Abstract
The hepatic glucose-sensing system is a functional network of enzymes and transcription factors that is critical for the maintenance of energy homeostasis and systemic glycemia. Here we review the recent literature on its components and metabolic actions. Glucokinase (GCK) is generally considered as the initial postprandial glucose-sensing component, which acts as the gatekeeper for hepatic glucose metabolism and provides metabolites that activate the transcription factor carbohydrate response element binding protein (ChREBP). Recently, liver receptor homolog 1 (LRH-1) has emerged as an upstream regulator of the central GCK–ChREBP axis, with a critical role in the integration of hepatic intermediary metabolism in response to glucose. Evidence is also accumulating that O-linked β-N-acetylglucosaminylation (O-GlcNAcylation) and acetylation can act as glucose-sensitive modifications that may contribute to hepatic glucose sensing by targeting regulatory proteins and the epigenome. Further elucidation of the components and functional roles of the hepatic glucose-sensing system may contribute to the future treatment of liver diseases associated with deregulated glucose sensors.
Keywords: Acetylation, ChREBP, Glucokinase, Glucose sensing, Hepatocytes, LRH-1, O-linked β-N-acetylglucosaminylation
Introduction
Glucose is a simple sugar carbohydrate that serves as a fundamental fuel for most species and provides precursors for biomolecule synthesis. In order to control metabolism, differentiation, and growth, cells possess evolutionary conserved glucose-sensitive signaling pathways [1]. These glucose-sensing systems ensure efficient adaptation to changes in environmental glucose availability in unicellular organisms and allow for homeostatic maintenance of internal glucose pools in multicellular organisms. In higher species, the internal pool is represented by glucose circulating in the bloodstream. From here, glucose is further distributed to different tissues and organs to meet local needs.
The liver plays a central role in metabolic homeostasis by coordinating the breakdown, synthesis, storage, and redistribution of nutrients. Hepatocytes possess multiple nutrient-sensing systems that interact to modulate biochemical pathways in order to accommodate systemic fuel requirements and availability. These systems enable the body to maintain its functions during periods of feeding and fasting and upon excessive energy demands such as exercise. Blood glucose concentrations fluctuate during the feeding and fasting cycles [2], and one of the liver’s primary functions is to maintain blood glucose concentrations within a physiological range [3]. Hepatocytes are among the few cell types that possess the ability to both consume and produce glucose [4]. Glycemic control, which is coordinated by both extrahepatic and intrahepatic factors, is hence the result of a balancing act between these two processes. Most reviews have focused on extrahepatic glucose-sensing systems such as hormonal regulation by insulin and glucagon [3, 5, 6]. In contrast, this review will provide an overview of the regulatory components within the liver that are activated by glucose metabolites in response to glucose availability. We provide an overview of these regulatory components and discuss the role of this intrahepatic glucose-sensing system in health and disease.
Hepatic glucose metabolism
The concentration of glucose in the blood is a primary determinant of glucose availability to the liver. During the postprandial phase, which in humans lasts about 2 h after the intake of a meal, blood glucose levels rise and approximately 10–25 % of ingested glucose is taken up by hepatocytes [7–10]. Facilitated transport of glucose across cellular membranes is mediated by members of the glucose transporter (GLUT) family [11]. GLUT2 is the major glucose transporter in the hepatocytes [11, 12] and its physiological role has been studied extensively [13–15]. GLUT2 is also expressed in pancreatic islets, intestine, kidney, and brain [11, 12]. The rate of GLUT2-mediated glucose transport into the liver is high and only saturates at glucose concentrations above 30 mM [11] allowing efficient glucose transport and extremely rapid equilibration of glucose across the hepatocyte membrane [16]. Once in the cytoplasm, glucose is phosphorylated to glucose-6-phosphate (G6P) by glucokinase (GCK; also known as hexokinase IV) [17, 18]. G6P lies at the crossroads of different biochemical pathways and has multiple biochemical fates. Elevated G6P synthesis allosterically activates glycogen synthase while inhibiting glycogen phosphorylase [19–21]. G6P is also oxidized for energy supply via glycolysis, which involves several steps including the production of fructose-6-phosphate (F6P) and triose phosphates. The pentose phosphate pathway represents a third route of G6P utilization that involves the production of ribose-5-phosphate, an intermediate of nucleotide synthesis, and the biological reductant NADPH. Excess pentose phosphates can ultimately enter the glycolytic pathway by their conversion into F6P and triose phosphates. Pyruvate produced by glycolysis is transported into the mitochondria, where it is decarboxylated to acetyl-coenzyme A (acetyl-CoA), which subsequently enters the tricarboxylic acid (TCA) cycle, a central metabolic hub that is involved in both energy production and biomolecule synthesis. To keep TCA cycle intermediates at a constant level, reactions that extract TCA metabolites for biosynthesis (cataplerotic reactions) are balanced by those that replenish TCA intermediates (anaplerotic reactions) [22]. In the TCA cycle, acetyl-CoA becomes further metabolized to generate reducing equivalents used for ATP production through oxidative phosphorylation. The TCA cycle intermediates also serve as precursors for non-essential amino acids, which serve as substrates for protein synthesis. Citrate produced in the TCA cycle is partly shuttled from the mitochondria into the cytosol where it is converted into oxaloacetate and acetyl-CoA, the latter of which can be used as a substrate for lipid synthesis. These hepatic glucose oxidation and storage pathways are summarized in Fig. 1a.
Fig. 1.
Pathways of hepatic glucose metabolism. a Simplified scheme depicting the major biochemical pathways activated during postprandial glucose consumption and storage. b Simplified scheme depicting the major biochemical pathways activated during postabsorptive glucose production. Glycerol, lactate, and alanine are used as gluconeogenic substrates upon their conversion into triose phosphate and pyruvate. Acetyl-CoA acetyl-coenzyme A, GCK glucokinase, GLUT2 glucose transporter 2, TCA tricarboxylic acid
Hepatic glucose uptake and metabolism decrease as soon as the intestinal absorption of glucose is completed. During this period, which is often referred to as the postabsorptive phase, most tissues reduce their glucose consumption by switching to alternate energy sources. Endogenous glucose production by the liver now represents the major route of glucose supply to the bloodstream. The maintenance of glucose homeostasis is particularly important for cells that partly or fully rely on glucose as energetic substrate such as neurons and erythrocytes. The liver contributes to endogenous glucose production via two G6P-generating pathways. In the initial postabsorptive phase, hepatic G6P is derived from glycogen breakdown while gluconeogenesis becomes the major source of G6P after prolonged fasting. G6P generated through glycogen breakdown and gluconeogenesis is first translocated from the cytosol into the endoplasmic reticulum by the glucose-6-phosphate transporter (G6Pt; also known as SLC37A4), and subsequently dephosphorylated into glucose by glucose-6-phosphatase (G6Pc). Glucose is finally released into the bloodstream, presumably through the concerted action of GLUT2 and a membrane traffic-based mechanism [13, 14, 23]. These glucose-production pathways in the liver are summarized in Fig. 1b.
Postprandial glucose sensing in the liver
When blood glucose concentrations rise, hepatic glucose sensors induce adaptive responses to shift the balance toward hepatic glucose consumption and storage. GLUT2 is a high-capacity glucose transporter that allows glucose to flow into hepatocytes in response to increasing glycemia [11]. However, its activity does not appear to be critical for postprandial glucose sensing in the liver, as was recently reported [13]. In this study, hepatic GLUT2 deficiency did not result in major perturbations in hepatic glucose metabolism in fed and refed mice, suggesting that alternate mechanisms compensate for the reduction in glucose transport. GCK, on the contrary, is a major component of the hepatic glucose-sensing system. By converting glucose into G6P, GCK catalyzes the first step of intrahepatic glucose metabolism [17]. In contrast to hexokinases (HKs) I-II, GCK exhibits low affinity for glucose, is not feedback-inhibited by its product G6P [17, 24], and its activity increases sigmoidal with increasing glycemia [18, 25]. High glucose concentrations furthermore inhibit the interaction of GCK with its regulatory protein (GCKR), hence promoting the translocation of free GCK to the cytoplasm where it can access glucose and convert it into G6P [26]. GCK consequently acts as a glucose-sensitive enzyme that remains active over a wide range of glucose concentrations and enables hepatocytes to efficiently trap glucose in response to glycemic fluctuations. Lack of hepatic GCK expression in mice perturbs intrahepatic glucose metabolism [27, 28] while overexpression of GCK, but not HK-I, markedly induces glycogen storage and glycolysis in hepatocytes [29, 30]. These fundamental differences of GCK versus HK-mediated G6P synthesis illustrate the unique role of hepatocytes as compared to other cells.
Further downstream metabolism of G6P generates metabolites that act as signaling molecules to regulate the activity of enzymes within seconds to minutes after hepatic glucose exposure [19, 20, 31–35]. Glucose-mediated control of gene transcription in hepatocytes translates into adaptive responses on longer timescales, i.e., within a timeframe of minutes to hours [36–38]. The expression of many glucose-sensitive genes is regulated by the carbohydrate response element binding protein (ChREBP; also known as Mondo B or Max-like protein X interacting protein-like, MLXIPL) [39, 40], a transcription factor that recognizes conserved carbohydrate response elements (ChoREs) in gene promoters [41, 42]. ChREBP is a member of the Mondo family, which forms heterodimers with Max-like protein X (Mlx) to induce transcriptional responses [43–48]. Mondo-Mlx-dependent glucose sensing is evolutionary conserved among worms, flies, and vertebrates [49–55]. ChREBP has been identified as the major mediator of ChoRE-dependent gene transcription in the liver [40, 48], while its paralog MondoA (or Mlx interacting protein, MLXIP) has been proposed to act predominantly in extrahepatic tissue [45, 56]. However, a recent study showed that MondoA also regulates transcription of specific glucose-responsive genes in hepatocytes [49]. ChREBP is best-known for its effects on the expression of enzymes involved in glycolysis and fatty acid synthesis [57]. In addition, ChREBP suppresses sirtuin 1, thereby likely reducing PGC-1α-dependent gluconeogenesis under glucose abundant conditions [58]. Somewhat counter-intuitively, ChREBP also induces G6Pc expression, a response that may serve to maintain the intracellular G6P homeostasis [59]. ChIP-seq analysis indicated that ChREBP not only regulates metabolism, but also targets genes related to transport, development, and cell motility [39].
Several studies have shown that hepatic ChREBP activation requires GCK-dependent glucose metabolism [28, 60]. Early work showed that the pentose phosphate pathway intermediate xylulose-5-phosphate (X5P) induces ChREBP dephosphorylation, thereby promoting its nuclear translocation and transcriptional activity [61]. However, this model has been challenged, based on the finding that pentose phosphate pathway inhibition leads to a decrease rather than an increase in ChREBP activity [62, 63]. Instead, G6P was suggested to be the major signaling metabolite responsible for ChREBP activation [62, 63]. Finally, fructose-2,6-bisphosphate (F2,6bisP), another glucose metabolite, has also been proposed to induce ChREBP-mediated transcription in hepatocytes [49, 64]. The mechanisms through which these three glucose derivatives act remain to be resolved, but likely involve changes in allosteric regulation and post-translational modifications [53, 65, 66]. In this respect, it should be noted that ChREBP activity is increased by acetylation and O-linked β-N-acetylglucosaminylation (O-GlcNAcylation) [67, 68], two enzyme-catalyzed post-translational modifications that use glucose metabolites as substrates, as will be discussed in more detail below [69–71]. The fact that several independent glucose metabolites (X5P, G6P, F2,6bisP, acetyl-CoA, and O-GlcNAc) activate hepatic ChREBP illustrates the unique glucose-sensing ability of this transcription factor in hepatocytes [57]. A recent study furthermore showed that glucose promotes the binding of full-length ChREBP-α to a ChoRE located in an alternative promoter region of the Chrebp gene thereby inducing transcription of a potent, short ChREBP isoform (ChREBP-β) [72]. Future work should identify the specific glucose-dependent pathways that induce and activate these different isoforms in hepatocytes, and reveal whether ChREBP-α and ChREBP-β regulate different target genes.
Regulation of the central hepatic glucose-sensing axis
The GCK–ChREBP axis can be considered as the central glucose-sensing system in the liver. Because GCK acts as a gatekeeper for hepatic glucose metabolism and ChREBP activation [60, 73], regulation of its expression and activity will significantly impact hepatic glucose sensing. Interestingly, glucose increases GCKR expression while it inhibits GCK transcription in cultured hepatocytes [59]. However, in vivo GCK expression is induced in response to an oral glucose load [60]. Because insulin is a major regulator of GCK expression in the liver [31], the discrepancy between these findings can be explained by the lack of a concomitant insulin-mediated GCK transcription under in vitro conditions [74]. The mechanistic basis of insulin-dependent GCK induction is incompletely understood [31, 75]. Several transcription factors, i.e., hepatocyte nuclear factor 4 (HNF-4), hypoxia-inducible factor 1 (HIF-1), sterol regulatory binding protein-1c (SREBP-1c), liver x receptor (LXR), peroxisome proliferator activated receptor gamma (PPARγ), Kruppel-like factor 6 (KLF-6) and transcription factor E3 (TCFE3) have been shown to control hepatic GCK transcription [60, 76–82]. Studies from our laboratory have indicated that the nuclear receptor liver receptor homolog 1 (LRH-1) coordinates multiple aspects of hepatic intermediary metabolism by regulating GCK-dependent G6P synthesis [60, 83]. While initially identified as a transcriptional regulator of cholesterol and bile salt homeostasis [84, 85], LRH-1 has recently emerged as a key integrator of hepatic glucose and fatty acid metabolism [60, 83, 86, 87]. LRH-1 contributes to basal GCK expression under fed and fasted conditions and its activity is not dependent on glucose. This was based on the finding that ectopic LRH-1 expression is sufficient to induce Gck expression in hepatoma cells, and that increasing glycemia fails to amplify LRH-1-mediated transcription [60]. Hepatic LRH-1 deficiency significantly perturbed the hepatic response to feeding, as illustrated by delayed glycogen synthesis, as well as reduced ChREBP expression and activity, which resulted in a strong attenuation of glycolysis and de novo fatty acid synthesis upon refeeding [60]. Importantly, these perturbations occurred secondary to reduced GCK activity, as GCK reconstitution restored ChREBP target gene expression in hepatocyte-specific LRH-1 knockout mice [60]. LRH-1-dependent glucose sensing in the liver also affected systemic glucose homeostasis. In liver-specific LRH-1 knockout mice impaired GCK-mediated glucose consumption triggered the pancreas to release more insulin, leading to elevated insulin levels and increased glucose disposal [60]. These findings place LRH-1 upstream of the central glucose-sensing system in the liver (Fig. 2).
Fig. 2.
LRH-1 is an upstream regulator of the central glucose-sensing system in the liver. ChREBP activation requires GCK-dependent synthesis of glucose metabolites. Because LRH-1 is a transcriptional regulator of GCK, it impacts postprandial G6P synthesis and ChREBP activity. Acetyl-CoA acetyl-coenzyme A, ChREBP carbohydrate response element binding protein, F2,6bisP fructose-2,6-bisphosphate, F6P fructose-6-phosphate, G6P glucose-6-phosphate, GCK glucokinase, LRH-1 liver receptor homolog 1, UDP-GlcNAc UDP-N-acetylglucosamine, X5P xylulose-5-phosphate
Similar functions have been attributed to LXR. Although LXR has been identified as a transcriptional regulator of both GCK and ChREBP [76, 88–92], its deficiency does not impair the hepatic response to carbohydrate refeeding or ChREBP activity [93, 94]. Further work will be necessary to establish whether LXR is essentially required for postprandial glucose sensing in the liver.
Glucose-sensitive modifications as potential glucose sensors in the liver
Post-translational modifications of regulatory proteins allow for adaptive responses to a variety of metabolic cues [95, 96]. Interestingly, some post-translational modifications are closely linked to glucose metabolism and target metabolic enzymes, components of cellular signal transduction pathways as well as transcription factors and their co-regulators (reviewed in [97, 98]). These modifications are typically enzyme-catalyzed, but can also occur through non-enzymatic interaction between metabolites and proteins. Although enzyme-mediated transfer of glucose metabolites has been investigated most intensively, a very recent study has identified a glucose-sensitive and enzyme-independent post-translational modification that controls hepatocyte function [99]. It is now also increasingly recognized that glucose metabolism can induce epigenetic changes through glucose-dependent post-translational modification of histone proteins (reviewed in [96, 100–102]). Because the composition of the histone code determines the degree of chromatin condensation, glucose-dependent modification of histones may alter the accessibility for transcription factors and regulatory enzymes that may ultimately translate into changes in transcriptional activity. In this section, we will discus two enzyme-mediated glucose-sensitive post-translational modifications that target regulatory proteins and epigenome, and may as such contribute to glucose sensing in the liver. The metabolic origins, enzymatics, and hepatic targets of these post-translational modifications are summarized in Fig. 3.
Fig. 3.
Working model depicting the metabolic origins, enzymatics, and targets of glucose-sensitive post-translational modifications in the liver. a The hexosamine biosynthesis pathway uses F6P, glutamine, and UTP for O-linked β-N-acetylglucosaminylation. b Glycolysis can link glucose metabolism to acetylation. Acetyl-CoA acetyl-coenzyme A, ChREBP carbohydrate response element binding protein, CRTC2 cAMP-regulated transcriptional co-activator 2, FOXA2 forkhead box protein A2, FOXO1 forkhead box protein O1, F6P fructose-6-phosphate, FXR farnesoid x receptor, HDAC histone deacetylase, KAT lysine acetyltransferases, LXR liver x receptor, OGA O-GlcNAcase, OGT O-GlcNAc transferase, PGC-1α peroxisome proliferator-activated receptor gamma coactivator 1-alpha, SREBP-1c sterol regulatory element binding protein-1c, TCA tricarboxylic acid, UTP uridine triphosphate
O-GlcNAcylation of serine and threonine residues is a modification that occurs in the cytoplasm, nucleus, and mitochondria [103]. The substrate, UDP-N-acetylglucosamine (UDP-GlcNAc), is generated by the hexosamine biosynthesis pathway, a branch of hepatic glucose metabolism that uses F6P, glutamine, acetyl-CoA, and uridine triphosphate (UTP) [104]. The addition and removal of UDP-GlcNAc is catalyzed by two enzymes. O-GlcNAc transferase (OGT) mediates the addition of UDP-GlcNAc to target proteins while O-GlcNAcase (OGA) catalyzes its removal [105, 106]. Both OGT and OGA are encoded by single genes that are alternatively spliced in mammals, and the different isoforms are located in separate subcellular compartments [105, 107–110]. Their activities are regulated by protein–protein interactions and post-translational modifications including O-GlcNAcylation, however this domain is as yet largely unexplored [111]. O-GlcNAcylation is considered as a unique glucose-sensitive post-translational modification [112] and has wide-ranging effects on transcription, protein activity, and stability as well as on epigenetic and genomic imprinting (reviewed in [113]). In hepatocytes, O-GlcNAcylation has mainly been studied in relation to its role in metabolism. Recent work has shown that hepatic OGT is required to maintain circadian control of glucose homeostasis by regulating the clock system in the liver [114, 115]. OGT also targets metabolic transcriptional regulators such as LXR [90] and cAMP-regulated transcriptional co-activator 2 (CRTC2), a coregulator of the gluconeogenic transcription factor cyclic AMP-responsive element binding protein (CREB) [116]. Moreover, the activity of two other key gluconeogenic regulators, i.e., forkhead box protein O (FOXO1) and peroxisome proliferator-activated receptor gamma coactivator 1-alpha (PGC-1α), is regulated by O-GlcNAcylation [117–119]. Another key finding is that OGT modifies multiple nodes of the insulin signaling pathway [108, 120]. Interestingly, under normoglycemic conditions O-GlcNAcylation contributes to insulin signaling [121], while it induces insulin resistance when chronically activated [120]. Although these studies point to a general role for O-GlcNAcylation in regulating glucose homeostasis, strong evidence for a more specific role in glucose-sensing stems from the fact that O-GlcNAcylation activates hepatic ChREBP [68]. Finally, it should be mentioned that despite the fact that O-GlcNAcylation is emerging as a histone-modifying post-translational modification [122], there is currently no evidence that O-GlcNAcylation also contributes to hepatic glucose sensing via epigenetic regulation. As methodologies for high-throughput O-GlcNAc profiling are emerging [123, 124], more insight into the hepatic targets of O-GlcNAcylation and its potential contribution to hepatic glucose sensing is expected in the near future.
Acetylation is another post-translational modification that potentially reflects glucose availability. This modification involves the enzymatic transfer of acetyl-CoA, and is facilitated by lysine acetyltransferases (KATs) [125]. These enzymes act on the lysine residues of both histones and non-histone proteins in different cellular compartments. The reverse reaction is mediated by deacetylases, which can be divided into four classes. Class I, II, and IV deacetylases are considered as the classical histone deacetylases (HDACs). Class III deacetylases, better known as sirtuins, are structurally unrelated to HDACs. HDACs and sirtuins are localized in the mitochondria or cytoplasm, and are able to shuttle between the nucleus and the cytosol [126–129]. High-throughput analysis of human liver biopsies and liver cells has shown that many metabolic enzymes are acetylated [69, 130], either to modulate their activities or to direct them towards proteosomal or lysosomal degradation [131, 132]. Moreover, the activity of several transcriptional regulators of hepatic metabolism including ChREBP, LXR, CRTC2, PGC-1α, FOXO1, SREBP-1c, forkhead box protein A2 (FOXA2), and farnesoid x receptor (FXR), is known to be modified by acetylation [67, 133–140], in some cases in coordination with phosphorylation [134, 141]. Studies in yeast and mammalian cell cultures have shown that histone acetylation is dependent on subcellular acetyl-CoA concentrations [142–146]. Notably, glucose was shown to promote histone acetylation via ATP citrate lyase (ACL), the enzyme that generates acetyl-CoA from TCA-derived citrate in mammalian cell lines [142]. The existence of a similar mechanism in hepatocytes challenged with glucose would point to a glucose-sensing role of histone acetylation in liver but needs to be confirmed. The observation that both histones [147] and non-histone proteins [130] are dynamically acetylated in response to feeding/fasting cycles is also suggestive of glucose-dependent acetylation in liver. Moreover, it has been reported that hepatic acetyl-CoA levels increase upon short-term refeeding as compared to fasted conditions [148]. It should however be noted that besides being produced by decarboxylation of glycolytic pyruvate, hepatic acetyl-CoA can also be derived from fatty acid oxidation and amino acid metabolism. A dedicated analysis of acetylation profiles in glucose-challenged hepatocytes is therefore warranted to establish the impact of glucose metabolism on protein acetylation, as well as the potential contribution of protein acetylation to glucose sensing in the liver.
Metabolic liver diseases associated with aberrant glucose sensing
Glucose sensors enable the liver to respond to dynamic changes in glucose availability. However, when these sensors are chronically activated, they may predispose to the development of liver diseases.
During poorly controlled diabetes, the liver is frequently exposed to hyperglycemic episodes. In type 2 diabetes (T2D), GCK is constitutively active and GCK flux is increased secondary to elevated glucose concentrations [149, 150]. This leads to sustained activation of glucose sensors in the liver. For example, the hexosamine biosynthesis pathway normally accounts for less than 5 % of the hepatic glucose flux, yet its activity is markedly increased by hyperglycemia [151, 152]. Aberrant glucose sensing in T2D results in triglyceride accumulation and excessive glucose production in the liver [116, 153]. While triglyceride accumulation contributes to the development of liver steatosis, increased hepatic glucose output leads to a further increase in glycemia.
A clear association exists between hepatic steatosis and the pathogenesis of T2D, cardiovascular disease, and steatohepatitis [154]. It is, however, increasingly recognized that, up to a certain threshold, the accumulation of triglycerides may serve as a buffering system that would actually protect the liver against metabolic dysfunction [154]. In mice, ChREBP plays a key role in the development of hepatic steatosis in T2D [153], and hepatic ChREBP function is perturbed in obese and (pre-)diabetic subjects [155, 156]. Interestingly, a recent study revealed that ChREBP overexpression protects against diet-induced glucose intolerance and insulin resistance [157]. This finding indicates that under conditions of dietary fat overload, ChREBP-mediated lipogenesis likely contributes to a metabolically benign state by promoting mono-unsaturated fatty acid synthesis [154, 157]. It was furthermore shown that diabetic steatosis is associated with ChREBP hyperacetylation [67] and that hepatic lipid accumulation can be prevented when ChREBP O-GlcNAcylation is reduced [68]. Increased O-GlcNAcylation also contributes to uncontrolled hepatic glucose production under diabetic conditions. T2D is associated with increased O-GlcNAcylation levels of the gluconeogenic co-regulator CRTC2 and removal of O-GlcNAc from CRTC2 normalizes glycemia in diabetic mice [116]. Whether sustained CRTC2 acetylation levels also promote hepatic glucose production in diabetics remains to be established. Likewise, it is as yet unknown whether aberrant acetylation and O-GlcNAcylation of the gluconeogenic regulators FOXO1 and PGC-1α [117–119, 135, 158] directly contribute to hyperglycemia in T2D.
Glucose sensors may also become deregulated by inherited loss-of-function mutations in enzymes that regulate intrahepatic glucose metabolism. An example of such an “inborn error of metabolism” is Glycogen Storage Disease type 1 (GSD-1) [159] which is caused by loss of either G6Pc or G6Pt activity [160–162]. The primary consequences of perturbed hepatic G6Pase activity in GSD-1 are hypoglycemia and the accumulation of G6P in the liver [163–165]. In addition, GSD-1 is characterized by excessive glycogen and lipid storage in the liver [163, 165–167] as well as hyperlipidemia [165, 167–170]. Interestingly, GSD-1 is associated with a ChREBP-dependent increase in de novo fatty acid synthesis [163, 167, 170]. Combined, these observations indicate that sustained activation of hepatic glucose sensors by extrahepatic (diabetes) or intrahepatic (GSD-1) changes in glucose homeostasis predisposes to development of hepatic steatosis [171].
Another consequence of both T2D and GSD-1 is the increased incidence of liver tumorigenesis [172–175]. Although steatosis has been proposed as a predisposing factor for liver cancer [176, 177], altered metabolism may be the actual driving force for tumor development. It is well known that tumors require specific metabolic adaptations to support the bioenergetic and biosynthetic demands of growth and proliferation [178]. More specifically, a switch to non-oxidative glucose metabolism combined with a predominant anabolic role of the TCA cycle are considered as major hallmarks of cancer metabolism [179]. T2D and GSD-1 are characterized by a high flux from hepatic G6P towards glycolysis and lipid- and nucleotide biosynthesis. The exact mechanisms by which these metabolic adaptations confer a preneoplastic status to hepatocytes and direct them towards tumorigenesis are incompletely understood [165, 172, 179]. Glucose sensors likely play an important role here. In support of this hypothesis, ChREBP mediates the switch towards pro-oncogenic metabolism in proliferating cells [180]. Moreover, ChREBP functionally interacts with the pro-oncogenic transcription factor c-Myc, which is critical for ChREBP-dependent glucose sensing in the liver [45, 181, 182]. Because mouse models of T2D and GSD-1 exhibit increased hepatic ChREBP activity [153, 163], ChREBP may play a key role in the pathophysiology of liver tumor development in these diseased states. The potential existence of such a mechanism urges for the exploration of a potential oncogenic role of hepatic LRH-1, a potent upstream regulator of the GCK–ChREBP axis in the liver ([60] and Fig. 2) and a key player in the development of colorectal, breast and pancreatic cancers [183–185]. Finally, acetylation and O-GlcNAcylation have recently emerged as critical modifiers of the activity of metabolic enzymes as well as of oncogenes and tumor suppressors in cancer cells [123, 186–190]. These glucose-sensing post-translational modifications may therefore direct hepatocytes towards a pro-oncogenic state under conditions of excessive hepatic glucose metabolism [142, 191, 192].
Conclusions and future directions
Hepatic glucose sensing is critical for an adequate postprandial response and the maintenance of glycemic control. However, it may also contribute to liver pathology under conditions of excessive intrahepatic glucose metabolism. Research in the past years has identified GCK–ChREBP as the central glucose-sensing system in the liver. Further exploration of the mechanisms by which different glucose metabolites activate hepatic ChREBP, and the function of the different ChREBP isoforms are required to unravel the mechanistic basis of the glucose-sensing axis. In addition, it remains to be established whether ChREBP’s paralog MondoA, which can be activated by G6P and F2,6bisP [49, 56], also contributes to postprandial glucose sensing in the liver.
Because LRH-1 has recently emerged as a potent upstream regulator of the GCK–ChREBP axis, modulation of its activity may provide opportunities for the treatment of diseases that are characterized by aberrant hepatic glucose sensing. LRH-1 transcriptional activity can be modified by post-translational modifications or by agonists/antagonist binding, and depends on its interaction with co-regulators [83, 183, 193–204]. Detailed insight into these processes is therefore needed to define strategies that target LRH-1-dependent glucose sensing.
Finally, dedicated studies are required to uncover the exact role of O-GlcNAcylation and acetylation in hepatic glucose sensing. Systematic analysis of glucose-dependent responses in the absence of OGT or KAT activity will establish to what extent, and via which mechanisms these post-translational modifications contribute to glucose-sensing system in the liver. Moreover, there is extensive crosstalk between post-translational modifications, and different combinations of post-translational modifications on a single target may lead to distinct biological outcomes [112, 205]. Future research will likely uncover novel interplays between GlcNAcylation/acetylation and other post-translational modifications including protein ubiquitination and methylation (reviewed in [206, 207]). Such crosstalk may in turn unveil unexpected functions and consequences of chronically activated hepatic glucose sensors that go beyond metabolism.
Acknowledgments
We thank Albert K. Groen for reading and commenting on the manuscript. The work in the laboratory of the authors is supported by the Ecole Polytechnique Fédérale de Lausanne (EPFL), the Swiss National Science Foundation, the Swiss Cancer League and the University Medical Center Groningen (UMCG).
Abbreviations
- Acetyl-CoA
Acetyl-coenzyme A
- ACL
ATP citrate lyase
- ChoRE
Carbohydrate response element
- ChREBP
Carbohydrate response element binding protein
- CREB
Cyclic AMP-responsive element binding protein
- CRTC2
cAMP-regulated transcriptional co-activator 2
- F2
6bisP, fructose-2,6-bisphosphate
- F6P
Fructose-6-phosphate
- FOXA2
Forkhead box protein A2
- FOXO1
Forkhead box protein O1
- FXR
Farnesoid x receptor
- G6P
Glucose-6-phosphate
- G6Pc
Glucose-6-phosphatase
- G6Pt
Glucose-6-phosphate transporter
- GCK
Glucokinase
- GCKR
GCK regulatory protein
- GLUT
Glucose transporter
- GSD-1
Glycogen storage disease type 1
- HDAC
Histone deacetylase
- HIF-1
Hypoxia-inducible factor 1
- HK
Hexokinase
- HNF-4
Hepatocyte nuclear factor 4
- KAT
Lysine acetyltransferase
- KLF-6
Kruppel-like factor 6
- LRH-1
Liver receptor homolog 1
- LXR
Liver x receptor
- Mlx
Max-like protein X
- MLXIP
Mlx interacting protein
- MLXIPL
Max-like protein X interacting protein-like
- OGA
O-GlcNAcase
- O-GlcNAcylation
O-linked β-N-acetylglucosaminylation
- OGT
O-GlcNAc transferase
- PGC-1α
Peroxisome proliferator-activated receptor gamma coactivator 1-alpha
- PPARγ
Peroxisome proliferator activated receptor gamma
- SREBP-1c
Sterol regulatory binding protein-1c
- T2D
Type 2 diabetes
- TCA
Tricarboxylic acid
- TCFE3
Transcription factor E3
- UDP-GlcNAc
UDP-N-acetylglucosamine
- UTP
Uridine triphosphate
- X5P
Xylulose-5-phosphate
References
- 1.Towle HC. Glucose as a regulator of eukaryotic gene transcription. Trends Endocrinol Metab. 2005;16(10):489–494. doi: 10.1016/j.tem.2005.10.003. [DOI] [PubMed] [Google Scholar]
- 2.Daly ME, Vale C, Walker M, Littlefield A, Alberti KG, Mathers JC. Acute effects on insulin sensitivity and diurnal metabolic profiles of a high-sucrose compared with a high-starch diet. Am J Clin Nutr. 1998;67(6):1186–1196. doi: 10.1093/ajcn/67.6.1186. [DOI] [PubMed] [Google Scholar]
- 3.Klover PJ, Mooney RA. Hepatocytes: critical for glucose homeostasis. Int J Biochem Cell Biol. 2004;36(5):753–758. doi: 10.1016/j.biocel.2003.10.002. [DOI] [PubMed] [Google Scholar]
- 4.Mithieux G. Brain, liver, intestine: a triumvirate to coordinate insulin sensitivity of endogenous glucose production. Diabetes Metab. 2010;36(Suppl 3):S50–S53. doi: 10.1016/S1262-3636(10)70467-5. [DOI] [PubMed] [Google Scholar]
- 5.Taniguchi CM, Emanuelli B, Kahn CR. Critical nodes in signalling pathways: insights into insulin action. Nat Rev Mol Cell Biol. 2006;7(2):85–96. doi: 10.1038/nrm1837. [DOI] [PubMed] [Google Scholar]
- 6.Ramnanan CJ, Edgerton DS, Kraft G, Cherrington AD. Physiologic action of glucagon on liver glucose metabolism. Diabetes Obes Metab. 2011;13(Suppl 1):118–125. doi: 10.1111/j.1463-1326.2011.01454.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Capaldo B, Gastaldelli A, Antoniello S, Auletta M, Pardo F, Ciociaro D, Guida R, Ferrannini E, Sacca L. Splanchnic and leg substrate exchange after ingestion of a natural mixed meal in humans. Diabetes. 1999;48(5):958–966. doi: 10.2337/diabetes.48.5.958. [DOI] [PubMed] [Google Scholar]
- 8.Ferrannini E, Bjorkman O, Reichard GA, Jr, Pilo A, Olsson M, Wahren J, DeFronzo RA. The disposal of an oral glucose load in healthy subjects. A quantitative study. Diabetes. 1985;34(6):580–588. doi: 10.2337/diab.34.6.580. [DOI] [PubMed] [Google Scholar]
- 9.Woerle HJ, Meyer C, Dostou JM, Gosmanov NR, Islam N, Popa E, Wittlin SD, Welle SL, Gerich JE. Pathways for glucose disposal after meal ingestion in humans. Am J Physiol Endocrinol Metab. 2003;284(4):E716–E725. doi: 10.1152/ajpendo.00365.2002. [DOI] [PubMed] [Google Scholar]
- 10.Moore MC, Pagliassotti MJ, Swift LL, Asher J, Murrell J, Neal D, Cherrington AD. Disposition of a mixed meal by the conscious dog. Am J Physiol. 1994;266(4 Pt 1):E666–E675. doi: 10.1152/ajpendo.1994.266.4.E666. [DOI] [PubMed] [Google Scholar]
- 11.Mueckler M, Thorens B. The SLC2 (GLUT) family of membrane transporters. Mol Aspects Med. 2013;34(2–3):121–138. doi: 10.1016/j.mam.2012.07.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Aschenbach JR, Steglich K, Gabel G, Honscha KU. Expression of mRNA for glucose transport proteins in jejunum, liver, kidney and skeletal muscle of pigs. J Physiol Biochem. 2009;65(3):251–266. doi: 10.1007/BF03180578. [DOI] [PubMed] [Google Scholar]
- 13.Seyer P, Vallois D, Poitry-Yamate C, Schutz F, Metref S, Tarussio D, Maechler P, Staels B, Lanz B, Grueter R, Decaris J, Turner S, da Costa A, Preitner F, Minehira K, Foretz M, Thorens B. Hepatic glucose sensing is required to preserve beta cell glucose competence. J Clin Investig. 2013 doi: 10.1172/JCI65538. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Burcelin R, del Carmen Munoz M, Guillam MT, Thorens B. Liver hyperplasia and paradoxical regulation of glycogen metabolism and glucose-sensitive gene expression in GLUT2-null hepatocytes. Further evidence for the existence of a membrane-based glucose release pathway. J Biol Chem. 2000;275(15):10930–10936. doi: 10.1074/jbc.275.15.10930. [DOI] [PubMed] [Google Scholar]
- 15.Burcelin R, Dolci W, Thorens B. Glucose sensing by the hepatoportal sensor is GLUT2-dependent: in vivo analysis in GLUT2-null mice. Diabetes. 2000;49(10):1643–1648. doi: 10.2337/diabetes.49.10.1643. [DOI] [PubMed] [Google Scholar]
- 16.Williams TF, Exton JH, Park CR, Regen DM. Stereospecific transport of glucose in the perfused rat liver. Am J Physiol. 1968;215(5):1200–1209. doi: 10.1152/ajplegacy.1968.215.5.1200. [DOI] [PubMed] [Google Scholar]
- 17.Wilson JE. Isozymes of mammalian hexokinase: structure, subcellular localization and metabolic function. J Exp Biol. 2003;206(Pt 12):2049–2057. doi: 10.1242/jeb.00241. [DOI] [PubMed] [Google Scholar]
- 18.Cardenas ML, Cornish-Bowden A, Ureta T. Evolution and regulatory role of the hexokinases. Biochim Biophys Acta. 1998;1401(3):242–264. doi: 10.1016/s0167-4889(97)00150-x. [DOI] [PubMed] [Google Scholar]
- 19.Wera S, Bollen M, Moens L, Stalmans W. Time-dependent pseudo-activation of hepatic glycogen synthase b by glucose 6-phosphate without involvement of protein phosphatases. Biochem J. 1996;315(Pt 1):91–96. doi: 10.1042/bj3150091. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Aiston S, Green A, Mukhtar M, Agius L. Glucose 6-phosphate causes translocation of phosphorylase in hepatocytes and inactivates the enzyme synergistically with glucose. Biochem J. 2004;377(Pt 1):195–204. doi: 10.1042/BJ20031191. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.von Wilamowitz-Moellendorff A, Hunter RW, Garcia-Rocha M, Kang L, Lopez-Soldado I, Lantier L, Patel K, Peggie MW, Martinez-Pons C, Voss M, Calbo J, Cohen PT, Wasserman DH, Guinovart JJ, Sakamoto K. Glucose-6-phosphate-mediated activation of liver glycogen synthase plays a key role in hepatic glycogen synthesis. Diabetes. 2013 doi: 10.2337/db13-0880. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Owen OE, Kalhan SC, Hanson RW. The key role of anaplerosis and cataplerosis for citric acid cycle function. J Biol Chem. 2002;277(34):30409–30412. doi: 10.1074/jbc.R200006200. [DOI] [PubMed] [Google Scholar]
- 23.Guillam MT, Burcelin R, Thorens B. Normal hepatic glucose production in the absence of GLUT2 reveals an alternative pathway for glucose release from hepatocytes. Proc Natl Acad Sci USA. 1998;95(21):12317–12321. doi: 10.1073/pnas.95.21.12317. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Van Schaftingen E, Detheux M, Veiga da Cunha M. Short-term control of glucokinase activity: role of a regulatory protein. FASEB J. 1994;8(6):414–419. doi: 10.1096/fasebj.8.6.8168691. [DOI] [PubMed] [Google Scholar]
- 25.Heredia VV, Thomson J, Nettleton D, Sun S. Glucose-induced conformational changes in glucokinase mediate allosteric regulation: transient kinetic analysis. Biochemistry. 2006;45(24):7553–7562. doi: 10.1021/bi060253q. [DOI] [PubMed] [Google Scholar]
- 26.Iynedjian PB, Marie S, Gjinovci A, Genin B, Deng SP, Buhler L, Morel P, Mentha G. Glucokinase and cytosolic phosphoenolpyruvate carboxykinase (GTP) in the human liver. Regulation of gene expression in cultured hepatocytes. J Clin Invest. 1995;95(5):1966–1973. doi: 10.1172/JCI117880. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Postic C, Niswender KD, Decaux JF, Parsa R, Shelton KD, Gouhot B, Pettepher CC, Granner DK, Girard J, Magnuson MA. Cloning and characterization of the mouse glucokinase gene locus and identification of distal liver-specific DNase I hypersensitive sites. Genomics. 1995;29(3):740–750. doi: 10.1006/geno.1995.9943. [DOI] [PubMed] [Google Scholar]
- 28.Dentin R, Pegorier JP, Benhamed F, Foufelle F, Ferre P, Fauveau V, Magnuson MA, Girard J, Postic C. Hepatic glucokinase is required for the synergistic action of ChREBP and SREBP-1c on glycolytic and lipogenic gene expression. J Biol Chem. 2004;279(19):20314–20326. doi: 10.1074/jbc.M312475200. [DOI] [PubMed] [Google Scholar]
- 29.O’Doherty RM, Lehman DL, Seoane J, Gomez-Foix AM, Guinovart JJ, Newgard CB. Differential metabolic effects of adenovirus-mediated glucokinase and hexokinase I overexpression in rat primary hepatocytes. J Biol Chem. 1996;271(34):20524–20530. doi: 10.1074/jbc.271.34.20524. [DOI] [PubMed] [Google Scholar]
- 30.Seoane J, Gomez-Foix AM, O’Doherty RM, Gomez-Ara C, Newgard CB, Guinovart JJ. Glucose 6-phosphate produced by glucokinase, but not hexokinase I, promotes the activation of hepatic glycogen synthase. J Biol Chem. 1996;271(39):23756–23760. doi: 10.1074/jbc.271.39.23756. [DOI] [PubMed] [Google Scholar]
- 31.Iynedjian PB. Molecular physiology of mammalian glucokinase. Cell Mol Life Sci. 2009;66(1):27–42. doi: 10.1007/s00018-008-8322-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Sommercorn J, Steward T, Freedland RA. Activation of phosphofructokinase from rat tissues by 6-phosphogluconate and fructose 2,6-bisphosphate. Arch Biochem Biophys. 1984;232(2):579–584. doi: 10.1016/0003-9861(84)90576-9. [DOI] [PubMed] [Google Scholar]
- 33.Sawada M, Mitsui Y, Sugiya H, Furuyama S. Ribose 1,5-bisphosphate is a putative regulator of fructose 6-phosphate/fructose 1,6-bisphosphate cycle in liver. Int J Biochem Cell Biol. 2000;32(4):447–454. doi: 10.1016/s1357-2725(99)00137-5. [DOI] [PubMed] [Google Scholar]
- 34.Wallace JC, Jitrapakdee S, Chapman-Smith A. Pyruvate carboxylase. Int J Biochem Cell Biol. 1998;30(1):1–5. doi: 10.1016/s1357-2725(97)00147-7. [DOI] [PubMed] [Google Scholar]
- 35.Schrenk DF, Bisswanger H. Measurements of electron spin resonance with the pyruvate dehydrogenase complex from Escherichia coli. Studies on the allosteric binding site of acetyl-coenzyme A. Eur J Biochem. 1984;143(3):561–566. doi: 10.1111/j.1432-1033.1984.tb08406.x. [DOI] [PubMed] [Google Scholar]
- 36.Desvergne B, Michalik L, Wahli W. Transcriptional regulation of metabolism. Physiol Rev. 2006;86(2):465–514. doi: 10.1152/physrev.00025.2005. [DOI] [PubMed] [Google Scholar]
- 37.Francis GA, Fayard E, Picard F, Auwerx J. Nuclear receptors and the control of metabolism. Annu Rev Physiol. 2003;65:261–311. doi: 10.1146/annurev.physiol.65.092101.142528. [DOI] [PubMed] [Google Scholar]
- 38.McKenna NJ, Cooney AJ, DeMayo FJ, Downes M, Glass CK, Lanz RB, Lazar MA, Mangelsdorf DJ, Moore DD, Qin J, Steffen DL, Tsai MJ, Tsai SY, Yu R, Margolis RN, Evans RM, O’Malley BW. Minireview: evolution of NURSA, the nuclear receptor signaling atlas. Mol Endocrinol. 2009;23(6):740–746. doi: 10.1210/me.2009-0135. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Jeong YS, Kim D, Lee YS, Kim HJ, Han JY, Im SS, Chong HK, Kwon JK, Cho YH, Kim WK, Osborne TF, Horton JD, Jun HS, Ahn YH, Ahn SM, Cha JY. Integrated expression profiling and genome-wide analysis of ChREBP targets reveals the dual role for ChREBP in glucose-regulated gene expression. PLoS One. 2011;6(7):e22544. doi: 10.1371/journal.pone.0022544. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Ma L, Robinson LN, Towle HC. ChREBP*Mlx is the principal mediator of glucose-induced gene expression in the liver. J Biol Chem. 2006;281(39):28721–28730. doi: 10.1074/jbc.M601576200. [DOI] [PubMed] [Google Scholar]
- 41.Thompson KS, Towle HC. Localization of the carbohydrate response element of the rat L-type pyruvate kinase gene. J Biol Chem. 1991;266(14):8679–8682. [PubMed] [Google Scholar]
- 42.Shih HM, Liu Z, Towle HC. Two CACGTG motifs with proper spacing dictate the carbohydrate regulation of hepatic gene transcription. J Biol Chem. 1995;270(37):21991–21997. doi: 10.1074/jbc.270.37.21991. [DOI] [PubMed] [Google Scholar]
- 43.Yamashita H, Takenoshita M, Sakurai M, Bruick RK, Henzel WJ, Shillinglaw W, Arnot D, Uyeda K. A glucose-responsive transcription factor that regulates carbohydrate metabolism in the liver. Proc Natl Acad Sci USA. 2001;98(16):9116–9121. doi: 10.1073/pnas.161284298. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Billin AN, Eilers AL, Coulter KL, Logan JS, Ayer DE. MondoA, a novel basic helix-loop-helix-leucine zipper transcriptional activator that constitutes a positive branch of a max-like network. Mol Cell Biol. 2000;20(23):8845–8854. doi: 10.1128/mcb.20.23.8845-8854.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Peterson CW, Ayer DE. An extended Myc network contributes to glucose homeostasis in cancer and diabetes. Front Biosci. 2011;16:2206–2223. doi: 10.2741/3848. [DOI] [PubMed] [Google Scholar]
- 46.Billin AN, Eilers AL, Queva C, Ayer DE. Mlx, a novel Max-like BHLHZip protein that interacts with the Max network of transcription factors. J Biol Chem. 1999;274(51):36344–36350. doi: 10.1074/jbc.274.51.36344. [DOI] [PubMed] [Google Scholar]
- 47.Stoeckman AK, Ma L, Towle HC. Mlx is the functional heteromeric partner of the carbohydrate response element-binding protein in glucose regulation of lipogenic enzyme genes. J Biol Chem. 2004;279(15):15662–15669. doi: 10.1074/jbc.M311301200. [DOI] [PubMed] [Google Scholar]
- 48.Ma L, Tsatsos NG, Towle HC. Direct role of ChREBP.Mlx in regulating hepatic glucose-responsive genes. J Biol Chem. 2005;280(12):12019–12027. doi: 10.1074/jbc.M413063200. [DOI] [PubMed] [Google Scholar]
- 49.Petrie JL, Al-Oanzi ZH, Arden C, Tudhope SJ, Mann J, Kieswich J, Yaqoob MM, Towle HC, Agius L. Glucose induces protein targeting to glycogen in hepatocytes by fructose 2,6-bisphosphate-mediated recruitment of MondoA to the promoter. Mol Cell Biol. 2013;33(4):725–738. doi: 10.1128/MCB.01576-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.McFerrin LG, Atchley WR. Evolution of the Max and Mlx networks in animals. Genome Biol Evol. 2011;3:915–937. doi: 10.1093/gbe/evr082. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Yuan J, Tirabassi RS, Bush AB, Cole MD. The C. elegans MDL-1 and MXL-1 proteins can functionally substitute for vertebrate MAD and MAX. Oncogene. 1998;17(9):1109–1118. doi: 10.1038/sj.onc.1202036. [DOI] [PubMed] [Google Scholar]
- 52.McFerrin LG, Atchley WR. A novel N-terminal domain may dictate the glucose response of Mondo proteins. PLoS One. 2012;7(4):e34803. doi: 10.1371/journal.pone.0034803. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Li MV, Chang B, Imamura M, Poungvarin N, Chan L. Glucose-dependent transcriptional regulation by an evolutionarily conserved glucose-sensing module. Diabetes. 2006;55(5):1179–1189. doi: 10.2337/db05-0822. [DOI] [PubMed] [Google Scholar]
- 54.Sassu ED, McDermott JE, Keys BJ, Esmaeili M, Keene AC, Birnbaum MJ, DiAngelo JR. Mio/dChREBP coordinately increases fat mass by regulating lipid synthesis and feeding behavior in Drosophila. Biochem Biophys Res Commun. 2012;426(1):43–48. doi: 10.1016/j.bbrc.2012.08.028. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Havula E, Teesalu M, Hyotylainen T, Seppala H, Hasygar K, Auvinen P, Oresic M, Sandmann T, Hietakangas V. Mondo/ChREBP-Mlx-regulated transcriptional network is essential for dietary sugar tolerance in Drosophila. PLoS Genet. 2013;9(4):e1003438. doi: 10.1371/journal.pgen.1003438. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Stoltzman CA, Peterson CW, Breen KT, Muoio DM, Billin AN, Ayer DE. Glucose sensing by MondoA:Mlx complexes: a role for hexokinases and direct regulation of thioredoxin-interacting protein expression. Proc Natl Acad Sci USA. 2008;105(19):6912–6917. doi: 10.1073/pnas.0712199105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Filhoulaud G, Guilmeau S, Dentin R, Girard J, Postic C. Novel insights into ChREBP regulation and function. Trends Endocrinol Metab. 2013 doi: 10.1016/j.tem.2013.01.003. [DOI] [PubMed] [Google Scholar]
- 58.Noriega LG, Feige JN, Canto C, Yamamoto H, Yu J, Herman MA, Mataki C, Kahn BB, Auwerx J. CREB and ChREBP oppositely regulate SIRT1 expression in response to energy availability. EMBO Rep. 2011;12(10):1069–1076. doi: 10.1038/embor.2011.151. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59.Arden C, Petrie JL, Tudhope SJ, Al-Oanzi Z, Claydon AJ, Beynon RJ, Towle HC, Agius L. Elevated glucose represses liver glucokinase and induces its regulatory protein to safeguard hepatic phosphate homeostasis. Diabetes. 2011;60(12):3110–3120. doi: 10.2337/db11-0061. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Oosterveer MH, Mataki C, Yamamoto H, Harach T, Moullan N, van Dijk TH, Ayuso E, Bosch F, Postic C, Groen AK, Auwerx J, Schoonjans K. LRH-1-dependent glucose sensing determines intermediary metabolism in liver. J Clin Invest. 2012;122(8):2817–2826. doi: 10.1172/JCI62368. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Kabashima T, Kawaguchi T, Wadzinski BE, Uyeda K. Xylulose 5-phosphate mediates glucose-induced lipogenesis by xylulose 5-phosphate-activated protein phosphatase in rat liver. Proc Natl Acad Sci USA. 2003;100(9):5107–5112. doi: 10.1073/pnas.0730817100. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62.Li MV, Chen W, Harmancey RN, Nuotio-Antar AM, Imamura M, Saha P, Taegtmeyer H, Chan L. Glucose-6-phosphate mediates activation of the carbohydrate responsive binding protein (ChREBP) Biochem Biophys Res Commun. 2010;395(3):395–400. doi: 10.1016/j.bbrc.2010.04.028. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Dentin R, Tomas-Cobos L, Foufelle F, Leopold J, Girard J, Postic C, Ferre P. Glucose 6-phosphate, rather than xylulose 5-phosphate, is required for the activation of ChREBP in response to glucose in the liver. J Hepatol. 2012;56(1):199–209. doi: 10.1016/j.jhep.2011.07.019. [DOI] [PubMed] [Google Scholar]
- 64.Arden C, Tudhope SJ, Petrie JL, Al-Oanzi ZH, Cullen KS, Lange AJ, Towle HC, Agius L. Fructose 2,6-bisphosphate is essential for glucose-regulated gene transcription of glucose-6-phosphatase and other ChREBP target genes in hepatocytes. Biochem J. 2012;443(1):111–123. doi: 10.1042/BJ20111280. [DOI] [PubMed] [Google Scholar]
- 65.Sakiyama H, Wynn RM, Lee WR, Fukasawa M, Mizuguchi H, Gardner KH, Repa JJ, Uyeda K. Regulation of nuclear import/export of carbohydrate response element-binding protein (ChREBP): interaction of an alpha-helix of ChREBP with the 14-3-3 proteins and regulation by phosphorylation. J Biol Chem. 2008;283(36):24899–24908. doi: 10.1074/jbc.M804308200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.Davies MN, O’Callaghan BL, Towle HC. Glucose activates ChREBP by increasing its rate of nuclear entry and relieving repression of its transcriptional activity. J Biol Chem. 2008;283(35):24029–24038. doi: 10.1074/jbc.M801539200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67.Bricambert J, Miranda J, Benhamed F, Girard J, Postic C, Dentin R. Salt-inducible kinase 2 links transcriptional coactivator p300 phosphorylation to the prevention of ChREBP-dependent hepatic steatosis in mice. J Clin Invest. 2010;120(12):4316–4331. doi: 10.1172/JCI41624. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Guinez C, Filhoulaud G, Rayah-Benhamed F, Marmier S, Dubuquoy C, Dentin R, Moldes M, Burnol AF, Yang X, Lefebvre T, Girard J, Postic C. O-GlcNAcylation increases ChREBP protein content and transcriptional activity in the liver. Diabetes. 2011;60(5):1399–1413. doi: 10.2337/db10-0452. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Zhao S, Xu W, Jiang W, Yu W, Lin Y, Zhang T, Yao J, Zhou L, Zeng Y, Li H, Li Y, Shi J, An W, Hancock SM, He F, Qin L, Chin J, Yang P, Chen X, Lei Q, Xiong Y, Guan KL. Regulation of cellular metabolism by protein lysine acetylation. Science. 2010;327(5968):1000–1004. doi: 10.1126/science.1179689. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70.Hanover JA, Cohen CK, Willingham MC, Park MK. O-linked N-acetylglucosamine is attached to proteins of the nuclear pore. Evidence for cytoplasmic and nucleoplasmic glycoproteins. J Biol Chem. 1987;262(20):9887–9894. [PubMed] [Google Scholar]
- 71.Holt GD, Hart GW. The subcellular distribution of terminal N-acetylglucosamine moieties. Localization of a novel protein-saccharide linkage, O-linked GlcNAc. J Biol Chem. 1986;261(17):8049–8057. [PubMed] [Google Scholar]
- 72.Herman MA, Peroni OD, Villoria J, Schon MR, Abumrad NA, Bluher M, Klein S, Kahn BB. A novel ChREBP isoform in adipose tissue regulates systemic glucose metabolism. Nature. 2012;484(7394):333–338. doi: 10.1038/nature10986. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.Postic C, Shiota M, Niswender KD, Jetton TL, Chen Y, Moates JM, Shelton KD, Lindner J, Cherrington AD, Magnuson MA. Dual roles for glucokinase in glucose homeostasis as determined by liver and pancreatic beta cell-specific gene knock-outs using Cre recombinase. J Biol Chem. 1999;274(1):305–315. doi: 10.1074/jbc.274.1.305. [DOI] [PubMed] [Google Scholar]
- 74.Iynedjian PB. Mammalian glucokinase and its gene. Biochem J. 1993;293(Pt 1):1–13. doi: 10.1042/bj2930001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75.Nelson JD, LeBoeuf RC, Bomsztyk K. Direct recruitment of insulin receptor and ERK signaling cascade to insulin-inducible gene loci. Diabetes. 2011;60(1):127–137. doi: 10.2337/db09-1806. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76.Kim TH, Kim H, Park JM, Im SS, Bae JS, Kim MY, Yoon HG, Cha JY, Kim KS, Ahn YH. Interrelationship between liver X receptor alpha, sterol regulatory element-binding protein-1c, peroxisome proliferator-activated receptor gamma, and small heterodimer partner in the transcriptional regulation of glucokinase gene expression in liver. J Biol Chem. 2009;284(22):15071–15083. doi: 10.1074/jbc.M109.006742. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 77.Kim MY, Jo SH, Park JM, Kim TH, Im SS, Ahn YH. Adenovirus-mediated overexpression of Tcfe3 ameliorates hyperglycaemia in a mouse model of diabetes by upregulating glucokinase in the liver. Diabetologia. 2013;56(3):635–643. doi: 10.1007/s00125-012-2807-7. [DOI] [PubMed] [Google Scholar]
- 78.Bechmann LP, Gastaldelli A, Vetter D, Patman GL, Pascoe L, Hannivoort RA, Lee UE, Fiel I, Munoz U, Ciociaro D, Lee YM, Buzzigoli E, Miele L, Hui KY, Bugianesi E, Burt AD, Day CP, Mari A, Agius L, Walker M, Friedman SL, Reeves HL. Glucokinase links Kruppel-like factor 6 to the regulation of hepatic insulin sensitivity in nonalcoholic fatty liver disease. Hepatology. 2012;55(4):1083–1093. doi: 10.1002/hep.24793. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79.Roth U, Jungermann K, Kietzmann T. Modulation of glucokinase expression by hypoxia-inducible factor 1 and upstream stimulatory factor 2 in primary rat hepatocytes. Biol Chem. 2004;385(3–4):239–247. doi: 10.1515/BC.2004.018. [DOI] [PubMed] [Google Scholar]
- 80.Roth U, Curth K, Unterman TG, Kietzmann T. The transcription factors HIF-1 and HNF-4 and the coactivator p300 are involved in insulin-regulated glucokinase gene expression via the phosphatidylinositol 3-kinase/protein kinase B pathway. J Biol Chem. 2004;279(4):2623–2631. doi: 10.1074/jbc.M308391200. [DOI] [PubMed] [Google Scholar]
- 81.Roth U, Jungermann K, Kietzmann T. Activation of glucokinase gene expression by hepatic nuclear factor 4alpha in primary hepatocytes. Biochem J. 2002;365(Pt 1):223–228. doi: 10.1042/BJ20020340. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82.Ganjam GK, Dimova EY, Unterman TG, Kietzmann T. FoxO1 and HNF-4 are involved in regulation of hepatic glucokinase gene expression by resveratrol. J Biol Chem. 2009;284(45):30783–30797. doi: 10.1074/jbc.M109.045260. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83.Lee JM, Lee YK, Mamrosh JL, Busby SA, Griffin PR, Pathak MC, Ortlund EA, Moore DD. A nuclear-receptor-dependent phosphatidylcholine pathway with antidiabetic effects. Nature. 2011;474(7352):506–510. doi: 10.1038/nature10111. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84.Mataki C, Magnier BC, Houten SM, Annicotte JS, Argmann C, Thomas C, Overmars H, Kulik W, Metzger D, Auwerx J, Schoonjans K. Compromised intestinal lipid absorption in mice with a liver-specific deficiency of liver receptor homolog 1. Mol Cell Biol. 2007;27(23):8330–8339. doi: 10.1128/MCB.00852-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 85.Lee YK, Schmidt DR, Cummins CL, Choi M, Peng L, Zhang Y, Goodwin B, Hammer RE, Mangelsdorf DJ, Kliewer SA. Liver receptor homolog-1 regulates bile acid homeostasis but is not essential for feedback regulation of bile acid synthesis. Mol Endocrinol. 2008;22(6):1345–1356. doi: 10.1210/me.2007-0565. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86.Matsukuma KE, Wang L, Bennett MK, Osborne TF. A key role for orphan nuclear receptor liver receptor homologue-1 in activation of fatty acid synthase promoter by liver X receptor. J Biol Chem. 2007;282(28):20164–20171. doi: 10.1074/jbc.M702895200. [DOI] [PubMed] [Google Scholar]
- 87.Chong HK, Biesinger J, Seo YK, Xie X, Osborne TF. Genome-wide analysis of hepatic LRH-1 reveals a promoter binding preference and suggests a role in regulating genes of lipid metabolism in concert with FXR. BMC Genomics. 2012;13:51. doi: 10.1186/1471-2164-13-51. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88.Mitro N, Mak PA, Vargas L, Godio C, Hampton E, Molteni V, Kreusch A, Saez E. The nuclear receptor LXR is a glucose sensor. Nature. 2007;445(7124):219–223. doi: 10.1038/nature05449. [DOI] [PubMed] [Google Scholar]
- 89.Cha JY, Repa JJ. The liver X receptor (LXR) and hepatic lipogenesis. The carbohydrate-response element-binding protein is a target gene of LXR. J Biol Chem. 2007;282(1):743–751. doi: 10.1074/jbc.M605023200. [DOI] [PubMed] [Google Scholar]
- 90.Anthonisen EH, Berven L, Holm S, Nygard M, Nebb HI, Gronning-Wang LM. Nuclear receptor liver X receptor is O-GlcNAc-modified in response to glucose. J Biol Chem. 2010;285(3):1607–1615. doi: 10.1074/jbc.M109.082685. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91.Gauthier K, Billon C, Bissler M, Beylot M, Lobaccaro JM, Vanacker JM, Samarut J. Thyroid hormone receptor beta (TRbeta) and liver X receptor (LXR) regulate carbohydrate-response element-binding protein (ChREBP) expression in a tissue-selective manner. J Biol Chem. 2010;285(36):28156–28163. doi: 10.1074/jbc.M110.146241. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 92.Beaven SW, Matveyenko A, Wroblewski K, Chao L, Wilpitz D, Hsu TW, Lentz J, Drew B, Hevener AL, Tontonoz P. Reciprocal regulation of hepatic and adipose lipogenesis by liver x receptors in obesity and insulin resistance. Cell Metab. 2013;18(1):106–117. doi: 10.1016/j.cmet.2013.04.021. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 93.Denechaud PD, Bossard P, Lobaccaro JM, Millatt L, Staels B, Girard J, Postic C. ChREBP, but not LXRs, is required for the induction of glucose-regulated genes in mouse liver. J Clin Invest. 2008;118(3):956–964. doi: 10.1172/JCI34314. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 94.Oosterveer MH, van Dijk TH, Grefhorst A, Bloks VW, Havinga R, Kuipers F, Reijngoud DJ. Lxralpha deficiency hampers the hepatic adaptive response to fasting in mice. J Biol Chem. 2008;283(37):25437–25445. doi: 10.1074/jbc.M801922200. [DOI] [PubMed] [Google Scholar]
- 95.Walsh CT, Garneau-Tsodikova S, Gatto GJ., Jr Protein posttranslational modifications: the chemistry of proteome diversifications. Angew Chem Int Ed Engl. 2005;44(45):7342–7372. doi: 10.1002/anie.200501023. [DOI] [PubMed] [Google Scholar]
- 96.Kaelin WG, Jr, McKnight SL. Influence of metabolism on epigenetics and disease. Cell. 2013;153(1):56–69. doi: 10.1016/j.cell.2013.03.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 97.Wellen KE, Thompson CB. A two-way street: reciprocal regulation of metabolism and signalling. Nat Rev Mol Cell Biol. 2012;13(4):270–276. doi: 10.1038/nrm3305. [DOI] [PubMed] [Google Scholar]
- 98.Ruan HB, Singh JP, Li MD, Wu J, Yang X. Cracking the O-GlcNAc code in metabolism. Trends Endocrinol Metab. 2013;24(6):301–309. doi: 10.1016/j.tem.2013.02.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 99.Moellering RE, Cravatt BF. Functional lysine modification by an intrinsically reactive primary glycolytic metabolite. Science. 2013;341(6145):549–553. doi: 10.1126/science.1238327. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 100.Lu C, Thompson CB. Metabolic regulation of epigenetics. Cell Metab. 2012;16(1):9–17. doi: 10.1016/j.cmet.2012.06.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101.Donohoe DR, Bultman SJ. Metaboloepigenetics: interrelationships between energy metabolism and epigenetic control of gene expression. J Cell Physiol. 2012;227(9):3169–3177. doi: 10.1002/jcp.24054. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 102.Katada S, Imhof A, Sassone-Corsi P. Connecting threads: epigenetics and metabolism. Cell. 2012;148(1–2):24–28. doi: 10.1016/j.cell.2012.01.001. [DOI] [PubMed] [Google Scholar]
- 103.Hanover JA, Yu S, Lubas WB, Shin SH, Ragano-Caracciola M, Kochran J, Love DC. Mitochondrial and nucleocytoplasmic isoforms of O-linked GlcNAc transferase encoded by a single mammalian gene. Arch Biochem Biophys. 2003;409(2):287–297. doi: 10.1016/s0003-9861(02)00578-7. [DOI] [PubMed] [Google Scholar]
- 104.Hanover JA, Krause MW, Love DC. The hexosamine signaling pathway: O-GlcNAc cycling in feast or famine. Biochim Biophys Acta. 2010;1800(2):80–95. doi: 10.1016/j.bbagen.2009.07.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 105.Kreppel LK, Blomberg MA, Hart GW. Dynamic glycosylation of nuclear and cytosolic proteins. Cloning and characterization of a unique O-GlcNAc transferase with multiple tetratricopeptide repeats. J Biol Chem. 1997;272(14):9308–9315. doi: 10.1074/jbc.272.14.9308. [DOI] [PubMed] [Google Scholar]
- 106.Gao Y, Wells L, Comer FI, Parker GJ, Hart GW. Dynamic O-glycosylation of nuclear and cytosolic proteins: cloning and characterization of a neutral, cytosolic beta-N-acetylglucosaminidase from human brain. J Biol Chem. 2001;276(13):9838–9845. doi: 10.1074/jbc.M010420200. [DOI] [PubMed] [Google Scholar]
- 107.Kreppel LK, Hart GW. Regulation of a cytosolic and nuclear O-GlcNAc transferase. Role of the tetratricopeptide repeats. J Biol Chem. 1999;274(45):32015–32022. doi: 10.1074/jbc.274.45.32015. [DOI] [PubMed] [Google Scholar]
- 108.Lubas WA, Hanover JA. Functional expression of O-linked GlcNAc transferase. Domain structure and substrate specificity. J Biol Chem. 2000;275(15):10983–10988. doi: 10.1074/jbc.275.15.10983. [DOI] [PubMed] [Google Scholar]
- 109.Keembiyehetty CN, Krzeslak A, Love DC, Hanover JA. A lipid-droplet-targeted O-GlcNAcase isoform is a key regulator of the proteasome. J Cell Sci. 2011;124(Pt 16):2851–2860. doi: 10.1242/jcs.083287. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 110.Gloster TM, Vocadlo DJ. Mechanism, structure, and inhibition of O-GlcNAc processing enzymes. Curr Signal Transduct Ther. 2010;5(1):74–91. doi: 10.2174/157436210790226537. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 111.Vocadlo DJ. O-GlcNAc processing enzymes: catalytic mechanisms, substrate specificity, and enzyme regulation. Curr Opin Chem Biol. 2012;16(5–6):488–497. doi: 10.1016/j.cbpa.2012.10.021. [DOI] [PubMed] [Google Scholar]
- 112.Hart GW, Slawson C, Ramirez-Correa G, Lagerlof O. Cross talk between O-GlcNAcylation and phosphorylation: roles in signaling, transcription, and chronic disease. Annu Rev Biochem. 2011;80:825–858. doi: 10.1146/annurev-biochem-060608-102511. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 113.Bond MR, Hanover JA. O-GlcNAc cycling: a link between metabolism and chronic disease. Annu Rev Nutr. 2013 doi: 10.1146/annurev-nutr-071812-161240. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 114.Li MD, Ruan HB, Hughes ME, Lee JS, Singh JP, Jones SP, Nitabach MN, Yang X. O-GlcNAc signaling entrains the circadian clock by inhibiting BMAL1/CLOCK ubiquitination. Cell Metab. 2013;17(2):303–310. doi: 10.1016/j.cmet.2012.12.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 115.Kaasik K, Kivimae S, Allen JJ, Chalkley RJ, Huang Y, Baer K, Kissel H, Burlingame AL, Shokat KM, Ptacek LJ, Fu YH. Glucose sensor O-GlcNAcylation coordinates with phosphorylation to regulate circadian clock. Cell Metab. 2013;17(2):291–302. doi: 10.1016/j.cmet.2012.12.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 116.Dentin R, Hedrick S, Xie J, Yates J, 3rd, Montminy M. Hepatic glucose sensing via the CREB coactivator CRTC2. Science. 2008;319(5868):1402–1405. doi: 10.1126/science.1151363. [DOI] [PubMed] [Google Scholar]
- 117.Housley MP, Udeshi ND, Rodgers JT, Shabanowitz J, Puigserver P, Hunt DF, Hart GW. A PGC-1alpha-O-GlcNAc transferase complex regulates FoxO transcription factor activity in response to glucose. J Biol Chem. 2009;284(8):5148–5157. doi: 10.1074/jbc.M808890200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 118.Housley MP, Rodgers JT, Udeshi ND, Kelly TJ, Shabanowitz J, Hunt DF, Puigserver P, Hart GW. O-GlcNAc regulates FoxO activation in response to glucose. J Biol Chem. 2008;283(24):16283–16292. doi: 10.1074/jbc.M802240200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 119.Ruan HB, Han X, Li MD, Singh JP, Qian K, Azarhoush S, Zhao L, Bennett AM, Samuel VT, Wu J, Yates JR, 3rd, Yang X. O-GlcNAc transferase/host cell factor C1 complex regulates gluconeogenesis by modulating PGC-1alpha stability. Cell Metab. 2012;16(2):226–237. doi: 10.1016/j.cmet.2012.07.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 120.Yang X, Ongusaha PP, Miles PD, Havstad JC, Zhang F, So WV, Kudlow JE, Michell RH, Olefsky JM, Field SJ, Evans RM. Phosphoinositide signalling links O-GlcNAc transferase to insulin resistance. Nature. 2008;451(7181):964–969. doi: 10.1038/nature06668. [DOI] [PubMed] [Google Scholar]
- 121.Soesanto YA, Luo B, Jones D, Taylor R, Gabrielsen JS, Parker G, McClain DA. Regulation of Akt signaling by O-GlcNAc in euglycemia. Am J Physiol Endocrinol Metab. 2008;295(4):E974–E980. doi: 10.1152/ajpendo.90366.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 122.Sakabe K, Wang Z, Hart GW. Beta-N-acetylglucosamine (O-GlcNAc) is part of the histone code. Proc Natl Acad Sci USA. 2010;107(46):19915–19920. doi: 10.1073/pnas.1009023107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 123.Hahne H, Sobotzki N, Nyberg T, Helm D, Borodkin VS, van Aalten DM, Agnew B, Kuster B. Proteome wide purification and identification of O-GlcNAc-modified proteins using click chemistry and mass spectrometry. J Proteome Res. 2013;12(2):927–936. doi: 10.1021/pr300967y. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 124.Teo CF, Ingale S, Wolfert MA, Elsayed GA, Not LG, Chatham JC, Wells L, Boons GJ. Glycopeptide-specific monoclonal antibodies suggest new roles for O-GlcNAc. Nat Chem Biol. 2010;6(5):338–343. doi: 10.1038/nchembio.338. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 125.Allis CD, Berger SL, Cote J, Dent S, Jenuwien T, Kouzarides T, Pillus L, Reinberg D, Shi Y, Shiekhattar R, Shilatifard A, Workman J, Zhang Y. New nomenclature for chromatin-modifying enzymes. Cell. 2007;131(4):633–636. doi: 10.1016/j.cell.2007.10.039. [DOI] [PubMed] [Google Scholar]
- 126.Blander G, Guarente L. The Sir2 family of protein deacetylases. Annu Rev Biochem. 2004;73:417–435. doi: 10.1146/annurev.biochem.73.011303.073651. [DOI] [PubMed] [Google Scholar]
- 127.Bakin RE, Jung MO. Cytoplasmic sequestration of HDAC7 from mitochondrial and nuclear compartments upon initiation of apoptosis. J Biol Chem. 2004;279(49):51218–51225. doi: 10.1074/jbc.M409271200. [DOI] [PubMed] [Google Scholar]
- 128.Michishita E, Park JY, Burneskis JM, Barrett JC, Horikawa I. Evolutionarily conserved and nonconserved cellular localizations and functions of human SIRT proteins. Mol Biol Cell. 2005;16(10):4623–4635. doi: 10.1091/mbc.E05-01-0033. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 129.Houtkooper RH, Pirinen E, Auwerx J. Sirtuins as regulators of metabolism and healthspan. Nat Rev Mol Cell Biol. 2012;13(4):225–238. doi: 10.1038/nrm3293. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 130.Yang L, Vaitheesvaran B, Hartil K, Robinson AJ, Hoopmann MR, Eng JK, Kurland IJ, Bruce JE. The fasted/fed mouse metabolic acetylome: N6-acetylation differences suggest acetylation coordinates organ-specific fuel switching. J Proteome Res. 2011;10(9):4134–4149. doi: 10.1021/pr200313x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 131.Xiong Y, Guan KL. Mechanistic insights into the regulation of metabolic enzymes by acetylation. J Cell Biol. 2012;198(2):155–164. doi: 10.1083/jcb.201202056. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 132.Xiong Y, Lei QY, Zhao S, Guan KL. Regulation of glycolysis and gluconeogenesis by acetylation of PKM and PEPCK. Cold Spring Harb Symp Quant Biol. 2011;76:285–289. doi: 10.1101/sqb.2011.76.010942. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 133.Li X, Zhang S, Blander G, Tse JG, Krieger M, Guarente L. SIRT1 deacetylates and positively regulates the nuclear receptor LXR. MolCell. 2007;28(1):91–106. doi: 10.1016/j.molcel.2007.07.032. [DOI] [PubMed] [Google Scholar]
- 134.Liu Y, Dentin R, Chen D, Hedrick S, Ravnskjaer K, Schenk S, Milne J, Meyers DJ, Cole P, Yates J, 3rd, Olefsky J, Guarente L, Montminy M. A fasting inducible switch modulates gluconeogenesis via activator/coactivator exchange. Nature. 2008;456(7219):269–273. doi: 10.1038/nature07349. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 135.Rodgers JT, Lerin C, Haas W, Gygi SP, Spiegelman BM, Puigserver P. Nutrient control of glucose homeostasis through a complex of PGC-1alpha and SIRT1. Nature. 2005;434(7029):113–118. doi: 10.1038/nature03354. [DOI] [PubMed] [Google Scholar]
- 136.Mihaylova MM, Vasquez DS, Ravnskjaer K, Denechaud PD, Yu RT, Alvarez JG, Downes M, Evans RM, Montminy M, Shaw RJ. Class IIa histone deacetylases are hormone-activated regulators of FOXO and mammalian glucose homeostasis. Cell. 2011;145(4):607–621. doi: 10.1016/j.cell.2011.03.043. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 137.von Meyenn F, Porstmann T, Gasser E, Selevsek N, Schmidt A, Aebersold R, Stoffel M. Glucagon-induced acetylation of Foxa2 regulates hepatic lipid metabolism. Cell Metab. 2013;17(3):436–447. doi: 10.1016/j.cmet.2013.01.014. [DOI] [PubMed] [Google Scholar]
- 138.Kemper JK, Xiao Z, Ponugoti B, Miao J, Fang S, Kanamaluru D, Tsang S, Wu SY, Chiang CM, Veenstra TD. FXR acetylation is normally dynamically regulated by p300 and SIRT1 but constitutively elevated in metabolic disease states. Cell Metab. 2009;10(5):392–404. doi: 10.1016/j.cmet.2009.09.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 139.Ponugoti B, Kim DH, Xiao Z, Smith Z, Miao J, Zang M, Wu SY, Chiang CM, Veenstra TD, Kemper JK. SIRT1 deacetylates and inhibits SREBP-1C activity in regulation of hepatic lipid metabolism. J Biol Chem. 2010;285(44):33959–33970. doi: 10.1074/jbc.M110.122978. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 140.Walker AK, Yang F, Jiang K, Ji JY, Watts JL, Purushotham A, Boss O, Hirsch ML, Ribich S, Smith JJ, Israelian K, Westphal CH, Rodgers JT, Shioda T, Elson SL, Mulligan P, Najafi-Shoushtari H, Black JC, Thakur JK, Kadyk LC, Whetstine JR, Mostoslavsky R, Puigserver P, Li X, Dyson NJ, Hart AC, Naar AM. Conserved role of SIRT1 orthologs in fasting-dependent inhibition of the lipid/cholesterol regulator SREBP. Genes Dev. 2010;24(13):1403–1417. doi: 10.1101/gad.1901210. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 141.Canto C, Gerhart-Hines Z, Feige JN, Lagouge M, Noriega L, Milne JC, Elliott PJ, Puigserver P, Auwerx J. AMPK regulates energy expenditure by modulating NAD+ metabolism and SIRT1 activity. Nature. 2009;458(7241):1056–1060. doi: 10.1038/nature07813. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 142.Wellen KE, Hatzivassiliou G, Sachdeva UM, Bui TV, Cross JR, Thompson CB. ATP-citrate lyase links cellular metabolism to histone acetylation. Science. 2009;324(5930):1076–1080. doi: 10.1126/science.1164097. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 143.Takahashi H, McCaffery JM, Irizarry RA, Boeke JD. Nucleocytosolic acetyl-coenzyme a synthetase is required for histone acetylation and global transcription. Mol Cell. 2006;23(2):207–217. doi: 10.1016/j.molcel.2006.05.040. [DOI] [PubMed] [Google Scholar]
- 144.Madiraju P, Pande SV, Prentki M, Madiraju SR. Mitochondrial acetylcarnitine provides acetyl groups for nuclear histone acetylation. Epigenetics. 2009;4(6):399–403. doi: 10.4161/epi.4.6.9767. [DOI] [PubMed] [Google Scholar]
- 145.Tu BP, Kudlicki A, Rowicka M, McKnight SL. Logic of the yeast metabolic cycle: temporal compartmentalization of cellular processes. Science. 2005;310(5751):1152–1158. doi: 10.1126/science.1120499. [DOI] [PubMed] [Google Scholar]
- 146.Cai L, Sutter BM, Li B, Tu BP. Acetyl-CoA induces cell growth and proliferation by promoting the acetylation of histones at growth genes. Mol Cell. 2011;42(4):426–437. doi: 10.1016/j.molcel.2011.05.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 147.Feng D, Liu T, Sun Z, Bugge A, Mullican SE, Alenghat T, Liu XS, Lazar MA. A circadian rhythm orchestrated by histone deacetylase 3 controls hepatic lipid metabolism. Science. 2011;331(6022):1315–1319. doi: 10.1126/science.1198125. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 148.Tokutake Y, Onizawa N, Katoh H, Toyoda A, Chohnan S. Coenzyme A and its thioester pools in fasted and fed rat tissues. Biochem Biophys Res Commun. 2010;402(1):158–162. doi: 10.1016/j.bbrc.2010.10.009. [DOI] [PubMed] [Google Scholar]
- 149.Bandsma RH, Grefhorst A, van Dijk TH, van der Sluijs FH, Hammer A, Reijngoud DJ, Kuipers F. Enhanced glucose cycling and suppressed de novo synthesis of glucose-6-phosphate result in a net unchanged hepatic glucose output in ob/ob mice. Diabetologia. 2004;47(11):2022–2031. doi: 10.1007/s00125-004-1571-8. [DOI] [PubMed] [Google Scholar]
- 150.Yen TT, Stamm NB. Constitutive hepatic glucokinase activity in db/db and ob/ob mice. Biochim Biophys Acta. 1981;657(1):195–202. doi: 10.1016/0005-2744(81)90143-1. [DOI] [PubMed] [Google Scholar]
- 151.Srinivasan V, Sandhya N, Sampathkumar R, Farooq S, Mohan V, Balasubramanyam M. Glutamine fructose-6-phosphate amidotransferase (GFAT) gene expression and activity in patients with type 2 diabetes: inter-relationships with hyperglycaemia and oxidative stress. Clin Biochem. 2007;40(13–14):952–957. doi: 10.1016/j.clinbiochem.2007.05.002. [DOI] [PubMed] [Google Scholar]
- 152.Buse MG. Hexosamines, insulin resistance, and the complications of diabetes: current status. Am J Physiol Endocrinol Metab. 2006;290(1):E1–E8. doi: 10.1152/ajpendo.00329.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 153.Dentin R, Benhamed F, Hainault I, Fauveau V, Foufelle F, Dyck JR, Girard J, Postic C. Liver-specific inhibition of ChREBP improves hepatic steatosis and insulin resistance in ob/ob mice. Diabetes. 2006;55(8):2159–2170. doi: 10.2337/db06-0200. [DOI] [PubMed] [Google Scholar]
- 154.Stefan N, Haring HU. The metabolically benign and malignant fatty liver. Diabetes. 2011;60(8):2011–2017. doi: 10.2337/db11-0231. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 155.Kursawe R, Caprio S, Giannini C, Narayan D, Lin A, D’Adamo E, Shaw M, Pierpont B, Cushman SW, Shulman GI. Decreased transcription of ChREBP-alpha/beta isoforms in abdominal subcutaneous adipose tissue of obese adolescents with prediabetes or early type 2 diabetes: associations with insulin resistance and hyperglycemia. Diabetes. 2013;62(3):837–844. doi: 10.2337/db12-0889. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 156.Eissing L, Scherer T, Todter K, Knippschild U, Greve JW, Buurman WA, Pinnschmidt HO, Rensen SS, Wolf AM, Bartelt A, Heeren J, Buettner C, Scheja L. De novo lipogenesis in human fat and liver is linked to ChREBP-beta and metabolic health. Nat Commun. 2013;4:1528. doi: 10.1038/ncomms2537. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 157.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, Postic C. The lipogenic transcription factor ChREBP dissociates hepatic steatosis from insulin resistance in mice and humans. J Clin Invest. 2012;122(6):2176–2194. doi: 10.1172/JCI41636. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 158.Lerin C, Rodgers JT, Kalume DE, Kim SH, Pandey A, Puigserver P. GCN5 acetyltransferase complex controls glucose metabolism through transcriptional repression of PGC-1alpha. Cell Metab. 2006;3(6):429–438. doi: 10.1016/j.cmet.2006.04.013. [DOI] [PubMed] [Google Scholar]
- 159.Sever S, Weinstein DA, Wolfsdorf JI, Gedik R, Schaefer EJ. Glycogen storage disease type Ia: linkage of glucose, glycogen, lactic acid, triglyceride, and uric acid metabolism. J Clin Lipidol. 2012;6(6):596–600. doi: 10.1016/j.jacl.2012.08.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 160.Shelly LL, Lei KJ, Pan CJ, Sakata SF, Ruppert S, Schutz G, Chou JY. Isolation of the gene for murine glucose-6-phosphatase, the enzyme deficient in glycogen storage disease type 1A. J Biol Chem. 1993;268(29):21482–21485. [PubMed] [Google Scholar]
- 161.Ihara K, Kuromaru R, Hara T. Genomic structure of the human glucose 6-phosphate translocase gene and novel mutations in the gene of a Japanese patient with glycogen storage disease type Ib. Hum Genet. 1998;103(4):493–496. doi: 10.1007/s004390050856. [DOI] [PubMed] [Google Scholar]
- 162.Gerin I, Veiga-da-Cunha M, Achouri Y, Collet JF, Van Schaftingen E. Sequence of a putative glucose 6-phosphate translocase, mutated in glycogen storage disease type Ib. FEBS Lett. 1997;419(2–3):235–238. doi: 10.1016/s0014-5793(97)01463-4. [DOI] [PubMed] [Google Scholar]
- 163.Grefhorst A, Schreurs M, Oosterveer MH, Cortes VA, Havinga R, Herling AW, Reijngoud DJ, Groen AK, Kuipers F. Carbohydrate-response-element-binding protein (ChREBP) and not the liver X receptor alpha (LXRalpha) mediates elevated hepatic lipogenic gene expression in a mouse model of glycogen storage disease type 1. Biochem J. 2010;432(2):249–254. doi: 10.1042/BJ20101225. [DOI] [PubMed] [Google Scholar]
- 164.Van Dijk TH, van der Sluijs FH, Wiegman CH, Baller JF, Gustafson LA, Burger HJ, Herling AW, Kuipers F, Meijer AJ, Reijngoud DJ. Acute inhibition of hepatic glucose-6-phosphatase does not affect gluconeogenesis but directs gluconeogenic flux toward glycogen in fasted rats. A pharmacological study with the chlorogenic acid derivative S4048. J Biol Chem. 2001;276(28):25727–25735. doi: 10.1074/jbc.M101223200. [DOI] [PubMed] [Google Scholar]
- 165.Mutel E, Abdul-Wahed A, Ramamonjisoa N, Stefanutti A, Houberdon I, Cavassila S, Pilleul F, Beuf O, Gautier-Stein A, Penhoat A, Mithieux G, Rajas F. Targeted deletion of liver glucose-6 phosphatase mimics glycogen storage disease type 1a including development of multiple adenomas. J Hepatol. 2011;54(3):529–537. doi: 10.1016/j.jhep.2010.08.014. [DOI] [PubMed] [Google Scholar]
- 166.McAdams AJ, Hug G, Bove KE. Glycogen storage disease, types I to X: criteria for morphologic diagnosis. Hum Pathol. 1974;5(4):463–487. doi: 10.1016/s0046-8177(74)80024-9. [DOI] [PubMed] [Google Scholar]
- 167.Bandsma RH, Wiegman CH, Herling AW, Burger HJ, ter Harmsel A, Meijer AJ, Romijn JA, Reijngoud DJ, Kuipers F. Acute inhibition of glucose-6-phosphate translocator activity leads to increased de novo lipogenesis and development of hepatic steatosis without affecting VLDL production in rats. Diabetes. 2001;50(11):2591–2597. doi: 10.2337/diabetes.50.11.2591. [DOI] [PubMed] [Google Scholar]
- 168.Fernandes J, Pikaar NA. Hyperlipemia in children with liver glycogen disease. Am J Clin Nutr. 1969;22(5):617–627. doi: 10.1093/ajcn/22.5.617. [DOI] [PubMed] [Google Scholar]
- 169.Jakovcic S, Khachadurian AK, Hsia DY. The hyperlipidemia in glycogen storage disease. J Lab Clin Med. 1966;68(5):769–779. [PubMed] [Google Scholar]
- 170.Bandsma RH, Prinsen BH, van Der Velden Mde S, Rake JP, Boer T, Smit GP, Reijngoud DJ, Kuipers F. Increased de novo lipogenesis and delayed conversion of large VLDL into intermediate density lipoprotein particles contribute to hyperlipidemia in glycogen storage disease type 1a. Pediatr Res. 2008;63(6):702–707. doi: 10.1203/PDR.0b013e31816c9013. [DOI] [PubMed] [Google Scholar]
- 171.Postic C, Girard J. The role of the lipogenic pathway in the development of hepatic steatosis. Diabetes Metab. 2008;34(6 Pt 2):643–648. doi: 10.1016/S1262-3636(08)74599-3. [DOI] [PubMed] [Google Scholar]
- 172.Rajas F, Labrune P, Mithieux G. Glycogen storage disease type 1 and diabetes: learning by comparing and contrasting the two disorders. Diabetes Metab. 2013 doi: 10.1016/j.diabet.2013.03.002. [DOI] [PubMed] [Google Scholar]
- 173.Wang P, Kang D, Cao W, Wang Y, Liu Z. Diabetes mellitus and risk of hepatocellular carcinoma: a systematic review and meta-analysis. Diabetes Metab Res Rev. 2012;28(2):109–122. doi: 10.1002/dmrr.1291. [DOI] [PubMed] [Google Scholar]
- 174.Seshasai SR, Kaptoge S, Thompson A, Di Angelantonio E, Gao P, Sarwar N, Whincup PH, Mukamal KJ, Gillum RF, Holme I, Njolstad I, Fletcher A, Nilsson P, Lewington S, Collins R, Gudnason V, Thompson SG, Sattar N, Selvin E, Hu FB, Danesh J. Diabetes mellitus, fasting glucose, and risk of cause-specific death. N Engl J Med. 2011;364(9):829–841. doi: 10.1056/NEJMoa1008862. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 175.Lee PJ. Glycogen storage disease type I: pathophysiology of liver adenomas. Eur J Pediatr. 2002;161(Suppl 1):S46–S49. doi: 10.1007/s00431-002-1002-0. [DOI] [PubMed] [Google Scholar]
- 176.Ertle J, Dechene A, Sowa JP, Penndorf V, Herzer K, Kaiser G, Schlaak JF, Gerken G, Syn WK, Canbay A. Non-alcoholic fatty liver disease progresses to hepatocellular carcinoma in the absence of apparent cirrhosis. Int J Cancer. 2011;128(10):2436–2443. doi: 10.1002/ijc.25797. [DOI] [PubMed] [Google Scholar]
- 177.Paradis V, Zalinski S, Chelbi E, Guedj N, Degos F, Vilgrain V, Bedossa P, Belghiti J. Hepatocellular carcinomas in patients with metabolic syndrome often develop without significant liver fibrosis: a pathological analysis. Hepatology. 2009;49(3):851–859. doi: 10.1002/hep.22734. [DOI] [PubMed] [Google Scholar]
- 178.Hanahan D, Weinberg RA. Hallmarks of cancer: the next generation. Cell. 2011;144(5):646–674. doi: 10.1016/j.cell.2011.02.013. [DOI] [PubMed] [Google Scholar]
- 179.DeBerardinis RJ, Thompson CB. Cellular metabolism and disease: what do metabolic outliers teach us? Cell. 2012;148(6):1132–1144. doi: 10.1016/j.cell.2012.02.032. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 180.Tong X, Zhao F, Mancuso A, Gruber JJ, Thompson CB. The glucose-responsive transcription factor ChREBP contributes to glucose-dependent anabolic synthesis and cell proliferation. Proc Natl Acad Sci USA. 2009;106(51):21660–21665. doi: 10.1073/pnas.0911316106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 181.Zimonjic DB, Popescu NC. Role of DLC1 tumor suppressor gene and MYC oncogene in pathogenesis of human hepatocellular carcinoma: potential prospects for combined targeted therapeutics (review) Int J Oncol. 2012;41(2):393–406. doi: 10.3892/ijo.2012.1474. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 182.Zhang P, Metukuri MR, Bindom SM, Prochownik EV, O’Doherty RM, Scott DK. c-Myc is required for the CHREBP-dependent activation of glucose-responsive genes. Mol Endocrinol. 2010;24(6):1274–1286. doi: 10.1210/me.2009-0437. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 183.Safi R, Kovacic A, Gaillard S, Murata Y, Simpson ER, McDonnell DP, Clyne CD. Coactivation of liver receptor homologue-1 by peroxisome proliferator-activated receptor gamma coactivator-1alpha on aromatase promoter II and its inhibition by activated retinoid X receptor suggest a novel target for breast-specific antiestrogen therapy. Cancer Res. 2005;65(24):11762–11770. doi: 10.1158/0008-5472.CAN-05-2792. [DOI] [PubMed] [Google Scholar]
- 184.Schoonjans K, Dubuquoy L, Mebis J, Fayard E, Wendling O, Haby C, Geboes K, Auwerx J. Liver receptor homolog 1 contributes to intestinal tumor formation through effects on cell cycle and inflammation. Proc Natl Acad Sci USA. 2005;102(6):2058–2062. doi: 10.1073/pnas.0409756102. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 185.Benod C, Vinogradova MV, Jouravel N, Kim GE, Fletterick RJ, Sablin EP. Nuclear receptor liver receptor homologue 1 (LRH-1) regulates pancreatic cancer cell growth and proliferation. Proc Natl Acad Sci USA. 2011;108(41):16927–16931. doi: 10.1073/pnas.1112047108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 186.Yi W, Clark PM, Mason DE, Keenan MC, Hill C, Goddard WA, 3rd, Peters EC, Driggers EM, Hsieh-Wilson LC. Phosphofructokinase 1 glycosylation regulates cell growth and metabolism. Science. 2012;337(6097):975–980. doi: 10.1126/science.1222278. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 187.Zhao D, Zou SW, Liu Y, Zhou X, Mo Y, Wang P, Xu YH, Dong B, Xiong Y, Lei QY, Guan KL. Lysine-5 acetylation negatively regulates lactate dehydrogenase a and is decreased in pancreatic cancer. Cancer Cell. 2013;23(4):464–476. doi: 10.1016/j.ccr.2013.02.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 188.Lv L, Li D, Zhao D, Lin R, Chu Y, Zhang H, Zha Z, Liu Y, Li Z, Xu Y, Wang G, Huang Y, Xiong Y, Guan KL, Lei QY. Acetylation targets the M2 isoform of pyruvate kinase for degradation through chaperone-mediated autophagy and promotes tumor growth. Mol Cell. 2011;42(6):719–730. doi: 10.1016/j.molcel.2011.04.025. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 189.Slawson C, Hart GW. O-GlcNAc signalling: implications for cancer cell biology. Nat Rev Cancer. 2011;11(9):678–684. doi: 10.1038/nrc3114. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 190.Brooks CL, Gu W. The impact of acetylation and deacetylation on the p53 pathway. Protein Cell. 2011;2(6):456–462. doi: 10.1007/s13238-011-1063-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 191.Wellen KE, Lu C, Mancuso A, Lemons JM, Ryczko M, Dennis JW, Rabinowitz JD, Coller HA, Thompson CB. The hexosamine biosynthetic pathway couples growth factor-induced glutamine uptake to glucose metabolism. Genes Dev. 2010;24(24):2784–2799. doi: 10.1101/gad.1985910. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 192.Morrish F, Noonan J, Perez-Olsen C, Gafken PR, Fitzgibbon M, Kelleher J, VanGilst M, Hockenbery D. Myc-dependent mitochondrial generation of acetyl-CoA contributes to fatty acid biosynthesis and histone acetylation during cell cycle entry. J Biol Chem. 2010;285(47):36267–36274. doi: 10.1074/jbc.M110.141606. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 193.Bouchard MF, Taniguchi H, Viger RS. Protein kinase A-dependent synergism between GATA factors and the nuclear receptor, liver receptor homolog-1, regulates human aromatase (CYP19) PII promoter activity in breast cancer cells. Endocrinology. 2005;146(11):4905–4916. doi: 10.1210/en.2005-0187. [DOI] [PubMed] [Google Scholar]
- 194.Chalkiadaki A, Talianidis I. SUMO-dependent compartmentalization in promyelocytic leukemia protein nuclear bodies prevents the access of LRH-1 to chromatin. Mol Cell Biol. 2005;25(12):5095–5105. doi: 10.1128/MCB.25.12.5095-5105.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 195.Goodwin B, Jones SA, Price RR, Watson MA, McKee DD, Moore LB, Galardi C, Wilson JG, Lewis MC, Roth ME, Maloney PR, Willson TM, Kliewer SA. A regulatory cascade of the nuclear receptors FXR, SHP-1, and LRH-1 represses bile acid biosynthesis. Mol Cell. 2000;6(3):517–526. doi: 10.1016/s1097-2765(00)00051-4. [DOI] [PubMed] [Google Scholar]
- 196.Hsieh HT, Wang CH, Wu ML, Yang FM, Tai YC, Hu MC. PIASy inhibits LRH-1-dependent CYP11A1 expression by competing for SRC-1 binding. Biochem J. 2009;419(1):201–209. doi: 10.1042/BJ20081402. [DOI] [PubMed] [Google Scholar]
- 197.Qin J, Gao DM, Jiang QF, Zhou Q, Kong YY, Wang Y, Xie YH. Prospero-related homeobox (Prox1) is a corepressor of human liver receptor homolog-1 and suppresses the transcription of the cholesterol 7-alpha-hydroxylase gene. Mol Endocrinol. 2004;18(10):2424–2439. doi: 10.1210/me.2004-0009. [DOI] [PubMed] [Google Scholar]
- 198.Sablin EP, Woods A, Krylova IN, Hwang P, Ingraham HA, Fletterick RJ. The structure of corepressor Dax-1 bound to its target nuclear receptor LRH-1. Proc Natl Acad Sci USA. 2008;105(47):18390–18395. doi: 10.1073/pnas.0808936105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 199.Steffensen KR, Holter E, Bavner A, Nilsson M, Pelto-Huikko M, Tomarev S, Treuter E. Functional conservation of interactions between a homeodomain cofactor and a mammalian FTZ-F1 homologue. EMBO Rep. 2004;5(6):613–619. doi: 10.1038/sj.embor.7400147. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 200.Xu PL, Kong YY, Xie YH, Wang Y. Corepressor SMRT specifically represses the transcriptional activity of orphan nuclear receptor hB1F/hLRH-1. Sheng Wu Hua Xue Yu Sheng Wu Wu Li Xue Bao (Shanghai) 2003;35(10):897–903. [PubMed] [Google Scholar]
- 201.Lee YK, Choi YH, Chua S, Park YJ, Moore DD. Phosphorylation of the hinge domain of the nuclear hormone receptor LRH-1 stimulates transactivation. J Biol Chem. 2006;281(12):7850–7855. doi: 10.1074/jbc.M509115200. [DOI] [PubMed] [Google Scholar]
- 202.Ohno M, Komakine J, Suzuki E, Nishizuka M, Osada S, Imagawa M. Repression of the promoter activity mediated by liver receptor homolog-1 through interaction with ku proteins. Biol Pharm Bull. 2010;33(5):784–791. doi: 10.1248/bpb.33.784. [DOI] [PubMed] [Google Scholar]
- 203.Brendel C, Gelman L, Auwerx J. Multiprotein bridging factor-1 (MBF-1) is a cofactor for nuclear receptors that regulate lipid metabolism. Mol Endocrinol. 2002;16(6):1367–1377. doi: 10.1210/mend.16.6.0843. [DOI] [PubMed] [Google Scholar]
- 204.Xu PL, Liu YQ, Shan SF, Kong YY, Zhou Q, Li M, Ding JP, Xie YH, Wang Y. Molecular mechanism for the potentiation of the transcriptional activity of human liver receptor homolog 1 by steroid receptor coactivator-1. Mol Endocrinol. 2004;18(8):1887–1905. doi: 10.1210/me.2003-0334. [DOI] [PubMed] [Google Scholar]
- 205.Nussinov R, Tsai CJ, Xin F, Radivojac P. Allosteric post-translational modification codes. Trends Biochem Sci. 2012;37(10):447–455. doi: 10.1016/j.tibs.2012.07.001. [DOI] [PubMed] [Google Scholar]
- 206.Yang XJ, Seto E. Lysine acetylation: codified crosstalk with other posttranslational modifications. Mol Cell. 2008;31(4):449–461. doi: 10.1016/j.molcel.2008.07.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 207.Ruan HB, Nie Y, Yang X. Regulation of protein degradation by O-GlcNAcylation: crosstalk with ubiquitination. Mol Cell Proteomics. 2013 doi: 10.1074/mcp.R113.029751. [DOI] [PMC free article] [PubMed] [Google Scholar]



