Skip to main content
Journal of Cellular and Molecular Medicine logoLink to Journal of Cellular and Molecular Medicine
. 2018 Sep 14;22(12):5787–5800. doi: 10.1111/jcmm.13877

PCAF fine‐tunes hepatic metabolic syndrome, inflammatory disease, and cancer

Tongxin Wang 1, Weilei Yao 1, Yafei Shao 1, Ruilong Zheng 1, Feiruo Huang 1,
PMCID: PMC6237576  PMID: 30216660

Abstract

The P300/CBP‐associating factor (PCAF), a histone acetyltransferase, is involved in metabolic and pathogenic diseases, particularly of the liver. The effects of PCAF on fine‐tuning liver diseases are extremely complex and vary according to different pathological conditions. This enzyme has dichotomous functions, depending on differently modified sites, which regulate the activities of various enzymes, metabolic functions, and gene expression. Here, we summarize the most recent findings on the functions and targets of PCAF in various metabolic and immunological processes in the liver and review these new discoveries and models of PCAF biology in three areas: hepatic metabolic syndrome, inflammatory disease, and cancer. Finally, we discuss the potential implications of these findings for therapeutic interventions in liver diseases.

Keywords: cancer, fine‐tuning, hepatic metabolic syndrome, inflammatory disease, PCAF

1. INTRODUCTION

Obesity, type 2 diabetes mellitus (T2DM), and other disturbances of chronic metabolic syndrome are now worldwide health problems.1, 2, 3 There is a parallel trend of the incidence and prevalence of immune diseases and tumorigenesis.4, 5 The liver, the largest visceral organ in the body, has been intimately associated with various diseases including chronic metabolic syndrome, inflammation, and cancer.6, 7, 8, 9 Under pathological conditions, the activation of immune cells and inflammatory pathways lead to liver complications, including damage to hepatic metabolic homeostasis and a cluster of interrelated metabolic risk factors such as raised fasting glucose, central obesity, dyslipoproteinaemia, and hypertension.10, 11, 12, 13 In addition, the damage of hepatic metabolic homeostasis and inflammatory disease has gained increasing attention for its relationship with end‐stage liver disease: primary liver cancer and hepatocellular carcinoma (HCC).14, 15, 16

Several transcription factors are concurrently critical mediators in the regulation of biological processes, especially under illness stress.17, 18 These transcription factors are subjected to posttranslational modifications (PTMs) that affect their activity, stability, intracellular distribution, and interaction with other proteins.19, 20, 21, 22, 23 During the past decades, the inventory of acetylation, demonstrated by the number of modification sites, is fast catching up with other major PTMs, such as phosphorylation and ubiquitylation, highlighting the regulatory potential of this modification at the proteomics scale.24 Recent studies have reported that overall cellular metabolism can be regulated through acetylation.25, 26, 27, 28 Predictably, as a pivotal metabolic organ, the liver is largely subjected to lysine acetylation, which occurs in most of the metabolic enzymes in human liver cells involved in glycolysis, tricarboxylic acid (TCA) and urea cycles, and fatty acid and glycogen metabolism.29, 30, 31, 32, 33 The acetylation status of these enzymes may alter their activities in order to respond to any changes in metabolic pathological demands.34, 35

The P300/CBP‐associating factor (PCAF) is a histone acetyltransferase (HAT) that primarily acetylates H3 histones and has a strong association with tumour initiation and progression, and it is similar to other HAT family member‐GCN5 eukaryotes.36 Using various nutritional, genetic, and pharmacological model systems, the roster of PCAF‐acetylated lysine sites has rapidly expanded. An emergent theory indicates that PCAF is also involved in multiple hepatic metabolic and pathogenic diseases such as metabolic syndrome, inflammation, apoptosis, injury, and cancer. For example, PCAF can acetylate nonhistone proteins, including, phosphoglycerate kinase 1 (PGK1) (K323), ATP‐citrate lyase (ACLY) (K540, K546, and K554), Peroxisome proliferator‐activated receptor gamma coactivator 1‐alpha (PGC1‐α) (K328 and K450), forkhead box P3 (FOXP3), and p53, etc.37, 38, 39, 40, 41 Additionally, increasing evidence demonstrates that PCAF is not only a HAT but also shows other effects such as ubiquitination.42 This review primarily focuses on the different mechanisms by which PCAF fine‐tunes hepatic metabolic syndrome, inflammatory disease, and tumour growth.

2. PCAF IN HEPATIC METABOLIC SYNDROME

As an associated factor of p300/CBP, PCAF and has been identified based on its activity in histone acetylation and is implicated in various cellular processes including proliferation and apoptosis.43, 44 More recently, a growing body of evidence has implicated PCAF in hepatic metabolic homeostasis, including lipogenesis, fatty acid oxidation, gluconeogenesis, and the regulation of insulin action, suggesting that PCAF activity may ensure the coordinated regulation of several distinct metabolic functions in the liver. Further study of PCAF will likely contribute to the development of treatments for hepatic metabolic syndrome and other nonalcoholic fatty liver diseases.

2.1. Hepatic lipid metabolism

2.1.1. Hepatic de novo lipogenesis and hyperlipidaemia

Hyperlipidaemia is a prevalent disease and a major component of metabolic syndromes.45, 46 In a physiological study, using a diabetic pregnant rat model, Abraham and colleagues showed that increased hepatic de novo lipogenesis directly contributes to hyperglycaemia and hyperlipidaemia.47 Apolipoprotein apoC‐III is important in lipogenesis and triglyceride (TG) metabolism. Hepatic apoC‐III production is subject to insulin inhibition, and is effectively up‐regulated by forkhead box O1 (FoxO1). A previous study on mice demonstrated that FoxO1 deregulation is associated with insulin deficiency or insulin resistance. Elevated FoxO1 production in the liver augments hepatic apoC‐III expression, resulting in increased plasma TG levels and impaired fat tolerance.48 In addition, transgenic mice expressing a constitutively nuclear FoxO1 allele directly exhibited hypertriglyceridaemia.48 These results provide several lines of evidence that FoxO1 functions as a regulator of lipogenesis and TG metabolism through different pathways. Remarkably, Yoshimochi et al recently showed that PCAF repressed FoxO1‐induced transcription in an enzymatic activity‐independent manner.49 PCAF bound to the forkhead domain of FoxO1 and acetylated FoxO1 at the K242 and K245 residues, while Akt‐induced phosphorylation of FoxO1 is required for its binding to PCAF, and this binding inhibits FoxO1‐induced transcription in the nucleus.49 Hence, the binding between PCAF and FoxO1 reduces lipid accumulation via the inhibition of TG synthesis (Figure 1A), thus providing a novel therapeutic strategy for the treatment of FoxO1‐induced diabetic hypertriglyceridaemia.

Figure 1.

Figure 1

Potential contributions of PCAF‐induced acetylation to the development of hepatic lipid metabolism. A, Hepatic de novo lipogenesis and hyperlipidaemia in PCAF regulation: the scheme indicates the up‐ and downstream signalling regulatory actions of the inducible PCAF acetylation and their effects on de novo lipogenesis both in enzymatic and transcriptional events. B, Nonalcoholic fatty liver disease and ER stress in PCAF regulation: lipid overboard, starvation, increased protein synthesis will lead to ER stress. The scheme indicates the enzyme and transcription factors that regulate ER stress and hepatic steatosis controlled by PCAF. PCAF has both positive and negative effects on occurrence of hepatic steatosis in different pathways. ACLY, ATP‐citrate lyase; FoxO1, forkhead box O1; USF‐1, upstream stimulatory factor‐1; SREBP‐1c, sterol regulatory element‐binding protein‐1c; DNA‐PK, DNA‐dependent protein kinase; apoC‐III, apolipoprotein C‐III; UPR, unfolded protein response; XBP‐1, X‐box‐binding protein1; AC, acetylation; ER, endoplasmic reticulum

Fatty acid synthase (FAS) is another key enzyme in hepatic de novo lipogenesis. It is also transcriptionally activated in response to feeding and insulin signalling.50 Two decades ago, Moustaïd et al showed that binding of the upstream stimulatory factor‐1/2 (USF‐1/2) heterodimer to the ‐65 E‐box is required for FAS promoter activation.51, 52, 53 Additionally, several trait‐mapping studies in humans have identified USF‐1 as a candidate gene for familial combined hyperlipidemia.54 Recently, Wong and Sul showed that during fasting USF‐1 recruits HDAC9, which deacetylates USF‐1 to repress transcription despite its binding to the ‐65 E‐box.55 Intriguingly, after feeding, the DNA‐dependent protein kinase (DNA‐PK) phosphorylates USF‐1, which subsequently recruits SREBP‐1c and PCAF and is acetylated at K237. This binding directly results in FAS promoter activation in the liver and subsequently promotes lipogenesis. In their study, K237 acetylation is dependent on S262 phosphorylation in response to feeding/insulin via preferential interaction with PCAF.55 Thus, the phosphorylation‐dependent acetylation of USF‐1 by PCAF and DNA‐PK functions as a dynamic molecular switch in sensing the nutritional transition from fasting to feeding. Such a multistep switch provides a mechanism to fine‐tune the transcription of lipogenic genes in response to different nutritional states (Figure 1A).

2.1.2. Nonalcoholic fatty liver disease and endoplasmic reticulum stress

The long‐term state of hyperlipidaemia leads to liver pathologies referred to as nonalcoholic fatty liver disease (NAFLD), which represents a large spectrum of diseases, including hepatic steatosis, fatty liver, and nonalcoholic steato hepatitis (NASH).56, 57, 58, 59 Deciphering the specific activities and substrates of PCAF mediating its different physiological functions in NAFLD has become an important area of research, as these functions may not only be limited to acetylation. According to one previous study, Lew et al showed that PCAF mediates X‐box‐binding protein1 (XBP‐1)‐dependent transcription through interaction with XBP‐1S. XBP‐1S is the necessary transcriptional activator of unfolded protein response, which is triggered when the endoplasmic reticulum (ER) is under stress.60 ER stress has been closely associated with hepatic steatosis, as hepatic lipid overload has been implicated in the initiation of the chronic ER stress in steatosis.61, 62 Steatosis is the first step in the progression of NAFLD, and is characterized by lipid accumulation in hepatocytes and is highly prevalent in people with obesity and hyperlipidaemia.63, 64 The PCAF‐induced translation of the spliced XBP1 mRNA generates a very potent form of the XBP1 transcription factor, which in turn increases the expression of ER chaperones and ER‐associated degradation under stress.65 Although the liver only undergoes transient ER stress under physiological conditions, this ER stress becomes chronic in NAFLD, as demonstrated by the detection of ER stress responses in the fatty liver.66 Therefore, these studies indicate that PCAF‐binding XBP1 is involved in activating ER stress‐induced hepatic steatosis/fatty liver (Figure 1B).

Cyclic AMP‐responsive element‐binding protein‐like 3(CREBH) is a hepatocyte‐specific ER‐anchored transcription factor involved in the initiation of NAFLD.67 Numerous evidence indicates that CREBH functions as a key regulator of energy homeostasis.68, 69, 70 Kim and colleagues observed that CREBH interacts with Peroxisome proliferator‐activated receptor alpha (PPARα) to synergistically activate the metabolic hormone fibroblast growth factor 21 to regulate lipolysis, fatty acid oxidation, and ketogenesis upon fasting or under an atherogenic high‐fat diet.69 Similar results were reported in another study using metabolic stress mice: defects in CREBH directly lead to nonalcoholic steatohepatitis and hyperlipidaemia under an atherogenic high‐fat diet or fasting conditions.68, 70 More recently, an emerging research by Kim demonstrated that PCAF and SIRT1 are the acetyltransferase and deacetylase of CREBH, respectively, under fasting conditions. CREBH is acetylated at lysine 294 within the CREBH bZIP domain by PCAF. Acetylation at this site is required for CREBH transcriptional activity.71 PCAF‐mediated acetylation and the transcriptional activities of CREBH are required for the interaction and synergy between CREBH and PPARα in activating their target gene upon fasting (Figure 1B). CREBH acetylation at lysine 294 is critical to maintain hepatic lipid homeostasis in fasting states.

Considering that ER stress promotes fat accumulation in hepatocytes and induces NAFLD, Baiceanu et al reported that this condition primarily occurs via the induction of de novo lipogenesis.72 ACLY is a lipogenic enzyme that catalyses the conversion of cytosolic citrate to acetyl‐CoA, which is the building block for de novo lipid synthesis.73 A recent study provided evidence that under high‐glucose conditions, ACLY is acetylated in both cells and mouse liver by PCAF acetyltransferase, which increases its stability and promotes de novo lipid synthesis.38 This finding has important implications in the context of metabolic regulation, as PCAF‐dependent ACLY acetylation could provide a strategy to improve lipogenesis (Figure 1B). This observation is consistent with the results of a previous study in which liver‐specific ACLY down‐regulation in leptin receptor‐deficient db/db mice led to the inhibition of hepatic de novo lipogenesis and protection against hepatic steatosis.74 Therefore, in this case, approaches to reduce hepatic PCAF activity or expression could serve as potential therapeutic strategies for the treatment of fatty liver disease.

2.2. Hepatic glucose homeostasis

2.2.1. PGC1‐α and Foxo1 acetylation in hepatic gluconeogenesis

In 1898, Naunyn coined the term hepatogenous diabetes to describe the coincidence between diabetes and liver glucose disorder, emphasizing the pathological importance of the liver in regulating glucose metabolism.75 Several transcription factors and co‐activators are involved in the nutritional and hormonal control of gluconeogenesis, including PGC‐1α and FoxO1.

Accumulating evidence reveals that PGC‐1α increases the activity and increases the expression of many proteins involved in fatty acid β‐oxidation, the TCA cycle, and the electron transport chain, such as PPARα, and the glucocorticoid receptor.76, 77, 78 Notably, the adenoviral‐mediated expression of PGC‐1α in the hepatocytes strongly activates the entire programme of gluconeogenesis and increases glucose output.79 These results robustly demonstrate that PGC‐1α is a key transcriptional coactivator in hepatic gluconeogenesis, which is implicated in the onset of type‐2 diabetes. It was recently observed that PGC‐1α is acetylated and deacetylated by acetyltransferase GCN5 and deacetylase Sirt1, respectively, for regulation of its transcriptional activity in response to nutrients supply. When nutrient availability is high, PGC‐1α is acetylated by GCN5. As nutrients are exhausted, sirt1 activity is enhanced and PGC‐1α is deacetylated to maintain energy metabolism balance.80, 81 Intriguingly, another recent finding revealed that PCAF is a pivotal HAT that acetylates PGC‐1α in both fasting and diabetic states, and acetylates K328 and K450 residues in PGC‐1α, leading to its proteasomal degradation.39 It was demonstrated that adenoviral‐mediated expression of PCAF improves glucose homeostasis in obese mice by attenuating PGC‐1α‐driven hepatic gluconeogenesis: in an in vivo experiment liver‐specific knockdown of PCAF enhanced the transcriptional activity of PGC‐1α and stimulated hepatic gluconeogenesis39 (Figure 2A). These interesting results provide insight on the different roles for the two crucial HATs, GCN5 and PCAF, in acetylating PGC‐1α under physiological and pathological conditions, respectively, and enhance the current understanding of the acetylation of PGC‐1α in controlling glucose homeostasis in nutrient state and diabetes.

Figure 2.

Figure 2

Potential contributions of PCAF‐induced acetylation to the development of hepatic Glucose homeostasis. A, PGC1‐α acetylation and hepatic gluconeogenesis. PCAF acetylate at lysine 328 and 450 residues, leading to its proteasomal degradation and attenuates PGC‐1α‐driven hepatic gluconeogenesis; (B) PI3K‐Akt/PKB signalling pathway and hepatic insulin resistance: ER stress activates TRB3; PTEN and TRB3 are both acetylated by PCAF increasing their activity to inhibit the PI3K‐Akt/PKB signalling pathway, therefore, leading to hepatic insulin resistance; (C) FoxO1 activity in diabetes. PCAF might regulate gluconeogenesis through binding and acetylating FoxO1 and decreasing both Pck1 and G6pc gene expression, resulting in decreased gluconeogenesis, therefore mitigating diabetes. PTEN, phosphatase and tensin homologue deleted on chromosome ten; TRB3, tribbles 3; Pck1, phosphoenolpyruvate carboxykinase; G6pc, glucose 6‐phosphatase

Given that the acetylation of FoxO1 at K242 and K245 by PCAF inhibits its transcriptional activity, it is likely that PCAF may modulate gluconeogenesis through additional mechanisms. Studies have shown that S‐phase kinase‐associated protein 2 (Skp2), which is involved in the regulation of glucose metabolism, interacts with ubiquitinases and promotes the degradation of FoxO1 through the Akt‐specific phosphorylation of Serine 256, thus suppressing the effects of this protein.82 This result suggests that the ubiquitination and degradation of FoxO1 by Skp2 may contribute to the onset of insulin resistance and diabetes. As the acetylation of FoxO1 by PCAF inhibits FoxO1‐induced transcription in the nucleus, there is also cooperation between PCAF and Skp2 in modulating FoxO1, thereby regulating glucose metabolism.49 Additionally, the genes involved in gluconeogenesis, including glucose 6‐phosphatase (G6pc) and phosphoenolpyruvate carboxykinase (Pck1), were recently demonstrated as target genes of FoxO1.83 In an in vivo experiment employing the adenoviral‐mediated transfer of a dominant negative FoxO1 into Lepr db/db mice, the reduction in nuclear FoxO1 decreased both Pck1 and G6pc gene expression, reducing gluconeogenesis and fasting blood glucose.84 Furthermore, antisense oligonucleotides of FoxO1 inhibit Pck1 and G6pc gene expression in primary hepatocytes which were stimulated with glucagons.85 These observations support the previous finding that FoxO1 increases hepatic gluconeogenesis through G6pc and Pck1 gene expression. Thus, PCAF might inhibit gluconeogenesis through additional mechanisms via the acetylation of FoxO1 (Figure 2C).

2.2.2. PI3k‐Akt/Pkb signalling pathway and hepatic insulin resistance

The liver is a significant insulin‐sensitive organ in the regulation of glucose homeostasis.86, 87 Thus, insulin resistance in the liver was suggested as an underlying cause of metabolic syndrome, including hyperglycaemia, dyslipidaemia, and increased inflammatory factors.87 PI3‐kinase is a downstream target of insulin signalling, and its inhibition leads to hepatic insulin resistance. PI3‐kinase is mediated through the serine/threonine kinases, Akt/protein kinase B (PKB) and Protein kinase C, zeta (PKCζ), which are activated by PI3‐kinase via phosphoinositide‐dependent protein kinase‐1.88, 89, 90, 91 Notably, on the one hand, both Akt and PKCζ are under the inhibitory control of the lipid phosphatase and tensin homologue deleted on chromosome ten (PTEN), on the other hand, Akt is also inhibited by tribbles 3 (TRB3).92, 93 Moreover, Altomonte et al suggested that as a result of both the increased gene expression and the reduced acetylation of PTEN and TRB3 by PCAF, insulin stimulation of Akt and PKCζ is impaired along with increased expression of PTEN and TRB3 in rats and by the mechanisms leading to hepatic insulin resistance.94 This observation is consistent with another study showing that the acetylation of PTEN (Lys125 and Lys128) under the control of PCAF reduces its enzymatic activity.95 Moreover, TRB3 interacts with PCAF, but whether PCAF is a TRB3 acetylase remains unknown. Taken together, as PTEN and TRB3 inhibit the PI3K‐Akt/PKB signalling pathway, the acetylation status of PTEN and TRB3 and the balance between PCAF and HDAC activities also explain their role in insulin resistance and increased gluconeogenesis, which may indicate an involvement of PCAF‐induced acetylation and hepatic insulin resistance (Figure 2B).

3. PCAF IN LIVER INJURY AND MEDIATORS OF INFLAMMATION

Metabolism and immunity are closely linked. Long‐term hepatic metabolic disorders result in systemic low‐grade chronic inflammation of the liver. To a great extent, the inflammatory response contributes to the development of acute and chronic liver diseases.96, 97, 98 Recently, PCAF was demonstrated to influence various aspects of liver inflammation through multiple mechanisms, thus the targets of PCAF in proinflammatory reaction, oxidative stress, and liver injury will be discussed in this review.

3.1. Proinflammatory reaction and liver epithelial inflammatory response

Decades of research have confirmed that hepatocytes and cholangiocytes are involved in the initiation, regulation, and resolution of inflammatory reactions in the liver. Previous studies have shown that both cell types express receptors for tumour necrosis factor (TNF)‐α and the activation of downstream signalling cascades of TNF‐α receptors initiates a series of epithelial inflammatory reactions.99 Such hepatic epithelial cell responses are finely controlled under physiological conditions and reflect a delicate balance between effector functions and their potential to cause subsequent damage to liver tissues.99 Recently, Zhao et al reported that PCAF is a target for miR‐181a/b, play an important role in the regulation of inflammatory reactions in liver epithelial cells.100 A previous study established that the transcription of miRNA genes in cholangiocytes can be elaborately controlled through nuclear transcription factors associated with inflammation, such as NF‐κB.101 Functionally, miRNAs may activate epithelial inflammatory responses, including the production and release of cytokines/chemokines, the expression of adhesion and costimulatory molecules, and feedback regulation of epithelial homeostasis.102 Zhao et al showed that TNF‐α down‐regulates PCAF expression in liver epithelial cells, and cells pretreated with TNF‐α inhibited the transcription of inflammatory genes in response to subsequent TNF‐α stimulation. These authors further observed that overexpression of PCAF or inhibition of miR‐181a/b using anti‐miRs attenuated the inhibitory effects of TNF‐α pretreatment on epithelial inflammatory response to subsequent TNF‐α stimulation both in vitro and in vivo.100 Hence, fine‐tuning the inflammatory reactions in hepatocytes in response to TNF‐α stimulation may involve the miR‐181a/b‐mediated suppression of PCAF. Such a negative feedback regulatory loop may function in concert with other regulatory mechanisms to ensure finely controlled inflammatory responses in the liver. As a target for miR‐181a/b, PCAF down‐regulation by TNF‐α maintains a negative feedback regulation to inflammatory reactions in liver epithelial cell responses, a process that may be relevant to the epigenetic fine‐tuning of epithelial inflammatory processes in general (Figure 3D).

Figure 3.

Figure 3

Potential contributions of PCAF‐induced acetylation to the development of liver injury and mediators of inflammation. A, Oxidative stress increases the level of PCAF; acetylation of TyrRS by PCAF promote the nuclear translocation of TyrRS, which may cause hepatocellular DNA damage under oxidative stress; (B) PCAF is recruited by KLF10 to the FOXP3 transcriptional regulatory regions that are critical for the induction of this gene which will amplify the chronic liver damage; (C) The EMT and MET events which are involved in liver fibrosis occurrence are regulated by PCAF acetylation on ZEB1; (D) As a target for miR‐181a/b, PCAF down‐regulation by TNF‐α provides negative feedback regulation to inflammatory reactions in liver epithelial cells responses, a process that may be relevant to the epigenetic fine‐tuning of epithelial inflammatory processes in general; (E) HMGB1 is acetylated by PCAF, this modification is associated with nonclassical vesicle‐mediated pathway which is an inflammatory response to liver injury. TyrRS, Tyrosyl‐tRNA synthetase; KLF10, transforming growth factorβ (TGFβ)‐inducible Kruppel‐like factor; FoxP3, forkhead box P3; Tregs, CD4+ ‐T cell; APC, antigen‐presenting cell; ZEB1, zinc finger E‐box‐binding homeobox 1, HMGB1, Serum high‐mobility group box 1; EMT, epithelial‐mesenchymal transition; MET, mesenchymal‐epithelial transition

3.2. Liver inflammation and oxidative stress

Oxidative stress and liver inflammation/injury are strongly associated. It occurs as a consequence of imbalance between the production of reactive oxygen species and the body's ability to detoxify reactive intermediates.103, 104 More recently, Tyrosyl‐tRNA synthetase (TyrRS), a member of the 20‐enzyme family of aminoacyl‐tRNA synthetases, has been shown to translocate to the nucleus and protect against DNA damage because of oxidative stress, and the nuclear translocation of TyrRS is promoted by lysine acetylation in response to oxidative stress.105 In addition, PCAF and SIRT1 function as the acetyltransferase and deacetylase of TyrRS, respectively. Oxidative stress increases the level of PCAF and decreases the level of SIRT1 and deacetylase activity, all of which promote the nuclear translocation of hyperacetylated TyrRS. Furthermore, TyrRS is primarily acetylated on the K244 residue near the nuclear localization signal, and acetylation inhibits the aminoacylation activity of TyrRS.105 Notably, the acetylation of TyrRS by PCAF may cause hepatocellular DNA damage under oxidative stress, thus leading to immunoreaction in the liver (Figure 3A).

3.3. Liver injury and hepatic fibrosis

Drug‐induced liver injury (DILI) is a major cause of adverse drug reactions and is the leading cause of acute liver failure. Recently, Jonathan et al demonstrated that serum high‐mobility group box 1 (HMGB1) isoforms show promise as circulating biomarkers of hepatotoxin acetaminophen (APAP)‐induced DILI. HMGB1 is a nuclear nonhistone chromosomal protein that binds the DNA minor groove and is involved in DNA replication, repair, and energy homeostasis.106 Decades of research have shown that upon cell death, HMGB1 is released in a nonacetylated form, but in activated immune cells, HMGB1 is secreted in acetylated form.107, 108 Notably, HMGB1 is also acetylated by the HATs PCAF.108 PCAF‐induced acetylation of key lysine residues within the lysine‐rich NLSs of HMGB1 results in the cytoplasmic translocation of HMGB1 and directs HMGB1 for secretion via a nonclassical vesicle‐mediated pathway. Lea et al revealed the hyperacetylation of HMGB1 by PCAF in serum, suggesting that hepatotoxicity was associated with an inflammatory response.106 Thus, these authors showed histological evidence of DILI in APAP‐treated mice as a sensitive marker of APAP liver injury (Figure 3E), suggesting that serum HMGB1 isoforms, such as circulating biomarkers of hepatotoxin acetaminophen APAP‐induced DILI, are strongly correlated with the levels of HMGB1.

Similarly, many other factors show pronounced interactions with PCAF that are positively correlated with the intensity of liver injury and inflammation. A recent study in a chronic HBV hepatitis disease model demonstrated that the FOXP3 expression in the liver is associated with liver inflammation and is an underlying cause of tissue damage.40 Previous studies have established that the chronic liver damage is amplified by nonspecific liver‐infiltrating cells and CD4+ ‐T regulatory cell (Tregs) interaction pathways, T cell expression of programmed death 1, and inhibition of T effector cells through FOXP3+ Tregs are among the most powerful mechanisms for achieving a balanced immune response.109, 110 In addition, PCAF is recruited by transforming growth factor β‐inducible Kruppel‐like factor (KLF10) to FOXP3 in the transcriptional regulatory regions that are critical for the induction of this gene.111 Notably, Xiong and colleagues demonstrated that KLF10 possesses the dual capacity to either positively or negatively regulate FOXP3 via its differential association with PCAF or the histone deacetylase‐binding protein Sin3, respectively.40 The authors showed that in the absence of this Sin3‐binding event, KLF10 may couple with PCAF to facilitate the activation state of FOXP3. These studies suggested a novel mechanism of reversible chromatin‐dependent silencing of the key immunoregulatory gene FOXP3 through differential coupling of KLF10 to Sin3‐HDAC and PCAF (Figure 3B).

As discussed above, hepatic epithelial cell responses are finely controlled under physiological conditions and reflect a delicate balance between effector functions and their potential to cause subsequent damage to liver tissues.99 At present, research is focused on the mechanisms that potentially delay, and even reverse, the process of liver fibrosis. Accumulating evidence suggests that proliferating biliary epithelial cells directly respond via the epithelial‐mesenchymal transition (EMT) during the induction of fibrosis.112 The origin of fibrogenic cells in liver fibrosis remains controversial. As PCAF is a target for microRNA 181a/b, another member of miRNAs microRNA 200c/141 has also been associated with PCAF and to contribute to EMT.113 Recent studies have shown that the miR‐200 family and the transcriptional suppressor ZEB1 are important contributors to EMT. In addition, expression levels of miR‐200 family members and ZEB1/ZEB2 are closely and inversely associated. MiR‐200 down‐regulates the expression of the transcription factors ZEB1 and ZEB2 by binding to the 3′ untranslated region of the mRNA and preventing translation.114, 115 Mizuguchi et al showed that the acetyltransferases P300 and PCAF activate miR200c/141 transcription by interacting at its promoter region via the cysteine‐histidine‐rich (CH3) domain of P300.113 Furthermore, acetyltransferases overcome ZEB1 transcriptional suppression of miR200c, most likely through lysine acetylation of ZEB1, indicating a PCAF/P300 cooperative function in acetylating ZEB1, and leading to liver pathology113 (Figure 3C). These findings show a novel and unifying mechanism for the effect of acetyltransferases on miRNA transcription and the potential for morphological consequences. Moreover, this epigenetic mechanism will provide a potential therapeutic strategy to treat liver fibrosis.

4. PCAF IN LIVER CANCER AND TUMOUR GROWTH

An association between inflammation and cancer has long been suspected. Indeed, pathological changes in the metabolic and physiological status of the liver such as NAFLD, NASH, injury, and cirrhosis, result in the onset of liver cancer, including HCC.116, 117 Here, we review and discuss recent advances in the elucidation of cellular and molecular alterations, signalling pathways associated with PCAF, and their effects on hepatocarcinogenesis and tumour growth.

4.1. PCAF‐induced acetylation and HCC

As one of the end‐stage liver diseases, HCC ranks as the third leading cause of cancer mortality worldwide.118 Recently, with the rise of epigenetics, HCC metastasis and recurrence was demonstrated to involve lysine acetylation, especially, PCAF‐induced acetylation. Mounting evidence over the last 5 years implies different effects of PCAF on HCC. PCAF is likely expressed at low levels in most HCC cell lines. In addition, its overexpression induces HCC cell apoptosis and autophagy.119 PCAF has also been proposed to act as a metastasis suppressor of HCC by restraining the activity of transcription factor glioma‐associated oncogene homologue‐1 (Gli1), thus inhibiting the epithelial‐to‐mesenchymal transition (EMT) of HCC cells.120 EMT is controlled by a group of transcriptional repressors, including Gli1. Studies have suggested that the targeted inhibition of Gli1 promotes the growth inhibition of HCC with activated hedgehog signalling which has been observed in various human malignancies.121, 122 Li et al observed that PCAF was down‐regulated in HCC tissues compared with the adjacent nontumour tissues and significantly associated with malignant portal vein invasion. A functional study demonstrated that Gli1, as a direct negative target of PCAF, induced EMT and promoted tumour metastasis and invasions.120 As an antioncogene, PCAF works significantly in the development of HCC by suppressing HCC cell metastasis and EMT by targeting Gli1, indicating the value of PCAF for the suppression of HCC metastasis (Figure 4A). Moreover, PCAF can directly acetylate cytoplasmic Gli1 protein at lysine 518, preventing its nuclear translocation and promoter occupancy, consequently suppressing Hedgehog signalling in HCC. Furthermore, Gli1 can increase B‐cell lymphoma‐2 (Bcl‐2) expression and down‐regulate Bcl‐2‐Associated X Protein (BAX) while the forced expression of PCAF reduces Bcl‐2 expression, up‐regulates BAX and represses cell apoptosis.123 PCAF induces cell apoptosis by acetylating Gli1 and modulating a GLI1/Bcl‐2/BAX axis, which in turn suppresses HCC progression (Figure 4B). Thus, as a target of PCAF, Gli1 acts as a suppressor of HCC progression, and the two different binding effects of PCAF and Gli1 provide a new insight into the various target treatments of HCC.

Figure 4.

Figure 4

Potential contributions of PCAF‐induced acetylation to the development of hepatocellular carcinoma and liver cancer tumour growth. A, Gli1 is a direct negative target of PCAF‐induced EMT promoting tumour metastasis and invasions; (B) PCAF can directly acetylate cytoplasmic Gli1 proteinat, preventing its nuclear translocation and promoter occupancy, and consequently suppressing Hedgehog signalling in HCC; (C) PCAF acetylate PTEN at lysine 125 and 128 to inhibit PTEN regulation of PI3K/AKT signalling, therefore, inhibiting the PTEN‐regulated cell cycle arrest expression depends on the presence of growth factors, which results in the progression of HCC; (D) The three enzymes in the glycolysis pathway are acetylated by PCAF, accelerating aerobic glycolysis and increasing the Warburg effect in HCC; (E) P53 is acetylated by PCAF, and this modification increases the transcription of P53, a tumour suppressor, therefore inhibiting liver tumour growth. Gli1, glioma‐associated oncogene homologue‐1; Bcl‐2, B‐cell lymphoma‐2; BAX, Bcl‐2‐Associated X Protein; GAPDH, Glyceraldehyde‐3‐phosphate dehydrogenase; PKM2, pyruvate kinase M2; PGK1, Phosphoglycerate kinase 1; Ub, ubiquitination

It has been suggested that PCAF is linked with HCC through the regulation of autophagic processes. In vivo experiments have confirmed that PCAF‐induced autophagy inhibits tumour growth.119 Notably, subsequent in vitro experiments have shown that PCAF promotes autophagy by inhibiting Akt/mTOR signalling pathway, which was well studied in the past decades as a negative pathway of autophagy.119 In this study, PCAF expression inhibited the phosphorylation of Akt and mTOR kinase, whereas knockdown of PCAF stimulated Akt and mTOR kinase activity. These results demonstrated that PCAF down‐regulates Akt/mTOR signalling pathway to activate autophagy (in addition to apoptosis) and mediate cancer cell death. Similarly, PCAF affects HCC cell apoptosis through acetylating histone H4 and inactivating AKT signalling. PCAF overexpression induced cell apoptosis and growth arrest with the increased Histone H4 acetylation and inactivation of AKT signalling in Huh7 and HepG2 cells. Moreover, coimmunoprecipitation assay confirmed that PCAF protein functions as a HCC repressor, binding to histone H4 protein in the nucleus of Hep3B cells, and eventually promoting cell apoptosis and functioning as a HCC repressor124 (Figure 4C). Establishing that PCAF promotes autophagy in HCC makes this novel molecular mechanism an attractive therapeutic strategy of HCC treatment.

Given that PTEN protein can be acetylated by PCAF, this mechanism can also contribute to the occurrence of HCC.90 The tumour suppressor activity of PTEN principally exerts its antagonistic effects on the antiapoptotic, proliferative, and hypertrophic activity of PI3K.125 In cancer biology, PTEN is frequently mutated or deleted in a wide variety of tumours. In patients with HCC, mutations, decreased promoter activity, and decreased PTEN expression are common occurrences.126 Indeed, the expression of PCAF increases the acetylation of lysine residues (Lys125 and 128) within the catalytic cleft of PTEN.90 The acetylation of PTEN caused by PCAF was directly associated with the inhibition of PTEN regulation of PI3K signalling, and the inhibition of PTEN‐regulated cell cycle arrest expression depends on the presence of growth factors, resulting in the progression of HCC.90 As inflammation, EMT, and genomic alterations are typical features of HCC, impaired PTEN expression or activity can thus represent an important step in progression of NAFLD towards understanding HCC.127, 128 These results indicate that a PTEN acetylation mechanism supports a correlation between distinct cancer‐relevant pathways that are essential to the control of growth factor signalling and gene expression (Figure 4C).

4.2. Warburg effect and tumour growth

Almost all cancer cells mainly rely on an aerobic glycolysis pathway to utilize glucose, the so‐called “Warburg Effect” first described by Dr. Otto Warburg.129, 130 The two enzymes that control the production of ATP during aerobic glycolysis in cancer cells are PGK1 and pyruvate kinase M2 (PKM2). Both enzymes were acetylated by PCAF.37, 131 Using immunoprecipitation of endogenous proteins and in vitro pull‐down assay with recombinant, a study identified and validated that PCAF as the acetyltransferase acetylating PGK1 K323 acetylation in liver cancer, this modification is characterized as an important positive regulation of PGK1 enzymatic activity and its oncogenic function.37 In addition, they further identified and validated that SIRT7 is the deacetylase of PGK1 at K323. And it can be proposed that PCAF‐induced K323 acetylation of PGK1 enhances its activity and promotes cancer cell proliferation and tumorigenesis (Figure 4D).

Similarly, the enzyme PKM2 was also associated with PCAF.131 The K305 site of PKM2 can be acetylated by PCAF, enhancing its association with the heat shock cognate protein 70 (HSC70) thereby promoting its lysosomal‐dependent molecular chaperone‐mediated autophagy process of degradation.131 High‐glucose concentrations increased the acetylation of PKM2 at Lys305, decreased PKM2 abundance, and increased association of PKM2 with the acetyltransferase PCAF. Ectopic expression of PCAF increased acetylation of PKM2 at K305 and decreased PKM2 activity. The glucose‐induced acetylation of PKM2 by PCAF, which inhibits its activity and triggers its autophagic degradation, promotes increased amounts of glycolytic intermediates that enable tumour growth and cell proliferation (Figure 4D).

Glyceraldehyde‐3‐phosphate dehydrogenase (GAPDH), a housekeeping glycolytic enzyme, has also shown to be acetylated by PCAF.132 Lysing acetylation activates GAPDH, which will alter along with the dynamic changes of external environment, such as glucose concentration. A recent study demonstrated that GAPDH K254 was acetylated by PCAF and deacetylated by HDAC5. Acetylation of GAPDH by PCAF promoted cell proliferation and tumour growth132 Most enzymes in glycolysis, including GAPDH, are acetylated in the liver, suggesting a positive role for PCAF‐induced acetylation of GAPDH in liver tumour growth and providing a potential therapy target for cancer. Furthermore, K117 and K251 are the putative GAPDH residues that could also be acetylated by PCAF. These modifications enable the entry of GAPDH into the nucleus for gene transcription and DNA repair133 (Figure 4D).

Additionally, as a tumour suppressor, p53 has also shown to be acetylated by PCAF in response to DNA damage.41 CBP/p300 acetylates p53 to achieve full transcriptional activity. This mechanism raised the question of whether PCAF can also acetylate p53 and if so, whether the substrate specificity would be distinct from that of CBP/p300. To address this question, an in vitro study suggested that PCAF and p300 acetylated full‐length p53 (p53 FL), and both enzymes had nearly equal acetylation activities. Lysine 320 in p53 is acetylated by PCAF, and PCAF acetylation activates the sequence‐specific DNA binding of p53. The acetylation site at K320 is targeted by PCAF and the acetylation at K373 is targeted by p300. Both p300 and PCAF increase the affinity of p53 to bind its cognate DNA site.41 These findings suggested that p53 PCAF‐induced acetylation is of great importance for DNA damage and cancer (Figure 4E). As p53 serves as a tumour suppressor protein it might have significant connection to liver cancer. This mechanism could provide us with a new pathway for the treatment of liver cancer.

5. CONCLUDING DISCUSSION AND PERSPECTIVE

The development of PTMs in liver disease prompted considerable research interest on novel perspectives in the past 20 years. As one of the PTMs, lysine acetylation has been associated with immunological and metabolic pathways. Most recent studies have started to shed light on the roles of different HATs in diverse biological functions. Intriguingly, similar to other HATs, PCAF is now emerging as a regulator of diverse nutrition metabolism functions and diseases, especially in the liver. Studies using improved methods to optimize the detection of substrate‐specific acetylation suggest that PCAF, through acetylation of complex interactive network substrates, enables a wide range of interactions with components of important metabolic and immunological pathways in the liver, such as lipogenesis, gluconeogenesis, insulin sensitivity, the PI3K‐Akt/PKB signalling pathway, and epithelial inflammatory response, among others. These interactions are closely associated with the main liver diseases including hepatogenous diabetes, NALFD, liver fibrosis, and cancer (Table 1).

Table 1.

Overview of the fine‐tuning mechanism of PCAF in liver diseases via the regulation of different targets

Disease Activity Targets Effect Metabolic response Reference
Liver cancer Ubiquitination Gli‐1 Inhibitory HCC metastasis↓ 120
Acetylation PGK1 Stimulatory Cell proliferation and tumorigenesis↑ 37
Acetylation PKM2 Inhibitory Cell proliferation and tumorigenesis↑ 131
Acetylation GAPDH Stimulatory Cell proliferation and tumorigenesis↑;DNA repair process↑ 132, 133
Acetylation Histone H4/AKT signalling Inhibitory Cell apoptosis↑ and HCC↓ 124
Acetylation Gli‐1 Inhibitory GLI1/BCL2/BAX axis↓ and HCC↓ 123
Induction AKT/mTOR Inhibitory Autophagy↑ and HCC↓ 119
Acetylation PTEN Inhibitory Tumorigenesis↑ and HCC↑ 90, 127, 128
Acetylation P53 Stimulatory Apoptosis↓ 41
Hepatic metabolic syndrome Acetylation CREBH Stimulatory Fasting‐induced hepatic steatosis and hyperlipidaemia↓ 71
Acetylation USF‐1 Stimulatory Lipogenesis↑ 55
Acetylation ACLY Inhibitory Lipogenesis↑ 38
Binding/Acetylation FoxO1 Inhibitory FFA oxidation↑
TG synthesis↓
49
Interaction XBP‐1S Stimulatory Hepatic steatosis↑ 60
Acetylation PGC1‐α Inhibitory Glucose production↓ Insulin sensitivity↑ obesity and diabetes↓ 39
Acetylation PTEN and TRB3 Inhibitory Gluconeogenesis↓ Insulin resistance↓ 89, 90
Acetylation FoxO1 Inhibitory Gluconeogenesis↓
diabetes↓
49
Immune disease Acetylation TyrRS Inhibitory Repair pathways for damaged DNA↓ 105
Binding miR‐181a/b Inhibitory Proinflammatory Genes↓ 100
Induction FOXP3 Stimulatory Immune Activation↑ 40
Acetylation miR‐200c/141 (ZEB1) Stimulatory Liver fibrosis↑ 113
Acetylation HMGB1 Stimulatory Hepatocyte apoptosis↑ 106, 108

To date, the molecular events by which PCAF exerts a protective role at the hepatic organelle level, regulating the hyperglycaemia, ER stress to oxidative stress, and Warburg effect are still not fully understood. However, based on this review, it is increasingly clear that PCAF fine‐tunes the progress of liver diseases ranging from metabolic syndrome to inflammatory disease and tumour growth. Common to other HATs, PCAF not only acts on liver cellular targets for metabolism regulation including p53, FoxO1, PGC1‐α, ACLY, PTEN, CREBH, but also shows differences among the other HATs in their roles on diseases. For example, PCAF acetylates PGC1‐α and decreases gluconeogenesis in hepatogenous diabetes, while another HATs GCN5, exerts its effect only in response to nutrition alterations. In addition, PCAF exerts its different effects from other HATs on regulating liver injury, fibrosis, and especially HCC. For example, on one hand, PCAF acts as a suppressor of HCC progression by acetylating Gli1, histone H4, and PTEN. On the other hand, PCAF also acetylates PGK1, PKM2, and GAPDH to induce a Warburg effect in liver tumour growth. This contradictory phenomenon suggests the value of PCAF for research on liver diseases.

The discovery of novel functions for PCAF‐mediated acetylation of transcription factors and enzymes on liver diseases opened new avenues of research into these less‐understood modifications. It also raised several important questions that may be crucial for assessing the potential specificity and efficacy of PCAF modulation as a therapeutic strategy of liver diseases:

  1. Given the importance of PCAF in the occurrence of Warburg effect, will the down‐regulation, a general phenomenon during diseases, and inactivation of PCAF in such contexts ease the degenerative changes in aerobic glycolysis and cancer?

  2. How does the interaction between the pleiotropic effects of PCAF on impact tumour biology, and which of these mechanisms are most relevant in the occurrence of cancer?

  3. PCAF shows either positive or negative effects on the regulation of different sites in hepatic steatosis: which is the dominating function?

  4. Can modified targets of PCAF in different stages of liver diseases at the same site act as the biomarkers or therapeutic targets of different liver diseases, and how do we balance the interaction between PCAF and HDACs (eg SIRTs) in the liver?

CONFLICT OF INTEREST

The authors declare that they have no conflict of interest.

ACKNOWLEDGEMENTS

The research is supported by Hubei Provincial Natural Science Foundation of China (grant no. 2018CFA071), National Natural Science Foundation of China (grant no. 31572409), the National Key Research and Development Program (grant no. 2016YFD0500506).

Wang T, Yao W, Shao Y, Zheng R, Huang F. PCAF fine‐tunes hepatic metabolic syndrome, inflammatory disease, and cancer. J Cell Mol Med. 2018;22:5787–5800. 10.1111/jcmm.13877

REFERENCES

  • 1. Chen L, Magliano DJ, Zimmet PZ. The worldwide epidemiology of type 2 diabetes mellitus–present and future perspectives. Nat Rev Endocrinol. 2011;8:228. [DOI] [PubMed] [Google Scholar]
  • 2. Eaton SB, Eaton SB. Physical inactivity, obesity, and type 2 diabetes: an evolutionary perspective. Res Q Exerc Sport. 2017;88:1. [DOI] [PubMed] [Google Scholar]
  • 3. Genser L, Casella Mariolo JR, Castagneto‐Gissey L, Panagiotopoulos S, Rubino F. Obesity, type 2 diabetes, and the metabolic syndrome: pathophysiologic relationships and guidelines for surgical intervention. Surg Clin North Am. 2016;96:681. [DOI] [PubMed] [Google Scholar]
  • 4. Dart A. Tumorigenesis: cancer goes tick tock. Nat Rev Cancer. 2016;16:409. [DOI] [PubMed] [Google Scholar]
  • 5. Wang H. The gut microbiota, tumorigenesis, and liver diseases. Engineering. 2017;3:110‐114. [Google Scholar]
  • 6. Russo GI, Cimino S, Castelli T, et al. Benign prostatic hyperplasia, metabolic syndrome and non‐alcoholic fatty liver disease: is metaflammation the link? Prostate. 2016;76:1528. [DOI] [PubMed] [Google Scholar]
  • 7. Ballestri S, Zona S, Targher G, et al. Nonalcoholic fatty liver disease is associated with an almost twofold increased risk of incident type 2 diabetes and metabolic syndrome. Evidence from a systematic review and meta‐analysis. J Gastroenterol Hepatol. 2016;31:936. [DOI] [PubMed] [Google Scholar]
  • 8. Koyama Y, Brenner DA. Liver inflammation and fibrosis. J Clin Invest. 2017;127:55. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9. Mokdad AA, Singal AG, Yopp AC. JAMA PATIENT PAGE. Treatment of liver cancer. JAMA. 2016;315:100. [DOI] [PubMed] [Google Scholar]
  • 10. Mehta M, Slaughter C, Xanthakos SA, Kohli R. High prevalence of hepatitis B non‐immunity in paediatric non‐alcoholic fatty liver disease patients. Dig Liver Dis. 2014;46:760‐761. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11. Ferrara LA, Wang H, Umans JG, et al. Does serum uric acid predict incident metabolic syndrome in a population with high prevalence of obesity? Nutr Metab Cardiovasc Dis. 2014;24:1360‐1364. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12. Guarner V, Rubioruiz ME. Low‐grade systemic inflammation connects aging, metabolic syndrome and cardiovascular disease. Interdiscip Top Gerontol. 2015;40:99‐106. [DOI] [PubMed] [Google Scholar]
  • 13. Meli R, Raso GM, Calignano A. Role of innate immune response in non‐alcoholic fatty liver disease: metabolic complications and therapeutic tools. Front Immunol. 2014;5:177. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14. Marengo A, Rosso C, Bugianesi E. Liver cancer: connections with obesity, fatty liver, and cirrhosis. Annu Rev Med. 2015;67:103. [DOI] [PubMed] [Google Scholar]
  • 15. Rosselli M, Lotersztajn S, Vizzutti F, Arena U, Pinzani M, Marra F. The metabolic syndrome and chronic liver disease. Curr Pharm Des. 2014;20:5010‐5024. [DOI] [PubMed] [Google Scholar]
  • 16. LukacsKornek V. The role of lymphatic endothelial cells in liver injury and tumor development. Front Immunol. 2016;7:548. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17. Li Y, Varala K, Coruzzi GM. From milliseconds to lifetimes: tracking the dynamic behavior of transcription factors in gene networks. Trends Genet. 2015;31:509. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18. Papadakis MA, Workman CT. Oxidative stress response pathways: fission yeast as archetype. Crit Rev Microbiol. 2015;41:520‐535. [DOI] [PubMed] [Google Scholar]
  • 19. Wang YC, Peterson SE, Loring JF. Protein post‐translational modifications and regulation of pluripotency in human stem cells. Cell Res. 2014;24:143. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20. Park JM, Jo SH, Kim MY, Kim TH, Ahn YH. Role of transcription factor acetylation in the regulation of metabolic homeostasis. Protein Cell. 2015;6:804‐813. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21. Jing F, Krautkramer KA, Feldman JL, Denu JM. Metabolic regulation of histone post‐translational modifications. ACS Chem Biol. 2015;10:95‐108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22. Mathias RA, Guise AJ, Cristea IM. Post‐translational modifications regulate class IIa histone deacetylase (HDAC) function in health and disease. Mol Cell Proteomics. 2015;14:456. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23. Duan G, Walther D. The roles of post‐translational modifications in the context of protein interaction networks. PLoS Comput Biol. 2015;11:e1004049. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24. Choudhary C, Weinert BT, Nishida Y, Verdin E, Mann M. The growing landscape of lysine acetylation links metabolism and cell signalling. Nat Rev Mol Cell Biol. 2014;15:536‐550. [DOI] [PubMed] [Google Scholar]
  • 25. Wagner GR, Hirschey MD. Nonenzymatic protein acylation as a carbon stress regulated by sirtuin deacylases. Mol Cell. 2014;54:5‐16. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26. Baeza J, Smallegan MJ, Denu JM. Mechanisms and dynamics of protein acetylation in mitochondria. Trends Biochem Sci. 2016;41:231‐244. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27. Li L, Zhang P, Bao Z, Wang T, Liu S, Huang F. PGC‐1α promotes ureagenesis in mouse periportal hepatocytes through SIRT3 and SIRT5 in response to glucagon. Sci Rep. 2016;6:24156. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28. Zhao S, Xu W, Jiang W, et al. Regulation of cellular metabolism by protein lysine acetylation. Science. 2010;327:1000‐1004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29. Nishida Y, Rardin M, Carrico C, et al. SIRT5 regulates both cytosolic and mitochondrial protein malonylation with glycolysis as a major target. Mol Cell. 2015;59:321. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30. Schug ZT, Vande VJ, Gottlieb E. The metabolic fate of acetate in cancer. Nat Rev Cancer. 2016;16:708‐717. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31. Hallows WC, Yu W, Smith BC, et al. Sirt3 promotes the urea cycle and fatty acid oxidation during dietary restriction. Mol Cell. 2011;41:139‐149. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32. Hirschey MD, Shimazu T, Goetzman E, et al. SIRT3 regulates mitochondrial fatty‐acid oxidation by reversible enzyme deacetylation. Nature. 2010;464:121‐125. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33. Zhang T, Wang S, Lin Y, et al. Acetylation negatively regulates glycogen phosphorylase by recruiting protein phosphatase 1. Cell Metab. 2012;15:75. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34. Mcdonnell E, Peterson BS, Bomze HM, Hirschey MD. SIRT3 regulates progression and development of diseases of aging. Trends Endocrinol Metab. 2015;26:486. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35. Iyer A, Fairlie DP, Brown L. Lysine acetylation in obesity, diabetes and metabolic disease. Immunol Cell Biol. 2012;90:39‐46. [DOI] [PubMed] [Google Scholar]
  • 36. Yang XJ, Ogryzko VV, Nishikawa J, Howard BH, Nakatani Y. A p300/CBP‐associated factor that competes with the adenoviral oncoprotein E1A. Nature. 1996;382:319‐324. [DOI] [PubMed] [Google Scholar]
  • 37. Hu H, Zhu W, Qin J, et al. Acetylation of PGK1 promotes liver cancer cell proliferation and tumorigenesis. Hepatology. 2016;65:515. [DOI] [PubMed] [Google Scholar]
  • 38. Lin R, Tao R, Gao X, et al. Acetylation stabilizes ATP‐citrate lyase to promote lipid biosynthesis and tumor growth: molecular cell. Mol Cell. 2013;51:506. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39. Cheng W, Meihong L, Xiaoyu L, Li K. PCAF improves glucose homeostasis by suppressing the gluconeogenic activity of PGC‐1α. Cell Rep. 2014;9:2250‐2262. [DOI] [PubMed] [Google Scholar]
  • 40. Xiong Y, Svingen PA, Sarmento OO, et al. Differential coupling of KLF10 to Sin3‐HDAC and PCAF regulates the inducibility of the FOXP3 gene. Am J Physiol Regul Integr Comp Physiol. 2014;307:R608. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41. Liu L, Scolnick DM, Trievel RC, et al. p53 sites acetylated in vitro by PCAF and p300 are acetylated in vivo in response to DNA damage. Mol Cell Biol. 1999;19:1202‐1209. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42. Mazzà D, Infante P, Colicchia V, et al. PCAF ubiquitin ligase activity inhibits Hedgehog|[sol]|Gli1 signaling in p53‐dependent response to genotoxic stress. Cell Death Differ. 2013;20:1688‐1697. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43. Kawasaki H, Eckner R, Yao TP, et al. Distinct roles of the co‐activators p300 and CBP in retinoic‐acid‐induced F9‐cell differentiation. Nature. 1998;393:284‐289. [DOI] [PubMed] [Google Scholar]
  • 44. Yao TP, Oh SP, Fuchs M, et al. Gene dosage‐dependent embryonic development and proliferation defects in mice lacking the transcriptional integrator p300. Cell. 1998;93:361. [DOI] [PubMed] [Google Scholar]
  • 45. Ricci C, Gaeta M, Rausa E, Asti E, Bandera F, Bonavina L. Long‐term effects of bariatric surgery on type II diabetes, hypertension and hyperlipidemia: a meta‐analysis and meta‐regression study with 5‐year follow‐up. Obes Surg. 2015;25:397‐405. [DOI] [PubMed] [Google Scholar]
  • 46. Yin Y, Li X, Sha X, et al. Early hyperlipidemia promotes endothelial activation via a caspase‐1‐sirtuin 1 pathway. Arterioscler Thromb Vasc Biol. 2015;35:804. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47. Abraham P, Rabi S, Francis DV, Mohana PD, Natarajan K, Amaladass A. Increased hepatic gluconeogenesis and decreased glucose uptake, and increased hepatic de novo lipogenesis in rat model of maternal diabetes. Biomed Res. 2016;27:632‐640. [Google Scholar]
  • 48. Altomonte J, Cong L, Harbaran S, et al. Foxo1 mediates insulin action on apoC‐III and triglyceride metabolism. J Clin Invest. 2004;114:1493‐1503. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49. Yoshimochi K, Daitoku H, Fukamizu A. PCAF represses transactivation function of FOXO1 in an acetyltransferase‐independent manner. J Recept Signal Transduct. 2010;30:43‐49. [DOI] [PubMed] [Google Scholar]
  • 50. Paulauskis JD, Sul HS. Hormonal regulation of mouse fatty acid synthase gene transcription in liver. J Biol Chem. 1989;264:574‐577. [PubMed] [Google Scholar]
  • 51. Moustaïd N, Sul HS. Regulation of expression of the fatty acid synthase gene in 3T3‐L1 cells by differentiation and triiodothyronine. J Biol Chem. 1991;266:18550‐18554. [PubMed] [Google Scholar]
  • 52. Moustaïd N, Sakamoto K, Clarke S, Beyer RS, Sul HS. Regulation of fatty acid synthase gene transcription. Sequences that confer a positive insulin effect and differentiation‐dependent expression in 3T3‐L1 preadipocytes are present in the 332 bp promoter. Biochem J. 1993;292(Pt 3):767‐772. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53. Moustaïd N, Beyer RS, Sul HS. Identification of an insulin response element in the fatty acid synthase promoter. J Biol Chem. 1994;269:5629‐5634. [PubMed] [Google Scholar]
  • 54. Pajukanta P, Lilja HE, Sinsheimer JS, et al. Familial combined hyperlipidemia is associated with upstream transcription factor 1 (USF1). Nat Genet. 2004;36:371. [DOI] [PubMed] [Google Scholar]
  • 55. Wong RH, Chang I, Hudak CS, Hyun S, Kwan HY, Sul HS. A role of DNA‐PK for the metabolic gene regulation in response to insulin. Cell. 2009;136:1056‐1072. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56. Clark JM, Brancati FL, Diehl AM. Nonalcoholic fatty liver disease. Gastroenterology. 2002;122:1649‐1657. [DOI] [PubMed] [Google Scholar]
  • 57. Targher G, Day CP, Bonora E. Risk of cardiovascular disease in patients with nonalcoholic fatty liver disease. N Engl J Med. 2010;363:1341. [DOI] [PubMed] [Google Scholar]
  • 58. Sass DA, Chang P, Chopra KB. Nonalcoholic fatty liver disease: a clinical review. Dig Dis Sci. 2005;50:171‐180. [DOI] [PubMed] [Google Scholar]
  • 59. Yoon H, Cha BS. Pathogenesis and therapeutic approaches for non‐alcoholic fatty liver disease. World J Hepatol. 2014;6:800‐811. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60. Lew QJ, Chu KL, Lee J, et al. PCAF interacts with XBP‐1S and mediates XBP‐1S‐dependent transcription. Nucleic Acids Res. 2011;39:429‐439. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61. Kharroubi I, Ladrière L, Cardozo AK, Dogusan Z, Cnop M, Eizirik DL. Free fatty acids and cytokines induce pancreatic β‐cell apoptosis by different mechanisms: role of nuclear factor‐κB and endoplasmic reticulum stress. Endocrinology. 2004;145:5087‐5096. [DOI] [PubMed] [Google Scholar]
  • 62. Wei Y, Wang D, Topczewski F, Pagliassotti MJ. Saturated fatty acids induce endoplasmic reticulum stress and apoptosis independently of ceramide in liver cells. Am J Physiol Endocrinol Metab. 2006;291:E275. [DOI] [PubMed] [Google Scholar]
  • 63. Leroux A, Ferrere G, Godie V, et al. Toxic lipids stored by Kupffer cells correlates with their pro‐inflammatory phenotype at an early stage of steatohepatitis. J Hepatol. 2012;57:141‐149. [DOI] [PubMed] [Google Scholar]
  • 64. Luukkonen PK, Zhou Y, Sädevirta S, et al. Hepatic ceramides dissociate steatosis and insulin resistance in patients with non‐alcoholic fatty liver disease. J Hepatol. 2016;64:1167. [DOI] [PubMed] [Google Scholar]
  • 65. Puri P, Baillie RA, Wiest MM, et al. A lipidomic analysis of nonalcoholic fatty liver disease. Hepatology. 2007;46:1081. [DOI] [PubMed] [Google Scholar]
  • 66. Ozcan U, Cao Q, Yilmaz E, et al. Endoplasmic reticulum stress links obesity, insulin action, and type 2 diabetes. Science. 2004;306:457‐461. [DOI] [PubMed] [Google Scholar]
  • 67. Omori Y, Imai J, Watanabe M, et al. CREB‐H: a novel mammalian transcription factor belonging to the CREB/ATF family and functioning via the box‐B element with a liver‐specific expression. Nucleic Acids Res. 2001;29:2154‐2162. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68. Zhang C, Wang G, Zheng Z, et al. Endoplasmic reticulum‐tethered transcription factor cAMP responsive element‐binding protein, hepatocyte specific, regulates hepatic lipogenesis, fatty acid oxidation, and lipolysis upon metabolic stress in mice. Hepatology. 2012;55:1070‐1082. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69. Kim H, Mendez R, Zheng Z, et al. Liver‐enriched transcription factor CREBH interacts with peroxisome proliferator‐activated receptor α to regulate metabolic hormone FGF21. Endocrinology. 2014;155:769. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70. Lee JH, Giannikopoulos P, Duncan SA, et al. The transcription factor cyclic AMP‐responsive element‐binding protein H regulates triglyceride metabolism. Nat Med. 2011;17:812. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71. Kim H, Mendez R, Chen X, Fang D, Zhang K. Lysine acetylation of CREBH regulates fasting‐induced hepatic lipid metabolism. Mol Cell Biol. 2015;35:4121‐4134. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72. Baiceanu A, Mesdom P, Lagouge M, Foufelle F. Endoplasmic reticulum proteostasis in hepatic steatosis. Nat Rev Endocrinol. 2016;12:710. [DOI] [PubMed] [Google Scholar]
  • 73. Icard P, Poulain L, Lincet H. Understanding the central role of citrate in the metabolism of cancer cells. Biochem Biophys Acta. 2012;1825:111. [DOI] [PubMed] [Google Scholar]
  • 74. Wang Q, Jiang L, Wang J, et al. Abrogation of hepatic ATP‐citrate lyase protects against fatty liver and ameliorates hyperglycemia in leptin receptor‐deficient mice. Hepatology. 2009;49:1166‐1175. [DOI] [PubMed] [Google Scholar]
  • 75. Vega R, Huss J, Kelly DP. The coactivator PGC1 cooperates with peroxisome proliferator‐activated receptor a in transcriptional control of nuclear genes encoding mitochondrial fatty acid oxidation enzymes. Mol Cell Biol. 2000;20:1868. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76. Puigserver P, Rhee J, Donovan J, et al. Insulin‐regulated hepatic gluconeogenesis through FOXO1‐PGC‐1alpha interaction. Nature. 2003;423:550. [DOI] [PubMed] [Google Scholar]
  • 77. Huss JM, Kopp RP, Kelly DP. Peroxisome proliferator‐activated receptor coactivator‐1α (PGC‐1α) coactivates the cardiac‐enriched nuclear receptors estrogen‐related receptor‐α and ‐γ identification of novel leucine‐rich interaction motif within PGC‐1α. J Biol Chem. 2002;277:40265‐40274. [DOI] [PubMed] [Google Scholar]
  • 78. Yoon JC, Puigserver P, Chen G, et al. Control of hepatic gluconeogenesis through the transcriptional coactivator PGC‐1. Nature. 2001;413:131. [DOI] [PubMed] [Google Scholar]
  • 79. 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;434:113. [DOI] [PubMed] [Google Scholar]
  • 80. 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;3:429. [DOI] [PubMed] [Google Scholar]
  • 81. Chen S, Zhao X, Ran L, et al. Resveratrol improves insulin resistance, glucose and lipid metabolism in patients with non‐alcoholic fatty liver disease: a randomized controlled trial. Dig Liver Dis. 2015;47:226. [DOI] [PubMed] [Google Scholar]
  • 82. Birkenfeld AL, Shulman GI. Nonalcoholic fatty liver disease, hepatic insulin resistance, and type 2 diabetes. Hepatology. 2014;59:713‐723. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83. Burgering BM, Coffer PJ. Protein kinase B (c‐Akt) in phosphatidylinositol‐3‐OH kinase signal transduction. Nature. 1995;376:599. [DOI] [PubMed] [Google Scholar]
  • 84. Whiteman EL, Cho H, Birnbaum MJ. Role of Akt/protein kinase B in metabolism. Trends Endocrinol Metab. 2002;13:444‐451. [DOI] [PubMed] [Google Scholar]
  • 85. Farese RV. Function and dysfunction of aPKC isoforms for glucose transport in insulin‐sensitive and insulin‐resistant states. Am J Physiol Endocrinol Metab. 2002;283:E1‐E11. [DOI] [PubMed] [Google Scholar]
  • 86. Idris I, Gray S, Donnelly R. Protein kinase C activation: isozyme‐specific effects on metabolism and cardiovascular complications in diabetes. Diabetologia. 2001;44:659‐673. [DOI] [PubMed] [Google Scholar]
  • 87. Lo YT, Tsao CJ, Liu IM, Liou SS, Cheng JT. Increase of PTEN gene expression in insulin resistance. Horn Metab Res. 2004;36:662‐666. [DOI] [PubMed] [Google Scholar]
  • 88. Du K, Herzig S, Kulkarni RN, Montminy M. TRB3: a tribbles homolog that inhibits Akt/PKB activation by insulin in liver. Science. 2003;300:1574. [DOI] [PubMed] [Google Scholar]
  • 89. Yao XH, Nyomba BL. Hepatic insulin resistance induced by prenatal alcohol exposure is associated with reduced PTEN and TRB3 acetylation in adult rat offspring. Am J Physiol Regul Integr Comp Physiol. 2008;294:R1797‐R1806. [DOI] [PubMed] [Google Scholar]
  • 90. Okumura K, Mendoza M, Bachoo RM, Depinho RA, Cavenee WK, Furnari FB. PCAF modulates PTEN activity. J Biol Chem. 2006;281:26562. [DOI] [PubMed] [Google Scholar]
  • 91. Gundling F, Schepp W. Diabetes und Leberzirrhose. Der Diabetologe. 2013;9:541‐550. [Google Scholar]
  • 92. Huang H, Regan KM, Wang F, et al. Skp2 inhibits FOXO1 in tumor suppression through ubiquitin‐mediated degradation. Proc Natl Acad Sci USA. 2005;102:1649‐1654. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93. Nakae J, Rd BW, Kitamura T, et al. Regulation of insulin action and pancreatic beta‐cell function by mutated alleles of the gene encoding forkhead transcription factor Foxo1. Nat Genet. 2002;32:245. [DOI] [PubMed] [Google Scholar]
  • 94. Altomonte J, Richter A, Harbaran S, et al. Inhibition of Foxo1 function is associated with improved fasting glycemia in diabetic mice. Am J Physiol Endocrinol Metab. 2003;285:E718. [DOI] [PubMed] [Google Scholar]
  • 95. Matsumoto M, Pocai A, Rossetti L, Depinho RA, Accili D. Impaired regulation of hepatic glucose production in mice lacking the forkhead transcription factor Foxo1 in liver. Cell Metab. 2007;6:208‐216. [DOI] [PubMed] [Google Scholar]
  • 96. Domingueti CP, Dusse LM, Md C, de Sousa LP, Gomes KB, Fernandes AP. Diabetes mellitus: the linkage between oxidative stress, inflammation, hypercoagulability and vascular complications. J Diabetes Complications. 2016;30:738. [DOI] [PubMed] [Google Scholar]
  • 97. Rolando N, Wade J, Davalos M, Wendon J, Philpott‐Howard J, Williams R. The systemic inflammatory response syndrome in acute liver failure. Hepatology. 2000;32:734‐739. [DOI] [PubMed] [Google Scholar]
  • 98. Arkan MC, Hevener AL, Greten FR, et al. Arkan, M. C. et al. IKK‐β links inflammation to obesity‐induced insulin resistance. Nat. Med. 11, 191‐198. Nat Med. 2005;11:191‐198. [DOI] [PubMed] [Google Scholar]
  • 99. Guicciardi ME, Deussing J, Miyoshi H, et al. Cathepsin B contributes to TNF‐alpha‐mediated hepatocyte apoptosis by promoting mitochondrial release of cytochrome c. J Clin Invest. 2000;106:1127‐1137. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100. Zhao J, Gong AY, Zhou R, Liu J, Eischeid AN, Chen XM. Downregulation of PCAF by miR‐181a/b provides feedback regulation to TNF‐α‐induced transcription of pro‐inflammatory genes in liver epithelial cells. J Immunol. 2012;188:1266‐1274. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101. Zhou R, Hu G, Liu J, Gong AY, Drescher KM, Chen XM. NF‐kappaB p65‐dependent transactivation of miRNA genes following Cryptosporidium parvum infection stimulates epithelial cell immune responses. PLoS Pathog. 2009;5:e1000681. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102. Hu G, Zhou R, Liu J, et al. MicroRNA‐98 and let‐7 confer cholangiocyte expression of cytokine‐inducible Src homology 2‐containing protein in response to microbial challenge. J Immunol. 2009;183:1617‐1624. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103. Kostopanagiotou GG, Grypioti AD, Matsota P, et al. Acetaminophen‐induced liver injury and oxidative stress: protective effect of propofol. Eur J Anaesthesiol. 2009;26:548‐553. [DOI] [PubMed] [Google Scholar]
  • 104. Singh R, Czaja MJ. Regulation of hepatocyte apoptosis by oxidative stress. J Gastroenterol Hepatol. 2007;22:S45‐S48. [DOI] [PubMed] [Google Scholar]
  • 105. Cao X, Li C, Xiao S, et al. Acetylation promotes TyrRS nuclear translocation to prevent oxidative damage. Proc Natl Acad Sci USA. 2017;114:687. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106. Lea JD, Clarke JI, Mcguire N, Antoine D. Redox‐dependent HMGB1 isoforms as pivotal co‐ordinators of drug‐induced liver injury ‐ mechanistic biomarkers and therapeutic targets. Antioxid Redox Signal. 2016;24:652‐665. [DOI] [PubMed] [Google Scholar]
  • 107. Bonaldi T, Talamo F, Scaffidi P, et al. Monocytic cells hyperacetylate chromatin protein HMGB1 to redirect it towards secretion. EMBO J. 2003;22:5551‐5560. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108. Wong LC, Sharpe DJ, Wong SS. High‐mobility group and other nonhistone substrates for nuclear histone N‐acetyltransferase. Biochem Genet. 1991;29:461. [DOI] [PubMed] [Google Scholar]
  • 109. Bertoletti A, Ferrari C. Innate and adaptive immune responses in chronic hepatitis B virus infections: towards restoration of immune control of viral infection. Postgrad Med J. 2013;61:1754. [DOI] [PubMed] [Google Scholar]
  • 110. Chisari FV. Cytotoxic T cells and viral hepatitis. J Clin Invest. 1997;99:1472. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111. Xiong Y, Khanna S, Grzenda AL, et al. Polycomb antagonizes p300/CREB‐binding protein‐associated factor to silence FOXP3 in a Kruppel‐like factor‐dependent manner. J Biol Chem. 2012;287:34372‐34385. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112. Zhao Y, Zhu R, Sun Y. Epithelial‐mesenchymal transition in liver fibrosis. Biomed Rep. 2016;4:269. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113. Mizuguchi Y, Specht S, Rd LJ, et al. Cooperation of p300 and PCAF in the control of microRNA 200c/141 transcription and epithelial characteristics. PLoS ONE. 2012;7:e32449. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114. Gregory PA, Bert AG, Paterson EL, et al. The miR‐200 family and miR‐205 regulate epithelial to mesenchymal transition by targeting ZEB1 and SIP1. Nat Cell Biol. 2008;10:593‐601. [DOI] [PubMed] [Google Scholar]
  • 115. Bracken CP, Gregory PA, Kolesnikoff N, et al. A double‐negative feedback loop between ZEB1‐SIP1 and the microRNA‐200 family regulates epithelial‐mesenchymal transition. Can Res. 2008;68:7846‐7854. [DOI] [PubMed] [Google Scholar]
  • 116. Mohamad B, Shah V, Onyshchenko M, et al. Characterization of hepatocellular carcinoma (HCC) in non‐alcoholic fatty liver disease (NAFLD) patients without cirrhosis. Hep Intl. 2016;10:632‐639. [DOI] [PubMed] [Google Scholar]
  • 117. Roncalli M, Bianchi P, Bruni B, et al. Methylation framework of cell cycle gene inhibitors in cirrhosis and associated hepatocellular carcinoma. Hepatology. 2002;36:427. [DOI] [PubMed] [Google Scholar]
  • 118. Welzel TM, Graubard BI, Zeuzem S, El‐Serag HB, Davila JA, Mcglynn KA. Metabolic syndrome increases the risk of primary liver cancer in the United States: a study in the SEER‐Medicare database. J Hepatol. 2011;54:463‐471. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119. Jia YL, Xu M, Dou CW, et al. P300/CBP‐associated factor (PCAF) inhibits the growth of hepatocellular carcinoma by promoting cell autophagy. Cell Death Dis. 2016;7:e2400. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120. Li Q, Liu Z, Xu M, et al. PCAF inhibits hepatocellular carcinoma metastasis by inhibition of epithelial‐mesenchymal transition by targeting Gli‐1. Cancer Lett. 2016;375:190‐198. [DOI] [PubMed] [Google Scholar]
  • 121. Amakye D, Jagani Z, Dorsch M. Unraveling the therapeutic potential of the Hedgehog pathway in cancer. Nat Med. 2013;19:1410‐1422. [DOI] [PubMed] [Google Scholar]
  • 122. Briscoe J, Thérond PP. The mechanisms of Hedgehog signalling and its roles in development and disease. Nat Rev Mol Cell Biol. 2013;14:416‐429. [DOI] [PubMed] [Google Scholar]
  • 123. Gai X, Tu K, Li C, Lu Z, Roberts LR, Zheng X. Histone acetyltransferase PCAF accelerates apoptosis by repressing a GLI1|[sol]|BCL2|[sol]|BAX axis in hepatocellular carcinoma. Cell Death Dis. 2015;6:e1712. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124. Zheng X, Gai X, Ding F, et al. Histone acetyltransferase PCAF Up‐regulated cell apoptosis in hepatocellular carcinoma via acetylating histone H4 and inactivating AKT signaling. Mol Cancer. 2013;12:96. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125. Yuan TL, Cantley LC. PI3K pathway alterations in cancer: variations on a theme. Oncogene. 2008;27:5497. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126. Shearn CT, Petersen DR. Understanding the tumor suppressor PTEN in chronic alcoholism and hepatocellular carcinoma. Adv Exp Med Biol. 2015;815:173‐184. [DOI] [PubMed] [Google Scholar]
  • 127. Laurent PP, Zucman RJ. Genetics of hepatocellular tumors. Oncogene. 2006;25:3778. [DOI] [PubMed] [Google Scholar]
  • 128. Lee TK, Poon RT, Yuen AP, et al. Twist overexpression correlates with hepatocellular carcinoma metastasis through induction of epithelial‐mesenchymal transition. Clin Cancer Res. 2006;12:5369‐5376. [DOI] [PubMed] [Google Scholar]
  • 129. Warburg O. The metabolism of carcinoma cells. J Cancer Res. 1925;9:148‐163. [Google Scholar]
  • 130. Koppenol WH, Bounds PL, Dang CV. Otto Warburg's contributions to current concepts of cancer metabolism. Nat Rev Cancer. 2011;11:325. [DOI] [PubMed] [Google Scholar]
  • 131. Lv L, Li D, Zhao D, et al. Acetylation targets the M2 isoform of pyruvate kinase for degradation through chaperone‐mediated autophagy and promotes tumor growth. Mol Cell. 2011;42:719. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132. Li T, Liu M, Feng X, et al. Glyceraldehyde‐3‐phosphate dehydrogenase is activated by lysine 254 acetylation in response to glucose signal. J Biol Chem. 2014;289:3775‐3785. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133. Ventura M, Mateo F, Serratosa J, et al. Nuclear translocation of glyceraldehyde‐3‐phosphate dehydrogenase is regulated by acetylation. Int J Biochem Cell Biol. 2010;42:1672‐1680. [DOI] [PubMed] [Google Scholar]

Articles from Journal of Cellular and Molecular Medicine are provided here courtesy of Blackwell Publishing

RESOURCES