Open Access

Epigenetics in inflammatory liver diseases: A clinical perspective (Review)

  • Authors:
    • Teodora Isac
    • Sebastian Isac
    • Razvan Rababoc
    • Mihail Cotorogea
    • Laura Iliescu
  • View Affiliations

  • Published online on: April 4, 2022     https://doi.org/10.3892/etm.2022.11293
  • Article Number: 366
  • Copyright: © Isac 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

Inflammatory liver diseases are, nowadays, multifactorial and wide‑spread, thus having an important socio‑economic impact. Although the therapeutic algorithms are well‑known in hepatitis, regardless of etiology, strategies to identify inflammatory hepatic lesions in early stages and to develop new epigenetic therapies should be prioritized. The main entities of inflammatory liver disease are: alcoholic and non‑alcoholic fatty liver disease, autoimmune hepatitis, viral hepatitis and Wilson disease. The main epigenetic processes include: DNA methylation/demethylation, which imply changes in DNA tertiary structure; post‑translational histone covalent changes (methylation/demethylation, acetylation/deacetylation, ubiquitination), that cause DNA‑histone instability; synthesis of small, non‑coding RNA molecules, called microRNAs, that modulate translational potential of transcripts (mRNAs) and post‑translational modification of polypeptide chains. Consequently, the epigenetic interactions aforementioned, play an important modulatory role in disease progression and response to conventional therapies The present review focused on the main epigenetic changes in inflammatory liver conditions, considering a new perspective: Epigenetic therapy. This approach is more than welcomed, taking into consideration that conventional therapeutic strategies are almost exhausted.

1. Introduction

DNA represents the main molecule that incorporates the genetic information of the human cells independent of the organ tissue. It is well known that this molecule, consisting of a specific nucleotide sequence, forms the nuclear chromatin in a cell. The functional unit of chromatin is called a nucleosome, whose structure consists of a DNA double helix and adjacent histone proteins. Furthermore, all individual variations of genes form the genotype, which is the same in all cell types within an organism. The interactions between the genotype and the environment form variable phenotypes, depending on the cell type (1).

As a result, different gene suppression and activation mechanisms determine consistent phenotypic differences between cells from different organs and even within the same organ.

Subsequently, modifying the transcriptional potential of DNA without changing its sequence or genetic information, will change the chromatin tertiary structure, due to histone-DNA interaction. However, this process will not change the amino-acid sequence in the polypeptide chain (this represents the unchanged genetic information). Remodeling of the chromatin conformation, epigenetically, at the nuclear level, results in the synthesis of the mRNA species or in suppressing this transcriptional process (1). Furthermore, in cytoplasm, the mRNA transcripts are controlled by other epigenetic factors: The microRNA (miRNA or miR) species, the transcripts of lesser dimensions (2-22 nucleotides), of the noncoding repetitive DNA. The final product, the polypeptide chain, may therefore be translated or not, depending on the impact of such a post-transcriptional interference RNA network. Finally, there are certain active proteins such as enzymes, for example, which may be formed only following the post-translational covalent modifications of the polypeptide chains (2).

All these processes are examples of natural, physiological, epigenetic regulations of gene function or expression. Epigenetics studies the variation of gene expression that is independent of genetic information or nucleotide sequences. It refers to gene function control through both nuclear chromatin covalent modification and remodeling and cytoplasmic activity of the interference RNA involving miRNAs along with post-translational covalent modification of the newly synthetized polypeptides (1).

Various epigenetic mechanisms could explain how a static genome interacts with a dynamic environment. According to Cold Spring Harbor Meeting-2009, ‘an epigenetic trait’ represents a term designed to define ‘a stably heritable phenotype resulting from changes in a chromosome without alterations in the DNA sequence’ (3). Over time, different definitions of epigenetics were suggested. Waddington was the first to introduce the term epigenetics in 1942, as a variety of mechanisms which promotes gene expression changes without DNA mutation (4). In 1968, he defined epigenesis as ‘the branch of biology which studies the causal interactions between genes and their products which bring the phenotype into being’ (5).

The current view of epigenetics includes the following processes: i) Active DNA methylation, demethylation, which imply changes in DNA tertiary structure; ii) also post-translational histone covalent changes (methylation/demethylation, acetylation/deacetylation, ubiquitination), that cause broader and complex fluctuations in DNA-histone interactions; iii) synthesis of small, non-coding RNA molecules, called miRNAs, that modulate translational potential of transcripts (mRNAs) at the ribosomal level; and iv) post-translational modification of polypeptide chains (1).

DNA methylation involves the addition of a methyl group to a cytosine major base through a covalent interaction, particularly at 5'-CpG-3' dinucleotide sites in DNA substrates. If this change is symmetrical in both DNA chains, then the DNA conformational structure will change and the replication will be delayed (6). Usually there are specific gene areas rich in 5'-CpG-3' nucleotides, which can be preferentially methylated. These areas are called ‘CpG islands’, in the case of the housekeeping genes (which have to stay active independent of the cell types and the CpG sites should not be methylated) (7). The extensive methylation of the CpG repetitive regions is linked to chromatin silencing in genes presenting tissue type expression. Over 60% of the genes in the human genome have a high proportion of CpG dinucleotides in the promoter region, thus being potentially influenced by this epigenetic mechanism (8).

DNA methylation is controlled by enzymes that transfer methyl groups from the methyl-donor, S-adenosil-methionine (SAM), to cytosine. They are called DNA N-methyl transferases (DNMT) (Fig. 1).

This epigenetic process is carried out by different isoforms of DNMT: DNMT 1 has a specific role in maintaining the pre-existent methylation pattern, while DNMT 3a and DNMT 3b promote de novo DNA methylation (6,9). In addition, environmental factors such as nutrition, exercise and particular chemical substances are able to modify DNMT expression and function with consecutive changes in DNA methylation degree and distribution, and all of these have a variable transcriptional effect upon the gene function (10).

Another epigenetic nuclear mechanism involves histone-DNA interaction and is represented by covalent binding of acetyl groups at lysine residues within histones forming the nucleosome core. Consequently, histone chains around which DNA molecules are wrapped, become more relaxed, easier exposing DNA to transcriptional factors. Conversely, if acetyl groups are removed from the acetylated histones, the nucleosomes will appear more compact and resistant to transcriptional factors (1) (Fig. 2).

The acetylation process is regulated by histone acetyltransferases (HATs), while histone deacetylation is regulated by histone deacetylases (HDACs) (11). At present, histone deacetylation is extensively studied (12). HDACs are classified according to their structural and functional similarities into four classes: class I (HDACs: 1, 2, 3, 8), class IIa (HDACs: 4, 5, 7, 9), class IIb (HDACs: 6, 10), class III (sirtuins 1-7), and class IV (HDAC11). It is well known that different classes of HDACs have specific intracellular locations (HDACs from class I are predominantly located intranuclearly, while class II HDACs shuttle between the cytoplasm and the nucleus) (13).

HATs are represented by a vast family of proteins such as cAMP-response element binding (protein) (CREB) binding protein (CBP), that acts in a phosphorylation dependent manner: Once phosphorylated, a HAT molecule will be activated, while dephosphorylating will lead to HAT inactivation. The equilibrium between HATs and HDACs is termed ‘acetylation homeostasis’ and will finally dictate the degree of DNA exposure to transcriptional factors inside the nucleosome (14). Histone acetylation and deacetylation regulate cellular processes such as aging and oncogenesis.

Conversely, miRNAs are small non-coding single-stranded RNA species, composed of 15-30 bases, that are involved in the post-transcriptional control of gene function or expression in the cytoplasm (15). Its effect of silencing is achieved by altering mRNA stability and blocking the mRNA elongation, thus terminating protein synthesis (Fig. 3).

Briefly, following maturation of pre-miRNA (formed in the nucleus and exported and processed in the cytoplasm) into miRNA by enzymatic cleavage of the hairpin structures, the leading strands of the miRNA are integrated into the protein RNA-induced silencing complex (RISC). The leading strand is thermodynamically unstable relative to the passenger strand and directs the RISC to the complementary strand of mRNA. miRNAs usually recognize binding sites in the 3' untranslated region (UTR) of mRNA transcripts. Perfect base complementarity between miRNA and mRNA cleaves the mRNA by the slicing activity of Argonaute-2 (AGO2), whereas imperfect binding leads to translational suppression and slicer-independent mRNA damage (16).

Since miRNA represents epigenetic markers that modulate terminal cell differentiation and developmental changes, there has been great interest in achieving new miRNA-targeted epigenetic therapies and identifying new epigenetic markers for various clinical paradigms (17). At present, there are numerous miRNA species recognized, being independently and organ-specifically coded which can modulate the protein synthesis and function in health and disease.

Although the screening tests and therapeutic algorithms are well-established in various hepatic diseases, including liver transplantation as etiologic therapy in end-stage liver disease, strategies to identify reversible lesions in early stages and to develop new epigenetic therapies should be prioritized (18-20).

2. Methods: Selection of studies

A systematic review was performed on articles in English published in databases including PubMed, Elsevier or Scopus until August 2021, using the following key word: ‘epigenetics in liver diseases’. Articles referring to any of the following hepatic diseases were included: Non-alcoholic fatty liver disease (NAFLD), alcoholic fatty liver disease (AFLD), autoimmune hepatitis (AIH), viral hepatitis (VH) and Wilson disease (WD). Studies focusing on end-stage liver disease were excluded. The most relevant articles published in the last 15 years were added.

Some relevant articles from the domain of epigenetics, in general, were further included, independent of the year of publication, for a better description of fundamental epigenetic mechanisms. After duplication removal and screening for eligibility, 65 articles consisting of 19 clinical studies, 22 experimental studies and 24 review articles were included.

3. Epigenetics in various liver diseases

Epigenetics in NAFLD

NAFLD represents a serious condition linked with an inappropriate high-fat diet, which is more obvious in developed countries where obesity and unhealthy dietary habits represent public health issues. These liver changes are in correlation with other multisystemic changes such as type II diabetes, various cardiovascular or renal diseases, as consequences of the interaction between the environment (due to various dietary habits) and the organism.

In NAFLD, epigenetic DNA changes have been observed, altering the insulin metabolism or producing dysregulations, at the cellular level of the various metabolic pathways. Ahrens et al observed that genes encoding insulin-like growth-factor 1 (IGF-1) and insulin-like factor binding protein 2 (IGFBP-2) could be hypermethylated in NAFLD, inducing gene silencing and consequent impairments in glucose metabolism. Other genes including pyruvate carboxylases and ATP citrate lyase involved in the glucose cycle could also be epigenetically silenced (21).

Histone deacetylation was also observed to interfere with lipid metabolism. Silent information regulator factor 2-related enzyme 1 (SIRT-1) improves hepatic steatosis and circadian rhythm (22,23). Other histone deacetylases such as HDAC-3 and HDAC-8 promote triglyceride metabolism and insulin sensitivity (24). De novo liver lipogenesis could also be epigenetically controlled due to certain histone changes: The interaction between host cell factor 1 (HCF-1) and carbohydrate response element binding protein (ChREBP) regulates hepatic lipogenic genes (25).

In NAFLD, miRNAs species have been identified as epigenetic markers for liver injury. miR-122 is one of the most abundant small non-coding RNAs expressed in the liver. In NAFLD, miR-122 is downregulated. Experimental study models have revealed that the downregulation of this marker promotes lipogenesis and liver inflammation (26). Conversely, upregulation of miR-21 has deleterious effects on the degree of hepatic steatosis and glucose metabolism (27). Other overexpressed liver miRNAs in NAFLD are miR-24, miR-34a and miR-124, which could interfere with lipid metabolism and insulin sensitivity (26,28). miR-155 modulates the crosstalk between adipose tissue and the liver in NAFLD induced by a high-fat diet (29).

In conclusion, NAFLD is a highly epigenetically controlled liver pathology, whose complex mechanisms involve factors associated with dietary habits, alterations in tertiary DNA and histone structure and specific miRNA expression.

Epigenetics in (AIH)

AIH exhibits various phenotypes, reflecting the complexity of underlying immune mechanisms. There are two forms: AIH type I, present particularly in middle-aged women and AIH type 2, which is more common in children. AIH type 1 is characterized by an increased titer of antinuclear antibodies, soluble liver antigen/liver pancreas antibodies, and smooth muscle antibodies, while AIH type 2 exhibits large amounts of the liver kidney microsomal 1 antibodies (30). As the main immune mechanism of the disease, the imbalance between pro and anti-inflammatory promoting T-cells is extensively studied. Treg cells are T-cells with anti-inflammatory properties, whose presence is linked with the disease activity and liver inflammation (31).

The data in the literature is rather focused on the miRNA-mediated epigenetic changes in AIH than on DNA or histone changes. Similar to other inflammatory hepatic diseases, miR-122, the most abundant liver miRNA species, is upregulated in AIH as well, serving as a marker of the disease activity (32). Consequently, this epigenetic marker could serve as biomarker for therapy response or disease control. miR-21 is also upregulated and is inversely correlated with the degree of fibrosis (33). miR-223 suppresses proinflammatory liver activity via the NF-κB pathway, inhibiting the macrophage function. In an AIH experimental model, the overexpression of miR-223 was revealed to have a liver-protective effect (34). miR-155 could affect AIH progression as well. The literature offers, however, contradictory data. miR-155 regulates the inflammatory response by influencing the Th17 cells, with no effect on IL-10-mediated Treg response (35).

Other epigenetic markers with an undefined role in AIH are: miR-218, miR-363, miR-518f, miR-628-5p, miR-888, miR-523, miR-141, miR-302b, miR-643 and miR-573(36).

Although there is insufficient data to characterize epigenetic histone changes in AIH, certain studies have revealed anti-histone auto-antibodies in AIH, whose interaction with the histones could be reduced due to the epigenetic aforementioned potential structural change (37). This paradigm should be, however, further explored.

A previous clinical study reported a correlation between the DNA methylation status in certain immune cells such as CD4+ and CD19+ lymphocytes and disease activity. Altered expression of enzymes involved in DNA methylation, TET1 and DNMT3A, characterizes lymphocytes in AIH (38). Further studies are required in order to confirm this association.

From an epigenetic perspective, AIH still represents an open field for research: To date, miRNAs represent the only area which have offered a perspective regarding epigenetic modulating mechanisms in this disease.

Epigenetics in VH

The interplay between a hepatic virus and the liver leaves, in the majority of cases, an epigenetic signature. Due to complex modulatory mechanisms, these signatures act either as a prognostic tool or as a therapeutic response (39). Subsequently, epigenetic therapies are evoked as novel therapies against these viruses. For example DNMT inhibitors could be useful in VHC-associated HCC, while HDAC inhibitors could reduce the VHB replication (39).

Hepatitis B virus (HBV) is a DNA virus, whose genetic structure consists of covalently closed circular (ccc)DNA incorporated into hepatocytes. The translational process is realized using the host nuclear enzymes. The cccDNA methylation in the region of GpG islands could, however, reduce the translational potential of the viral DNA. There are three CpG regions defined in the viral DNA: i) The start site of the S gene; ii) the region surrounding enhancer I, the HBx gene promoter (Xp), and the core promoter (Cp); and iii) the region harboring the Sp1 promoter and the start codon of the Pol gene (40). According to Zhang et al, the start site of the S gene is variably methylated among the different HBV genotypes, while the other two regions are more stably methylated (41). The methylation of the second island is linked with a decreased viremia, while the methylation of the third island influences carcinogenesis (41). This epigenetic change is observed mainly in the nuclear cccDNA, integrated in the host cells and not in the histone-free cytoplasmic DNA or circulating virions. The role of DNMTs in HBV infection is not fully understood. According to the literature, DNMT1, DNMT2 and DNMT3, are upregulated in HBV infection leading to hypermethylation in host cells and, consequently, to a reduction in virus replication (42). cccDNA replication is also modulated by epigenetic changes of the histones. Hypomethylation of H3 and H4 histones and the recruitment of HDAC 1 nearby cccDNA, could reduce the replication potential of the virus (43). Furthermore, epigenetic therapies targeting upregulation of DNMTs and histone hypomethylation linked with immunomodulatory therapy could represent the future in chronic HBV infection.

Considering the miRNA species as potential epigenetic biomarkers in HBV, miR-146 predicts the evolution to fibrosis in HBV-infected patients (44).

Conversely, hepatitis C virus (HCV) virus is an RNA virus, whose particular epigenetic signature reveals the risk of carcinogenesis even in the presence of the sustained viral response. Specific histone changes, H3K4Me3 and H3K9Ac, promote the persistence of the virus following successful direct antiviral therapy, acting as an epigenetic signature (45).

DNA methylation is another epigenetic change, influencing carcinogenesis in HCV-infected patients. The two most common repetitive elements in humans, long interspersed nuclear element-1 (LINE-1) and Alu element (Alu), have been linked with carcinogenesis. HCV may cause hepatocellular carcinoma (HCC) by suppressing host defenses through DNA methylation that controls the mobilization of repetitive elements (46).

Certain miRNA species are used as prognostic tools, being particularly associated with the risk for HCV patients to develop various complications such as HCC, fibrosis or cirrhosis. miR-122-5p, miR-486-5p and miRNA-142-3p could predict the development of HCC in HCV-infected patients (47).

According to Shrivastava et al, miR-20a and miR-92a are epigenetic biomarkers, which promote the evolution from an acute to a chronic state (48). Let-7c is another epigenetic biomarker, which could predict the evolution to fibrosis (49).

miR-494 is associated to the therapeutic response, while miR-34a is upregulated in fatty liver compared with chronic HCV (50,51).

Epigenetics in alcoholic fatty liver disease (ALFD)

Diet-induced epigenetic changes are common and one of the first described modulatory factors which could lead to epigenetic changes according to the target tissue. The most exposed tissues developing epigenetic modulatory mechanisms secondary to various dietary factors are the brain, hematopoietic system or liver (52,53). Alcoholic liver disease (ALD) has a significant influence on the life quality, having a very important impact on various health systems.

Alcohol-induced oxidative stress in hepatocytes interferes with all main mechanisms of chromosomal epigenetic control including DNA hypomethylation, histone acetylation and phosphorylation and miRNA alteration.

Histone H3 acetylation at Lys 9 (H3AcK9) in alcohol-exposed hepatocytes was observed in experimental studies (54,55). HDAC inhibition, particularly SIRT1, and HAT enzyme activation are responsible for this change (54). Phosphorylation of the histone H3 at serine residues could synergistically influence the epigenetic signature in hepatocytes exposed to alcohol (55). Another experimental study revealed a differential methylation pattern of the histone H3 and H4 in alcohol-exposed hepatocytes (56). Whether chronic or acute exposure to alcohol could induce these histone alterations is, however, controversial.

S-adenosil-methionine (SAM) is an important methyl donor involved in CpG island DNA methylation. According to Dannenberg et al, SAM concentration is reduced in ALD, which consequently interferes with the DNA methylation process (57). This hypomethylation state promotes DNA damage and strand breaks in ALD (57). DNA hypomethylation also disturbs alcohol-metabolizing enzymes, such as alcohol dehydrogenase 1B (ADH1) (58). An interplay between histone acetylation and DNA methylation is also possible in ALD: DNA hypomethylation in promoter regions triggers histone deacetylation. The exact effect of this epigenetic crosstalk is, however, unknown.

miRNA species represent other epigenetic markers in ALD. miR-155 is increased in liver macrophages secondary to alcohol intake (59). miR-212 is upregulated secondary to alcohol intake, affecting the intestinal permeability by downregulating tight junction protein zonula occludens 1(60). Further studies are required, in order to understand the exact prognostic and therapeutic potential of these epigenetic markers in ALD.

Epigenetics in WD

WD is an autosomal recessive disease characterized by accumulation of copper in the liver and brain. The key gene causing this disease is the copper-transporting gene ATP7B, which blocks copper extraction through the biliary tract. The knowledge of the disease physiopathology is continuously evolving. At present, it is widely accepted that alteration of this gene interferes with at least eight transmembrane active transporters of copper from the hepatocytes (61).

There is a large variability in WD considering the onset, sex, severity, response to treatment and target organ (the liver vs. the brain). This disease inconsistency is partially explained through certain epigenetic modulatory mechanisms. WD impairs the methionine metabolism, which is the main methyl supplier for DNA and histone methylation. The aberrant SAM/S-adenosil-homocysteine (SAH) ratio impairs the methylation process (62).

Furthermore, traces of heavy metals impair the mitochondrial metabolism, thus increasing the amount of reactive oxygen species (ROS). This metabolic change dysregulates the activity of TET enzymes, inhibiting the DNA demethylation (63).

In animal models, the hepatic accumulation of copper is inversely correlated with DNMT3a and DNMT3b levels, impairing the DNA methylation process (64). This alteration is particularly important in utero. Consequently, choline and penicillamine treatment, differentially modify the methylation status of the DNA in mice according to the sex (65).

4. Final considerations

Liver pathology is a vast field with incomplete knowledge, which requires profound expertise. The outcome in advanced and irreversible chronic hepatic injury in the absence of a liver transplant (cirrhosis and end-stage liver disease) is poor, with short and long-term socio-economic consequences. Therefore, new strategies to identify, stratify and treat these hepatic diseases while still being in a reversible state are highly required.

The epigenetic approach of all these diseases is more than welcomed, taking into consideration that conventional therapeutic strategies are almost exhausted. This is generally valid for ALD and NAFLD, viral hepatitis, AIH as well as other metabolic conditions.

Analysis of gene function and expression independent on the ‘heritable’ character of the genome, consisting in analysis of histone-DNA interactions and small non-coding RNA synthesis, is an extremely valuable tool for future diagnostic and therapeutic strategies of hepatic diseases, whose molecular etiologies are far from being completely elucidated.

Acknowledgements

The authors would like to thank Ms Irina Matache, researcher at the Department of Physiology II and Neurosciences of ‘Carol Davila’ University of Medicine and Pharmacy (Bucharest, Romania), for her help with manuscript reviewing.

Funding

Funding: No funding was received.

Availability of data and materials

Not applicable.

Authors' contributions

TI contributed to literature research, manuscript writing and reviewing. SI contributed to manuscript writing and reviewing. RR contributed to manuscript reviewing and design of figures. MC contributed to manuscript reviewing. LI coordinated and reviewed the manuscript. All authors read and approved the manuscript and agree to be accountable for all aspects of the research in ensuring that the accuracy or integrity of any part of the work are appropriately investigated and resolved. 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.

References

1 

Margueron R and Reinberg D: Chromatin structure and the inheritance of epigenetic information. Nat Rev Genet. 11:285–296. 2010.PubMed/NCBI View Article : Google Scholar

2 

Ryšlavá H, Doubnerová V, Kavan D and Vaněk O: Effect of posttranslational modifications on enzyme function and assembly. J Proteomics. 92:80–109. 2013.PubMed/NCBI View Article : Google Scholar

3 

Berger SL, Kouzarides T, Shiekhattar R and Shilatifard A: An operational definition of epigenetics. Genes Dev. 23:781–783. 2009.PubMed/NCBI View Article : Google Scholar

4 

Waddington CH: Towards a Theoretical Biology. Edinburgh University Press, Edinburgh, 1968.

5 

Jeltsch A: On the enzymatic properties of Dnmt1: Specificity, processivity, mechanism of linear diffusion and allosteric regulation of the enzyme. Epigenetics. 1:63–66. 2006.PubMed/NCBI View Article : Google Scholar

6 

Fischer A, Sananbenesi F, Mungenast A and Tsai LH: Targeting the correct HDAC(s) to treat cognitive disorders. Trends Pharmacol Sci. 31:605–617. 2010.PubMed/NCBI View Article : Google Scholar

7 

Zhu J, He F, Hu S and Yu J: On the nature of human housekeeping genes. Trends Genet. 24:481–484. 2008.PubMed/NCBI View Article : Google Scholar

8 

Bibikova M: Chapter 2. DNA Methylation. Microarrays. In: Epigenomics in Health and Disease. Fraga MF and Fernández AF (eds). Academic Press, Boston, MA, pp19-46, 2016.

9 

Goll MG and Bestor TH: Eukaryotic cytosine methyltransferases. Annu Rev Biochem. 74:481–514. 2005.PubMed/NCBI View Article : Google Scholar

10 

Keil KP and Lein PJ: DNA methylation: A mechanism linking environmental chemical exposures to risk of autism spectrum disorders? Environ Epigenet. 2(dvv012)2016.PubMed/NCBI View Article : Google Scholar

11 

Mielcarek M, Zielonka D, Carnemolla A, Marcinkowski JT and Guidez F: HDAC4 as a potential therapeutic target in neurodegenerative diseases: A summary of recent achievements. Front Cell Neurosci. 9(42)2015.PubMed/NCBI View Article : Google Scholar

12 

Saha RN and Pahan K: HATs and HDACs in neurodegeneration: A tale of disconcerted acetylation homeostasis. Cell Death Differ. 13:539–550. 2006.PubMed/NCBI View Article : Google Scholar

13 

Peserico A and Simone C: Physical and functional HAT/HDAC interplay regulates protein acetylation balance. J Biomed Biotechnol. 2011(371832)2011.PubMed/NCBI View Article : Google Scholar

14 

Varela MA, Roberts TC and Wood MJ: Epigenetics and ncRNAs in brain function and disease: Mechanisms and prospects for therapy. Neurotherapeutics. 10:621–631. 2013.PubMed/NCBI View Article : Google Scholar

15 

Kim VN: MicroRNA biogenesis: Coordinated cropping and dicing. Nat Rev Mol Cell Biol. 6:376–385. 2005.PubMed/NCBI View Article : Google Scholar

16 

Pasquinelli AE: MicroRNAs and their targets: Recognition, regulation and an emerging reciprocal relationship. Nat Rev Genet. 13:271–282. 2012.PubMed/NCBI View Article : Google Scholar

17 

Isac S, Panaitescu AM, Iesanu MI, Zeca V, Cucu N, Zagrean L, Peltecu G and Zagrean AM: Maternal citicoline-supplemented diet improves the response of the immature hippocampus to perinatal asphyxia in rats. Neonatology. 117:729–735. 2020.PubMed/NCBI View Article : Google Scholar

18 

Isac T, Isac S, Ioanitescu S, Mihaly E, Tanasescu MD, Balan DG, Tulin A and Iliescu L: Dynamics of serum α-fetoprotein in viral hepatitis C without hepatocellular carcinoma. Exp Ther Med. 22(749)2021.PubMed/NCBI View Article : Google Scholar

19 

Wu K, Ye C, Lin L, Chu Y, Ji M, Dai W, Zeng X and Lin Y: Inhibiting miR-21 attenuates experimental hepatic fibrosis by suppressing both the ERK1 pathway in HSC and hepatocyte EMT. Clin Sci (Lond). 130:1469–1480. 2016.PubMed/NCBI View Article : Google Scholar

20 

Popescu I, Ionescu M, Braşoveanu V, Hrehoreţ D, Copca N, Lupaşcu C, Botea F, Dorobanţu B, Alexandrescu S, Grigorie M, et al: The Romanian National program for liver transplantation -852 procedures in 815 patients over 17 years (2000-2017): A continuous evolution to success. Chirurgia (Bucur). 112:229–243. 2017.PubMed/NCBI View Article : Google Scholar

21 

Ahrens M, Ammerpohl O, von Schönfels W, Kolarova J, Bens S, Itzel T, Teufel A, Herrmann A, Brosch M, Hinrichsen H, et al: DNA methylation analysis in nonalcoholic fatty liver disease suggests distinct disease-specific and remodeling signatures after bariatric surgery. Cell Metab. 18:296–302. 2013.PubMed/NCBI View Article : Google Scholar

22 

Cao Y, 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.PubMed/NCBI View Article : Google Scholar

23 

Nakahata Y, Kaluzova M, Grimaldi B, Sahar S, Hirayama J, Chen D, Guarente LP and Sassone-Corsi P: The NAD+-dependent deacetylase SIRT1 modulates CLOCK-mediated chromatin remodeling and circadian control. Cell. 134:329–340. 2008.PubMed/NCBI View Article : Google Scholar

24 

Xu F and Guo W: The progress of epigenetics in the development and progression of non-alcoholic fatty liver disease. Liver Res. 4:118–123. 2020.

25 

Lane EA, Choi DW, Garcia-Haro L, Levine ZG, Tedoldi M, Walker S and Danial NN: HCF-1 regulates de novo lipogenesis through a nutrient-sensitive complex with ChREBP. Mol Cell 25:. 75:357–371.e7. 2019.PubMed/NCBI View Article : Google Scholar

26 

Hsu SH, Wang B, Kota J, Yu J, Costinean S, Kutay H, Yu L, Bai S, La Perle K, Chivukula RR, et al: Essential metabolic, anti-inflammatory, and anti-tumorigenic functions of miR-122 in liver. J Clin Invest. 122:2871–2883. 2012.PubMed/NCBI View Article : Google Scholar

27 

Calo N, Ramadori P, Sobolewski C, Romero Y, Maeder C, Fournier M, Rantakari P, Zhang FP, Poutanen M, Dufour JF, et al: Stress-activated miR-21/miR-21* in hepatocytes promotes lipid and glucose metabolic disorders associated with high-fat diet consumption. Gut. 65:1871–1881. 2016.PubMed/NCBI View Article : Google Scholar

28 

Liu X, Zhao J, Liu Q, Xiong X, Zhang Z, Jiao Y, Li X, Liu B, Li Y and Lu Y: MicroRNA-124 promotes hepatic triglyceride accumulation through targeting tribbles homolog 3. Sci Rep. 6(37170)2016.PubMed/NCBI View Article : Google Scholar

29 

Ying W, Riopel M, Bandyopadhyay G, Dong Y, Birmingham A, Seo JB, Ofrecio JM, Wollam J, Hernandez-Carretero A, Fu W, et al: Adipose tissue macrophage-derived exosomal miRNAs can modulate in vivo and in vitro insulin sensitivity. Cell. 171:372–384.e12. 2017.PubMed/NCBI View Article : Google Scholar

30 

Kerkar N and Vergani D: De novo autoimmune hepatitis -is this different in adults compared to children? J Autoimmun. 95:26–33. 2018.PubMed/NCBI View Article : Google Scholar

31 

Behairy BE, El-Araby HA, Abd El Kader HH, Ehsan NA, Salem ME, Zakaria HM and Khedr MA: Assessment of intrahepatic regulatory T cells in children with autoimmune hepatitis. Ann Hepatol. 15:682–690. 2016.PubMed/NCBI View Article : Google Scholar

32 

Bandiera S, Pfeffer S, Baumert TF and Zeisel MB: MiR-122-a key factor and therapeutic target in liver disease. J Hepatol. 62:448–457. 2015.PubMed/NCBI View Article : Google Scholar

33 

Migita K, Komori A, Kozuru H, Jiuchi Y, Nakamura M, Yasunami M, Furukawa H, Abiru S, Yamasaki K, Nagaoka S, et al: Circulating microRNA profiles in patients with type-1 autoimmune hepatitis. PLoS One. 10(e0136908)2015.PubMed/NCBI View Article : Google Scholar

34 

Chen L, Lu FB, Chen DZ, Wu JL, Hu ED, Xu LM, Zheng MH, Li H, Huang Y, Jin XY, et al: BMSCs-derived miR-223-containing exosomes contribute to liver protection in experimental autoimmune hepatitis. Mol Immunol. 93:38–46. 2018.PubMed/NCBI View Article : Google Scholar

35 

Xia G, Wu S, Wang X and Fu M: Inhibition of microRNA-155 attenuates concanavalin-A-induced autoimmune hepatitis by regulating Treg/Th17 cell differentiation. Can J Physiol Pharmacol. 96:1293–1300. 2018.PubMed/NCBI View Article : Google Scholar

36 

Liu Q, Li Y, Xiong M and Tang R: Epigenetics of autoimmune liver diseases: Current progress and future directions. J Bio-X Res. 2:46–55. 2019.

37 

Chen M, Shirai M, Czaja AJ, Kurokohchi K, Arichi T, Arima K, Kodama T and Nishioka M: Characterization of anti-histone antibodies in patients with type 1 autoimmune hepatitis. J Gastroenterol Hepatol. 13:483–489. 1998.PubMed/NCBI View Article : Google Scholar

38 

Zachou K, Arvaniti P, Lyberopoulou A, Sevdali E, Speletas M, Ioannou M, Koukoulis GK, Renaudineau Y and Dalekos GN: Altered DNA methylation pattern characterizes the peripheral immune cells of patients with autoimmune hepatitis. Liver Int: Feb 2, 2022 (Epub ahead of print).

39 

Nehme Z, Pasquereau S and Herbein G: Control of viral infections by epigenetic-targeted therapy. Clin Epigenetics. 11(55)2019.PubMed/NCBI View Article : Google Scholar

40 

Jain S, Chang TT, Chen S, Boldbaatar B, Clemens A, Lin SY, Yan R, Hu CT, Guo H, Block TM, et al: Comprehensive DNA methylation analysis of hepatitis B virus genome in infected liver tissues. Sci Rep. 5(10478)2015.PubMed/NCBI View Article : Google Scholar

41 

Zhang Y, Li C, Zhang Y, Zhu H, Kang Y, Liu H, Wang J, Qin Y, Mao R, Xie Y, et al: Comparative analysis of CpG islands among HBV genotypes. PLoS One. 8(e56711)2013.PubMed/NCBI View Article : Google Scholar

42 

Hong X, Kim ES and Guo H: Epigenetic regulation of hepatitis B virus covalently closed circular DNA: Implications for epigenetic therapy against chronic hepatitis B. Hepatology. 66:2066–2077. 2017.PubMed/NCBI View Article : Google Scholar

43 

Pollicino T, Belloni L, Raffa G, Pediconi N, Squadrito G, Raimondo G and Levrero M: Hepatitis B virus replication is regulated by the acetylation status of hepatitis B virus cccDNA-bound H3 and H4 histones. Gastroenterology. 130:823–837. 2006.PubMed/NCBI View Article : Google Scholar

44 

Wang TZ, Lin DD, Jin BX, Sun XY and Li N: Plasma microRNA: A novel non-invasive biomarker for HBV-associated liver fibrosis staging. Exp Ther Med. 17:1919–1929. 2019.PubMed/NCBI View Article : Google Scholar

45 

Perez S, Kaspi A, Domovitz T, Davidovich A, Lavi-Itzkovitz A, Meirson T, Alison Holmes J, Dai CY, Huang CF, Chung RT, et al: Hepatitis C virus leaves an epigenetic signature post cure of infection by direct-acting antivirals. PLoS Genet. 15(e1008181)2019.PubMed/NCBI View Article : Google Scholar

46 

Zheng Y, Hlady RA, Joyce BT, Robertson KD, He C, Nannini DR, Kibbe WA, Achenbach CJ, Murphy RL, Roberts LR and Hou L: DNA methylation of individual repetitive elements in hepatitis C virus infection-induced hepatocellular carcinoma. Clin Epigenetics. 11(145)2019.PubMed/NCBI View Article : Google Scholar

47 

Weis A, Marquart L, Calvopina DA, Genz B, Ramm GA and Skoien R: Serum MicroRNAs as biomarkers in hepatitis C: Preliminary evidence of a MicroRNA panel for the diagnosis of hepatocellular carcinoma. Int J Mol Sci. 20(864)2019.PubMed/NCBI View Article : Google Scholar

48 

Shrivastava S, Petrone J, Steele R, Lauer GM, Di Bisceglie AM and Ray RB: Up-regulation of circulating miR-20a is correlated with hepatitis C virus-mediated liver disease progression. Hepatology. 58:863–871. 2013.PubMed/NCBI View Article : Google Scholar

49 

Matsuura K, De Giorgi V, Schechterly C, Wang RY, Farci P, Tanaka Y and Alter HJ: Circulating let-7 levels in plasma and extracellular vesicles correlate with hepatic fibrosis progression in chronic hepatitis C. Hepatology. 64:732–745. 2016.PubMed/NCBI View Article : Google Scholar

50 

El-Diwany R, Wasilewski LN, Witwer KW, Bailey JR, Page K, Ray SC, Cox AL, Thomas DL and Balagopal A: Acute hepatitis C virus infection induces consistent changes in circulating MicroRNAs that are associated with nonlytic hepatocyte release. J Virol. 89:9454–9464. 2015.PubMed/NCBI View Article : Google Scholar

51 

Liu XL, Pan Q, Zhang RN, Shen F, Yan SY, Sun C, Xu ZJ, Chen YW and Fan JG: Disease-specific miR-34a as diagnostic marker of non-alcoholic steatohepatitis in a Chinese population. World J Gastroenterol. 22:9844–9852. 2016.PubMed/NCBI View Article : Google Scholar

52 

Isac S, Panaitescu AM, Iesanu M, Grigoras IF, Totan A, Udriste A, Cucu N, Peltecu G, Zagrean L and Zagrean AM: Maternal high-fat diet modifies the immature hippocampus vulnerability to perinatal asphyxia in rats. Neonatology. 114:355–361. 2018.PubMed/NCBI View Article : Google Scholar

53 

Lieber CS, Leo MA, Wang X and Decarli LM: Effect of chronic alcohol consumption on Hepatic SIRT1 and PGC-1alpha in rats. Biochem Biophys Res Commun. 370:44–48. 2008.PubMed/NCBI View Article : Google Scholar

54 

Lee YJ and Shukla SD: Histone H3 phosphorylation at serine 10 and serine 28 is mediated by p38 MAPK in rat hepatocytes exposed to ethanol and acetaldehyde. Eur J Pharmacol. 573:29–38. 2007.PubMed/NCBI View Article : Google Scholar

55 

Pal-Bhadra M, Bhadra U, Jackson DE, Mamatha L, Park PH and Shukla SD: Distinct methylation patterns in histone H3 at Lys-4 and Lys-9 correlate with up- & down-regulation of genes by ethanol in hepatocytes. Life Sci. 81:979–987. 2007.PubMed/NCBI View Article : Google Scholar

56 

Lu SC, Huang ZZ, Yang H, Mato JM, Avila MA and Tsukamoto H: Changes in methionine adenosyltransferase and S-adenosylmethionine homeostasis in alcoholic rat liver. Am J Physiol Gastrointest Liver Physiol. 279:G178–G185. 2000.PubMed/NCBI View Article : Google Scholar

57 

Dannenberg LO, Chen HJ, Tian H and Edenberg HJ: Differential regulation of the alcohol dehydrogenase 1B (ADH1B) and ADH1C genes by DNA methylation and histone deacetylation. Alcohol Clin Exp Res. 30:928–937. 2006.PubMed/NCBI View Article : Google Scholar

58 

Mandrekar P: Epigenetic regulation in alcoholic liver disease. World J Gastroenterol. 17:2456–2464. 2011.PubMed/NCBI View Article : Google Scholar

59 

Bala S, Marcos M, Kodys K, Csak T, Catalano D, Mandrekar P and Szabo G: Up-regulation of microRNA-155 in macrophages contributes to increased tumor necrosis factor {alpha} (TNF{alpha}) production via increased mRNA half-life in alcoholic liver disease. J Biol Chem. 286:1436–1444. 2011.PubMed/NCBI View Article : Google Scholar

60 

Tang Y, Banan A, Forsyth CB, Fields JZ, Lau CK, Zhang LJ and Keshavarzian A: Effect of alcohol on miR-212 expression in intestinal epithelial cells and its potential role in alcoholic liver disease. Alcohol Clin Exp Res. 32:355–364. 2008.PubMed/NCBI View Article : Google Scholar

61 

Medici V and LaSalle JM: Genetics and epigenetic factors of Wilson disease. Ann Transl Med. 7 (Suppl 2)(S58)2019.PubMed/NCBI View Article : Google Scholar

62 

Li M, Li Y, Chen J, Wei W, Pan X, Liu J, Liu Q, Leu W, Zhang L, Yang X, et al: Copper ions inhibit S-adenosylhomocysteine hydrolase by causing dissociation of NAD+ cofactor. Biochemistry. 46:11451–11458. 2007.PubMed/NCBI View Article : Google Scholar

63 

Yin R, Mo J, Dai J and Wang H: Nickel(II) inhibits tet-mediated 5-methylcytosine oxidation by high affinity displacement of the cofactor iron(II). ACS Chem Biol. 12:1494–1498. 2017.PubMed/NCBI View Article : Google Scholar

64 

Le A, Shibata NM, French SW, Kim K, Kharbanda KK, Islam MS, LaSalle JM, Halsted CH, Keen CL and Medici V: Characterization of timed changes in hepatic copper concentrations, methionine metabolism, gene expression, and global DNA methylation in the Jackson toxic milk mouse model of Wilson disease. Int J Mol Sci. 15:8004–8023. 2014.PubMed/NCBI View Article : Google Scholar

65 

Medici V, Kieffer DA, Shibata NM, Chima H, Kim K, Canovas A, Medrano JF, Islas-Trejo AD, Kharbanda KK, Olson K, et al: Wilson disease: Epigenetic effects of choline supplementation on phenotype and clinical course in a mouse model. Epigenetics. 11:804–818. 2016.PubMed/NCBI View Article : Google Scholar

Related Articles

Journal Cover

May-2022
Volume 23 Issue 5

Print ISSN: 1792-0981
Online ISSN:1792-1015

Sign up for eToc alerts

Recommend to Library

Copy and paste a formatted citation
x
Spandidos Publications style
Isac T, Isac S, Rababoc R, Cotorogea M and Iliescu L: Epigenetics in inflammatory liver diseases: A clinical perspective (Review). Exp Ther Med 23: 366, 2022
APA
Isac, T., Isac, S., Rababoc, R., Cotorogea, M., & Iliescu, L. (2022). Epigenetics in inflammatory liver diseases: A clinical perspective (Review). Experimental and Therapeutic Medicine, 23, 366. https://doi.org/10.3892/etm.2022.11293
MLA
Isac, T., Isac, S., Rababoc, R., Cotorogea, M., Iliescu, L."Epigenetics in inflammatory liver diseases: A clinical perspective (Review)". Experimental and Therapeutic Medicine 23.5 (2022): 366.
Chicago
Isac, T., Isac, S., Rababoc, R., Cotorogea, M., Iliescu, L."Epigenetics in inflammatory liver diseases: A clinical perspective (Review)". Experimental and Therapeutic Medicine 23, no. 5 (2022): 366. https://doi.org/10.3892/etm.2022.11293