Open Access

Histone deacetylase‑2: A potential regulator and therapeutic target in liver disease (Review)

  • Authors:
    • Ya-Ru Liu
    • Jie-Quan Wang
    • Zhao-Gang Huang
    • Ruo-Nan Chen
    • Xi Cao
    • Dong-Chun Zhu
    • Hai-Xia Yu
    • Xiu-Rong Wang
    • Hai-Yun Zhou
    • Quan Xia
    • Jun Li
  • View Affiliations

  • Published online on: May 18, 2021     https://doi.org/10.3892/ijmm.2021.4964
  • Article Number: 131
  • Copyright: © Liu et al. This is an open access article distributed under the terms of Creative Commons Attribution License.

Metrics: Total Views: 0 (Spandidos Publications: | PMC Statistics: )
Total PDF Downloads: 0 (Spandidos Publications: | PMC Statistics: )


Abstract

Histone acetyltransferases are responsible for histone acetylation, while histone deacetylases (HDACs) counteract histone acetylation. An unbalanced dynamic between histone acetylation and deacetylation may lead to aberrant chromatin landscape and chromosomal function. HDAC2, a member of class I HDAC family, serves a crucial role in the modulation of cell signaling, immune response and gene expression. HDAC2 has emerged as a promising therapeutic target for liver disease by regulating gene transcription, chromatin remodeling, signal transduction and nuclear reprogramming, thus receiving attention from researchers and clinicians. The present review introduces biological information of HDAC2 and its physiological and biochemical functions. Secondly, the functional roles of HDAC2 in liver disease are discussed in terms of hepatocyte apoptosis and proliferation, liver regeneration, hepatocellular carcinoma, liver fibrosis and non‑alcoholic steatohepatitis. Moreover, abnormal expression of HDAC2 may be involved in the pathogenesis of liver disease, and its expression levels and pharmacological activity may represent potential biomarkers of liver disease. Finally, research on selective HDAC2 inhibitors and non‑coding RNAs relevant to HDAC2 expression in liver disease is also reviewed. The aim of the present review was to improve understanding of the multifunctional role and potential regulatory mechanism of HDAC2 in liver disease.

1. Introduction

The level of histone acetylation is of importance for nuclear stability, chromatin structure, gene expression and physiological functions in hepatocytes (1,2). Histone acetyltransferases (HATs) and histone deacetylases (HDACs) are antagonistic proteases that serve regulatory roles in the balance of histone acetylation and deacetylation in nucleosomes (3). HATs, which favor of histone acetylation, transfer acetyl groups from acetyl-CoA to the ε-NH2 group of lysine residue side chains and neutralize the positive charge of histone tails, thus making chromatin structure more loose and conducive to active transcription (46). HDACs, which facilitate histone deacetylation, remove the acetyl group from the ε-amino group of lysine residue side chains, reconstitute positive charge on the surface of lysine and increase binding affinity with the negatively charged surface of DNA (5,6). Furthermore, interactions between histone and DNA result in the formation of compacted and inactive chromatin that restrains gene transcription (710).

The dysfunction of histone deacetylation is associated with the occurrence and development of liver disease (11). HDACs are emerging as next-generation drug targets, thus gaining increasing attention and recognition (1215). HDAC2 is responsible for the deacetylation of the N-terminal of histone H3 and H4, leading to a more compacted chromatin structure and transcriptional gene silencing (16,17). HDAC2 participates in the genesis and development of renal, cardiovascular, neurological and lung disease (1823). Furthermore, small molecular compounds, peptides and other biological agents that inhibit HDAC2 show potential in the treatment of cancer (2426), as well as degenerative and inflammatory immune disease (2729). In particular, evidence has also highlighted the key role of HDAC2 in the pathological process of liver disease (3032).

The present review introduces gene localization, structural information and biological functions of HDAC2, as well as the pharmacological role of HDAC2, its expression level in a variety of liver diseases and effects on hepatocyte apoptosis and proliferation, liver regeneration, hepatocellular carcinoma (HCC), liver fibrosis and non-alcoholic steatohepatitis (NASH). Finally, a number of selective HDAC2 inhibitors and non-coding (nc)RNAs relevant for HDAC2 expression in liver disease are reviewed (Fig. 1).

2. HDAC2

Characterization and classification of HDACs

Histone proteins serve structural and functional roles almost in all nuclear processes (33,34). Histones, DNA and a number of different protein complexes form chromatin to facilitate dynamic changes that occur during DNA replication, cell-cycle progression, transcription process and post-transcription events (35). Changes in chromatin that do not involve a change in DNA sequence are defined as epigenetic modification. One of the earliest known types of chromatin epigenetic modifications is histone acetylation, although its potential role in cell fate determination has not been fully elucidated (36). Acetylation has been widely studied and its potential roles and regulatory mechanisms have been revealed (37,38).

Like HATs, HDACs are rich in structural diversity and serve multiple functions (39), making them potential targets for pharmacological intervention and drug development (40). HDACs are part of a multiprotein family in which each member of HDAC has its own specialized function. In mammals, the HDAC family can be divided into four distinct subfamilies according to their structure, enzyme function, subcellular localization, expression pattern and homology with a typical HDAC in yeast (Table I): Class I (HDACs 1–3 and 8), II (HDACs 4–7, 9 and 10), III [Sirtuins (Sirs/SIRTs)] and IV (HDAC11) (4144). The class I HDAC family consists of HDAC1-3 and 8, which have a relatively high similarity to HDAC2. They are ~50 kDa in size, are ubiquitously expressed and have homology with yeast reduced potassium dependency 3 (37,45,46). Class II HDACs comprise HDAC4–7, 9 and 10, which have a high degree of homology to yeast HAD-1 (4749). They are ~120–150 kDa in size and are widely expressed in a tissue-specific manner (50,51). Class II HDACs can be separated into class IIa (HDACs 4, 5, 7 and 9) and IIb (HDACs 6 and 10), depending on the presence of tandem deacetylase domains (41,52). The class III HDACs, also known as Sir2 family, are named for their homology to the yeast Sir2gene. Sir2 family is a highly conservative gene family, including Sirt1–7 (5355). Among these, Sirt1, 2, 3 and 5 have a common NAD-dependent deacetylase domain, which can catalyze the deacetylation of histone as well as non-histone proteins. By contrast, Sirt4 and 6 have a NAD+-dependent ADP ribosylation domain, which is key for protein ribosylation (54,55). Class III HDACs share little homology with class I and II HDACs, and their enzyme activity is not inhibited by broad-spectrum HDAC inhibitors [such as butyrate, valproic acid, trichostatin A (TSA) or suberoyl anilide hydroxamic acid] (56). The class IV family HDAC11 shares some, but not sufficient, homology with class I and II HDACs (57,58). HDAC11 is enriched in the brain, heart, muscle, kidney and testis in a cell type-specific manner (48,59). Furthermore, class I, II and IV HDACs have a zinc-dependent active site that can be specifically targeted by(compounds containing hydroxamates), such as TSA. As many drugs targeting HDAC isotypes have shown positive effects, HDAC enzymes may represent a novel therapeutic target for liver disease (2,11).

Table I

Classification of classic HDACs.

Table I

Classification of classic HDACs.

ClassCo-factorsSequence homologyProteinSize (AA)Chromosomal locationSub-cellular localizationTissue expression
IZn+Yeast RPD3HDAC14821p34NucleusUbiquitous
HDAC24886q21NucleusUbiquitous
HDAC34285q31 Nucleus/cytoplasmUbiquitous
HDAC8377Xq13NucleusUbiquitous
IIaZn+Yeast HDA1HDAC410842q37 Nucleus/cytoplasmHeart, smooth muscle, brain
HDAC5112217q21 Nucleus/cytoplasmHeart, smooth muscle, brain
HDAC795212q13 Nucleus/cytoplasmHeart, placenta, pancreas, smooth muscle
HDAC91011/879/5907p15-p21 Nucleus/cytoplasmSmooth muscle, brain
IIbZn+Yeast HDA1HDAC61215Xp11 Nucleus/cytoplasmKidney, liver, heart, pancreas
HDAC1066922q13 Nucleus/cytoplasmSpleen, kidney, liver
III NAD+/NADYeast Sir2SIRT1564No dataEuchromatinUnknown
SIRT2377Cytoplasm
SIRT3396Mitochondrial
SIRT4320Mitochondrial
SIRT 5308Mitochondrial
SIRT6356 Heterochromatin
SIRT7407Nucleolus
IVZn+Yeast RPD3/HDA1HDAC113473p25NucleusHeart, smooth muscle, kidney, brain

[i] HDAC/HDA, histone deacetylase; RPD3, reduced potassium dependency 3; Sir/SIRT, sirtuin.

Structure and subcellular location of HDAC2

HDAC2 belongs to class I HDAC family, which is primarily located in the nucleus. HDAC2 is a specific enzyme with high activity and enantioselectivity to histone substrates (60). HDAC2 shares high structural homology and a common catalytic mechanism with other class I HDACs, particularly HDAC1. Similar to other class I HDACs, HDAC2 comprises a conserved deacetylase domain with short amino- and carboxy-terminal extensions, which are key for localization and maintaining their stability and function (48). HDAC1 and HDAC2 have notable amino acid homology. In large-scale gene expression analysis of brain and heart tissue, they affect different target gene sets by forming the same compressor complex (61). The crystal structure of human HDAC2 protein in the presence of hydroxamates has been revealed. The HDAC2 catalytic site is made up of a 'foot pocket', a lipophilic 'tube' and a catalytic Zn2+ 8 Å deep. More specifically, the 'foot pocket', tightly adjacent to the zinc binding site, is primarily formed by Tyr29, Met35, Phe114 and Leu144. The lipophilic tube, leading from the surface to the zinc binding site, is surrounded by Gly154, Phe155, His183, Phe210 and Leu276 (62,63). Furthermore, the zinc ion is accompanied by Asp181, His183, and Asp269 (6264). HDAC2 inhibitors typically have a pharmacophore comprising three sectors: A zinc-binding group, a linker portion and a hydrophobic cap group (6567). Based on molecular docking and virtual screening techniques, a series of compounds with novel skeletal structures have been identified as HDAC inhibitors, and their inhibitory activities and clinical therapeutic effects have been investigated (68,69). HDAC2 is the most thoroughly studied member of the HDAC family, which can be modulated by post-translational modifications, such as phosphorylation (70,71), acetylation (72), ubiquitination (73) and sumoylation (74). In particular, post-translational phosphorylation of HDAC2 negatively regulates its deacetylase activity and serves an active role in chronic inflammation (75). Furthermore, HDAC2 possesses several phosphorylation sites at the C-terminal, which are concentrated on serine residue (76).

3. Role of HDAC2 in hepatocyte apoptosis and proliferation

HDAC2 inhibits hepatocyte apoptosis

Due to its anatomical location and complex intersection, the liver is vulnerable to a variety of toxic, metabolic, inflammatory and necrotic stimuli (77,78). Hepatocytes are highly sensitive to death receptor-mediated apoptosis due to ubiquitous expression of death receptors in the liver (79). It has been demonstrated that the abnormal death of hepatocytes is a key trigger factor for the occurrence and development of both acute and chronic liver disease (77). It is therefore necessary to investigate the underlying pathogenesis and therapeutic targets of liver disease and the expression levels and pharmacological activity of HDAC2. There is increasing evidence that HDAC2 is involved in the regulation of hepatocyte death, thereby implying that HDAC2 serves a key role in the pathogenesis of liver disease (80,81). Here, experimental evidence for the effects of HDAC2 on hepatocyte apoptosis will be discussed.

TGF-β, a pleiotropic growth factor, has been shown to induce apoptosis in primary rat and AML-12 murine hepatocytes, consequently leading to liver injury and regeneration termination (80,81). Through transfection of HDAC2 RNA interference (RNAi) plasmid, Lei et al (80) found that HDAC2 serves as a significant anti-apoptotic factor in TGF-β1-induced apoptosis of AML-12. Lei et al further found that downregulation of HDAC2 significantly induces spontaneous cell apoptosis and increases the apoptotic response following TGF-β1 treatment. Moreover, ERK1/2 has been shown to be significantly inhibited in cells transfected with HDAC2i. In HDAC2i-transfected cells, low baseline levels of phosphorylated ERK1/2 were concomitant with decreased TGF-β1-induced apoptosis, suggesting that negative regulation of ERK1/2 is associated with the role of HDAC2 in apoptosis. In sum, these findings highlighted that silencing HDAC2 may induce spontaneous apoptosis of AML12 hepatocytes by promoting ERK1/2 expression and pharmacological activity. Taken together, this supports the important physiological role of HDAC2 in hepatocyte apoptosis. The anti-apoptotic effect of HDAC2 overexpression may be a promising therapeutic strategy for treatment of liver disease.

HDAC2 promotes proliferation of hepatocytes

Impaired proliferation of hepatocytes is associated with the occurrence and progression of liver disease (82). However, little is known about the underlying mechanisms that lead to defective hepatocyte proliferation. HDAC2 serves critical roles in cell proliferation and tissues regeneration (83). HDAC2 knockout inhibits proliferation and induces senescence of MCF7 cells by enhancing the binding activity and interaction of p53-DNA (84). HDAC2 deficiency usually results in different cellular phenotypes, suggesting that HDAC2 has a cell-type-specific role that may be relevant to the cell proliferative status (30). Turgeon et al (85) demonstrated that defects in tissue structure and perturbation of microenvironment homeostasis are accompanied by inhibition of cell proliferation when HDAC1/2 is knocked out. There is increasing evidence that HDAC2 promote the proliferation of liver cancer cell lines, although there is no indication of HDAC2 involvement in the proliferation of normal liver cells (8688). Ler et al (86) found that the combined knock out of HDAC2 and HDAC1 decreases cell proliferation and improves survival of patients with HCC. HDAC2 overexpression is routinely detected in cancer cells, and HDAC2 deficiency and inhibition lead to HCC cell apoptosis (87). The role of HDAC2 in cell proliferation was previously observed in the development of cardiac and B cells; HDAC2 and HDAC1 jointly inhibit cell cycle protein-dependent kinase p21 (WAF1/CIP1) and p57KIP2 transcription and promote progression from G1 to S phase (88). By contrast, HDAC2 suppresses transcription of p21WAF1/CIP1 via binding to Sp1-binding site enriched proximal region of the p21WAF1/CIP1 promoter (89).

4. HDAC2 in liver disease

HDAC2 promotes liver regeneration

Liver regeneration is of clinical significance in various types of liver disease (90,91). In the event of massive hepatocyte loss or damage, the intrinsic regenerative capacity of hepatocytes is activated by endogenous molecule-mediated signaling pathways (92). Rapid synchronous compensatory regeneration occurs following 2/3 partial hepatectomy (PH), and regenerated hepatocytes immediately enter the cell cycle and proliferate rapidly, restoring their original quality and function (93). Studies on the deficiency of HDAC2 in mouse liver regeneration have confirmed the key role of HDAC2 in liver regeneration (94,95). Following PH, the liver/body weight ratio is significantly lower in hepatocyte-selective HDAC2-/- mice compared with wild-type mice; HDAC2-/- mice also show more severe liver damage. Additionally, the expression of HDAC2 gradually increases within 0.5 to 3.0 days in mouse post-hepatectomy livers at an early stage of regeneration. Ki67, a mitotic marker, is decreased by ~30–70% in HDAC2 knockout mice, subsequently leading to defective mitosis. Decreased cyclinD1 and CDK2 in HDAC2-deficient hepatocytes suggests that HDAC2 liver-specific knockout triggers downregulation of cell cycle proteins and blocks cell cycle progression. Studies have shown that HDAC2 is expressed differently in male and female mice, and HDAC2 can directly bind to the promoter of B-myc (93,95). The expression of B-myc in the female liver is higher than in the male liver, which may be potential mechanism for the significantly slower rate of replication and quality reconstruction of individual female hepatocytes following PH. In conclusion, the altered metabolic pattern in HDAC2 knockout mice is consistent with the well-known regenerative characteristic of hepatocytes. This evidence also confirms a key role for HDAC2 in the metabolic response following PH (Fig. 2).

HDAC2 in liver fibrosis

Liver fibrosis, primarily characterized by excessive accumulation of extracellular matrix proteins, is a worldwide medical problem with increasing annual morbidity (96,97). The majority cases of liver fibrosis arise in the context of various etiology of liver damage, such as chronic viral infections (98), excessive alcohol consumption (99), metabolic disorder (100) or autoimmune disease (101). The regression and improvement of liver fibrosis are primarily attributed to inactivation and apoptosis of activated hepatic stellate cells (HSCs) (102). Notably, emerging evidence has revealed the potential features and roles of HDACs in the progression of liver fibrosis (103,104). It was also reported that several HDACs are involved in the activation of HSCs and the progression of hepatic fibrosis (105,106). In addition, accumulating evidence has highlighted the key role of HDAC2 in the development of renal fibrosis and pulmonary fibrosis (18,107,108). In this regard, it is worth verifying the functional role of HDAC2 in the occurrence and reversal of liver fibrosis.

In order to reveal the possible association between HDAC2 and liver fibrosis, a CCl4-induced mouse liver fibrosis and its spontaneous reversal model have been successively established (119); this supported the view that aberrant HDAC expression and activity participate in the occurrence and development of liver fibrosis. Expression of HDAC2 is increased during CCl4-induced liver fibrosis and significantly decreased during its reversal. Similarly, the expression of HDAC2 is also significantly increased in human hepatic fibrosis. Exposure of HSC-T6 cells to TGF-β1 results in increased HDAC2 expression in a dose- and time-dependent manner. Loss-of-function analyses have confirmed that loss of HDAC2 induces cell cycle arrest and inhibits the expression of collagen 1α1 and α-smooth muscle actin protein in HSC-T6 cells activated by TGF-β1 (110112). Mechanistically, it has been widely reported that HDAC2-small interfering (si)RNA leads to increased expression of SMAD7 compared with scramble siRNA-transfected groups (111,112). Collectively, these results suggest that HDAC2 activates HSCs and promotes the occurrence of liver fibrosis by suppressing SMAD7 expression. In conclusion, these findings may demonstrate the role of HDAC2 in the progression and reversal of liver fibrosis, and therefore have significant implications for the development of novel treatment strategies for liver fibrosis (Fig. 3).

HDAC2 in NASH

NASH, a more aggressive form of non-alcoholic fatty liver disease, is pathologically characterized by cell damage, inflammatory cell infiltration and hepatocyte ballooning (113,115). Sustained accumulation of reactive oxygen species (ROS) and resultant oxidative stress, mitochondrial dysfunction and accumulation of triglyceride and lipotoxic metabolites have been identified as contributing factors to NASH (115). To date, there are no current Food and Drug Administration (USA)-approved effective therapies to manage NASH (116). The inhibitory modulation of HDAC2 may contribute to the prevention of NASH (117).

Zhong et al (117) found that CD36 deficiency specifically upregulates monocyte chemotactic protein-1 (MCP-1) expression, thereby aggravating macrophage infiltration and hepatic inflammation. In addition, they also indicated that CD36 deficiency effectively suppresses nuclear HDAC2 expression by decreasing intracellular ROS and increasing the binding of acetyl histone 3 to MCP-1 promoter, which subsequently enhances expression of MCP-1, increases hepatic macrophage infiltration and promotes NASH development. Screening of 11 classic HDACs in CD36-/- mouse liver and CD36-deficient hepatocytes also revealed that CD36 deletion significantly inhibits nuclear expression of HDAC2 in hepatocytes, but not that of other HDACs. Taken together, this indicates that CD36 deficiency in hepatocytes promotes MCP-1 expression by inhibiting nuclear expression of HDAC2. Thus, the loss of CD36 results in decreased ROS levels, which lead to the development of NASH in mice by inhibiting the expression of HDAC2 and promoting that of MCP-1. Overall, maintaining a good balance between nuclear HDAC2 expression and hepatic ROS levels may be a potential novel therapeutic strategy for the prevention of NASH (Fig. 4).

HDAC2 promotes HCC

HCC is one of the most common types of solid malignancy and is driven by different molecular mechanisms (118,119). Researchers have linked gene expression signatures with the occurrence and prognosis of HCC and investigated gene expression patterns and potential therapeutic targets (120). Evidence suggests that HDAC2 is overexpressed in tumors, and HDAC2 downregulation leads to high expression levels of cell cycle circuit elements, including p21WAF1/Cip1, which is a well-characterized regulatory factor that serves a key role in cell senescence (121,122). With regard to liver cancer, HDAC2 promotes proliferation, and its aberrant expression may be a prognostic indicator of HCC (86).

A study found that HDAC2 overexpression is associated with poor survival of patients with low-grade and early-stage tumors, suggesting that HDAC2 is an independent and reliable predictor of survival of patients with HCC (123). Noh et al (89) assessed the tumorigenic potential of HDAC2, evaluated abnormal HDAC2 expression and investigated its regulatory mechanism in HCC; abnormal regulation of HDAC2 served a key role in HCC progression by regulating cell cycle regulatory components at the transcriptional level. Their data also showed that HDAC2 overexpression was not associated with Wnt and c-myc signaling pathways, which play an important role in malignant cell proliferation. Kim et al (124) investigated the underlying mechanism of HDAC2 in tumorigenesis; increased expression of casein kinase II (CK2α) was positively correlated with HDAC2. They proposed a regulatory mechanism whereby increased HDAC2 expression in HCC is primarily caused by the activation of CK2α/AKT pathways mediated by EGF. Lee et al (108) indicated that systemic delivery of HDAC2 siRNA encapsulated in lipid nanoparticles is sufficient to inhibit HCC progression. In addition, mTORC1 activation and NF-κB p50 nuclear translocation are essential for the transcriptional activation of oncogenic HDAC2 in HCC (125). Furthermore, 1,25(OH)2D3 inhibits the progression of HCC by downregulating HDAC2 (126,127). Consistently, Wang et al (128) also found that high levels of HDAC2 expression are negatively correlated with PTEN expression in HCC patients with poor prognosis (Fig. 5).

Merck60 selectively inhibits HDAC1 and HDAC2, thereby increasing histone acetylation and disrupting core gene regulatory architecture in rhabdomyosarcoma (129). Methot et al (130) investigated novel selective HDAC1/HDAC2 inhibitors (SHI-1:2), which incorporate a biaryl zinc-binding motif into a nicotinyl scaffold; the optimized SHI-1:2 structure exhibited notable inhibitory activity against HDAC1 and HDAC2, and its specific selectivity for HDAC1/HDAC2 was 100 times higher than that for other HDACs (131). N-(2-amino-5-substituted phenyl) benzamide significantly induced HCT116 cell death by specifically targeting HDAC2 (62). In addition, the effects of C15 urushiol and its triazole derivatives on the apoptosis of liver cancer cells have been qualitatively and quantitatively verified (132). Venturelli et al (133) reported that 6- and 8-prenylnaringenin enter into the 'foot pocket' of HDAC2 and combine with zinc ion of their catalytic center, subsequently inhibiting excessive proliferation of melanoma cells. N-[4-(Hydrazinecarbonyl)phenyl]-3,5,6-trimethylpyr- azine-2-carboxamide exhibits notable anticancer activity in vivo (IC50=1.60 μM) (134) Among squaramide-based derivatives, the lead compound 42 exhibits good druggability by specifically inhibiting HDAC2 (67). Isopropyl derivative 5 and tert-butyl derivative 6, is which derived from the lead compound NSC746457, exhibit a significantly inhibitory effect on HDAC2 (63). Novel indazole and pyrazolo(3,4-b) pyridine derivatives have been designed and synthesized via fragment-based virtual screening; biological evaluation showed that compounds 15k and 15m possess distinctly inhibitory effect towards HDAC2 (66). Rosmarinic acid has been demonstrated to downregulate HDAC2 expression, subsequently leading to cell cycle arrest and apoptosis (135). N-(2-aminophenyl)–4-[(4-fluorophenoxy)methyl] benzamide exhibits antitumor activity by inhibiting HDAC2 at an IC50 of 3.84 μM (65). In addition, a series of 2-aminobenzamide-based compounds exhibit highly inhibitory effects on solid cancer cell lines and low cytotoxicity against normal cells (Table II) (68).

Table II

Lead compounds of potential histone deacetylase 2 inhibitors.

Table II

Lead compounds of potential histone deacetylase 2 inhibitors.

InhibitorFunctionIC50, nMAnalysis method(Refs.)
Compound 12Anticancer44.0SAR(136)
Compound 5/8Anticancer27.0/39.0SAR(59)
K560 Neuroprotection520.0VS(138)
Compound 4bND40.6SAR(87)
RH01652Prevent ADNDMDs-QC(139)
BRD8430Neuroblastoma differentiationNDHTS(140)
Eight hits urushiol derivativesNDNDPM-VS(141)
C15 TrieneNDNDMD-MDs-VS(61)
TriazoleAnticancerNDSAR-MD(143)
ST088357ND16870.0Scaffold-Merging HQ(148)
6-PN/8-PNAnticancerNDSAR(145)
Compound 4g/6c/6gND 130.0/160.0/580.0MD(146)
LAQ824ND3.0MD(147)
Compound 7aAnticancer53.7QSAR(148)
Squaramide-based hydroxamic acidsAnticancerNDSAR(64)
ID5/TD6Anticancer22.0/18.0 Click-chemistry/MD(60)
Compound 15k/15mAnticancer4.2/3.6VS-MD(63)
Rosmarinic acidAnticancerNDActivity screening(149)
Compound 12aAnticancerNDMD-SAR(62)
Compounds12g/12hAnticancer205.0/144.0SAR(65)

[i] ND, no description available; SAR, structure-activity relationship; VS, virtual screening; MDs, molecular dynamics; QC, quantum chemistry; HTS, high-throughput screening; PM, pharmacophore modeling; MD, molecular docking; HQ, hybrid query; QSAR, quantitative SAR.

ncRNAs mediate HDAC2 expression at the posttranscriptional level (136,137). A meta-analysis of 1,258 HCC samples showed that downregulation of microRNA (miRNA or miR)-100-5p contributes to the progression and prognosis of HCC by negatively regulating HDAC2 expression (138). Noh et al (87) found that miR-145 functions as a tumor suppressor by directly targeting HDAC2 in liver cancer. In addition, another study demonstrated that miR-31 significantly decreases HDAC2 expression by suppressing mRNA translation in HCC cells (139). In sum, these results suggest that promotion or suppression of certain miRNAs cause aberrant expression of HDAC2, which be involved in HCC tumorigenesis. Dai et al (140) found that long ncRNA SNHG15 upregulates HDAC2 expression by sponging miR-490–3p, which further promotes HCC progression.

In summary, these results demonstrate that HDAC2 possesses carcinogenic properties. These studies also suggest that the development of novel compounds or ncRNAs may be a promising therapeutic modality for liver cancer by selectively targeting HDAC2.

5. Conclusion

Epigenetic modifications serve prime regulatory roles in genetic events, such as transcriptional activation and silencing (141). The effects of epigenetic modifications have been recognized, although their specific roles may still be controversial. Histone acetylation contributes to gene expression, while histone deacetylation leads to suppression of gene transcription (142). Studies have shown that relative levels of histone acetylation and deacetylation are of significance for the regulation of pathophysiological processes, including proliferation, cell-cycle progression, differentiation, immune evasion, inflammatory lesion, apoptosis and death (143145). Pharmacological inhibition of HDAC activity or expression alters chromatin acetylation levels, subsequently confusing boundaries between transcriptionally active and quiescent chromatin (146148).

The increasing incidence of liver disease requires novel effective therapeutic interventions. HDAC2 exhibits attractive pharmacological effects in hepatocyte loss or injury, HCC and NASH by modulating hepatocyte death and regulating cell cycle components. In the past decades, researchers have characterized HDAC classification, structure and subcellular localization (4750,52,54). The malignant or beneficial role of HDAC2 in liver fibrosis, non-alcoholic fatty liver disease and liver cancer has been revealed. Nonetheless, the mechanism of HDAC2 in the development of liver disease has not been elucidated and more investigations are needed in future. In summary, these properties of HDAC2 make it an appealing therapeutic target by regulating expression of HDAC2 and HDAC2-dependent signaling pathways.

HDAC2 presents favorable characteristics as a potential drug target. Consequently, pharmacological agents that inhibit HDAC2 may be a prospective treatment for liver ailments. However, the few known HDAC2 inhibitors are broad spectrum inhibitors that simultaneously inhibit several members of HDACs family and thus may have more potential adverse side effects. Therefore, it is necessary to develop novel HDAC2 inhibitors with higher targeting selectivity. High homology and cellular co-localization of multiple HDACs makes development and use of HDAC2 inhibitors difficult. The expression and pharmacological activity of HDAC2 is important for the prediction, diagnosis and prognosis of liver disease. NF-κB, c-Myc, Sp1 and Sp3 can bind to the HDAC2 promoter, thereby augmenting HDAC2 transcription (149151). These findings suggest that the HDAC2 promoter may also be a potential target for pharmacological intervention.

The present study reviewed the specific roles of HDAC2 and the potential application of HDAC2 inhibitors in liver disease. Increasing evidence has highlighted the key role of HDAC2 in the occurrence and development of liver disease and demonstrated that HDAC2 inhibitor therapy may be a therapeutic approach. Better understanding of the potential roles and regulatory mechanisms of HDAC2 in liver disease may improve the ability to predict the pace of liver disease progression and exploit specific targeted therapeutic strategies.

Availability of data and materials

Not applicable.

Authors' contributions

YRL, JL and JQW conceived and designed the study and wrote the manuscript. ZGH, RNC and XC prepared the figures. DCZ and HXY prepared the tables. QX, XRW and HYZ revised the manuscript. All authors read and approved the final version of the manuscript. Data authentication is not applicable.

Ethics approval and consent to participate

Not applicable.

Patient consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Acknowledgments

Not applicable.

Funding

The present study was supported by the National Natural Science Foundation of China (grant no. 81770609).

References

1 

Cai C, Yu H, Huang G, Du X, Yu X, Zhou Y and Shen W: Histone modifications in fatty acid synthase modulated by carbohydrate responsive element binding protein are associated with non-alcoholic fatty liver disease. Int J Mol Med. 42:1215–1228. 2018.PubMed/NCBI

2 

Ferriero R, Nusco E, De Cegli R, Carissimo A, Manco G and Brunetti-Pierri N: Pyruvate dehydrogenase complex and lactate dehydrogenase are targets for therapy of acute liver failure. J Hepatol. 69:325–335. 2018. View Article : Google Scholar : PubMed/NCBI

3 

Kanyal A, Rawat M, Gurung P, Choubey D, Anamika K and Karmodiya K: Genome-wide survey and phylogenetic analysis of histone acetyltransferases and histone deacetylases of Plasmodium falciparum. FEBS J. 285:1767–1782. 2018. View Article : Google Scholar

4 

Berger SL: The complex language of chromatin regulation during transcription. Nature. 447:407–412. 2007. View Article : Google Scholar : PubMed/NCBI

5 

Khangura RK, Bali A, Jaggi AS and Singh N: Histone acetylation and histone deacetylation in neuropathic pain: An unresolved puzzle? Eur J Pharmacol. 795:36–42. 2017. View Article : Google Scholar

6 

Kouzarides T: Chromatin modifications and their function. Cell. 128:693–705. 2007. View Article : Google Scholar : PubMed/NCBI

7 

Leipe DD and Landsman D: Histone deacetylases, acetoin utilization proteins and acetylpolyamine amidohydrolases are members of an ancient protein superfamily. Nucleic Acids Res. 25:3693–3697. 1997. View Article : Google Scholar : PubMed/NCBI

8 

West AC and Johnstone RW: New and emerging HDAC inhibitors for cancer treatment. J Clin Invest. 124:30–39. 2014. View Article : Google Scholar : PubMed/NCBI

9 

Ibi D and Gonzalez-Maeso J: Epigenetic signaling in schizophrenia. Cell Signal. 27:2131–2136. 2015. View Article : Google Scholar : PubMed/NCBI

10 

Levenson JM, O'Riordan KJ, Brown KD, Trinh MA, Molfese DL and Sweatt JD: Regulation of histone acetylation during memory formation in the hippocampus. J Biol Chem. 279:40545–40559. 2004. View Article : Google Scholar : PubMed/NCBI

11 

Cao YN, Xue Y, Xue L, Jiang X, Wang X, Zhang Z, Yang J, Lu J, Zhang C, Wang W and Ning G: Hepatic menin recruits SIRT1 to control liver steatosis through histone deacetylation. J Hepatol. 59:1299–1306. 2013. View Article : Google Scholar : PubMed/NCBI

12 

Simões-Pires C, Zwick V, Nurisso A, Schenker E, Carrupt PA and Cuendet M: HDAC6 as a target for neurodegenerative diseases: What makes it different from the other HDACs? Mol Neurodegener. 8:72013. View Article : Google Scholar : PubMed/NCBI

13 

Kim HJ and Bae SC: Histone deacetylase inhibitors: Molecular mechanisms of action and clinical trials as anti-cancer drugs. Am J Transl Res. 3:166–179. 2011.PubMed/NCBI

14 

Glauben R, Batra A, Stroh T, Erben U, Fedke I, Lehr HA, Leoni F, Mascagni P, Dinarello CA, Zeitz M and Siegmund B: Histone deacetylases: Novel targets for prevention of colitis-associated cancer in mice. Gut. 57:613–622. 2008. View Article : Google Scholar : PubMed/NCBI

15 

Dokmanovic M, Clarke C and Marks PA: Histone deacetylase inhibitors: Overview and perspectives. Mol Cancer Res. 5:981–989. 2007. View Article : Google Scholar : PubMed/NCBI

16 

Marchion DC, Bicaku E, Turner JG, Schmitt ML, Morelli DR and Munster PN: HDAC2 regulates chromatin plasticity and enhances DNA vulnerability. Mol Cancer Ther. 8:794–801. 2009. View Article : Google Scholar : PubMed/NCBI

17 

Jahan S, Sun JM, He S and Davie JR: Transcription-dependent association of HDAC2 with active chromatin. J Cell Physiol. 233:1650–1657. 2018. View Article : Google Scholar

18 

Noh H, Oh EY, Seo JY, Yu MR, Kim YO, Ha H and Lee HB: Histone deacetylase-2 is a key regulator of diabetes- and transforming growth factor-beta1-induced renal injury. Am J Physiol Renal Physiol. 297:F729–F739. 2009. View Article : Google Scholar : PubMed/NCBI

19 

Yang FQ, Liu M, Yang FP, Che J, Li W, Zhai W, Wang GC, Zheng JH and Li X: VPA inhibits renal cancer cell migration by targeting HDAC2 and down-regulating HIF–1α. Mol Biol Rep. 41:1511–1518. 2014. View Article : Google Scholar : PubMed/NCBI

20 

Fritzsche FR, Weichert W, Roske A, Gekeler V, Beckers T, Stephan C, Jung K, Scholman K, Denkert C, Dietel M and Kristiansen G: Class I histone deacetylases 1, 2 and 3 are highly expressed in renal cell cancer. BMC Cancer. 8:3812008. View Article : Google Scholar : PubMed/NCBI

21 

Shang L, Pin L, Zhu S, Zhong X, Zhang Y, Shun M, Liu Y and Hou M: Plantamajoside attenuates isoproterenol-induced cardiac hypertrophy associated with the HDAC2 and AKT/GSK–3β signaling pathway. Chem Biol Interact. 307:21–28. 2019. View Article : Google Scholar : PubMed/NCBI

22 

Datta M, Staszewski O, Raschi E, Frosch M, Hagemeyer N, Tay TL, Blank T, Kreutzfeldt M, Merkler D, Ziegler-Waldkirch S, et al: Histone deacetylases 1 and 2 regulate microglia function during development, homeostasis, and neurodegeneration in a context-dependent manner. Immunity. 48:514–529.e6. 2018. View Article : Google Scholar : PubMed/NCBI

23 

Bin YF, Wu LJ, Sun XJ, Liang Y, Bai J, Zhang JQ, Li MH, Zhong XN, Liang YJ and He ZY: Expression of GR-α and HDAC2 in steroid-Sensitive and steroid-Insensitive interstitial lung disease. Biomed Pharmacother. 118:1093802019. View Article : Google Scholar

24 

Mahady L, Nadeem M, Malek-Ahmadi M, Chen K, Perez SE and Mufson EJ: HDAC2 dysregulation in the nucleus basalis of Meynert during the progression of Alzheimer's disease. Neuropathol Appl Neurobiol. 45:380–397. 2019. View Article : Google Scholar

25 

Lin CL, Tsai ML, Lin CY, Hsu KW, Hsieh WS, Chi WM, Huang LC and Lee CH: HDAC1 and HDAC2 double knockout triggers cell apoptosis in advanced thyroid cancer. Int J Mol Sci. 20:4542019. View Article : Google Scholar :

26 

Stojanovic N, Hassan Z, Wirth M, Wenzel P, Beyer M, Schäfer C, Brand P, Kroemer A, Stauber RH, Schmid RM, et al: HDAC1 and HDAC2 integrate the expression of p53 mutants in pancreatic cancer. Oncogene. 36:1804–1815. 2017. View Article : Google Scholar

27 

Tang W, Zhou W, Xiang L, Wu X, Zhang P, Wang J, Liu G, Zhang W, Peng Y, Huang X, et al: The p300/YY1/miR-500a-5p/HDAC2 signalling axis regulates cell proliferation in human colorectal cancer. Nat Commun. 10:6632019. View Article : Google Scholar : PubMed/NCBI

28 

Lai T, Wu M, Zhang C, Che L, Xu F, Wang Y, Wu Y, Xuan N, Cao C, Du X, et al: HDAC2 attenuates airway inflammation by suppressing IL-17A production in HDM-challenged mice. Am J Physiol Lung Cell Mol Physiol. 316:L269–L279. 2019. View Article : Google Scholar

29 

Barnes PJ: Corticosteroid resistance in patients with asthma and chronic obstructive pulmonary disease. J Allergy Clin Immunol. 131:636–645. 2013. View Article : Google Scholar : PubMed/NCBI

30 

Wilting RH, Yanover E, Heideman MR, Jacobs H, Horner J, van der Torre J, DePinho RA and Dannenberg JH: Overlapping functions of Hdac1 and Hdac2 in cell cycle regulation and haematopoiesis. EMBO J. 29:2586–2597. 2010. View Article : Google Scholar : PubMed/NCBI

31 

Wang Y, Yang F, Jiao FZ, Chen Q, Zhang WB, Wang LW and Gong ZJ: Modulations of histone deacetylase 2 offer a protective effect through the mitochondrial apoptosis pathway in acute liver failure. Oxid Med Cell Longev. 2019:81730162019.PubMed/NCBI

32 

Wu J, Zhu P, Lu T, Du Y, Wang Y, He L, Ye B, Liu B, Yang L, Wang J, et al: The long non-coding RNA LncHDAC2 drives the self-renewal of liver cancer stem cells via activation of Hedgehog signaling. J Hepatol. 70:918–929. 2019. View Article : Google Scholar

33 

Verdone L, Agricola E, Caserta M and Di Mauro E: Histone acetylation in gene regulation. Brief Funct Genomic Proteomic. 5:209–221. 2006. View Article : Google Scholar : PubMed/NCBI

34 

Millard CJ, Fairall L, Ragan TJ, Savva CG and Schwabe JWR: The topology of chromatin-binding domains in the NuRD deacetylase complex. Nucleic Acids Res. 48:12972–12982. 2020. View Article : Google Scholar : PubMed/NCBI

35 

Verdone L, Caserta M and Di Mauro E: Role of histone acetylation in the control of gene expression. Biochem Cell Biol. 83:344–353. 2005. View Article : Google Scholar : PubMed/NCBI

36 

Brownell JE, Zhou J, Ranalli T, Kobayashi R, Edmondson DG, Roth SY and Allis CD: Tetrahymena histone acetyltransferase A: A homolog to yeast Gcn5p linking histone acetylation to gene activation. Cell. 84:843–851. 1996. View Article : Google Scholar : PubMed/NCBI

37 

Kurdistani SK and Grunstein M: Histone acetylation and deacetylation in yeast. Nat Rev Mol Cell Biol. 4:276–284. 2003. View Article : Google Scholar : PubMed/NCBI

38 

Wang D, Kon N, Lasso G, Jiang L, Leng W, Zhu WG, Qin J, Honig B and Gu W: Acetylation-regulated interaction between p53 and SET reveals a widespread regulatory mode. Nature. 538:118–122. 2016. View Article : Google Scholar : PubMed/NCBI

39 

Wang Z, Zang C, Cui K, Schones DE, Barski A, Peng W and Zhao K: Genome-wide mapping of HATs and HDACs reveals distinct functions in active and inactive genes. Cell. 138:1019–1031. 2009. View Article : Google Scholar : PubMed/NCBI

40 

Abel T and Zukin RS: Epigenetic targets of HDAC inhibition in neurodegenerative and psychiatric disorders. Curr Opin Pharmacol. 8:57–64. 2008. View Article : Google Scholar : PubMed/NCBI

41 

de Ruijter AJ, van Gennip AH, Caron HN, Kemp S and van Kuilenburg AB: Histone deacetylases (HDACs): Characterization of the classical HDAC family. Biochem J. 370(Pt 3): 737–749. 2003. View Article : Google Scholar

42 

Gregoretti IV, Lee YM and Goodson HV: Molecular evolution of the histone deacetylase family: Functional implications of phylogenetic analysis. J Mol Biol. 338:17–31. 2004. View Article : Google Scholar : PubMed/NCBI

43 

Kiweler N, Brill B, Wirth M, Breuksch I, Laguna T, Dietrich C, Strand S, Schneider G, Groner B, Butter F, et al: The histone deacetylases HDAC1 and HDAC2 are required for the growth and survival of renal carcinoma cells. Arch Toxicol. 92:2227–2243. 2018. View Article : Google Scholar : PubMed/NCBI

44 

Bush EW and McKinsey TA: Protein acetylation in the cardiorenal axis: The promise of histone deacetylase inhibitors. Circ Res. 106:272–284. 2010. View Article : Google Scholar : PubMed/NCBI

45 

Yang XJ and Seto E: The Rpd3/Hda1 family of lysine deacetylases: From bacteria and yeast to mice and men. Nat Rev Mol Cell Biol. 9:206–218. 2008. View Article : Google Scholar : PubMed/NCBI

46 

Yang WM, Tsai SC, Wen YD, Fejer G and Seto E: Functional domains of histone deacetylase-3. J Biol Chem. 277:9447–9454. 2002. View Article : Google Scholar : PubMed/NCBI

47 

Martin M, Kettmann R and Dequiedt F: Class IIa histone deacetylases: Regulating the regulators. Oncogene. 26:5450–5467. 2007. View Article : Google Scholar : PubMed/NCBI

48 

Haberland M, Montgomery RL and Olson EN: The many roles of histone deacetylases in development and physiology: Implications for disease and therapy. Nat Rev Genet. 10:32–42. 2009. View Article : Google Scholar

49 

Guardiola AR and Yao TP: Molecular cloning and characterization of a novel histone deacetylase HDAC10. J Biol Chem. 277:3350–3356. 2002. View Article : Google Scholar

50 

Grozinger CM, Hassig CA and Schreiber SL: Three proteins define a class of human histone deacetylases related to yeast Hda1p. Proc Natl Acad Sci USA. 96:4868–4873. 1999. View Article : Google Scholar : PubMed/NCBI

51 

Marks PA and Breslow R: Dimethyl sulfoxide to vorinostat: Development of this histone deacetylase inhibitor as an anti-cancer drug. Nat Biotechnol. 25:84–90. 2007. View Article : Google Scholar : PubMed/NCBI

52 

Johnstone RW: Histone-deacetylase inhibitors: Novel drugs for the treatment of cancer. Nat Rev Drug Discov. 1:287–299. 2002. View Article : Google Scholar : PubMed/NCBI

53 

Michan S and Sinclair D: Sirtuins in mammals: Insights into their biological function. Biochem J. 404:1–13. 2007. View Article : Google Scholar : PubMed/NCBI

54 

Abbas A and Gupta S: The role of histone deacetylases in prostate cancer. Epigenetics. 3:300–309. 2008. View Article : Google Scholar : PubMed/NCBI

55 

Yamamoto H, Schoonjans K and Auwerx J: Sirtuin functions in health and disease. Mol Endocrinol. 21:1745–1755. 2007. View Article : Google Scholar : PubMed/NCBI

56 

Xu WS, Parmigiani RB and Marks PA: Histone deacetylase inhibitors: Molecular mechanisms of action. Oncogene. 26:5541–5552. 2007. View Article : Google Scholar : PubMed/NCBI

57 

Gao L, Cueto MA, Asselbergs F and Atadja P: Cloning and functional characterization of HDAC11, a novel member of the human histone deacetylase family. J Biol Chem. 277:25748–25755. 2002. View Article : Google Scholar : PubMed/NCBI

58 

Verdin E and Ott M: 50 years of protein acetylation: From gene regulation to epigenetics, metabolism and beyond. Nat Rev Mol Cell Biol. 16:258–264. 2015. View Article : Google Scholar : PubMed/NCBI

59 

Gong F and Miller KM: Mammalian DNA repair: HATs and HDACs make their mark through histone acetylation. Mutat Res. 750:23–30. 2013. View Article : Google Scholar : PubMed/NCBI

60 

Brunmeir R, Lagger S and Seiser C: Histone deacetylase HDAC1/HDAC2-controlled embryonic development and cell differentiation. Int J Dev Biol. 53:275–289. 2009. View Article : Google Scholar : PubMed/NCBI

61 

Montgomery RL, Hsieh J, Barbosa AC, Richardson JA and Olson EN: Histone deacetylases 1 and 2 control the progression of neural precursors to neurons during brain development. Proc Natl Acad Sci USA. 106:7876–7881. 2009. View Article : Google Scholar : PubMed/NCBI

62 

Bressi JC, Jennings AJ, Skene R, Wu Y, Melkus R, De Jong R, O'Connell S, Grimshaw CE, Navre M and Gangloff AR: Exploration of the HDAC2 foot pocket: Synthesis and SAR of substituted N-(2-aminophenyl)benzamides. Bioorg Med Chem Lett. 20:3142–3145. 2010. View Article : Google Scholar : PubMed/NCBI

63 

Hou J, Feng C, Li Z, Fang Q, Wang H, Gu G, Shi Y, Liu P, Xu F, Yin Z, et al: Structure-based optimization of click-based histone deacetylase inhibitors. Eur J Med Chem. 46:3190–3200. 2011. View Article : Google Scholar : PubMed/NCBI

64 

Zhou H, Wang C, Ye J, Chen H and Tao R: Design, virtual screening, molecular docking and molecular dynamics studies of novel urushiol derivatives as potential HDAC2 selective inhibitors. Gene. 637:63–71. 2017. View Article : Google Scholar : PubMed/NCBI

65 

Xie R, Yao Y, Tang P, Chen G, Liu X, Yun F, Cheng C, Wu X and Yuan Q: Design, synthesis and biological evaluation of novel hydroxamates and 2-aminobenzamides as potent histone deacetylase inhibitors and antitumor agents. Eur J Med Chem. 134:1–12. 2017. View Article : Google Scholar : PubMed/NCBI

66 

Liu J, Zhou J, He F, Gao L, Wen Y, Gao L, Wang P, Kang D and Hu L: Design, synthesis and biological evaluation of novel indazole-based derivatives as potent HDAC inhibitors via fragment-based virtual screening. Eur J Med Chem. 192:112–189. 2020. View Article : Google Scholar

67 

Fournier JF, Bhurruth-Alcor Y, Musicki B, Aubert J, Aurelly M, Bouix-Peter C, Bouquet K, Chantalat L, Delorme M, Drean B, et al: Squaramides as novel class I and IIB histone deacetylase inhibitors for topical treatment of cutaneous t-cell lymphoma. Bioorg Med Chem Lett. 28:2985–2992. 2018. View Article : Google Scholar : PubMed/NCBI

68 

Yun F, Cheng C, Ullah S, He J, Zahi MR and Yuan Q: Thioether-based 2-aminobenzamide derivatives: Novel HDAC inhibitors with potent in vitro and in vivo antitumor activity. Eur J Med Chem. 176:195–207. 2019. View Article : Google Scholar : PubMed/NCBI

69 

Alsawalha M, Rao Bolla S, Kandakatla N, Srinivasadesikan V, Veeraraghavan VP and Surapaneni KM: Molecular docking and ADMET analysis of hydroxamic acids as HDAC2 inhibitors. Bioinformation. 15:380–387. 2019. View Article : Google Scholar : PubMed/NCBI

70 

Ford J, Ahmed S, Allison S, Jiang M and Milner J: JNK2-dependent regulation of SIRT1 protein stability. Cell cycle. 7:3091–3097. 2008. View Article : Google Scholar : PubMed/NCBI

71 

Sun JM, Chen HY and Davie JR: Differential distribution of unmodified and phosphorylated histone deacetylase 2 in chromatin. J Biol Chem. 282:33227–33236. 2007. View Article : Google Scholar : PubMed/NCBI

72 

Ashktorab H, Belgrave K, Hosseinkhah F, Brim H, Nouraie M, Takkikto M, Hewitt S, Lee EL, Dashwood RH and Smoot D: Global Histone H4 Acetylation and HDAC2 expression in colon adenoma and carcinoma. Dig Dis Sci. 54:2109–2117. 2009. View Article : Google Scholar :

73 

Krämer OH, Zhu P, Ostendorff HP, Golebiewski M, Tiefenbach J, Peters MA, Brill B, Groner B, Bach I, Heinzel T and Göttlicher M: The histone deacetylase inhibitor valproic acid selectively induces proteasomal degradation of HDAC2. EMBO J. 22:3411–3420. 2003. View Article : Google Scholar : PubMed/NCBI

74 

Brandl A, Wagner T, Uhlig KM, Knauer SK, Stauber RH, Melchior F, Schneider G, Heinzel T and Krämer OH: Dynamically regulated sumoylation of HDAC2 controls p53 deacetylation and restricts apoptosis following genotoxic stress. J Mol Cell Biol. 4:284–293. 2012. View Article : Google Scholar : PubMed/NCBI

75 

Adenuga D and Rahman I: Protein kinase CK2-mediated phosphorylation of HDAC2 regulates co-repressor formation, deacetylase activity and acetylation of HDAC2 by cigarette smoke and aldehydes. Arch Biochem Biophys. 498:62–73. 2010. View Article : Google Scholar : PubMed/NCBI

76 

Tsai SC and Seto E: Regulation of histone deacetylase 2 by protein kinase CK2. J Biol Chem. 277:31826–31833. 2002. View Article : Google Scholar : PubMed/NCBI

77 

Chen PJ, Cai SP, Huang C, Meng XM and Li J: Protein tyrosine phosphatase 1B (PTP1B): A key regulator and therapeutic target in liver diseases. Toxicology. 337:10–20. 2015. View Article : Google Scholar : PubMed/NCBI

78 

Kan C, Ungelenk L, Lupp A, Dirsch O and Dahmen U: Ischemia-Reperfusion injury in aged Livers-The energy metabolism, inflammatory response, and autophagy. Transplantation. 102:368–377. 2018. View Article : Google Scholar

79 

Guicciardi ME, Malhi H, Mott JL and Gores GJ: Apoptosis and necrosis in the liver. Compr Physiol. 3:977–1010. 2013.PubMed/NCBI

80 

Lei WW, Zhang KH, Pan XC, Wang DM, Hu Y, Yang YN and Song JG: Histone deacetylase 1 and 2 differentially regulate apoptosis by opposing effects on extracellular signal-regulated kinase 1/2. Cell Death Dis. 1:e442010. View Article : Google Scholar

81 

Romero-Gallo J, Sozmen EG, Chytil A, Russell WE, Whitehead R, Parks WT, Holdren MS, Her MF, Gautam S, Magnuson M, et al: Inactivation of TGF-beta signaling in hepatocytes results in an increased proliferative response after partial hepatectomy. Oncogene. 24:3028–3041. 2005. View Article : Google Scholar : PubMed/NCBI

82 

Raven A, Lu WY, Man TY, Ferreira-Gonzalez S, O'Duibhir E, Dwyer BJ, Thomson JP, Meehan RR, Bogorad R, Koteliansky V, et al: Cholangiocytes act as facultative liver stem cells during impaired hepatocyte regeneration. Nature. 547:350–354. 2017. View Article : Google Scholar : PubMed/NCBI

83 

Willis-Martinez D, Richards HW, Timchenko NA and Medrano EE: Role of HDAC1 in senescence, aging, and cancer. Exp Gerontol. 45:279–285. 2010. View Article : Google Scholar :

84 

Harms KL and Chen X: Histone deacetylase 2 modulates p53 transcriptional activities through regulation of p53-DNA binding activity. Cancer Res. 67:3145–3152. 2007. View Article : Google Scholar : PubMed/NCBI

85 

Turgeon N, Blais M, Gagne JM, Tardif V, Boudreau F, Perreault N and Asselin C: HDAC1 and HDAC2 restrain the intestinal inflammatory response by regulating intestinal epithelial cell differentiation. PLoS One. 8:e737852013. View Article : Google Scholar : PubMed/NCBI

86 

Ler SY, Leung CH, Khin LW, Lu GD, Salto-Tellez M, Hartman M, Iau PT, Yap CT and Hooi SC: HDAC1 and HDAC2 independently predict mortality in hepatocellular carcinoma by a competing risk regression model in a Southeast Asian population. Oncol Rep. 34:2238–2250. 2015. View Article : Google Scholar : PubMed/NCBI

87 

Noh JH, Chang YG, Kim MG, Jung KH, Kim JK, Bae HJ, Eun JW, Shen Q, Kim SJ, Kwon SH, et al: MiR–145 functions as a tumor suppressor by directly targeting histone deacetylase 2 in liver cancer. Cancer Lett. 335:455–462. 2013. View Article : Google Scholar : PubMed/NCBI

88 

Makar AB, Mcmartin KE, Palese M and Tephly TR: Formate assay in body fluids: Application in methanol poisoning App. Biochem Med. 13:117–126. 1975. View Article : Google Scholar : PubMed/NCBI

89 

Noh JH, Jung KH, Kim JK, Eun JW, Bae HJ, Xie HJ, Chang YG, Kim MG, Park WS, Lee JY and Nam SW: Aberrant Regulation of HDAC2 Mediates proliferation of hepatocellular carcinoma cells by deregulating expression of G1/S cell cycle proteins. PLoS One. 6:e281032011. View Article : Google Scholar : PubMed/NCBI

90 

Yuan X, Yan S, Zhao J, Shi D, Yuan B, Dai W, Jiao B, Zhang W and Miao M: Lipid metabolism and peroxisome proliferator-activated receptor signaling pathways participate in late-phase liver regeneration. J Proteome Res. 10:1179–1190. 2011. View Article : Google Scholar : PubMed/NCBI

91 

Michalopoulos GK and Bhushan B: Liver regeneration: Biological and pathological mechanisms and implications. Nat Rev Gastroenterol Hepatol. 18:40–55. 2021. View Article : Google Scholar

92 

Michalopoulos GK: Principles of liver regeneration and growth homeostasis. Compr Physiol. 3:485–513. 2013.PubMed/NCBI

93 

Li L, Guo J, Chen Y, Chang C and Xu C: Comprehensive CircRNA expression profile and selection of key CircRNAs during priming phase of rat liver regeneration. BMC Genomics. 18:802017. View Article : Google Scholar : PubMed/NCBI

94 

Xia J, Zhou Y, Ji H, Wang Y, Wu Q, Bao J, Ye F, Shi Y and Bu H: Loss of Histone Deacetylases 1 and 2 in hepatocytes impairs murine liver regeneration through Ki67 depletion. Hepatology. 58:2089–2098. 2013. View Article : Google Scholar : PubMed/NCBI

95 

Wang Y, Ye F, Ke Q, Wu Q, Yang R and Bu H: Gender-dependent histone deacetylases injury may contribute to differences in liver recovery rates of male and female mice. Transplant Proc. 45:463–473. 2013. View Article : Google Scholar : PubMed/NCBI

96 

Bansal R, Nagorniewicz B and Prakash J: Clinical advancements in the targeted therapies against liver fibrosis. Mediators Inflamm. 2016:76297242016. View Article : Google Scholar : PubMed/NCBI

97 

Aydın MM and Akçalı KC: Liver fibrosis. Turk J Gastroenterol. 29:14–21. 2018. View Article : Google Scholar

98 

Gounder PP, Haering C, Bruden DJ, Townshend-Bulson L, Simons BC, Spradling PR and McMahon BJ: Does incorporating change in APRI or FIB–4 indices over time improve the accuracy of a single index for identifying liver fibrosis in persons with chronic hepatitis C virus infection? J Clin Gastroenterol. 52:60–66. 2018. View Article : Google Scholar

99 

Bilal U, Lau B, Lazo M, McCaul ME, Hutton HE, Sulkowski MS, Moore RD and Chander G: Interaction between alcohol consumption patterns, antiretroviral therapy type, and liver fibrosis in persons living with HIV. AIDS Patient Care STDS. 30:200–207. 2016. View Article : Google Scholar : PubMed/NCBI

100 

Lainé F, Bendavid C, Moirand R, Tessier S, Perrin M, Guillygomarc'h A, Guyader D, Calon E, Renault A, Brissot P, et al: Prediction of liver fibrosis in patients with features of the metabolic syndrome regardless of alcohol consumption. Hepatology. 39:1639–1646. 2004. View Article : Google Scholar : PubMed/NCBI

101 

Sunami Y, Leithäuser F, Gul S, Fiedler K, Güldiken N, Espenlaub S, Holzmann KH, Hipp N, Sindrilaru A, Luedde T, et al: Hepatic activation of IKK/NFκB signaling induces liver fibrosis via macrophage-mediated chronic inflammation. Hepatology. 56:1117–1128. 2012. View Article : Google Scholar : PubMed/NCBI

102 

Zhang H, Wu P, Chen F, Hao Y, Lao Y, Ren L, Sun L, Sun W, Wei H, Chan DW, et al: SILAC-based quantitative proteomic analysis of secretome between activated and reverted hepatic stellate cells. Proteomics. 14:1977–1986. 2014. View Article : Google Scholar : PubMed/NCBI

103 

Mannaerts I, Eysackers N, Onyema OO, Van Beneden K, Valente S, Mai A, Odenthal M and van Grunsven LA: Class II HDAC inhibition hampers hepatic stellate cell activation by induction of microRNA–29. PLoS One. 8:e557862013. View Article : Google Scholar

104 

Pannem RR, Dorn C, Hellerbrand C and Massoumi R: Cylindromatosis gene CYLD regulates hepatocyte growth factor expression in hepatic stellate cells through interaction with histone deacetylase 7. Hepatology. 60:1066–1081. 2014. View Article : Google Scholar : PubMed/NCBI

105 

Mannaerts I, Nuytten NR, Rogiers V, Vanderkerken K, van Grunsven LA and Geerts A: Chronic administration of valproic acid inhibits activation of mouse hepatic stellate cells in vitro and in vivo. Hepatology. 51:603–614. 2010. View Article : Google Scholar

106 

Qin L and Han YP: Epigenetic repression of matrix metalloproteinases in myofibroblastic hepatic stellate cells through histone deacetylases 4: Implication in tissue fibrosis. Am J Pathol. 177:1915–1928. 2010. View Article : Google Scholar : PubMed/NCBI

107 

Huang SK, Scruggs AM, Donaghy J, Horowitz JC, Zaslona Z, Przybranowski S, White ES and Peters-Golden M: Histone modifications are responsible for decreased Fas expression and apoptosis resistance in fibrotic lung fibroblasts. Cell Death Dis. 4:e6212013. View Article : Google Scholar : PubMed/NCBI

108 

Lee YH, Seo D, Choi KJ, Andersen JB, Won MA, Kitade M, Gómez-Quiroz LE, Judge AD, Marquardt JU, Raggi C, et al: Antitumor effects in hepatocarcinoma of isoform-selective inhibition of HDAC2. Cancer Res. 74:4752–4761. 2014. View Article : Google Scholar : PubMed/NCBI

109 

Li X, Wu XQ, Xu T, Li XF, Yang Y, Li WX, Huang C, Meng XM and Li J: Role of histone deacetylases(HDACs) in progression and reversal of liver fibrosis. Toxicol Appl Pharmacol. 306:58–68. 2016. View Article : Google Scholar : PubMed/NCBI

110 

Dooley S, Hamzavi J, Breitkopf K, Wiercinska E, Said HM, Lorenzen J, Ten Dijke P and Gressner AM: Smad7 prevents activation of hepatic stellate cells and liver fibrosis in rats. Gastroenterology. 125:178–191. 2003. View Article : Google Scholar : PubMed/NCBI

111 

Dooley S, Hamzavi J, Ciuclan L, Godoy P, Ilkavets I, Ehnert S, Ueberham E, Gebhardt R, Kanzler S, Geier A, et al: Hepatocyte-specific Smad7 expression attenuates TGF-beta-mediated fibrogenesis and protects against liver damage. Gastroenterology. 135:642–659. 2008. View Article : Google Scholar : PubMed/NCBI

112 

Hamzavi J, Ehnert S, Godoy P, Ciuclan L, Weng H, Mertens PR, Heuchel R and Dooley S: Disruption of the Smad7 gene enhances CCI4-dependent liver damage and fibrogenesis in mice. J Cell Mol Med. 12(5B): 2130–2144. 2008. View Article : Google Scholar : PubMed/NCBI

113 

Oseini AM and Sanyal AJ: Therapies in non-alcoholic steatohepatitis (NASH). Liver Int. 37(Suppl 1): S97–S103. 2017. View Article : Google Scholar

114 

Utsunomiya H, Yamamoto Y, Takeshita E, Tokumoto Y, Tada F, Miyake T, Hirooka M, Abe M, Kumagi T, Matsuura B, et al: Upregulated absorption of dietary palmitic acids with changes in intestinal transporters in non-alcoholic steatohepatitis (NASH). J Gastroenterol. 52:940–954. 2017. View Article : Google Scholar : PubMed/NCBI

115 

Fukushima J, Kamada Y, Matsumoto H, Yoshida Y, Ezaki H, Takemura T, Saji Y, Igura T, Tsutsui S, Kihara S, et al: Adiponectin prevents progression of steatohepatitis in mice by regulating oxidative stress and Kupffer cell phenotype polarization. Hepatol Res. 39:724–738. 2009. View Article : Google Scholar : PubMed/NCBI

116 

Afrin R, Arumugam S, Rahman A, Wahed MI, Karuppagounder V, Harima M, Suzuki H, Miyashita S, Suzuki K, Yoneyama H, et al: Curcumin ameliorates liver damage and progression of NASH in NASH-HCC mouse model possibly by modulating HMGB1-NF-κB translocation. Int Immunopharmacol. 44:174–182. 2017. View Article : Google Scholar : PubMed/NCBI

117 

Zhong S, Zhao L, Wang Y, Zhang C, Liu J, Wang P, Zhou W, Yang P, Varghese Z, Moorhead JF, et al: CD36 deficiency aggravates macrophage infiltration and hepatic inflammation by up-regulating MCP–1 expression of hepatocytes through HDAC2-dependant pathway. Antioxid Redox Signal. Aug 1–2017.Epub ahead of print. View Article : Google Scholar

118 

Torre LA, Bray F, Siegel RL, Ferlay J, Lortet-Tieulent J and Jemal A: Global cancer statistics, 2012. CA Cancer J Clin. 65:87–108. 2015. View Article : Google Scholar : PubMed/NCBI

119 

Chen Z, Xie H, Hu M, Huang T, Hu Y, Sang N and Zhao Y: Recent progress in treatment of hepatocellular carcinoma. Am J Cancer Res. 10:2993–3036. 2020.PubMed/NCBI

120 

Lee JS, Chu IS, Heo J, Calvisi DF, Sun Z, Roskams T, Durnez A, Demetris AJ and Thorgeirsson SS: Classification and prediction of survival in hepatocellular carcinoma by gene expression profiling. Hepatology. 40:667–676. 2004. View Article : Google Scholar : PubMed/NCBI

121 

Ropero S and Esteller M: The role of histone deacetylases (HDACs) in human cancer. Mol Oncol. 1:19–25. 2007. View Article : Google Scholar : PubMed/NCBI

122 

Bayat S, Mansoori Derakhshan S, Mansoori Derakhshan N, Shekari Khaniani M and Alivand MR: Downregulation of HDAC2 and HDAC3 via oleuropein as a potent prevention and therapeutic agent in MCF–7 breast cancer cells. J Cell Biochem. 120:9172–9180. 2019. View Article : Google Scholar : PubMed/NCBI

123 

Quint K, Agaimy A, Di Fazio P, Montalbano R, Steindorf C, Jung R, Hellerbrand C, Hartmann A, Sitter H, Neureiter D and Ocker M: Clinical significance of histone deacetylases 1, 2, 3, and 7: HDAC2 is an independent predictor of survival in HCC. Virchows Arch. 459:129–139. 2011. View Article : Google Scholar : PubMed/NCBI

124 

Kim HS, Chang YG, Bae HJ, Eun JW, Shen Q, Park SJ, Shin WC, Lee EK, Park S, Ahn YM, et al: Oncogenic potential of CK2α and its regulatory role in EGF-induced HDAC2 expression in human liver cancer. FEBS J. 281:851–861. 2014. View Article : Google Scholar : PubMed/NCBI

125 

Noh JH, Bae HJ, Eun JW, Shen Q, Park SJ, Kim HS, Nam B, Shin WC, Lee EK, Lee K, et al: HDAC2 provides a critical support to malignant progression of hepatocellular carcinoma through feedback control of mTORC1 and AKT. Cancer Res. 74:1728–1738. 2014. View Article : Google Scholar : PubMed/NCBI

126 

Huang J, Yang G, Huang Y, Kong W and Zhang S: 1,25(OH)2D3 inhibits the progression of hepatocellular carcinoma via down-regulating HDAC2 and upregulating P21(WAFI/CIP1). Mol Med Re. 13:1373–1380. 2016. View Article : Google Scholar

127 

Huang J, Yang G, Huang Y and Zhang S: Inhibitory effects of 1,25(OH)2D3 on the proliferation of hepatocellular carcinoma cells through the downregulation of HDAC2. Oncol Rep. 38:1845–1850. 2017. View Article : Google Scholar : PubMed/NCBI

128 

Wang H, Kohashi K, Yoshizumi T, Okumura Y, Tanaka Y, Shimokawa M, Iwasaki T, Aishima S, Maehara Y and Oda Y: Coexpression of SALL4 with HDAC1 and/or HDAC2 is associated with underexpression of PTEN and poor prognosis in patients with hepatocellular carcinoma. Hum Pathol. 64:69–75. 2017. View Article : Google Scholar : PubMed/NCBI

129 

Gryder BE, Pomella S, Sayers C, Wu XS, Song Y, Chiarella AM, Bagchi S, Chou HC, Sinniah RS, Walton A, et al: Histone hyperacetylation disrupts core gene regulatory architecture in rhabdomyosarcoma. Nat Genet. 51:1714–1722. 2019. View Article : Google Scholar : PubMed/NCBI

130 

Methot JL, Hamblett CL, Mampreian DM, Jung J, Harsch A, Szewczak AA, Dahlberg WK, Middleton RE, Hughes B, Fleming JC, et al: SAR profiles of spirocyclic nicotinamide derived selective HDAC1/HDAC2 inhibitors (SHI–1:2). Bioorg Med Chem Lett. 18:6104–6109. 2008. View Article : Google Scholar : PubMed/NCBI

131 

Methot JL, Chakravarty PK, Chenard M, Close J, Cruz JC, Dahlberg WK, Fleming J, Hamblett CL, Hamill JE, Harrington P, et al: Exploration of the internal cavity of histone deacetylase (HDAC) with selective HDAC1/HDAC2 inhibitors (SHI-1:2). Bioorg Med Chem Lett. 18:973–978. 2008. View Article : Google Scholar : PubMed/NCBI

132 

Qi Z, Wang C, Jiang J and Wu C: Novel C15 Triene Triazole, D-A derivatives anti-HepG2, and as HDAC2 inhibitors: A synergy study. Int J Mol Sci. 19:31842018. View Article : Google Scholar :

133 

Venturelli S, Niessner H, Sinnberg T, Berger A, Burkard M, Urmann C, Donaubauer K, Böcker A, Leischner C, Riepl H, et al: 6– and 8-Prenylnaringenin, novel natural histone deacetylase inhibitors found in hops, exert antitumor activity on melanoma cells. Cell Physiol Biochem. 51:543–556. 2018. View Article : Google Scholar

134 

Al-Sanea MM, Gotina L, Mohamed MF, Grace Thomas Parambi D, Gomaa HA, Mathew B, Youssif BG, Alharbi KS, Elsayed ZM, Abdelgawad MA and Eldehna WM: Design, synthesis and biological evaluation of new HDAC1 and HDAC2 inhibitors endowed with ligustrazine as a novel cap moiety. Drug Des Devel Ther. 14:497–508. 2020. View Article : Google Scholar : PubMed/NCBI

135 

Jang YG, Hwang KA and Choi KC: Rosmarinic acid, a component of rosemary tea, induced the cell cycle arrest and apoptosis through modulation of HDAC2 expression in prostate cancer cell lines. Nutrients. 10:17842018. View Article : Google Scholar :

136 

Deng L, Tang J, Yang H, Cheng C, Lu S, Jiang R and Sun B: MTA1 modulated by miR–30e contributes to epithelial-to-mesenchymal transition in hepatocellular carcinoma through an ErbB2-dependent pathway. Oncogene. 36:3976–3985. 2017. View Article : Google Scholar : PubMed/NCBI

137 

Buurman R, Gürlevik E, Schäffer V, Eilers M, Sandbothe M, Kreipe H, Wilkens L, Schlegelberger B, Kühnel F and Skawran B: Histone deacetylases activate hepatocyte growth factor signaling by repressing MicroRNA–449 in hepatocellular carcinoma cells. Gastroenterology. 143:811–820.e15. 2012. View Article : Google Scholar

138 

He QL, Qin SY, Tao L, Ning HJ and Jiang HX: Prognostic value and prospective molecular mechanism of miR–100-5p in hepatocellular carcinoma: A comprehensive study based on 1,258 samples. Oncol Lett. 18:6126–6142. 2019.PubMed/NCBI

139 

Kim HS, Lee KS, Bae HJ, Eun JW, Shen Q, Park SJ, Shin WC, Yang HD, Park M, Park WS, et al: MicroRNA–31 functions as a tumor suppressor by regulating cell cycle and epithelial-mesenchymal transition regulatory proteins in liver cancer. Oncotarget. 6:8089–8102. 2015. View Article : Google Scholar : PubMed/NCBI

140 

Dai W, Dai JL, Tang MH, Ye MS and Fang S: lncRNA-SNHG15 accelerates the development of hepatocellular carcinoma by targeting miR–490-3p/histone deacetylase 2 axis. World J Gastroenterol. 25:5789–5799. 2019. View Article : Google Scholar : PubMed/NCBI

141 

Turner BM: Cellular memory and the histone code. Cell. 111:285–291. 2002. View Article : Google Scholar : PubMed/NCBI

142 

Khan SN and Khan AU: Role of histone acetylation in cell physiology and diseases: An update. Clin Chim Acta. 411:1401–1411. 2010. View Article : Google Scholar : PubMed/NCBI

143 

Budillon A, Di Gennaro E, Bruzzese F, Rocco M, Manzo G and Caraglia M: Histone deacetylase inhibitors: A new wave of molecular targeted anticancer agents. Recent Pat Anticancer Drug Discov. 2:119–134. 2007. View Article : Google Scholar

144 

Wade PA: Transcriptional control at regulatory checkpoints by histone deacetylases: Molecular connections between cancer and chromatin. Hum Mol Genet. 10:693–698. 2001. View Article : Google Scholar : PubMed/NCBI

145 

Forsberg EC and Bresnick EH: Histone acetylation beyond promoters: Long-range acetylation patterns in the chromatin world. Bioessays. 23:820–830. 2001. View Article : Google Scholar : PubMed/NCBI

146 

Zhang L, Qiu Z, Hu Y, Yang F, Yan S, Zhao L, Li B, He S, Huang M, Li J and Li L: ABA treatment of germinating maize seeds induces VP1 gene expression and selective promoter-associated histone acetylation. Physiol Plant. 143:287–296. 2011. View Article : Google Scholar : PubMed/NCBI

147 

Tian XL, Lu X, Feng JB, Cai TJ, Li S, Tian M and Liu QJ: Alterations in histone acetylation following exposure to 60Co ү-rays and their relationship with chromosome damage in human lymphoblastoid cells. Radiat Environ Biophys. 57:215–222. 2018. View Article : Google Scholar : PubMed/NCBI

148 

Heinz KS, Rapp A, Casas-Delucchi CS, Lehmkuhl A, Romero-Fernández I, Sánchez A, Krämer OH, Marchal JA and Cardoso MC: DNA replication dynamics of vole genome and its epigenetic regulation. Epigenetics Chromatin. 12:182019. View Article : Google Scholar : PubMed/NCBI

149 

Ibi D, de la Fuente Revenga M, Kezunovic N, Muguruza C, Saunders JM, Gaitonde SA, Moreno JL, Ijaz MK, Santosh V, Kozlenkov A, et al: Antipsychotic-induced Hdac2 transcription via NF-kB leads to synaptic and cognitive side effects. Nat Neurosci. 20:1247–1259. 2017. View Article : Google Scholar : PubMed/NCBI

150 

Bhandari DR, Seo KW, Jung JW, Kim HS, Yang SR and Kang KS: The regulatory role of c-MYC on HDAC2 and PcG expression in human multipotent stem cells. J Cell Mol Med. 15:1603–1614. 2011. View Article : Google Scholar

151 

Yang H, Salz T, Zajac-Kaye M, Liao D, Huang S and Qiu Y: Overexpression of histone deacetylases in cancer cells is controlled by interplay of transcription factors and epigenetic modulators. FASEB J. 28:4265–4279. 2014. View Article : Google Scholar : PubMed/NCBI

Related Articles

Journal Cover

July-2021
Volume 48 Issue 1

Print ISSN: 1107-3756
Online ISSN:1791-244X

Sign up for eToc alerts

Recommend to Library

Copy and paste a formatted citation
x
Spandidos Publications style
Liu Y, Wang J, Huang Z, Chen R, Cao X, Zhu D, Yu H, Wang X, Zhou H, Xia Q, Xia Q, et al: Histone deacetylase‑2: A potential regulator and therapeutic target in liver disease (Review). Int J Mol Med 48: 131, 2021
APA
Liu, Y., Wang, J., Huang, Z., Chen, R., Cao, X., Zhu, D. ... Li, J. (2021). Histone deacetylase‑2: A potential regulator and therapeutic target in liver disease (Review). International Journal of Molecular Medicine, 48, 131. https://doi.org/10.3892/ijmm.2021.4964
MLA
Liu, Y., Wang, J., Huang, Z., Chen, R., Cao, X., Zhu, D., Yu, H., Wang, X., Zhou, H., Xia, Q., Li, J."Histone deacetylase‑2: A potential regulator and therapeutic target in liver disease (Review)". International Journal of Molecular Medicine 48.1 (2021): 131.
Chicago
Liu, Y., Wang, J., Huang, Z., Chen, R., Cao, X., Zhu, D., Yu, H., Wang, X., Zhou, H., Xia, Q., Li, J."Histone deacetylase‑2: A potential regulator and therapeutic target in liver disease (Review)". International Journal of Molecular Medicine 48, no. 1 (2021): 131. https://doi.org/10.3892/ijmm.2021.4964