphatases. Biochim Biophys Acta 1831: 575–581, 2013. [PubMed:
23026159]
Author Manuscript

235. Smagris E, BasuRay S, Li J, Huang Y, Lai KM, Gromada J, Cohen JC, and Hobbs HH.
Pnpla3I148M knockin mice accumulate PNPLA3 on lipid droplets and develop hepatic steatosis.
Hepatology 61: 108–118, 2015. [PubMed: 24917523]
236. Smagris E, Gilyard S, BasuRay S, Cohen JC, and Hobbs HH. Inactivation of Tm6sf2, a Gene
Defective in Fatty Liver Disease, Impairs Lipidation but Not Secretion of Very Low Density
Lipoproteins. J Biol Chem 291: 10659–10676, 2016. [PubMed: 27013658]
237. Smathers RL, and Petersen DR. The human fatty acid-binding protein family: evolutionary
divergences and functions. Human Genomics 5: 170–191, 2011. [PubMed: 21504868]
238. Smith SJ, Cases S, Jensen DR, Chen HC, Sande E, Tow B, Sanan DA, Raber J, Eckel RH, and
Farese RV. Obesity resistance and multiple mechanisms of triglyceride synthesis in mice lacking
Dgat. Nat Genet 25: 87–90, 2000. [PubMed: 10802663]
239. Stahl A, Gimeno RE, Tartaglia LA, and Lodish HF. Fatty acid transport proteins: a current view
of a growing family. Trends Endocrinol Metab 12: 266–273, 2001. [PubMed: 11445444]
240. Stamps AC, Elmore MA, Hill ME, Kelly K, Makda AA, and Finnen MJ. A human cDNA
Author Manuscript

sequence with homology to non-mammalian lysophosphatidic acid acyltransferases. Biochem J


326: 455–461, 1997. [PubMed: 9291118]
241. Stolz A, Ernst A, and Dikic I. Cargo recognition and trafficking in selective autophagy. Nat Cell
Biol 16: 495–501, 2014 10.1038/ncb2979 [PubMed: 24875736]
242. Stone SJ, Levin MC, Zhou P, Han J, Walther TC, and Farese RVJ. The endoplasmic reticulum
enzyme DGAT2 is found in mitochondria-associated membranes and has a mitochondrial
targeting signal that promotes its association with mitochondria. J Biol Chem 284: 5352–5361,
2009. [PubMed: 19049983]
243. Stone SJ, Myers HM, Watkins SM, Brown BE, Feingold KR, Elias PM, and Farese RV. Lipopenia
and skin barrier abnormalities in DGAT2-deficient mice. J Biol Chem 279: 11767–11776, 2004.
[PubMed: 14668353]
244. Strable MS, and Ntambi JM. Genetic control of de novo lipogenesis: role in diet-induced obesity.
Crit Rev Biochem Mol Biol 45: 199–214, 2010. [PubMed: 20218765]
245. Straub BK, Stoeffel P, Heid H, Zimbelmann R, and Schirmacher P. Differential pattern of lipid
droplet-associated proteins and de novo perilipin expression in hepatocyte steatogenesis.
Author Manuscript

Hepatology 47: 1936–1946, 2008. [PubMed: 18393390]


246. Stremmel W, Diede HE, Rodilla-Sala E, Vyska K, Schrader M, Fitscher B, and Passarella S. The
membrane fatty acid-binding protein is not identical to mitochondrial glutamic oxaloacetic
transaminase (mGOT). Mol Cell Biochem 98: 191–199, 1990. [PubMed: 2266960]
247. Stremmel W, Pohl J, Ring A, and Herrmann T. A new concept of cellular uptake and intracellular
trafficking of long-chain fatty acids. Lipids 36: 981–989, 2001. [PubMed: 11724471]
248. Stremmel W, Staffer S, Wannhoff A, Pathil A, and Chamulitrat W. Plasma membrane
phospholipase A2 controls hepatocellular fatty acid uptake and is responsive to pharmacological
modulation: implications for nonalcoholic steatohepatitis. FASEB J 28: 159–170, 2014.

Compr Physiol. Author manuscript; available in PMC 2019 February 15.


Alves-Bezerra and Cohen Page 32

249. Stremmel W, Strohmeyer G, Borchard F, Kochwa S, and Berk PD. Isolation and partial
characterization of a fatty acid binding protein in rat liver plasma membranes. Proc Natl Acad Sci
Author Manuscript

U S A 82: 4–8, 1985. [PubMed: 3881757]


250. Su W, Wang Y, Jia X, Wu W, Li L, Tian X, Li S, Wang C, Xu H, Cao J, Han Q, Xu S, Chen Y,
Zhong Y, Zhang X, Liu P, Gustafsson JA, and Guan Y. Comparative proteomic study reveals
17beta-HSD13 as a pathogenic protein in nonalcoholic fatty liver disease. Proc Natl Acad Sci U
S A 111: 11437–11442, 2014 10.1073/pnas.1410741111 [PubMed: 25028495]
251. Subramanian S, and Chait A. Hypertriglyceridemia secondary to obesity and diabetes. Biochim
Biophys Acta 1821: 819–825, 2012. [PubMed: 22005032]
252. Takeuchi K, and Reue K. Biochemistry, physiology, and genetics of GPAT, AGPAT, and lipin
enzymes in triglyceride synthesis. Am J Physiol Endocrinol Metab 296: E1195–E1209, 2009.
[PubMed: 19336658]
253. Thompson BR, Lobo S, and Bernlohr DA. Fatty acid flux in adipocytes: The in’s and out’s of fat
cell lipid trafficking. Mol Cell Endocrinol 318: 24–33, 2010. [PubMed: 19720110]
254. Thumser AE, and Wilton DC. The binding of cholesterol and bile salts to recombinant rat liver
fatty acid-binding protein. Biochem J 320 ( Pt 3): 729–733, 1996. [PubMed: 9003356]
Author Manuscript

255. Thumser AE, and Wilton DC. The binding of natural and fluorescent lysophospholipids to wild-
type and mutant rat liver fatty acid-binding protein and albumin. Biochem J 307 ( Pt 1): 305–311,
1995. [PubMed: 7717990]
256. Tillander VE, Alexson SE, and Cohen DE. Deactivating fatty acids: acyl-CoA thioesterase-
mediated control of lipid metabolism. Trends Endocrinol Metab in press: 2017.
257. Tiwari S, Siddiqi S, and Siddiqi SA. CideB protein is required for the biogenesis of very low
density lipoprotein (VLDL) transport vesicle. J Biol Chem 288: 5157–5165, 2013. [PubMed:
23297397]
258. Tong FM, Black PN, Coleman RA, and DiRusso CC. Fatty acid transport by vectorial acylation in
mammals: Roles played by different isoforms of rat long-chain acyl-CoA synthetases. Arch
Biochem Biophys 447: 46–52, 2006. [PubMed: 16466685]
259. Tontonoz P, Nagy L, Alvarez JG, Thomazy VA, and Evans RM. PPARgamma promotes
monocyte/macrophage differentiation and uptake of oxidized LDL. Cell 93: 241–252, 1998.
[PubMed: 9568716]
Author Manuscript

260. Turpin SM, Nicholls HT, Willmes DM, Mourier A, Brodesser S, Wunderlich CM, Mauer J, Xu E,
Hammerschmidt P, Bronneke HS, Trifunovic A, LoSasso G, Wunderlich FT, Kornfeld JW,
Bluher M, Kronke M, and Bruning JC. Obesity-induced CerS6-dependent C16:0 ceramide
production promotes weight gain and glucose intolerance. Cell Metab 20: 678–686, 2014
10.1016/j.cmet.2014.08.002 [PubMed: 25295788]
261. Vasconcellos R, Alvarenga EC, Parreira RC, Lima SS, and Resende RR. Exploring the cell
signalling in hepatocyte differentiation. Cell Signal 28: 1773–1788, 2016 10.1016/j.cellsig.
2016.08.011 [PubMed: 27555287]
262. Vergnes L, Beigneux AP, Davis R, Watkins SM, Young SG, and Reue K. Agpat6 deficiency
causes subdermal lipodystrophy and resistance to obesity. J Lipid Res 47: 745–754, 2006.
[PubMed: 16436371]
263. Visser ME, Kastelein JJ, and Stroes ES. Apolipoprotein B synthesis inhibition: results from
clinical trials. Curr Opin Lipidol 21: 319–323, 2010 10.1097/MOL.0b013e32833af4c1 [PubMed:
20508521]
264. Viswakarma N, Yu S, Naik S, Kashireddy P, Matsumoto K, Sarkar J, Surapureddi S, Jia Y, Rao
Author Manuscript

MS, and Reddy JK. Transcriptional regulation of Cidea, mitochondrial cell death-inducing DNA
fragmentation factor alpha-like effector A, in mouse liver by peroxisome proliferator-activated
receptor alpha and gamma. J Biol Chem 282: 18613–18624, 2007. [PubMed: 17462989]
265. Volmer R, van der Ploeg K, and Ron D. Membrane lipid saturation activates endoplasmic
reticulum unfolded protein response transducers through their transmembrane domains. Proc Natl
Acad Sci U S A 110: 4628–4633, 2013 10.1073/pnas.1217611110 [PubMed: 23487760]
266. Wakil SJ, and Abu-Elheiga LA. Fatty acid metabolism: target for metabolic syndrome. J Lipid
Res 50: S138–S143, 2009. [PubMed: 19047759]

Compr Physiol. Author manuscript; available in PMC 2019 February 15.


Alves-Bezerra and Cohen Page 33

267. Walther TC, and Farese RVJ. Lipid droplets and cellular lipid metabolism. Annu Rev Biochem
81: 687–714, 2012. [PubMed: 22524315]
Author Manuscript

268. Wang C, Zhao Y, Gao X, Li L, Yuan Y, Liu F, Zhang L, Wu J, Hu P, Zhang X, Gu Y, Xu Y, Wang


Z, Li Z, Zhang H, and Ye J. Perilipin 5 improves hepatic lipotoxicity by inhibiting lipolysis.
Hepatology 61: 870–882, 2015. [PubMed: 25179419]
269. Wang H, Bell M, Sreenivasan U, Hu H, Liu J, Dalen K, Londos C, Yamaguchi T, Rizzo MA,
Coleman R, Gong D, Brasaemle D, and Sztalryd C. Unique regulation of adipose triglyceride
lipase (ATGL) by perilipin 5, a lipid droplet-associated protein. J Biol Chem 286: 15707–15715,
2011. [PubMed: 21393244]
270. Wang J, Bie J, and Ghosh S. Intracellular cholesterol transport proteins enhance hydrolysis of
HDL-delivered cholesteryl esters and facilitate preferential elimination of resulting cholesterol
into bile. J Lipid Res In press: 2016.
271. Wang S, Lee DP, Gong N, Schwerbrock NMJ, Mashek DG, Gonzalez-Baro MR, Stapleton C, Li
LO, Lewin TM, and Coleman RA. Cloning and functional characterization of a novel
mitochondrial N-ethylmaleimide-sensitive glycerol-3-phosphate acyltransferase (GPAT2). Arch
Biochem Biophys 465: 347–358, 2007. [PubMed: 17689486]
Author Manuscript

272. Waterman IJ, and Zammit VA. Activities of overt and latent diacylglycerol acyltransferases
(DGATs I and II) in liver microsomes of ob/ob mice. Int J Obes 26: 742–743, 2002.
273. Watkins PA. Very-long-chain acyl-CoA synthetases. J Biol Chem 283: 1773–1777, 2008 10.1074/
jbc.R700037200 [PubMed: 18024425]
274. Welty FK. Hypobetalipoproteinemia and abetalipoproteinemia. Curr Opin Lipidol 25: 161–168,
2014 10.1097/MOL.0000000000000072 [PubMed: 24751931]
275. Wendel AA, Cooper DE, Ilkayeva OR, Muoio DM, and Coleman RA. Glycerol-3-phosphate
acyltransferase (GPAT)-1, but not GPAT4, incorporates newly synthesized fatty acids into
triacylglycerol and diminishes fatty acid oxidation. J Biol Chem 288: 27299–27306, 2013.
[PubMed: 23908354]
276. Wendel AA, Lewin TM, and Coleman RA. Glycerol-3-phosphate acyltransferases: Rate limiting
enzymes of triacylglycerol biosynthesis. Biochimica Et Biophysica Acta-Molecular and Cell
Biology of Lipids 1791: 501–506, 2009.
277. West J, Tompkins CK, Balantac N, Nudelman E, Meengs B, White T, Bursten S, Coleman J,
Kumar A, Singer JW, and Leung DW. Cloning and expression of two human lysophosphatidic
Author Manuscript

acid acyltransferase cDNAs that enhance cytokine-induced signaling responses in cells. DNA
Cell Biol 16: 691–701, 1997. [PubMed: 9212163]
278. Wilfling F, Haas JT, Walther TC, and Farese RVJ. Lipid droplet biogenesis. Curr Opin Cell Biol
29: 39–45, 2014. [PubMed: 24736091]
279. Wilson CG, Tran JL, Erion DM, Vera NB, Febbraio M, and Weiss EJ. Hepatocyte-specific
disruption of CD36 attenuates fatty liver and improves insulin sensitivity in HFD-fed mice.
Endocrinology 157: 570–585, 2016. [PubMed: 26650570]
280. Wu JW, Wang SP, Alvarez F, Casavant S, Gauthier N, Abed L, Soni KG, Yang G, and Mitchell
GA. Deficiency of liver adipose triglyceride lipase in mice causes progressive hepatic steatosis.
Hepatology 54: 122–132, 2011. [PubMed: 21465509]
281. Wurie HR, Buckett L, and Zammit VA. Diacylglycerol acyltransferase 2 acts upstream of
diacylglycerol acyltransferase 1 and utilizes nascent diglycerides and de novo synthesized fatty
acids in HepG2 cells. FEBS Journal 279: 3033–3047, 2012. [PubMed: 22748069]
282. Xu W, Wu L, Yu M, Chen FJ, Arshad M, Xia X, Ren H, Yu J, Xu L, Xu D, Li JZ, Li P, and Zhou
Author Manuscript

L. Differential Roles of Cell Death-inducing DNA Fragmentation Factor-alpha-like Effector


(CIDE) Proteins in Promoting Lipid Droplet Fusion and Growth in Subpopulations of
Hepatocytes. J Biol Chem 291: 4282–4293, 2016 10.1074/jbc.M115.701094 [PubMed:
26733203]
283. Yang YY, Pritchard PH, Bhuiyan J, Seccombe DW, and Moghadasian MH. Overexpression of
Acyl-CoA binding protein and its effects on the flux of free fatty acids in McA-RH 7777 cells.
Lipids 36: 595–600, 2001. [PubMed: 11485163]

Compr Physiol. Author manuscript; available in PMC 2019 February 15.


Alves-Bezerra and Cohen Page 34

284. Ye J, Li JZ, Liu Y, Li X, Yang T, Ma X, Li Q, Yao Z, and Li P. Cideb, an ER- and lipid droplet-
associated protein, mediates VLDL lipidation and maturation by interacting with apolipoprotein
Author Manuscript

B. Cell Metabolism 9: 177–190, 2009. [PubMed: 19187774]


285. Yen CLE, Monetti M, Burri BJ, and Farese RV. The triacylglycerol synthesis enzyme DGAT1 also
catalyzes the synthesis of diacylglycerols, waxes, and retinyl esters. J Lipid Res 46: 1502–1511,
2005. [PubMed: 15834126]
286. Yen CLE, Stone SJ, Koliwad S, Harris C, and Farese RV. DGAT enzymes and triacylglycerol
biosynthesis. J Lipid Res 49: 2283–2301, 2008. [PubMed: 18757836]
287. Yeon JE, Choi KM, Baik SH, Kim KO, Lim HJ, Park KH, Kim JY, Park JJ, Kim JS, Bak YT,
Byun KS, and Lee CH. Reduced expression of peroxisome proliferator-activated receptor-alpha
may have an important role in the development of non-alcoholic fatty liver disease. J
Gastroenterol Hepatol 19: 799–804, 2004. [PubMed: 15209628]
288. Yu S, Matsusue K, Kashireddy P, Cao WQ, Yeldandi V, Yeldandi AV, Rao MS, Gonzalez FJ, and
Reddy JK. Adipocyte-specific gene expression and adipogenic steatosis in the mouse liver due to
peroxisome proliferator-activated receptor gamma1 (PPARgamma1) overexpression. J Biol Chem
278: 498–505, 2003. [PubMed: 12401792]
Author Manuscript

289. Zammit VA. Hepatic triacylglycerol synthesis and secretion: DGAT2 as the link between
glycaemia and triglyceridaemia. The Biochemical Journal 451: 1–12, 2013. [PubMed: 23489367]
290. Zhang L, Miyaki K, Nakayama T, and Muramatsu M. Cell death-inducing DNA fragmentation
factor alpha-like effector A (CIDEA) gene V115F (G–>T) polymorphism is associated with
phenotypes of metabolic syndrome in Japanese men. Metabolism 57: 502–505, 2008. [PubMed:
18328351]
291. Zhang L, and Wang HH. The essential functions of endoplasmic reticulum chaperones in hepatic
lipid metabolism. Dig Liver Dis 48: 709–716, 2016 10.1016/j.dld.2016.03.016 [PubMed:
27133206]
292. Zhang Y, Li Y, Niepel MW, Kawano Y, Han S, Liu S, Marsili A, Larsen PR, Lee CH, and Cohen
DE. Targeted deletion of thioesterase superfamily member 1 promotes energy expenditure and
protects against obesity and insulin resistance. Proc Natl Acad Sci U S A 109: 2012.
293. Zhou H, and Liu R. ER stress and hepatic lipid metabolism. Front Genet 5: 112, 2014 10.3389/
fgene.2014.00112 [PubMed: 24847353]
294. Zhou J, Zhai Y, Mu Y, Gong H, Uppal H, Toma D, Ren S, Evans RM, and Xie W. A novel
Author Manuscript

pregnane X receptor-mediated and sterol regulatory element-binding protein-independent


lipogenic pathway. J Biol Chem 281: 15013–15020, 2006. [PubMed: 16556603]
295. Zhou SL, Gordon RE, Bradbury M, Stump D, Kiang CL, and Berk PD. Ethanol up-regulates fatty
acid uptake and plasma membrane expression and export of mitochondrial aspartate
aminotransferase in HepG2 cells. Hepatology 27: 1064–1074, 1998. [PubMed: 9537447]
296. Zhuravleva E, Gut H, Hynx D, Marcellin D, Bleck CK, Genoud C, Cron P, Keusch JJ, Dummler
B, Esposti MD, and Hemmings BA. Acyl coenzyme A thioesterase Them5/Acot15 is involved in
cardiolipin remodeling and fatty liver development. Mol Cell Biol 32: 2685–2697, 2012.
[PubMed: 22586271]
297. Zoltowska M, Ziv E, Delvin E, Lambert M, Seidman E, and Levy E. Both insulin resistance and
diabetes in Psammomys obesus upregulate the hepatic machinery involved in intracellular VLDL
assembly. Arterioscler Thromb Vasc Biol 24: 118–123, 2004. [PubMed: 14604833]
Author Manuscript

Compr Physiol. Author manuscript; available in PMC 2019 February 15.


Alves-Bezerra and Cohen Page 35

Didactic Synopsis
Author Manuscript

Major teaching points:

• A variety of endogenous and exogenous sources provide fatty acids that can
be assembled into to triglycerides within the liver in health and disease;

• Multiple hepatocellular mechanisms regulate fatty acid uptake, synthesis,


transport, and oxidation;

• Triglycerides can be stored within hepatocytes, undergo lipolysis, or be


exported into the bloodstream depending upon physiological and pathological
conditions;

• Mitochondria, endoplasmic reticulum, Golgi apparatus, peroxisomes, and


lipid droplets are examples of organelles that participate in fatty acid and
Author Manuscript

triglyceride metabolism;

• The metabolism of fatty acids and triglyceride is regulated by transcriptional


and post-transcriptional mechanisms;

• Alterations in fatty acid and triglyceride metabolism lead to non-alcoholic


fatty liver disease (NAFLD), which is a common consequence of
overnutrition.
Author Manuscript
Author Manuscript

Compr Physiol. Author manuscript; available in PMC 2019 February 15.


Alves-Bezerra and Cohen Page 36
Author Manuscript
Author Manuscript
Author Manuscript

Figure 1. Major sources for hepatic fatty acids.


Author Manuscript

The three major sources for hepatic fatty acids (FA) are dietary lipids, adipose tissue
derived-FA and de novo-synthesized FA. After a meal, dietary lipids are hydrolyzed within
the intestinal lumen. Upon intestinal uptake, FA are re-esterified to form TG molecules,
which are packaged into chylomicrons and delivered primarily to muscle and adipose tissue.
The remaining TG present in chylomicron remnants is then transported to the liver and
processed intracellularly, leading to FA release within hepatocytes. Carbohydrates,
particularly glucose, are utilized in hepatic de novo lipogenesis (DNL) for the production of

Compr Physiol. Author manuscript; available in PMC 2019 February 15.


Alves-Bezerra and Cohen Page 37

FA. In order to be metabolized, FAs are activated to form acyl-CoA molecules, which can
Author Manuscript

undergo oxidation or be incorporated into complex lipids. Locally synthesized TG can be


stored in intracellular lipid droplets (LDs) or packed into VLDL and secreted into the
plasma. Upon fasting, intracellular TG stores are mobilized from adipocytes and hepatocytes
to release FA products. Hepatic DNL may also contribute to form an acyl-CoA pool
available for energy production, undergoing oxidation by mitochondria, or for VLDL-TG
production. In the setting of overnutrition and insulin resistance, hepatic FA levels are
increased due to enhanced lipolysis within adipocytes, which leads to increased circulating
levels of FA in the bloodstream, and increased hepatic DNL. Excess FA cannot be consumed
by oxidative pathways and FA are instead directed towards the synthesis of TG, leading to
increased hepatic TG storage and VLDL overproduction. Arrow thickness denotes rates of
metabolic activities.
Author Manuscript
Author Manuscript
Author Manuscript

Compr Physiol. Author manuscript; available in PMC 2019 February 15.


Alves-Bezerra and Cohen Page 38
Author Manuscript
Author Manuscript
Author Manuscript

Figure 2. Hepatic fatty acid transport and metabolism.


Within the plasma membrane, FA translocase (FAT)/CD36, plasma membrane FA-binding
protein (FABPpm), and Caveolin-1 mediate the uptake of fatty acid (FA) that is bound to
circulating albumin. Alternatively, hepatic FA can be obtained by the internalization of
chylomicron remnant or by de novo lipogenesis. The latter occurs through the activity of
three key enzymes: ATP-citrate lyase (ACLY), acetyl-CoA carboxylase (ACC), and fatty
acid synthase (FAS). In the cytosol, FAs are bound to the fatty acid-binding protein-1
(FABP1) and sterol carrier protein-2 (SCP2), which may control their cellular distribution.
Author Manuscript

FA transport proteins (FATP2, 4 and 5) and long-chain acyl-CoA synthetases (ACSL1, 3 and
5) mediate the activation of long-chain FA to acyl-CoA molecules and their channeling to
metabolic pathways. Although associated with mitochondria, ACSL5 may function to
promote triglyceride biosynthesis. In the cytosol, acyl-CoAs are bound to acyl-CoA-binding
protein (ACBP) or SCP2. Acyl-CoA thioesterases (ACOT)/thioesterase superfamily
members (Them1, 2 and 5) appear to counteract ACSL activity by catalyzing the hydrolysis
of acyl-CoA molecules into FA and CoA. This may provide additional means of controlling
the balance between FA and acyl-CoA within hepatocytes.

Compr Physiol. Author manuscript; available in PMC 2019 February 15.


Alves-Bezerra and Cohen Page 39
Author Manuscript
Author Manuscript
Author Manuscript

Figure 3. Hepatic triglyceride metabolism.


Author Manuscript

Acyl-CoA molecules can be esterified to glycerol-3-phosphate (G3P) by isoforms of


glycerol-3-phosphate acyltransferase (GPAT). In hepatocytes, the isoforms are
predominantly GPATs 1 and 4. The produced lysophosphatidic acid (LPA) is acylated by
acylglycerol-3-phosphate acyltransferases (AGPATs) to form phosphatidic acid (PA), which
can be dephosphorylated by phosphatidic acid phosphatase (PAP) to form diacylglycerol
(DG). Both PA and DG can be directed towards phospholipid (PL) synthesis. Additionally,
diacylglycerol acyltransferases (DGATs 1 and 2) synthesize triglyceride (TG) by acylation

Compr Physiol. Author manuscript; available in PMC 2019 February 15.


Alves-Bezerra and Cohen Page 40

of DG. (A) Fas synthesized de novo are likely to be channeled to VLDL-TG production
Author Manuscript

through DGAT1 activity. (B) Exogenous FA appear to be directed towards TG synthesis for
storage in lipid droplets by the activity of DGAT2. In addition, lipid droplet-TG can undergo
hydrolysis, with FA re-routed into VLDL by DGAT1.
Author Manuscript
Author Manuscript
Author Manuscript

Compr Physiol. Author manuscript; available in PMC 2019 February 15.


Alves-Bezerra and Cohen Page 41
Author Manuscript
Author Manuscript
Author Manuscript

Figure 4. Triglyceride storage and secretion in hepatocytes.


Triglyceride (TG) can be synthesized from diacylglycerol (DG) by diacylglycerol
acyltransferases (DGAT1 and 2). DGAT1 preferably provides TG to VLDL during lipidation
in the lumen of the endoplasmic reticulum. This process is mediated by microsomal
triglyceride transfer protein (MTP), which facilitates the association between TG and
apoB100. Transmembrane 6 superfamily member 2 (TM6SF2) may also contribute to
VLDL lipidation via yet unknown mechanisms. The nascent VLDL particle is then
transferred to the Golgi apparatus through the VLDL transfer vesicle (VTV), followed by a
Author Manuscript

second MTP-mediated lipidation step. VLDL particles are secreted via a vesicle-mediated
mechanism. TG can also be formed by the activity of DGAT2, which mainly contributes to
storage in lipid droplets. Lipid droplets are delimitated by proteins and a
phosphatidylcholine (PC)-enriched surfactant monolayer. Among the lipid-droplet
associated proteins, perilipins (PLIN2, 3 and 5) and comparative gene identification-58
(CGI-58) contribute to lipid droplet structure and/or the regulation of lipid droplet-
associated enzymes; CTP-phosphocholine cytidylyltransferase (CCT) and acyl-CoA
synthetase 3 (ACSL3) may be required for the biosynthesis of lipids; DFF45-like effector

Compr Physiol. Author manuscript; available in PMC 2019 February 15.


Alves-Bezerra and Cohen Page 42

(CIDE) proteins, Cidea, Cideb, and Fsp27, and the microsomal fat-inducing transmembrane
Author Manuscript

protein 2 (FIT2) are required for lipid droplet formation, however their specific cellular roles
are incompletely understood. Cideb also contributes to VLDL production and secretion via
its interaction with Sar1 and Sec24, which are present in VTV.
Author Manuscript
Author Manuscript
Author Manuscript

Compr Physiol. Author manuscript; available in PMC 2019 February 15.


Alves-Bezerra and Cohen Page 43
Author Manuscript
Author Manuscript
Author Manuscript

Figure 5. Lipolysis and fatty acid oxidation in hepatocytes.


The initiation of hepatic lipolysis depends on the activation of adipose triglyceride lipase
(ATGL) through the binding to the comparative gene identification-58 (CGI-58). ATGL
catalyzes the hydrolysis of triglyceride (TG), releasing diacylglycerol (DG). This lipid is
further hydrolyzed by hormone sensitive lipase (HSL), releasing monoacylglycerol (MG).
Monoacylglycerol lipase (MGL) mediates the breakdown of MG into fatty acid (FA) and
glycerol. Alternatively, autophagic pathways, such as macroautophagy and chaperone-
mediated autophagy (CMA), promote the hydrolysis of LD. FA can also be generated from
Author Manuscript

acyl-CoA by the activity of acyl-CoA thioesterase 13 (ACOT 13; synonym: thioesterase


superfamily member 2, Them2). In the mitochondria outer membrane, long-chain acyl-CoA
synthetase 1 (ACSL1) converts long-chain FAs to acyl-COAs. ACSL1 interacts physically
with carnitine palmitoyltransferase 1 (CPT1). It is likely that ACSL1 channels acyl-CoA to
CPT1 to control the availability of substrates for mitochondrial β-oxidation. Additionally,
acyl-CoAs can be directed to β-oxidation in the peroxisome, where fatty acyl-CoA oxidase
(AOX) represents a rate-limiting step, or ω-oxidation and α-oxidation in the endoplasmic
reticulum, mediated by P450 4A family members. Also, part of the acyl-CoA pool can be

Compr Physiol. Author manuscript; available in PMC 2019 February 15.


Alves-Bezerra and Cohen Page 44

directed for TG synthesis and VLDL assembly, sustaining the transfer of hepatic lipids to
Author Manuscript

other organs.
Author Manuscript
Author Manuscript
Author Manuscript

Compr Physiol. Author manuscript; available in PMC 2019 February 15.

You might also like