Next Article in Journal
Comparative Transcriptome Combined with Proteome Analyses Revealed Key Factors Involved in Alfalfa (Medicago sativa) Response to Waterlogging Stress
Next Article in Special Issue
Molecular Mechanisms of Pulmonary Fibrogenesis and Its Progression to Lung Cancer: A Review
Previous Article in Journal
Changes in Human Foetal Osteoblasts Exposed to the Random Positioning Machine and Bone Construct Tissue Engineering
Previous Article in Special Issue
Nano-Strategies to Target Breast Cancer-Associated Fibroblasts: Rearranging the Tumor Microenvironment to Achieve Antitumor Efficacy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Molecular Mechanisms Driving Progression of Liver Cirrhosis towards Hepatocellular Carcinoma in Chronic Hepatitis B and C Infections: A Review

1
Division of Gastroenterology and Hepatology, Department of Medicine, Nihon University School of Medicine, 30-1 Oyaguchi-kamicho, Itabashi-ku, Tokyo 173-8610, Japan
2
Lung Cancer and Respiratory Disease Center, Yamanashi Central Hospital, 1-1-1 Fujimi, Kofu, Yamanashi 400-8506, Japan
3
Genome Analysis Center, Yamanashi Central Hospital, Yamanashi 400-8506, Japan
4
The University of Tokyo, 7-3-1 Hongo, Bunkyo-ku, Tokyo 113-8655, Japan
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2019, 20(6), 1358; https://doi.org/10.3390/ijms20061358
Submission received: 19 January 2019 / Revised: 23 February 2019 / Accepted: 14 March 2019 / Published: 18 March 2019

Abstract

:
Almost all patients with hepatocellular carcinoma (HCC), a major type of primary liver cancer, also have liver cirrhosis, the severity of which hampers effective treatment for HCC despite recent progress in the efficacy of anticancer drugs for advanced stages of HCC. Here, we review recent knowledge concerning the molecular mechanisms of liver cirrhosis and its progression to HCC from genetic and epigenomic points of view. Because ~70% of patients with HCC have hepatitis B virus (HBV) and/or hepatitis C virus (HCV) infection, we focused on HBV- and HCV-associated HCC. The literature suggests that genetic and epigenetic factors, such as microRNAs, play a role in liver cirrhosis and its progression to HCC, and that HBV- and HCV-encoded proteins appear to be involved in hepatocarcinogenesis. Further studies are needed to elucidate the mechanisms, including immune checkpoints and molecular targets of kinase inhibitors, associated with liver cirrhosis and its progression to HCC.

1. Introduction

Almost all patients with hepatocellular carcinoma (HCC), a major type of primary liver cancer, have liver cirrhosis [1], the severity of which can prevent effective treatment for HCC despite recent progress in the efficacy of anticancer drugs for advanced stages of HCC [2,3,4,5,6,7]. Because most patients with liver cirrhosis are asymptomatic, it is difficult to diagnose early stages of HCC [8], and patients with hepatic symptoms and HCC are considered to have advanced-stage HCC [8,9]. These issues explain the prevalence of poor prognosis for HCC patients.
HCC is the 4th most common neoplasm and the 2nd commonest cause of cancer deaths in the world. Notably, HCC is a male-dominant disease, with the incidence of HCC ~3-fold higher in males than in females [10]. Hepatitis B virus (HBV) infection is associated with the higher HCC incidence in persons with cirrhosis, occurring in high endemic areas and in Western countries (5-year cumulative incidence, 15% and 10%, respectively) [11]. In hepatitis C virus (HCV)-related cirrhosis, the 5-year cumulative HCC risk is 30% in Japan and 17% in Western countries [11].
Histologically, liver fibrosis involves the deposition of extracellular matrix proteins, including collagen, in higher-order structures within hepatic parenchyma [12,13], with hepatic stellate cells and fibroblasts representing major producers of collagen [14]. The histological pattern of liver fibrosis is not unique, and the extent and distribution of liver fibrosis exhibits various patterns depending on different etiologies [12,15]. Excessive liver fibrosis often develops within portal tracts and extends into the hepatic parenchyma in viral hepatitis, with these activities appearing to be associated with persistent portal inflammation [12,16]. Bridging fibrosis appears following the development of periportal fibrosis and extends across lobules to connect mesenchymal structures (portal tracts and central veins) to different extents. Generally, these processes accompany intrahepatic portosystemic vascular shunting and regeneration of hepatocytes, thereby transforming from normal hepatic architecture to nodule formation and finally establishing the structure of cirrhosis [12].
In this review, we discuss the molecular mechanisms underlying the progression of liver cirrhosis to HCC. We expect that this review will help clinicians diagnose and treat patients with liver cirrhosis and/or HCC in their daily clinical practice. Notably, ~70% of HCC patients are afflicted with HBV or HCV infection [11]; therefore, we focused on the occurrence of HCC during HBV and HCV infection (Figure 1).

2. Liver Cirrhosis and Its Progression to HCC with HBV Infection

2.1. Development of Liver Cirrhosis in Patients with Chronic Hepatitis B Infection

HBV infection causes acute and chronic hepatitis, cirrhosis, and HCC. HBV-carrier rates are higher in African and Asian countries and globally are estimated to have resulted in 786,000 deaths in 2010, the majority of which were attributed to HCC (341,400 deaths) and cirrhosis (312,400 deaths) [17]. Annual rates of development from chronic hepatitis B to liver cirrhosis ranged from 2.1 to 6.0% [18,19,20,21], and annual rates of the development of liver cirrhosis in HBV e antigen (HBeAg)-positive or anti-HBe-positive patients were 2.4% and 1.3%, respectively [18].
HBeAg-positivity and elevated HBV DNA levels are risk factors for the development of liver cirrhosis in patients with chronic hepatitis B [19]. Sumi et al. [20] reported that progression to cirrhosis is slower in HBV genotype B than that in HBV genotype C infection. Additionally, for patients with chronic hepatitis B infection, coinfection with HCV or human immunodeficiency virus (HIV) is another risk factor for the development of liver cirrhosis [21,22].
Older age (≥55 years), male gender, chronic active hepatitis, higher alanine aminotransferase (ALT) levels, history of decompensation, ferredoxin-1-associated single-nucleotide polymorphism, HLA-DQA2 rs9276370 variants, and HLA-DQB2 rs7756516 variants are also respective risk factors for the development of liver cirrhosis [18,19,20,21,23,24,25].

2.2. Development of HCC in Patients with HBV-related Liver Cirrhosis

Annual rates of occurrence of HCC in patients with HBV-related cirrhosis are ~2.3% [1]. Higher HBV DNA levels, HBeAg-positivity, higher HBV surface antigen (HBsAg) levels, HBV genotype C, and basal core-promoter mutations are viral risk factors for the occurrence of HCC, and older age, male gender, chronic active hepatitis, higher ALT levels, and higher α-fetoprotein levels are non-viral risk factors for the occurrence of HCC [20,23,26]. Treatment with nucleos(t)ide analogs for HBV control decreases the occurrence rates of HCC in patients with HBV-related cirrhosis [27].

3. Liver Cirrhosis and Its Progression to HCC with HCV Infection

HCV infection causes acute and chronic hepatitis, cirrhosis, and HCC. In 2015, there were 170,000 new HCV infections [28], with annual incidence rates of cirrhosis accompanying HCV infection at 1.1% [29]. Forns et al. [30] reported that chronic HCV infection displayed a high rate of progression to liver cirrhosis over a prolonged follow-up period, and that an aspartate aminotransferase (AST) value >70 IU/L was associated with cirrhosis development [odds ratio (OR): 4.22, 95% confidence interval (CI): 1.3–13.8]. Additionally, there is an association between HCV genotype 3 and steatosis, which accelerates fibrosis development over time in HCV genotype 3-infected patients [31,32]. Moreover, previous studies reported that exposure to HCV at a young age is associated with a reduced rate of fibrosis progression [33,34,35,36,37], and that liver fibrosis progression was mainly dependent on the age and duration of infection [36]. The evidences are growing that liver steatosis and diabetes mellitus are factors affecting progression to HCC in HCV-infected patients with compensated cirrhosis [11]. These comorbidities typically affecting aged patients might explain the increased malignant progression of the cirrhotic liver damage. Ongoing alcohol consumption and severe inflammation according to liver histology are also associated with liver fibrosis progression [35,36,37]. A multivariate analysis of HCV-infected patients [38] revealed that only male sex (OR: 3.17, 95% CI: 1.152–8.773) and HIV infection (OR: 6.85, 95% CI: 2.93–16.005) were associated with advanced liver fibrosis. Furthermore, HBsAg-positive HCV-coinfected patients are at high risk of developing liver disease [39], and injected-drug users [40]; patients with insulin resistance or diabetes [41]; and patients with concurrent obesity, diabetes, and steatosis [42] are also at risk of advanced liver fibrosis. Following antiviral treatment, sustained virological response (SVR) can reduce liver fibrosis progression in most patients infected with HCV [43].
Annual rates of HCC occurrence in patients with HCV-related cirrhosis are ~4.5% [1]. Caporaso et al. [44] found that mean age and male/female ratio were significantly higher in patients with HCC plus liver cirrhosis than in those with liver cirrhosis alone. Additionally, HBsAg-positive HCV-coinfected patients with liver cirrhosis are at a high risk of developing HCC [45], with elevated ALT levels [46] and hepatic steatosis [47] risk factors for the development of HCC in patients with HCV-related liver cirrhosis. SVR can reduce HCC development in most patients infected with HCV-related liver cirrhosis [48], although changes in the risk of HCC development following HCV eradication with direct-acting antivirals (DAAs) is controversial [49,50,51]. Further studies will be needed in DAA-era.

4. Molecular Mechanisms of Liver Cirrhosis and Its Progression to HCC

Accumulation of genetic mutations occurs during HCC progression. Recent advances in next-generation sequencing technology to augment Sanger sequencing enabled whole-genome sequencing to allow critical insights into the molecular mechanisms associated with this activity.

4.1. Driver-Gene Candidates in HCC

Totoki et al. [52] found 30 candidate driver genes [telomerase reverse transcriptase (TERT), catenin β1 (CTNNB1), tumor protein p53 (TP53), AT-rich interaction domain 2 (ARID2), axin 1 (AXIN1), TSC complex subunit 2 (TSC2), retinoblastoma protein 1 (RB1), activin A receptor type 2A (ACVR2A), bromodomain containing 7 (BRD7), cyclin dependent kinase inhibitor (CDKN)1A, menin 1 (MEN1), polypeptide N-acetylgalactosaminyltransferase 11 (GALN11), fibroblast growth factor 19 (FGF19), cyclin (CCN)D1, AT-rich interaction domain 1A (ARID1A), CDKN2A, CDKN2B, ribosomal protein S6 kinase, 90 kDa, polypeptide 3 (RPS6KA3), nuclear factor, erythroid 2 like 2 (NFE2L2), nuclear receptor corepressor 1 (NCOR1), alcohol dehydrogenase 1B, β polypeptide (ADH1B), Snf2-related CREB binding protein (CREBBP) activator protein (SRCAP), Fc receptor like 1 (FCRL1), phosphatase and tensin homolog (PTEN), heterogeneous nuclear ribonucleoprotein A2/B1 (HNRNPA2B1), cytochrome P450 family 2 subfamily E member 1 (CYP2E1), mitogen-activated protein kinase 3 (MAP2K3), tuberous sclerosis (TSC)1, transmembrane protein 99 (TMEM99), and glucose-6-phosphatase, catalytic subunit (G6PC)] and 11 core pathway modules [β-catenin, chromatin remodeling, DNA repair, estrogen-related receptor beta (ERRB), fibrinogen, mechanistic target of rapamycin kinase (mTOR), synaptic connection, TERT, p53, NOTCH, and NCOR] through the collection of data from 503 HCC genomes from different populations (Table 1).
Fujimoto et al. [53] identified 15 significantly mutated genes, including TP53, ERBB-receptor feedback inhibitor 1, Zic family member 3, CTNNB1, glucoside xylosyltransferase 1, otopetrin 1, albumin (ALB), ATM serine/threonine kinase (ATM), zinc finger protein 226, ubiquitin specific peptidase (USP)25, WW-domain-containing E3 ubiquitin protein ligase 1, immunoglobulin superfamily member 10, ARID1A, ubiquitin protein ligase E3 component n-recognin 3, and bromodomain adjacent to zinc finger domain 2B, after sequencing and analyzing the whole genomes of 27 HCCs, including 25 with HBV- or HCV-associated HCC. Additionally, whole-genome sequencing analyses of HCCs demonstrated the influence of etiological backgrounds on mutation patterns and recurrent mutations in chromatin regulators [53]. The authors found that multiple chromatin regulators, including ARID1A, ARID1B, ARID2, lysine methyltransferase 2A, and lysine methyltransferase 2C, were mutated in ~50% of the tumors, with clonal integration of the HBV genome in the TERT gene frequently detected.

4.2. The p53-RB Pathway

TP53 mutations have been reported in HCC in Japan and aflatoxin B1-induced HCC [59,60]. In HCC, the p53-RB pathway is altered in 72% of cases, with 68% of these a result of significantly altered genes [52]. Somatic mutations and copy number alterations in TP53, which plays a central role in apoptosis [61], are present in 31% and 37% of cases, respectively. Among the upstream genes targeted by p53, somatic mutations and copy number alterations in ATM are present in 4% and 6% of cases, respectively; those in RPS6KA3 are present in 4% and 5% of cases, respectively, and those in CDKN2A are present in 2% and 20% of cases, respectively. Among downstream genes targeted by p53, somatic mutations and copy number alterations in CDKN1A are present in 1% and 1% of cases, respectively; those in FBXW7 are present in <1% and 16% of cases, respectively, and those in CCNE1, which suppress RB1 transcription and promote cell cycle progression, are present in 0% and 2% of cases, respectively. Among upstream genes targeted by RB, somatic mutations and copy number alterations in CCND1 are present in 0% and 8% of cases, respectively. Somatic mutations and copy number alterations in RB1, which inhibit cell cycle progression, are present in 4% and 20% of cases, respectively. Additionally, RB controls the levels of p21, which is associated with hepatocarcinogenesis [62]. Overall, 72% of HCC cases harbor alterations in component genes of either the p53 or RB pathway alone or the combined p53-RB pathway (Figure 2) [52].

4.3. The β-catenin Pathway (WNT Pathway)

In HCC, the β-catenin pathway is altered in 66% of cases, with 51% of these involving significantly altered genes [52]. Somatic mutations and copy number alterations in CTNNB1, which induce the expression of WNT target genes and transcription of CCND1 resulting in inhibited RB1 expression, are present in 31% and <1% of cases, respectively [52]. Among the upstream genes targeted by β-catenin, somatic mutations and copy number alterations in NCOR1 are present in 2% and 29% of cases, respectively; those in FGF19 are present in 0% and 6% of cases, respectively; those in AXIN1 are present in 6% and 15% of cases, respectively; and those in APC are present in 2% and 4% of cases, respectively [52]. Overall, 66% of HCC cases harbored WNT-pathway-related alterations.

4.4. Chromatin and Transcription Modulators

In HCC, chromatin-remodeling pathways are altered in 67% of cases, with 41% of these involving significantly altered genes [52]. Among upstream genes associated with the nucleosome-remodeling pathway, somatic mutations and copy number alterations in ARID1B are present in <1% and 15% of cases, respectively; those in ARID1A are present in 8% and 17% of cases, respectively; those in BRD7 are present in 2% and 16% of cases, respectively; those in ARID2 are present in 10% and 2% of cases, respectively; those in switch/sucrose non-fermentable (SWI/SNF)-related, matrix associated, actin-dependent regulator of chromatin (SMARC) subfamily C member 1 are present in <1% and 4% of cases, respectively; those in SMARC subfamily B member 1 are present in <1% and 6% of cases, respectively; those in SMARC subfamily A member 4 are present in 1% and 10%, respectively; those in SMARC subfamily E member 1 are present in <1% and 3% of cases, respectively; and those in SMARC subfamily A member 2 are present in 2% and 10% of cases, respectively [52]. Additionally, Li et al. reported that 18.2% of patients with HCV-associated HCC in the United States and Europe harbored ARID2-inactivation mutations [57]. Moreover, another study reported ARID1A mutation in 14 of 110 (13%) HBV-associated HCC specimens [54]. ARID1A functions as the epigenetic regulation of hepatic lipid homeostasis, and its suppression leads to nonalcoholic fatty liver disease and nonalcoholic steatohepatitis (NASH) [58].
Among upstream genes associated with the histone-modification pathway, somatic mutations and copy number alterations in NCOR1 are present in 2% and 29% of cases, respectively; those in SRCAP are present in 3% and 10% of cases, respectively; those in SET domain bifurcated histone lysine methyltransferase 1 (SETDB1) are present in 2% and <1% of cases, respectively; and those in lysine demethylase 6A (KDM6A) are present in 1% and 5% of cases, respectively [52]. Notably, the histone methyltransferase SETDB1 promotes HCC metastasis [63,64], and KDM6A is associated with the epithelial–mesenchymal transition in HCC [65].

4.5. Other Pathways

In HCC, the phosphoinositide 3-kinase (PI3K)–mTOR pathway is altered in 45% of cases, with 26% of these involving significantly altered genes [52]. Somatic mutations and copy number alterations of PI3KCA (encoding the p110α subunit of PI3K) are present in 1% and <1% of cases, respectively; those in neurofibromin 1 are present in 4% and 3% of cases, respectively; those in PTEN are present in 1% and 10% of cases, respectively; those in inositol polyphosphate-4-phosphatase, type IIB, are present in 1% and 16% of cases, respectively; those in serine/threonine kinase 11 are present in <1% and 11% of cases, respectively; those in TSC1 are present in 2% and 8% of cases, respectively; and those in TSC2 are present in 5% and 14% of cases, respectively [52]. TSC1 and TSC2 suppress mTOR expression, which promotes HCC proliferation and survival. In sorafenib-treated patients with HCC, oncogenic PI3K–mTOR-pathway alterations are associated with lower disease-control rates and decreased median progression-free survival and overall survival [66].
In HCC, the nuclear factor (erythroid-derived 2)-like 2 (NRF2)–kelch-like ECH-associated protein 1 (KEAP1) pathway is altered in 19% of cases, with 5% of these involving significantly altered genes [52]. The NRF2–KEAP1 pathway is associated with an oxidative-stress response, and persistent activation of NRF2 through the accumulation of p62 is involved in HCC development [67,68,69].
There are many protein-coding genes recurrently mutated at a frequency of <5% in HCC [55,70], whereas HBV integrations and frequent noncoding mutations in the TERT promoter represent prominent examples of noncoding mutations in HCC [52,53,54,55]. Nault et al. [56] identified frequent somatic mutations resulting in the activation of the TERT promoter in HCC (59%), cirrhotic preneoplastic macronodules (25%), and hepatocellular adenomas with malignant transformation in HCC (44%). Moreover, TERT-promoter mutation represents the most frequent genetic alteration in HCC arising from the cirrhotic or non-cirrhotic liver [56], resulting in reportedly enhanced TERT activity in HCC [71].

5. Molecular Mechanisms of HBV-Associated HCC

5.1. HBV Genome Integration Promotes HCC

The discovery that HBV integrates into host chromosomes calls into question whether HBV genome integration interacts with the activation of oncogenic processes in HCC [72]. Tokino et al. [72] could not detect specific HBV genome-integration sites in chromosomes. About half of HBV DNA-cell DNA junctions are located within the so-called cohesive end region that lies between two 11-bp direct repeats (DR1 and DR2) in the HBV genome where transcription and replication of the genome are initiated. All of the integrated HBV genomes were defective in at least one site around the cohesive end region, particularly within the HBx gene [73]. Additionally, Imazeki et al. [74] detected integrated HBV DNA in all nine HBsAg-seropositive HCCs and three of 25 (12%) HBsAg-seronegative HCCs in Japan. An analysis of breakpoints within the integration region showed that 40% of breakpoints were near the 1800th nucleotide of the HBV genome, which contains an enhancer, an HBx gene, and core promoters of the HBV genome [75].
TERT is located on chromosome 5p, which is one of the most common targets for amplification in non-small cell lung cancer [76] and is reportedly directly associated with HBV genome integration [77]. Recent data from whole-genome sequencing supports this observation and confirmed the frequent observation of HBV genome integration in the TERT locus in a high clonal proportion [52,53,54,55,56].
HBx expression might play a role in hepatocarcinogenesis by interfering with telomerase activity during hepatocyte proliferation [78], which is supported by breakpoint analysis of the HBV genome [75]. HBx functions as a transcriptional activator and suppressor and has effects on hepatocellular apoptosis [79,80]. HBx proteins may upregulate the transcriptional activation of human telomerase transcriptase [81]. Cis-activation of human TERT mRNA by HBx gene may also be the mechanism in hepatocarcinogenesis [82].
Therefore, HBV genome integration into host genomes is a mechanism involved in HCC associated with HBV infection. Moreover, whole-genome sequencing demonstrated that the integration of HBV DNA into the host hepatocyte genome can be detected in 80% to 90% of HCC and in ~30% of non-HCC liver tissue adjacent to HCC [83], and that this integration appears prior to the occurrence of HCC [75]. Thus, it is possible that occult HBV infection may also accelerate hepatocarcinogenesis in HBsAg-negative patients to some extent [84].

5.2. Inflammation Promotes HBV-Associated HCC

In general, a high number of HBV-DNA integrations randomly distributed among chromosomes has been detected in HBV-infected liver [85]. Chronic HBV infection progresses through multiple phases, including immune tolerance, immune activation, immune control, and, in a subset of patients who achieve immune control and immune reactivation [86]. Immune-mediated liver injury is often associated with elevated ALT levels, and elevations in tumor necrosis factor-α (TNF-α) and interleukin (IL)-1β levels are often observed in the sera of HBV-infected patients [87]. Additionally, long-lasting hepatic inflammation caused by host immune responses during chronic HBV infection can promote liver fibrosis, cirrhosis and HCC progression due to accelerated hepatocyte turnover rates and the accumulation of mutations [88].

5.3. Epigenetic Mechanisms Involved in HBV-Associated HCC

Epigenetic mechanisms play a role in HBV-associated hepatocarcinogenesis. MicroRNAs (miRs) are endogenous noncoding RNAs (18–22 nucleotides in length) that posttranscriptionally regulate the expression of target genes. miRs bind to the 3′-untranslated region (UTR) of target mRNA, thereby inhibiting translation [89].
Several miRNAs involved in the Toll-like receptor (TLR) signaling pathway play a critical role in innate immunity against HBV infection [89]. For example, miR-145 and miR-148a target TLR3, miR-200b, miR-200c, miR148a, miR-455, and let-7-family members target TLR4, let-7b and miR-155 target TLR7, and miR-148a targets TLR9, and all of which are involved in HBV infection [89,90]. HBV might influence miR expression and induce inflammatory cytokine production [91]. Previous reports indicated that some hepatic miRs are involved in the pathogenesis of HBV-associated HCC [92,93,94,95,96,97,98,99,100,101,102,103,104,105,106,107,108,109,110,111,112,113,114,115,116,117,118,119,120,121,122,123,124,125,126,127,128,129,130,131,132,133,134,135,136,137,138,139,140,141,142], with certain serum miRs involved in HBV-associated HCC also potentially useful as biomarkers (Table 2).
Other noncoding RNAs, such as long noncoding RNAs (lncRNAs) and circular RNA (circRNA), are also involved in the pathogenesis of HBV-associated HCC [143,144,145]. Functional studies reveal that lncRNAs (100–300 kb) contribute to the onset and progression of HBV-related HCC [143], and circRNAs, which form covalently closed continuous loops by means of unique non-canonical ‘head-to-tail’ splicing in the absence of free 3′ or 5′ ends, might promote the development of HBV-associated HCC [144,145].
Epigenetic silencing of genes, such as methylation of promoter regions, also regulates gene expression in HBV-associated HCC and could potentially serve as a diagnostic and prognostic marker of the disease [146,147]. Additionally, HBV can cause epigenetic changes by altering the methylation state of cellular DNA, the posttranslational modification of histones, and miR expression [146,148], all of which are also critical for the pathogenesis of HBV-associated HCC.

5.4. Roles of HBV-encoded Proteins

HBV is a partially double-stranded DNA virus (genome length: 3200 bp) and a member of the Orthohepadnavirus genus and the Hepadnaviridae family [149]. The HBV genome encodes at least four proteins [HBsAg, a core protein (splice variant: HBeAg), a DNA polymerase, and the HBx protein] that are translated from mRNAs transcribed from HBV covalently closed circular DNA and/or from HBV genome sequences integrated into the host genome [150]. Viral protein translation is initiated through binding of the PreS2 domain of HBV surface antigens to protein kinase C (PKC)α/β, which triggers PKC-dependent activation of Raf-1 proto-oncogene serine/threonine kinase (RAF-1)/MAP2K signaling and transcription factors, such as activator protein-1 (AP-1) and nuclear factor kappaB (NF-kB), resulting in the increased proliferation of hepatocytes [151]. PreS2-mediated activity subsequently upregulates the expression of the transcriptional coactivator with PDZ-binding motif by repressing miRNA-338-3p, which promotes HCC proliferation and migration [152]. Moreover, a previous study showed that a truncated mutant of HBsAg increases HBV-related tumorigenesis in a mechanism potentially associated with the downregulated expression of tumor growth factor (TGF)BI associated with the TGFβ–SMAD pathway [153]. Furthermore, HBsAg enhances the IL-6–STAT3 pathway, thereby increasing the HBsAg-mediated malignant potential of HBV-associated HCC [154].
The HBV core protein enhances cytokine production [155], and HBeAg is reportedly associated with the host immune response and cytokine production [156,157,158], both of which play roles in HBV-associated HCC.
Numerous studies reported an association between the HBx protein and hepatocarcinogenesis [159,160,161,162,163,164,165,166,167,168,169,170]. Integrated HBV DNA harbors the conserved sequences for genes encoding the core protein, surface antigens, and HBx along with their respective promoter sequences [159,160], suggesting that HBx is important for hepatocarcinogenesis. HBx represents a transactivating factor [161], and transgenic mice expressing HBx develop HCC [162]. HBx transactivates binding sites for the transcription factors AP-1 and NF-κB [163], activates the p53-RB [164,165] and β-catenin [164,165,166,167,168,169,170] pathways, and is involved in chromatin remodeling [171,172,173] and transcriptional modulation in hepatocarcinogenesis [174].
Overexpression of the HBV polymerase due to core-gene deletion enhances HCC-cell growth by inhibiting miR-100 [175]. A previous study showed that transgenic mice expressing the reverse-transcriptase domain of HBV polymerase in their livers developed early cirrhosis with steatosis by 18 months of age, with 10% subsequently developing HCC [176].

6. Molecular Mechanisms of HCV-Associated HCC

6.1. Inflammation Promotes HCV-Associated HCC

Cytokines reflect the degree of inflammation in the liver of patients with chronic hepatitis C, with this production possibly related to HCC development [177]. Takano et al. [1] prospectively investigated the incidence of HCC in 124 cases with HCV infection, finding that HCC occurred in 13 cases that included 12 cirrhotic livers and only 1 non-cirrhotic liver and suggesting that most HCC occurs in advanced liver diseases in HCV-infected individuals [178]. Because hepatic inflammation plays a role in the development of advanced liver diseases, inflammation is an important aspect in the development of HCC associated with chronic HCV infection.

6.2. Epigenetic Mechanisms Involved in HCV-Associated HCC

Epigenetic mechanisms also play a role in HCV-associated hepatocarcinogenesis. Previous studies reported the involvement of specific hepatic and serum miRs in the pathogenesis of HCV-associated HCC [179,180,181,182,183,184,185,186,187,188,189,190,191,192,193,194,195,196], with some of the serum miRs representing potentially useful biomarkers of the disease (Table 3). Notably, the number of reports of miRs involved in HCV-associated HCC is smaller than that related to HBV-associated HCC (Table 2).
Previous studies reported that lncRNAs are involved in the pathogenesis of HBV-associated HCC [143,144,145], with functional studies revealing that these lncRNAs contribute also to the onset and progression of HCV-related HCC [195,197]. Zhang et al. [198] reported LINC01419 transcripts expressed at higher levels in early stage HCC as compared with levels observed in dysplastic tissue. Moreover, this study also reported increased and decreased levels of AK021443 and AF070632 in advanced HCC, respectively, and that LINC01419 and AK021443 regulated the expression of genes associated with cell cycle progression, whereas AF070632 was associated with cofactor binding, oxidation–reduction activity and carboxylic acid catabolic processes [198].
Upregulated lncRNAs associated with urothelial carcinoma associated-1 (lncRNA-UCA1) and WD-repeat-containing antisense to TP53 (lncRNA-Wrap53) potentially serve as novel serum biomarkers for HCC diagnosis and prognosis [199]. Additionally, the lncRNA associated with activated by TGFβ (lncRNA-ATB) is a key regulator of TGFβ signaling and is positively correlated with the development of liver cirrhosis and HCC-specific vascular invasion [200]. Moreover, lncRNA-ATB might represent a novel diagnostic biomarker and potential therapeutic target for HCV-related hepatic fibrosis [200]. A previous study showed that HCV infection upregulates the expression of miR-373 and Wee1-like protein kinase (WEE1), a pivotal player in the G2/M transition in the cell cycle (although WEE1 is a direct target of miR-373). This study also showed that miR-373 forms a complex with the LINC00657, resulting in release of their common target, WEE1, in HCV-infected cells, and the promotion of uncontrolled cell growth [195]. Another study reported that hypermethylation of promoter regions suppressed mRNA expression, which played a role in the progression of HCV-associated HCC [201].

6.3. Roles of HCV-coding Proteins

We previously reported the roles for HCV-coding proteins in hepatocarcinogenesis [178], with the HCV core proteins and NS5A reportedly playing important roles in HCC development [202,203,204,205,206]. HCV-infected patients showed a higher prevalence of diabetes mellitus and insulin resistance (IR) relative to those with HBV infection [207]. IR measured using a homeostasis model assessment of IR (HOMA-IR) is significantly associated with HCC development in patients with or without chronic HCV infection. Moreover, patients with NASH, which is associated with elevated HOMA-IR, have an increased risk of liver fibrosis, cirrhosis, and HCC [208,209].
A previous study showed that the HCV core protein downregulates insulin receptor substrate (IRS)1 and IRS2 by upregulating SOCS3 levels [210]. Moreover, cross-talk between the HCV core protein and molecules regulating insulin signaling might affect HCV-associated hepatocarcinogenesis [210,211,212,213,214]. Insulin-like growth factors initiate tyrosyl phosphorylation of IRS1 and activate multiple signaling pathways essential for liver growth and HCC [215], with the activation of IRS1-mediated signaling potentially contributing to hepatic oncogenesis.
CCAAT/enhancer-binding protein (C/EBPβ) and HCV NS5A might be essential components promoting increased gluconeogenesis associated with HCV infection [216]. Previous studies reported the involvement of HCV NS5A in enhanced gluconeogenic gene expression associated with impaired insulin signaling [217,218].
Sphere-forming hepatocytes express several cancer stem-like cell (CSC) markers, including c-Kit. Previous studies showed that the HCV core protein significantly upregulates c-Kit expression at the transcriptional level, and that HCV infection potentiates CSC generation [219,220]. Additionally, a Western diet high in cholesterol and saturated fat (HCFD) in combination with the translation of HCV NS5A stimulates TLR4–Nanog homeobox (NANOG) and leptin receptor–phosphorylated STAT3 signaling, resulting in liver tumorigenesis through an exaggerated mesenchymal phenotype involving prominent Twist1-expressing tumor-initiating stem-like cells expressing NANOG [221].

7. HCC-specific Molecular Mechanisms from a Therapeutic Point of View

Sorafenib, regorafenib and lenvatinib are approved therapies as oral molecular-targeting drugs for advanced stages of HCC [3,4,5,6]. Sorafenib is an oral serine/threonine kinase inhibitor targeting the extracellular-signaling-regulated kinase/MAPK pathway, vascular endothelial growth factor receptor (VEGFR), platelet-derived growth factor receptor (PDGFR), and epithelial growth factor receptor (EGFR). Additionally, it acts as a tyrosine kinase inhibitor targeting VEGFR1, VEGFR2, VEGFR3, PDGFRβ, ret proto-oncogene (RET), and fms-related tyrosine kinase 3, resulting in the inhibition of tumor proliferation [222].
Regorafenib is an oral multiple kinase inhibitor of VEGFR1, VEGFR2, VEGFR3, TEK receptor tyrosine kinase, PDGFRβ, fibroblast growth factor receptor (FGFR), proto-oncogene receptor tyrosine kinase (KIT), RET, RAF-1, and B-RAF. Lenvatinib is an oral tyrosine kinase inhibitor targeting VEGFR1/2/3, FGFR1/2/3/4, PDGFRα, KIT, and RET to inhibit tumor angiogenesis and growth [222]. Therefore, sorafenib, regorafenib, and lenvatinib represent multiple kinase inhibitors that do not suppress a specific molecule, thereby restricting their use only in patients with advanced HCC and well-compensated liver function.
Nivolumab, pembrolizumab and tislelizumab are anti-programmed death-1 (PD-1) antibodies. The immune-checkpoint molecule PD-1 is a receptor that negatively regulates immune responses [222,223] via binding of its representative ligands PD-ligand (PD-L)1 and PD-L2. Inhibition of this pathway can eliminate tumors by recovering their immunosuppressive effects and restoring innate immune activity [222,224]. Cytotoxic T lymphocyte-associated antigen-4 (CTLA-4) is another counterreceptor for the B7 family of costimulatory molecules and a negative regulator of T cell activation. A blockade of the inhibitory effects of CTLA-4 is supposed to enhance immune responses against tumor cells [225] and also represents a type of immune checkpoint. The blockade of such immune checkpoints represents a promising therapeutic option for HCC; therefore, further studies are needed to determine other immune checkpoints, as well as potential targets of kinase inhibitors related to liver cirrhosis to HCC.

8. Conclusions

This review describes the molecular mechanisms of liver cirrhosis and its progression to HCC (Figure 3). Recent progress in next-generation sequencing has revealed several HCC-specific driver-gene candidates. Additionally, ~70% of HCC cases are caused by HBV and/or HCV infection. Although nucleos(t)ide analogs and direct-acting antivirals can control HBV and HCV replication, HCC occurrence is occasionally observed [49,50,51], and better biomarkers for early detection of HCC are needed. Furthermore, genetic and epigenetic factors, such as miRs, are involved in liver cirrhosis and its progression to HCC. Despite significant advances, additional studies are required to elucidate other molecular mechanisms, including immune checkpoints and molecular targets of kinase inhibitors associated with liver cirrhosis and its progression to HCC.

Author Contributions

Writing—Original Draft Preparation, T.K.; Writing—Review & Editing, T.G., Y.H., M.M.; Supervision, M.O.

Funding

This work was partly supported by JSPS KAKENHI GRANT Number 17K09404 (to T.K.).

Conflicts of Interest

The other authors declare no conflict of interest.

References

  1. Takano, S.; Yokosuka, O.; Imazeki, F.; Tagawa, M.; Omata, M. Incidence of hepatocellular carcinoma in chronic hepatitis B and C: A prospective study of 251 patients. Hepatology 1995, 21, 650–655. [Google Scholar] [CrossRef] [PubMed]
  2. Takayama, T.; Makuuchi, M. Segmental liver resections, present and future-caudate lobe resection for liver tumours. Hepatogastroenterology 1998, 45, 20–23. [Google Scholar]
  3. Llovet, J.M.; Ricci, S.; Mazzaferro, V.; Hilgard, P.; Gane, E.; Blanc, J.F.; de Oliveira, A.C.; Santoro, A.; Raoul, J.L.; Forner, A.; et al. Sorafenib in advanced hepatocellular carcinoma. N. Engl. J. Med. 2008, 359, 378–390. [Google Scholar] [CrossRef] [PubMed]
  4. Cheng, A.L.; Kang, Y.K.; Chen, Z.; Tsao, C.J.; Qin, S.; Kim, J.S.; Luo, R.; Feng, J.; Ye, S.; Yang, T.S.; et al. Efficacy and safety of sorafenib in patients in the Asia-Pacific region with advanced hepatocellular carcinoma: A phase III randomised, double-blind, placebo-controlled trial. Lancet Oncol. 2009, 10, 25–34. [Google Scholar] [CrossRef]
  5. Bruix, J.; Qin, S.; Merle, P.; Granito, A.; Huang, Y.H.; Bodoky, G.; Pracht, M.; Yokosuka, O.; Rosmorduc, O.; Breder, V.; et al. Regorafenib for patients with hepatocellular carcinoma who progressed on sorafenib treatment (RESORCE): A randomised, double-blind, placebo-controlled, phase 3 trial. Lancet 2017, 389, 56–66. [Google Scholar] [CrossRef]
  6. Kudo, M.; Finn, R.S.; Qin, S.; Han, K.H.; Ikeda, K.; Piscaglia, F.; Baron, A.; Park, J.W.; Han, G.; Jassem, J.; et al. Lenvatinib versus sorafenib in first-line treatment of patients with unresectable hepatocellular carcinoma: A randomised phase 3 non-inferiority trial. Lancet 2018, 391, 1163–1173. [Google Scholar] [CrossRef]
  7. Schachter, J.; Ribas, A.; Long, G.V.; Arance, A.; Grob, J.J.; Mortier, L.; Daud, A.; Carlino, M.S.; McNeil, C.; Lotem, M.; et al. Pembrolizumab versus ipilimumab for advanced melanoma: Final overall survival results of a multicentre, randomised, open-label phase 3 study (KEYNOTE-006). Lancet 2017, 390, 1853–1862. [Google Scholar] [CrossRef]
  8. Bruix, J.; Reig, M.; Sherman, M. Evidence-Based Diagnosis, Staging, and Treatment of Patients With Hepatocellular Carcinoma. Gastroenterology 2016, 150, 835–853. [Google Scholar] [CrossRef]
  9. Obi, S.; Yoshida, H.; Toune, R.; Unuma, T.; Kanda, M.; Sato, S.; Tateishi, R.; Teratani, T.; Shiina, S.; Omata, M. Combination therapy of intraarterial 5-fluorouracil and systemic interferon-alpha for advanced hepatocellular carcinoma with portal venous invasion. Cancer 2006, 106, 1990–1997. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Kanda, T.; Takahashi, K.; Nakamura, M.; Nakamoto, S.; Wu, S.; Haga, Y.; Sasaki, R.; Jiang, X.; Yokosuka, O. Androgen Receptor Could Be a Potential Therapeutic Target in Patients with Advanced Hepatocellular Carcinoma. Cancers 2017, 9, 43. [Google Scholar] [CrossRef] [PubMed]
  11. Fattovich, G.; Stroffolini, T.; Zagni, I.; Donato, F. Hepatocellular carcinoma in cirrhosis: Incidence and risk factors. Gastroenterology 2004, 127 (Suppl. 1), S35–S50. [Google Scholar] [CrossRef]
  12. Bataller, R.; Brenner, D.A. Liver fibrosis. J. Clin. Investig. 2005, 115, 209–218. [Google Scholar] [CrossRef] [PubMed]
  13. Patel, K.; Bedossa, P.; Castera, L. Diagnosis of liver fibrosis: Present and future. Semin. Liver Dis. 2015, 35, 166–183. [Google Scholar] [CrossRef]
  14. Xu, L.; Hui, A.Y.; Albanis, E.; Arthur, M.J.; O’Byrne, S.M.; Blaner, W.S.; Mukherjee, P.; Friedman, S.L.; Eng, F.J. Human hepatic stellate cell lines, LX-1 and LX-2: New tools for analysis of hepatic fibrosis. Gut 2005, 54, 142–151. [Google Scholar] [CrossRef]
  15. Okuda, K.; Nakashima, T.; Kojiro, M.; Kondo, Y.; Wada, K. Hepatocellular carcinoma without cirrhosis in Japanese patients. Gastroenterology 1989, 97, 140–146. [Google Scholar] [CrossRef]
  16. Clouston, A.D.; Powell, E.E.; Walsh, M.J.; Richardson, M.M.; Demetris, A.J.; Jonsson, J.R. Fibrosis correlates with a ductular reaction in hepatitis C: Roles of impaired replication, progenitor cells and steatosis. Hepatology 2005, 41, 809–818. [Google Scholar] [CrossRef] [Green Version]
  17. Lozano, R.; Naghavi, M.; Foreman, K.; Lim, S.; Shibuya, K.; Aboyans, V.; Abraham, J.; Adair, T.; Aggarwal, R.; Ahn, S.Y.; et al. Global and regional mortality from 235 causes of death for 20 age groups in 1990 and 2010: A systematic analysis for the Global Burden of Disease Study 2010. Lancet 2012, 380, 2095–2128. [Google Scholar] [CrossRef]
  18. Liaw, Y.F.; Tai, D.I.; Chu, C.M.; Chen, T.J. The development of cirrhosis in patients with chronic type B hepatitis: A prospective study. Hepatology 1988, 8, 493–496. [Google Scholar] [CrossRef]
  19. Fattovich, G.; Brollo, L.; Giustina, G.; Noventa, F.; Pontisso, P.; Alberti, A.; Realdi, G.; Ruol, A. Natural history and prognostic factors for chronic hepatitis type B. Gut 1991, 32, 294–298. [Google Scholar] [CrossRef] [PubMed]
  20. Sumi, H.; Yokosuka, O.; Seki, N.; Arai, M.; Imazeki, F.; Kurihara, T.; Kanda, T.; Fukai, K.; Kato, M.; Saisho, H. Influence of hepatitis B virus genotypes on the progression of chronic type B liver disease. Hepatology 2003, 37, 19–26. [Google Scholar] [CrossRef] [Green Version]
  21. Liaw, Y.F.; Chen, Y.C.; Sheen, I.S.; Chien, R.N.; Yeh, C.T.; Chu, C.M. Impact of acute hepatitis C virus superinfection in patients with chronic hepatitis B virus infection. Gastroenterology 2004, 126, 1024–1029. [Google Scholar] [CrossRef]
  22. Lacombe, K.; Massari, V.; Girard, P.M.; Serfaty, L.; Gozlan, J.; Pialoux, G.; Mialhes, P.; Molina, J.M.; Lascoux-Combe, C.; Wendum, D.; et al. Major role of hepatitis B genotypes in liver fibrosis during coinfection with HIV. AIDS 2006, 20, 419–427. [Google Scholar] [CrossRef]
  23. Stroffolini, T.; Esvan, R.; Biliotti, E.; Sagnelli, E.; Gaeta, G.B.; Almasio, P.L. Gender differences in chronic HBsAg carriers in Italy: Evidence for the independent role of male sex in severity of liver disease. J. Med. Virol. 2015, 87, 1899–1903. [Google Scholar] [CrossRef]
  24. Al-Qahtani, A.; Khalak, H.G.; Alkuraya, F.S.; Al-hamoudi, W.; Alswat, K.; Al Balwi, M.A.; Al Abdulkareem, I.; Sanai, F.M.; Abdo, A.A. Genome-wide association study of chronic hepatitis B virus infection reveals a novel candidate risk allele on 11q22.3. J. Med. Genet. 2013, 50, 725–732. [Google Scholar] [CrossRef] [PubMed]
  25. Chang, S.W.; Fann, C.S.; Su, W.H.; Wang, Y.C.; Weng, C.C.; Yu, C.J.; Hsu, C.L.; Hsieh, A.R.; Chien, R.N.; Chu, C.M.; et al. A genome-wide association study on chronic HBV infection and its clinical progression in male Han-Taiwanese. PLoS ONE 2014, 9, e99724. [Google Scholar] [CrossRef]
  26. Chu, C.M.; Liaw, Y.F. Hepatitis B virus-related cirrhosis: Natural history and treatment. Semin. Liver Dis. 2006, 26, 142–152. [Google Scholar] [CrossRef]
  27. Tawada, A.; Kanda, T.; Imazeki, F.; Yokosuka, O. Prevention of hepatitis B virus-associated liver diseases by antiviral therapy. Hepatol. Int. 2016, 10, 574–593. [Google Scholar] [CrossRef]
  28. World Health Organization. Global Hepatitis Report. 2017. Available online: www.who.int/hepatitis/publications/global-hepatitis-report2017/en/ (accessed on 27 December 2018).
  29. Yang, Y.H.; Chen, W.C.; Tsan, Y.T.; Chen, M.J.; Shih, W.T.; Tsai, Y.H.; Chen, P.C. Statin use and the risk of cirrhosis development in patients with hepatitis C virus infection. J. Hepatol. 2015, 63, 1111–1117. [Google Scholar] [CrossRef]
  30. Forns, X.; Ampurdanès, S.; Sanchez-Tapias, J.M.; Guilera, M.; Sans, M.; Sánchez-Fueyo, A.; Quintó, L.; Joya, P.; Bruguera, M.; Rodés, J. Long-term follow-up of chronic hepatitis C in patients diagnosed at a tertiary-care center. J. Hepatol. 2001, 35, 265–271. [Google Scholar] [CrossRef]
  31. Westin, J.; Nordlinder, H.; Lagging, M.; Norkrans, G.; Wejstål, R. Steatosis accelerates fibrosis development over time in hepatitis C virus genotype 3 infected patients. J. Hepatol. 2002, 37, 837–842. [Google Scholar] [CrossRef]
  32. Rubbia-Brandt, L.; Fabris, P.; Paganin, S.; Leandro, G.; Male, P.J.; Giostra, E.; Carlotto, A.; Bozzola, L.; Smedile, A.; Negro, F. Steatosis affects chronic hepatitis C progression in a genotype specific way. Gut 2004, 53, 406–412. [Google Scholar] [CrossRef] [Green Version]
  33. Locasciulli, A.; Testa, M.; Pontisso, P.; Benvegnù, L.; Fraschini, D.; Corbetta, A.; Noventa, F.; Masera, G.; Alberti, A. Prevalence and natural history of hepatitis C infection in patients cured of childhood leukemia. Blood 1997, 90, 4628–4633. [Google Scholar] [PubMed]
  34. Kenny-Walsh, E. Clinical outcomes after hpatitis C infection from contaminated anti-D immune globulin. Irish Hepatology Research Group. N. Engl. J. Med. 1999, 340, 1228–1233. [Google Scholar] [CrossRef]
  35. Poynard, T.; Bedossa, P.; Opolon, P. Natural history of liver fibrosis progression in patients with chronic hepatitis C. The OBSVIRC, METAVIR, CLINIVIR, and DOSVIRC groups. Lancet 1997, 349, 825–832. [Google Scholar] [CrossRef]
  36. Poynard, T.; Ratziu, V.; Charlotte, F.; Goodman, Z.; McHutchison, J.; Albrecht, J. Rates and risk factors of liver fibrosis progression in patients with chronic hepatitis c. J. Hepatol. 2001, 34, 730–739. [Google Scholar] [CrossRef]
  37. Shiffman, M.L. Natural history and risk factors for progression of hepatitis C virus disease and development of hepatocellular cancer before liver transplantation. Liver Transpl. 2003, 9, S14–S20. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Fuster, D.; Planas, R.; Muga, R.; Ballesteros, A.L.; Santos, J.; Tor, J.; Sirera, G.; Guardiola, H.; Salas, A.; Cabré, E.; et al. Advanced liver fibrosis in HIV/HCV-coinfected patients on antiretroviral therapy. AIDS Res. Hum. Retrovir. 2004, 20, 1293–1297. [Google Scholar] [CrossRef] [PubMed]
  39. Ribes, J.; Clèries, R.; Rubió, A.; Hernández, J.M.; Mazzara, R.; Madoz, P.; Casanovas, T.; Casanova, A.; Gallen, M.; Rodríguez, C.; et al. Cofactors associated with liver disease mortality in an HBsAg-positive Mediterranean cohort: 20 years of follow-up. Int. J. Cancer 2006, 119, 687–694. [Google Scholar] [CrossRef] [Green Version]
  40. Wilson, L.E.; Torbenson, M.; Astemborski, J.; Faruki, H.; Spoler, C.; Rai, R.; Mehta, S.; Kirk, G.D.; Nelson, K.; Afdhal, N.; et al. Progression of liver fibrosis among injection drug users with chronic hepatitis C. Hepatology 2006, 43, 788–795. [Google Scholar] [CrossRef] [Green Version]
  41. Petta, S.; Cammà, C.; Di Marco, V.; Alessi, N.; Cabibi, D.; Caldarella, R.; Licata, A.; Massenti, F.; Tarantino, G.; Marchesini, G.; et al. Insulin resistance and diabetes increase fibrosis in the liver of patients with genotype 1 HCV infection. Am. J. Gastroenterol. 2008, 103, 1136–1144. [Google Scholar] [CrossRef]
  42. Dyal, H.K.; Aguilar, M.; Bhuket, T.; Liu, B.; Holt, E.W.; Torres, S.; Cheung, R.; Wong, R.J. Concurrent Obesity, Diabetes, and Steatosis Increase Risk of Advanced Fibrosis Among HCV Patients: A Systematic Review. Dig. Dis. Sci. 2015, 60, 2813–2824. [Google Scholar] [CrossRef] [PubMed]
  43. Shiratori, Y.; Imazeki, F.; Moriyama, M.; Yano, M.; Arakawa, Y.; Yokosuka, O.; Kuroki, T.; Nishiguchi, S.; Sata, M.; Yamada, G.; et al. Histologic improvement of fibrosis in patients with hepatitis C who have sustained response to interferon therapy. Ann. Intern. Med. 2000, 132, 517–524. [Google Scholar] [CrossRef]
  44. Caporaso, N.; Romano, M.; Marmo, R.; de Sio, I.; Morisco, F.; Minerva, A.; Coltorti, M. Hepatitis C virus infection is an additive risk factor for development of hepatocellular carcinoma in patients with cirrhosis. J. Hepatol. 1991, 12, 367–371. [Google Scholar] [CrossRef]
  45. Benvegnù, L.; Fattovich, G.; Noventa, F.; Tremolada, F.; Chemello, L.; Cecchetto, A.; Alberti, A. Concurrent hepatitis B and C virus infection and risk of hepatocellular carcinoma in cirrhosis. A prospective study. Cancer 1994, 74, 2442–2448. [Google Scholar] [CrossRef]
  46. Ishikawa, T.; Ichida, T.; Yamagiwa, S.; Sugahara, S.; Uehara, K.; Okoshi, S.; Asakura, H. High viral loads, serum alanine aminotransferase and gender are predictive factors for the development of hepatocellular carcinoma from viral compensated liver cirrhosis. J. Gastroenterol. Hepatol. 2001, 16, 1274–1281. [Google Scholar] [CrossRef]
  47. Ohata, K.; Hamasaki, K.; Toriyama, K.; Matsumoto, K.; Saeki, A.; Yanagi, K.; Abiru, S.; Nakagawa, Y.; Shigeno, M.; Miyazoe, S.; et al. Hepatic steatosis is a risk factor for hepatocellular carcinoma in patients with chronic hepatitis C virus infection. Cancer 2003, 97, 3036–3043. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Yoshida, H.; Shiratori, Y.; Moriyama, M.; Arakawa, Y.; Ide, T.; Sata, M.; Inoue, O.; Yano, M.; Tanaka, M.; Fujiyama, S.; et al. Interferon therapy reduces the risk for hepatocellular carcinoma: National surveillance program of cirrhotic and noncirrhotic patients with chronic hepatitis C in Japan. IHIT Study Group. Inhibition of Hepatocarcinogenesis by Interferon Therapy. Ann. Intern. Med. 1999, 131, 174–181. [Google Scholar] [CrossRef] [PubMed]
  49. Alberti, A.; Piovesan, S. Increased incidence of liver cancer after successful DAA treatment of chronic hepatitis C: Fact or fiction? Liver Int. 2017, 37, 802–808. [Google Scholar] [CrossRef] [PubMed]
  50. El Kassas, M.; Funk, A.L.; Salaheldin, M.; Shimakawa, Y.; Eltabbakh, M.; Jean, K.; El Tahan, A.; Sweedy, A.T.; Afify, S.; Youssef, N.F.; et al. Increased recurrence rates of hepatocellular carcinoma after DAA therapy in a hepatitis C-infected Egyptian cohort: A comparative analysis. J. Viral Hepat. 2018, 25, 623–630. [Google Scholar] [CrossRef]
  51. Sasaki, R.; Kanda, T.; Kato, N.; Yokosuka, O.; Moriyama, M. Hepatitis C virus-associated hepatocellular carcinoma after sustained virologic response. World J. Hepatol. 2018, 10, 898–906. [Google Scholar] [CrossRef]
  52. Totoki, Y.; Tatsuno, K.; Covington, K.R.; Ueda, H.; Creighton, C.J.; Kato, M.; Tsuji, S.; Donehower, L.A.; Slagle, B.L.; Nakamura, H.; et al. Trans-ancestry mutational landscape of hepatocellular carcinoma genomes. Nat. Genet. 2014, 46, 1267–1273. [Google Scholar] [CrossRef]
  53. Fujimoto, A.; Totoki, Y.; Abe, T.; Boroevich, K.A.; Hosoda, F.; Nguyen, H.H.; Aoki, M.; Hosono, N.; Kubo, M.; Miya, F.; et al. Whole-genome sequencing of liver cancers identifies etiological influences on mutation patterns and recurrent mutations in chromatin regulators. Nat. Genet. 2012, 44, 760–764. [Google Scholar] [CrossRef] [PubMed]
  54. Huang, J.; Deng, Q.; Wang, Q.; Li, K.Y.; Dai, J.H.; Li, N.; Zhu, Z.D.; Zhou, B.; Liu, X.Y.; Liu, R.F.; et al. Exome sequencing of hepatitis B virus-associated hepatocellular carcinoma. Nat. Genet. 2012, 44, 1117–1121. [Google Scholar] [CrossRef]
  55. Schulze, K.; Imbeaud, S.; Letouzé, E.; Alexandrov, L.B.; Calderaro, J.; Rebouissou, S.; Couchy, G.; Meiller, C.; Shinde, J.; Soysouvanh, F.; et al. Exome sequencing of hepatocellular carcinomas identifies new mutational signatures and potential therapeutic targets. Nat. Genet. 2015, 47, 505–511. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Nault, J.C.; Mallet, M.; Pilati, C.; Calderaro, J.; Bioulac-Sage, P.; Laurent, C.; Laurent, A.; Cherqui, D.; Balabaud, C.; Zucman-Rossi, J. High frequency of telomerase reverse-transcriptase promoter somatic mutations in hepatocellular carcinoma and preneoplastic lesions. Nat. Commun. 2013, 4, 2218. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Li, M.; Zhao, H.; Zhang, X.; Wood, L.D.; Anders, R.A.; Choti, M.A.; Pawlik, T.M.; Daniel, H.D.; Kannangai, R.; Offerhaus, G.J.; et al. Inactivating mutations of the chromatin remodeling gene ARID2 in hepatocellular carcinoma. Nat. Genet. 2011, 43, 828–829. [Google Scholar] [CrossRef] [Green Version]
  58. Moore, A.; Wu, L.; Chuang, J.C.; Sun, X.; Luo, X.; Gopal, P.; Li, L.; Celen, C.; Zimmer, M.; Zhu, H. Arid1a loss drives non-alcoholic steatohepatitis in mice via epigenetic dysregulation of hepatic lipogenesis and fatty acid oxidation. Hepatology 2018. [Google Scholar] [CrossRef]
  59. Nose, H.; Imazeki, F.; Ohto, M.; Omata, M. p53 gene mutations and 17p allelic deletions in hepatocellular carcinoma from Japan. Cancer 1993, 72, 355–3560. [Google Scholar] [CrossRef]
  60. Imazeki, F.; Yokosuka, O.; Ohto, M.; Omata, M. Aflatoxin and p53 abnormality in duck hepatocellular carcinoma. J. Gastroenterol. Hepatol. 1995, 10, 646–649. [Google Scholar] [CrossRef]
  61. Ray, R.B.; Meyer, K.; Ray, R. Suppression of apoptotic cell death by hepatitis C virus core protein. Virology 1996, 226, 176–182. [Google Scholar] [CrossRef]
  62. Ray, R.B.; Steele, R.; Meyer, K.; Ray, R. Hepatitis C virus core protein represses p21WAF1/Cip1/Sid1 promoter activity. Gene 1998, 208, 331–336. [Google Scholar] [CrossRef]
  63. Fei, Q.; Shang, K.; Zhang, J.; Chuai, S.; Kong, D.; Zhou, T.; Fu, S.; Liang, Y.; Li, C.; Chen, Z.; et al. Histone methyltransferase SETDB1 regulates liver cancer cell growth through methylation of p53. Nat. Commun. 2015, 6, 8651. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Wong, C.M.; Wei, L.; Law, C.T.; Ho, D.W.; Tsang, F.H.; Au, S.L.; Sze, K.M.; Lee, J.M.; Wong, C.C.; Ng, I.O. Up-regulation of histone methyltransferase SETDB1 by multiple mechanisms in hepatocellular carcinoma promotes cancer metastasis. Hepatology 2016, 63, 474–487. [Google Scholar] [CrossRef] [PubMed]
  65. Kodama, T.; Newberg, J.Y.; Kodama, M.; Rangel, R.; Yoshihara, K.; Tien, J.C.; Parsons, P.H.; Wu, H.; Finegold, M.J.; Copeland, N.G.; et al. Transposon mutagenesis identifies genes and cellular processes driving epithelial-mesenchymal transition in hepatocellular carcinoma. Proc. Natl. Acad. Sci. USA 2016, 113, E3384–E3393. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Harding, J.J.; Nandakumar, S.; Armenia, J.; Khalil, D.N.; Albano, M.; Ly, M.; Shia, J.; Hechtman, J.F.; Kundra, R.; El Dika, I.; et al. Prospective Genotyping of Hepatocellular Carcinoma: Clinical Implications of Next Generation Sequencing for Matching Patients to Targeted and Immune Therapies. Clin. Cancer Res. 2018. [Google Scholar] [CrossRef] [PubMed]
  67. Inami, Y.; Waguri, S.; Sakamoto, A.; Kouno, T.; Nakada, K.; Hino, O.; Watanabe, S.; Ando, J.; Iwadate, M.; Yamamoto, M.; et al. Persistent activation of Nrf2 through p62 in hepatocellular carcinoma cells. J. Cell Biol. 2011, 193, 275–284. [Google Scholar] [CrossRef] [Green Version]
  68. Zavattari, P.; Perra, A.; Menegon, S.; Kowalik, M.A.; Petrelli, A.; Angioni, M.M.; Follenzi, A.; Quagliata, L.; Ledda-Columbano, G.M.; Terracciano, L.; et al. Nrf2, but not β-catenin, mutation represents an early event in rat hepatocarcinogenesis. Hepatology 2015, 62, 851–862. [Google Scholar] [CrossRef] [Green Version]
  69. Bartolini, D.; Dallaglio, K.; Torquato, P.; Piroddi, M.; Galli, F. Nrf2-p62 autophagy pathway and its response to oxidative stress in hepatocellular carcinoma. Transl. Res. 2018, 193, 54–71. [Google Scholar] [CrossRef]
  70. Fujimoto, A.; Furuta, M.; Totoki, Y.; Tsunoda, T.; Kato, M.; Shiraishi, Y.; Tanaka, H.; Taniguchi, H.; Kawakami, Y.; Ueno, M.; et al. Whole-genome mutational landscape and characterization of noncoding and structural mutations in liver cancer. Nat. Genet. 2016, 48, 500–509. [Google Scholar] [CrossRef]
  71. Kojima, H.; Yokosuka, O.; Imazeki, F.; Saisho, H.; Omata, M. Telomerase activity and telomere length in hepatocellular carcinoma and chronic liver disease. Gastroenterology 1997, 112, 493–500. [Google Scholar] [CrossRef]
  72. Tokino, T.; Matsubara, K. Chromosomal sites for hepatitis B virus integration in human hepatocellular carcinoma. J. Virol. 1991, 65, 6761–6764. [Google Scholar]
  73. Nagaya, T.; Nakamura, T.; Tokino, T.; Tsurimoto, T.; Imai, M.; Mayumi, T.; Kamino, K.; Yamamura, K.; Matsubara, K. The mode of hepatitis B virus DNA integration in chromosomes of human hepatocellular carcinoma. Genes Dev. 1987, 1, 773–782. [Google Scholar] [CrossRef]
  74. Imazeki, F.; Omata, M.; Yokosuka, O.; Okuda, K. Integration of hepatitis B virus DNA in hepatocellular carcinoma. Cancer 1986, 58, 1055–1060. [Google Scholar] [CrossRef]
  75. Wang, M.; Xi, D.; Ning, Q. Virus-induced hepatocellular carcinoma with special emphasis on HBV. Hepatol. Int. 2017, 11, 171–180. [Google Scholar] [CrossRef]
  76. Meyerson, M.; Counter, C.M.; Eaton, E.N.; Ellisen, L.W.; Steiner, P.; Caddle, S.D.; Ziaugra, L.; Beijersbergen, R.L.; Davidoff, M.J.; Liu, Q.; et al. hEST2, the putative human telomerase catalytic subunit gene, is up-regulated in tumor cells and during immortalization. Cell 1997, 90, 785–795. [Google Scholar] [CrossRef]
  77. Tokino, T.; Fukushige, S.; Nakamura, T.; Nagaya, T.; Murotsu, T.; Shiga, K.; Aoki, N.; Matsubara, K. Chromosomal translocation and inverted duplication associated with integrated hepatitis B virus in hepatocellular carcinomas. J. Virol. 1987, 61, 3848–3854. [Google Scholar]
  78. Koike, K.; Shirakata, Y.; Yaginuma, K.; Arii, M.; Takada, S.; Nakamura, I.; Hayashi, Y.; Kawada, M.; Kobayashi, M. Oncogenic potential of hepatitis B virus. Mol. Biol. Med. 1989, 6, 151–160. [Google Scholar]
  79. Kanda, T.; Yokosuka, O.; Imazeki, F.; Yamada, Y.; Imamura, T.; Fukai, K.; Nagao, K.; Saisho, H. Hepatitis B virus X protein (HBx)-induced apoptosis in HuH-7 cells: Influence of HBV genotype and basal core promoter mutations. Scand. J. Gastroenterol. 2004, 39, 478–485. [Google Scholar] [CrossRef]
  80. Liu, H.; Shi, W.; Luan, F.; Xu, S.; Yang, F.; Sun, W.; Liu, J.; Ma, C. Hepatitis B virus X protein upregulates transcriptional activation of human telomerase reverse transcriptase. Virus Genes. 2010, 40, 174–182. [Google Scholar] [CrossRef]
  81. Zou, S.Q.; Qu, Z.L.; Li, Z.F.; Wang, X. Hepatitis B virus X gene induces human telomerase reverse transcriptase mRNA expression in cultured normal human cholangiocytes. World J. Gastroenterol. 2004, 10, 2259–2262. [Google Scholar] [CrossRef] [Green Version]
  82. Kojima, H.; Kaita, K.D.; Xu, Z.; Ou, J.H.; Gong, Y.; Zhang, M.; Minuk, G.Y. The absence of up-regulation of telomerase activity during regeneration after partial hepatectomy in hepatitis B virus X gene transgenic mice. J. Hepatol. 2003, 39, 262–268. [Google Scholar] [CrossRef]
  83. Sung, W.K.; Zheng, H.; Li, S.; Chen, R.; Liu, X.; Li, Y.; Lee, N.P.; Lee, W.H.; Ariyaratne, P.N.; Tennakoon, C.; et al. Genome-wide survey of recurrent HBV integration in hepatocellular carcinoma. Nat. Genet. 2012, 44, 765–769. [Google Scholar] [CrossRef] [PubMed]
  84. Nakano, M.; Kawaguchi, T.; Nakamoto, S.; Kawaguchi, A.; Kanda, T.; Imazeki, F.; Kuromatsu, R.; Sumie, S.; Satani, M.; Yamada, S.; et al. Effect of occult hepatitis B virus infection on the early-onset of hepatocellular carcinoma in patients with hepatitis C virus infection. Oncol. Rep. 2013, 30, 2049–2055. [Google Scholar] [CrossRef] [PubMed]
  85. Mason, W.S.; Gill, U.S.; Litwin, S.; Zhou, Y.; Peri, S.; Pop, O.; Hong, M.L.; Naik, S.; Quaglia, A.; Bertoletti, A.; et al. HBV DNA Integration and Clonal Hepatocyte Expansion in Chronic Hepatitis B Patients Considered Immune Tolerant. Gastroenterology 2016, 151, 986–998. [Google Scholar] [CrossRef] [PubMed]
  86. Kennedy, P.T.F.; Litwin, S.; Dolman, G.E.; Bertoletti, A.; Mason, W.S. Immune Tolerant Chronic Hepatitis B: The Unrecognized Risks. Viruses 2017, 9, 96. [Google Scholar] [CrossRef] [PubMed]
  87. Wu, S.; Kanda, T.; Nakamoto, S.; Jiang, X.; Nakamura, M.; Sasaki, R.; Haga, Y.; Shirasawa, H.; Yokosuka, O. Cooperative effects of hepatitis B virus and TNF may play important roles in the activation of metabolic pathways through the activation of NF-κB. Int. J. Mol. Med. 2016, 38, 475–481. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Xie, Y. Hepatitis B Virus-Associated Hepatocellular Carcinoma. Adv. Exp. Med. Biol. 2017, 1018, 11–21. [Google Scholar] [CrossRef]
  89. Jiang, X.; Kanda, T.; Wu, S.; Nakamura, M.; Miyamura, T.; Nakamoto, S.; Banerjee, A.; Yokosuka, O. Regulation of microRNA by hepatitis B virus infection and their possible association with control of innate immunity. World J. Gastroenterol. 2014, 20, 7197–7206. [Google Scholar] [CrossRef]
  90. Sarkar, N.; Panigrahi, R.; Pal, A.; Biswas, A.; Singh, S.P.; Kar, S.K.; Bandopadhyay, M.; Das, D.; Saha, D.; Kanda, T.; et al. Expression of microRNA-155 correlates positively with the expression of Toll-like receptor 7 and modulates hepatitis B virus via C/EBP-β in hepatocytes. J. Viral Hepat. 2015, 22, 817–827. [Google Scholar] [CrossRef] [PubMed]
  91. Wang, W.; Bian, H.; Li, F.; Li, X.; Zhang, D.; Sun, S.; Song, S.; Zhu, Q.; Ren, W.; Qin, C.; et al. HBeAg induces the expression of macrophage miR-155 to accelerate liver injury via promoting production of inflammatory cytokines. Cell. Mol. Life Sci. 2018, 75, 2627–2641. [Google Scholar] [CrossRef] [PubMed]
  92. Wong, Q.W.; Lung, R.W.; Law, P.T.; Lai, P.B.; Chan, K.Y.; To, K.F.; Wong, N. MicroRNA-223 is commonly repressed in hepatocellular carcinoma and potentiates expression of Stathmin1. Gastroenterology 2008, 135, 257–269. [Google Scholar] [CrossRef] [PubMed]
  93. Zhang, X.; Liu, S.; Hu, T.; Liu, S.; He, Y.; Sun, S. Up-regulated microRNA-143 transcribed by nuclear factor kappa B enhances hepatocarcinoma metastasis by repressing fibronectin expression. Hepatology 2009, 50, 490–499. [Google Scholar] [CrossRef] [PubMed]
  94. Yang, L.; Ma, Z.; Wang, D.; Zhao, W.; Chen, L.; Wang, G. MicroRNA-602 regulating tumor suppressive gene RASSF1A is overexpressed in hepatitis B virus-infected liver and hepatocellular carcinoma. Cancer Biol. Ther. 2010, 9, 803–808. [Google Scholar] [CrossRef] [PubMed]
  95. Wang, Y.; Lu, Y.; Toh, S.T.; Sung, W.K.; Tan, P.; Chow, P.; Chung, A.Y.; Jooi, L.L.; Lee, C.G. Lethal-7 is down-regulated by the hepatitis B virus x protein and targets signal transducer and activator of transcription 3. J. Hepatol. 2010, 53, 57–66. [Google Scholar] [CrossRef]
  96. Huang, J.; Wang, Y.; Guo, Y.; Sun, S. Down-regulated microRNA-152 induces aberrant DNA methylation in hepatitis B virus-related hepatocellular carcinoma by targeting DNA methyltransferase 1. Hepatology 2010, 52, 60–70. [Google Scholar] [CrossRef] [PubMed]
  97. Li, L.M.; Hu, Z.B.; Zhou, Z.X.; Chen, X.; Liu, F.Y.; Zhang, J.F.; Shen, H.B.; Zhang, C.Y.; Zen, K. Serum microRNA profiles serve as novel biomarkers for HBV infection and diagnosis of HBV-positive hepatocarcinoma. Cancer Res. 2010, 70, 9798–9807. [Google Scholar] [CrossRef] [PubMed]
  98. Gao, P.; Wong, C.C.; Tung, E.K.; Lee, J.M.; Wong, C.M.; Ng, I.O. Deregulation of microRNA expression occurs early and accumulates in early stages of HBV-associated multistep hepatocarcinogenesis. J. Hepatol. 2011, 54, 1177–1184. [Google Scholar] [CrossRef] [PubMed]
  99. Jiang, R.; Deng, L.; Zhao, L.; Li, X.; Zhang, F.; Xia, Y.; Gao, Y.; Wang, X.; Sun, B. miR-22 promotes HBV-related hepatocellular carcinoma development in males. Clin. Cancer Res. 2011, 17, 5593–5603. [Google Scholar] [CrossRef] [PubMed]
  100. Wang, C.M.; Wang, Y.; Fan, C.G.; Xu, F.F.; Sun, W.S.; Liu, Y.G.; Jia, J.H. miR-29c targets TNFAIP3, inhibits cell proliferation and induces apoptosis in hepatitis B virus-related hepatocellular carcinoma. Biochem. Biophys. Res. Commun. 2011, 411, 586–592. [Google Scholar] [CrossRef]
  101. Qi, P.; Cheng, S.Q.; Wang, H.; Li, N.; Chen, Y.F.; Gao, C.F. Serum microRNAs as biomarkers for hepatocellular carcinoma in Chinese patients with chronic hepatitis B virus infection. PLoS ONE 2011, 6, e28486. [Google Scholar] [CrossRef]
  102. Cardin, R.; Piciocchi, M.; Sinigaglia, A.; Lavezzo, E.; Bortolami, M.; Kotsafti, A.; Cillo, U.; Zanus, G.; Mescoli, C.; Rugge, M.; et al. Oxidative DNA damage correlates with cell immortalization and mir-92 expression in hepatocellular carcinoma. BMC Cancer 2012, 12, 177. [Google Scholar] [CrossRef]
  103. Li, L.; Guo, Z.; Wang, J.; Mao, Y.; Gao, Q. Serum miR-18a: A potential marker for hepatitis B virus-related hepatocellular carcinoma screening. Dig. Dis. Sci. 2012, 57, 2910–2916. [Google Scholar] [CrossRef] [PubMed]
  104. Fu, X.; Tan, D.; Hou, Z.; Hu, Z.; Liu, G. miR-338-3p is down-regulated by hepatitis B virus X and inhibits cell proliferation by targeting the 3’-UTR region of CyclinD1. Int. J. Mol. Sci. 2012, 13, 8514–8539. [Google Scholar] [CrossRef] [PubMed]
  105. Yang, P.; Li, Q.J.; Feng, Y.; Zhang, Y.; Markowitz, G.J.; Ning, S.; Deng, Y.; Zhao, J.; Jiang, S.; Yuan, Y.; et al. TGF-β-miR-34a-CCL22 signaling-induced Treg cell recruitment promotes venous metastases of HBV-positive hepatocellular carcinoma. Cancer Cell 2012, 22, 291–303. [Google Scholar] [CrossRef] [PubMed]
  106. Wei, X.; Xiang, T.; Ren, G.; Tan, C.; Liu, R.; Xu, X.; Wu, Z. miR-101 is down-regulated by the hepatitis B virus x protein and induces aberrant DNA methylation by targeting DNA methyltransferase 3A. Cell. Signal. 2013, 25, 439–446. [Google Scholar] [CrossRef]
  107. Li, C.; Wang, Y.; Wang, S.; Wu, B.; Hao, J.; Fan, H.; Ju, Y.; Ding, Y.; Chen, L.; Chu, X.; et al. Hepatitis B virus mRNA-mediated miR-122 inhibition upregulates PTTG1-binding protein, which promotes hepatocellular carcinoma tumor growth and cell invasion. J. Virol. 2013, 87, 2193–2205. [Google Scholar] [CrossRef]
  108. Xu, X.; Fan, Z.; Kang, L.; Han, J.; Jiang, C.; Zheng, X.; Zhu, Z.; Jiao, H.; Lin, J.; Jiang, K.; et al. Hepatitis B virus X protein represses miRNA-148a to enhance tumorigenesis. J. Clin. Investig. 2013, 123, 630–645. [Google Scholar] [CrossRef] [Green Version]
  109. Shi, C.; Xu, X. MicroRNA-22 is down-regulated in hepatitis B virus-related hepatocellular carcinoma. Biomed. Pharmacother. 2013, 67, 375–380. [Google Scholar] [CrossRef]
  110. Lan, S.H.; Wu, S.Y.; Zuchini, R.; Lin, X.Z.; Su, I.J.; Tsai, T.F.; Lin, Y.J.; Wu, C.T.; Liu, H.S. Autophagy suppresses tumorigenesis of hepatitis B virus-associated hepatocellular carcinoma through degradation of microRNA-224. Hepatology 2014, 59, 505–517. [Google Scholar] [CrossRef]
  111. Fu, Y.; Wei, X.; Tang, C.; Li, J.; Liu, R.; Shen, A.; Wu, Z. Circulating microRNA-101 as a potential biomarker for hepatitis B virus-related hepatocellular carcinoma. Oncol. Lett. 2013, 6, 1811–1815. [Google Scholar] [CrossRef] [Green Version]
  112. Li, J.; Shi, W.; Gao, Y.; Yang, B.; Jing, X.; Shan, S.; Wang, Y.; Du, Z. Analysis of microRNA expression profiles in human hepatitis B virus-related hepatocellular carcinoma. Clin. Lab. 2013, 59, 1009–1015. [Google Scholar] [CrossRef] [PubMed]
  113. Zhang, T.; Zhang, J.; Cui, M.; Liu, F.; You, X.; Du, Y.; Gao, Y.; Zhang, S.; Lu, Z.; Ye, L.; et al. Hepatitis B virus X protein inhibits tumor suppressor miR-205 through inducing hypermethylation of miR-205 promoter to enhance carcinogenesis. Neoplasia 2013, 15, 1282–1291. [Google Scholar] [CrossRef]
  114. Sheng, Y.; Li, J.; Zou, C.; Wang, S.; Cao, Y.; Zhang, J.; Huang, A.; Tang, H. Downregulation of miR-101-3p by hepatitis B virus promotes proliferation and migration of hepatocellular carcinoma cells by targeting Rab5a. Arch. Virol. 2014, 159, 2397–2410. [Google Scholar] [CrossRef] [PubMed]
  115. Dang, Y.W.; Zeng, J.; He, R.Q.; Rong, M.H.; Luo, D.Z.; Chen, G. Effects of miR-152 on cell growth inhibition, motility suppression and apoptosis induction in hepatocellular carcinoma cells. Asian Pac. J. Cancer Prev. 2014, 15, 4969–4976. [Google Scholar] [CrossRef] [PubMed]
  116. Lan, S.H.; Wu, S.Y.; Zuchini, R.; Lin, X.Z.; Su, I.J.; Tsai, T.F.; Lin, Y.J.; Wu, C.T.; Liu, H.S. Autophagy-preferential degradation of MIR224 participates in hepatocellular carcinoma tumorigenesis. Autophagy 2014, 10, 1687–1689. [Google Scholar] [CrossRef] [Green Version]
  117. Meng, F.L.; Wang, W.; Jia, W.D. Diagnostic and prognostic significance of serum miR-24-3p in HBV-related hepatocellular carcinoma. Med. Oncol. 2014, 31, 177. [Google Scholar] [CrossRef]
  118. Bandopadhyay, M.; Banerjee, A.; Sarkar, N.; Panigrahi, R.; Datta, S.; Pal, A.; Singh, S.P.; Biswas, A.; Chakrabarti, S.; Chakravarty, R. Tumor suppressor micro RNA miR-145 and onco micro RNAs miR-21 and miR-222 expressions are differentially modulated by hepatitis B virus X protein in malignant hepatocytes. BMC Cancer 2014, 14, 721. [Google Scholar] [CrossRef] [PubMed]
  119. Xing, T.J.; Jiang, D.F.; Huang, J.X.; Xu, Z.L. Expression and clinical significance of miR-122 and miR-29 in hepatitis B virus-related liver disease. Genet. Mol. Res. 2014, 13, 7912–7918. [Google Scholar] [CrossRef]
  120. Zhao, Q.; Li, T.; Qi, J.; Liu, J.; Qin, C. The miR-545/374a cluster encoded in the Ftx lncRNA is overexpressed in HBV-related hepatocellular carcinoma and promotes tumorigenesis and tumor progression. PLoS ONE 2014, 9, e109782. [Google Scholar] [CrossRef]
  121. Gao, H.; Liu, C. miR-429 represses cell proliferation and induces apoptosis in HBV-related HCC. Biomed. Pharmacother. 2014, 68, 943–949. [Google Scholar] [CrossRef]
  122. Liu, F.Y.; Zhou, S.J.; Deng, Y.L.; Zhang, Z.Y.; Zhang, E.L.; Wu, Z.B.; Huang, Z.Y.; Chen, X.P. MiR-216b is involved in pathogenesis and progression of hepatocellular carcinoma through HBx-miR-216b-IGF2BP2 signaling pathway. Cell Death Dis. 2015, 6, e1670. [Google Scholar] [CrossRef] [Green Version]
  123. Yu, F.; Lu, Z.; Chen, B.; Dong, P.; Zheng, J. microRNA-150: A promising novel biomarker for hepatitis B virus-related hepatocellular carcinoma. Diagn. Pathol. 2015, 10, 129. [Google Scholar] [CrossRef] [PubMed]
  124. Cao, Y.; Chen, J.; Wang, D.; Peng, H.; Tan, X.; Xiong, D.; Huang, A.; Tang, H. Upregulated in Hepatitis B virus-associated hepatocellular carcinoma cells, miR-331-3p promotes proliferation of hepatocellular carcinoma cells by targeting ING5. Oncotarget 2015, 6, 38093–38106. [Google Scholar] [CrossRef] [Green Version]
  125. Gao, F.; Sun, X.; Wang, L.; Tang, S.; Yan, C. Downregulation of MicroRNA-145 Caused by Hepatitis B Virus X Protein Promotes Expression of CUL5 and Contributes to Pathogenesis of Hepatitis B Virus-Associated Hepatocellular Carcinoma. Cell. Physiol. Biochem. 2015, 37, 1547–1559. [Google Scholar] [CrossRef] [PubMed]
  126. Shao, J.; Cao, J.; Liu, Y.; Mei, H.; Zhang, Y.; Xu, W. MicroRNA-519a promotes proliferation and inhibits apoptosis of hepatocellular carcinoma cells by targeting FOXF2. FEBS Open Bio 2015, 5, 893–899. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Wang, Y.; Wang, C.M.; Jiang, Z.Z.; Yu, X.J.; Fan, C.G.; Xu, F.F.; Zhang, Q.; Li, L.I.; Li, R.F.; Sun, W.S.; et al. MicroRNA-34c targets TGFB-induced factor homeobox 2, represses cell proliferation and induces apoptosis in hepatitis B virus-related hepatocellular carcinoma. Oncol. Lett. 2015, 10, 3095–3102. [Google Scholar] [CrossRef] [Green Version]
  128. Ghosh, A.; Ghosh, A.; Datta, S.; Dasgupta, D.; Das, S.; Ray, S.; Gupta, S.; Datta, S.; Chowdhury, A.; Chatterjee, R.; et al. Hepatic miR-126 is a potential plasma biomarker for detection of hepatitis B virus infected hepatocellular carcinoma. Int. J. Cancer 2016, 138, 2732–2744. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  129. Chen, Y.; Dong, X.; Yu, D.; Wang, X. Serum miR-96 is a promising biomarker for hepatocellular carcinoma in patients with chronic hepatitis B virus infection. Int. J. Clin. Exp. Med. 2015, 8, 18462–18468. [Google Scholar] [PubMed]
  130. Chen, S.; Chen, H.; Gao, S.; Qiu, S.; Zhou, H.; Yu, M.; Tu, J. Differential expression of plasma microRNA-125b in hepatitis B virus-related liver diseases and diagnostic potential for hepatitis B virus-induced hepatocellular carcinoma. Hepatol. Res. 2017, 47, 312–320. [Google Scholar] [CrossRef] [PubMed]
  131. Yen, C.S.; Su, Z.R.; Lee, Y.P.; Liu, I.T.; Yen, C.J. miR-106b promotes cancer progression in hepatitis B virus-associated hepatocellular carcinoma. World J. Gastroenterol. 2016, 22, 5183–5192. [Google Scholar] [CrossRef]
  132. Kong, X.X.; Lv, Y.R.; Shao, L.P.; Nong, X.Y.; Zhang, G.L.; Zhang, Y.; Fan, H.X.; Liu, M.; Li, X.; Tang, H. HBx-induced MiR-1269b in NF-κB dependent manner upregulates cell division cycle 40 homolog (CDC40) to promote proliferation and migration in hepatoma cells. J. Transl. Med. 2016, 14, 189. [Google Scholar] [CrossRef] [Green Version]
  133. Liu, X.; Zhang, Y.; Wang, P.; Wang, H.; Su, H.; Zhou, X.; Zhang, L. HBX Protein-Induced Downregulation of microRNA-18a is Responsible for Upregulation of Connective Tissue Growth Factor in HBV Infection-Associated Hepatocarcinoma. Med. Sci. Monit. 2016, 22, 2492–2500. [Google Scholar] [CrossRef] [Green Version]
  134. Qiao, D.D.; Yang, J.; Lei, X.F.; Mi, G.L.; Li, S.L.; Li, K.; Xu, C.Q.; Yang, H.L. Expression of microRNA-122 and microRNA-22 in HBV-related liver cancer and the correlation with clinical features. Eur. Rev. Med. Pharmacol. Sci. 2017, 21, 742–747. [Google Scholar] [PubMed]
  135. Qin, X.; Chen, J.; Wu, L.; Liu, Z. MiR-30b-5p acts as a tumor suppressor, repressing cell proliferation and cell cycle in human hepatocellular carcinoma. Biomed. Pharmacother. 2017, 89, 742–750. [Google Scholar] [CrossRef] [PubMed]
  136. Bai, P.S.; Xia, N.; Sun, H.; Kong, Y. Pleiotrophin, a target of miR-384, promotes proliferation, metastasis and lipogenesis in HBV-related hepatocellular carcinoma. J. Cell. Mol. Med. 2017, 21, 3023–3043. [Google Scholar] [CrossRef] [PubMed]
  137. Li, G.; Zhang, W.; Gong, L.; Huang, X. MicroRNA-125a-5p Inhibits Cell Proliferation and Induces Apoptosis in Hepatitis B Virus-Related Hepatocellular Carcinoma by Downregulation of ErbB3. Oncol. Res. 2017. [Google Scholar] [CrossRef]
  138. Zhao, Q.; Sun, X.; Liu, C.; Li, T.; Cui, J.; Qin, C. Expression of the microRNA-143/145 cluster is decreased in hepatitis B virus-associated hepatocellular carcinoma and may serve as a biomarker for tumorigenesis in patients with chronic hepatitis B. Oncol. Lett. 2018, 15, 6115–6122. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Quoc, N.B.; Phuong, N.D.N.; Ngan, T.K.; Linh, N.T.M.; Cuong, P.H.; Chau, N.N.B. Expression of Plasma hsa-miR122 in HBV-Related Hepatocellular Carcinoma (HCC) in Vietnamese Patients. Microrna 2018, 7, 92–99. [Google Scholar] [CrossRef] [PubMed]
  140. Jones, K.R.; Nabinger, S.C.; Lee, S.; Sahu, S.S.; Althouse, S.; Saxena, R.; Johnson, M.S.; Chalasani, N.; Gawrieh, S.; Kota, J. Lower expression of tumor microRNA-26a is associated with higher recurrence in patients with hepatocellular carcinoma undergoing surgical treatment. J. Surg. Oncol. 2018, 118, 431–439. [Google Scholar] [CrossRef] [PubMed]
  141. Yang, L.; Guo, Y.; Liu, X.; Wang, T.; Tong, X.; Lei, K.; Wang, J.; Huang, D.; Xu, Q. The tumor suppressive miR-302c-3p inhibits migration and invasion of hepatocellular carcinoma cells by targeting TRAF4. J. Cancer 2018, 9, 2693–2701. [Google Scholar] [CrossRef]
  142. Chen, Y.; Zhao, Z.X.; Huang, F.; Yuan, X.W.; Deng, L.; Tang, D. MicroRNA-1271 functions as a potential tumor suppressor in hepatitis B virus-associated hepatocellular carcinoma through the AMPK signaling pathway by binding to CCNA1. J. Cell. Physiol. 2018. [Google Scholar] [CrossRef]
  143. Qiu, L.; Wang, T.; Xu, X.; Wu, Y.; Tang, Q.; Chen, K. Long Non-Coding RNAs in Hepatitis B Virus-Related Hepatocellular Carcinoma: Regulation, Functions, and Underlying Mechanisms. Int. J. Mol. Sci. 2017, 18, 2505. [Google Scholar] [CrossRef]
  144. Cui, S.; Qian, Z.; Chen, Y.; Li, L.; Li, P.; Ding, H. Screening of up- and downregulation of circRNAs in HBV-related hepatocellular carcinoma by microarray. Oncol. Lett. 2018, 15, 423–432. [Google Scholar] [CrossRef] [PubMed]
  145. Wang, S.; Cui, S.; Zhao, W.; Qian, Z.; Liu, H.; Chen, Y.; Lv, F.; Ding, H.G. Screening and bioinformatics analysis of circular RNA expression profiles in hepatitis B-related hepatocellular carcinoma. Cancer Biomark. 2018, 22, 631–640. [Google Scholar] [CrossRef]
  146. Matsuda, Y.; Ichida, T.; Genda, T.; Yamagiwa, S.; Aoyagi, Y.; Asakura, H. Loss of p16 contributes to p27 sequestration by cyclin D(1)-cyclin-dependent kinase 4 complexes and poor prognosis in hepatocellular carcinoma. Clin. Cancer Res. 2003, 9, 3389–3396. [Google Scholar] [PubMed]
  147. Pezzuto, F.; Buonaguro, L.; Buonaguro, F.M.; Tornesello, M.L. The Role of Circulating Free DNA and MicroRNA in Non-Invasive Diagnosis of HBV- and HCV-Related Hepatocellular Carcinoma. Int. J. Mol. Sci. 2018, 19, 1007. [Google Scholar] [CrossRef]
  148. Tian, Y.; Ou, J.H. Genetic and epigenetic alterations in hepatitis B virus-associated hepatocellular carcinoma. Virol. Sin. 2015, 30, 85–91. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  149. Wu, S.; Kanda, T.; Imazeki, F.; Arai, M.; Yonemitsu, Y.; Nakamoto, S.; Fujiwara, K.; Fukai, K.; Nomura, F.; Yokosuka, O. Hepatitis B virus e antigen downregulates cytokine production in human hepatoma cell lines. Viral Immunol. 2010, 23, 467–476. [Google Scholar] [CrossRef]
  150. Hadziyannis, E.; Laras, A. Viral Biomarkers in Chronic HBeAg Negative HBV Infection. Genes 2018, 9, 469. [Google Scholar] [CrossRef] [PubMed]
  151. Hildt, E.; Hofschneider, P.H. The PreS2 activators of the hepatitis B virus: Activators of tumour promoter pathways. Recent Results Cancer Res. 1998, 154, 315–329. [Google Scholar] [PubMed]
  152. Liu, P.; Zhang, H.; Liang, X.; Ma, H.; Luan, F.; Wang, B.; Bai, F.; Gao, L.; Ma, C. HBV preS2 promotes the expression of TAZ via miRNA-338-3p to enhance the tumorigenesis of hepatocellular carcinoma. Oncotarget 2015, 6, 29048–29059. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Wang, M.L.; Wu, D.B.; Tao, Y.C.; Chen, L.L.; Liu, C.P.; Chen, E.Q.; Tang, H. The truncated mutant HBsAg expression increases the tumorigenesis of hepatitis B virus by regulating TGF-β/Smad signaling pathway. Virol. J. 2018, 15, 61. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Song, J.; Zhang, X.; Ge, Q.; Yuan, C.; Chu, L.; Liang, H.F.; Liao, Z.; Liu, Q.; Zhang, Z.; Zhang, B. CRISPR/Cas9-mediated knockout of HBsAg inhibits proliferation and tumorigenicity of HBV-positive hepatocellular carcinoma cells. J. Cell. Biochem. 2018, 119, 8419–8431. [Google Scholar] [CrossRef] [PubMed]
  155. Kanda, T.; Wu, S.; Sasaki, R.; Nakamura, M.; Haga, Y.; Jiang, X.; Nakamoto, S.; Yokosuka, O. HBV Core Protein Enhances Cytokine Production. Diseases 2015, 3, 213–220. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Wu, S.; Kanda, T.; Imazeki, F.; Nakamoto, S.; Tanaka, T.; Arai, M.; Roger, T.; Shirasawa, H.; Nomura, F.; Yokosuka, O. Hepatitis B virus e antigen physically associates with receptor-interacting serine/threonine protein kinase 2 and regulates IL-6 gene expression. J. Infect. Dis. 2012, 206, 415–420. [Google Scholar] [CrossRef] [PubMed]
  157. Chen, M.T.; Billaud, J.N.; Sällberg, M.; Guidotti, L.G.; Chisari, F.V.; Jones, J.; Hughes, J.; Milich, D.R. A function of the hepatitis B virus precore protein is to regulate the immune response to the core antigen. Proc. Natl. Acad. Sci. USA 2004, 101, 14913–14918. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Chen, M.; Sällberg, M.; Hughes, J.; Jones, J.; Guidotti, L.G.; Chisari, F.V.; Billaud, J.N.; Milich, D.R. Immune tolerance split between hepatitis B virus precore and core proteins. J. Virol. 2005, 79, 3016–3027. [Google Scholar] [CrossRef]
  159. Yaginuma, K.; Kobayashi, H.; Kobayashi, M.; Morishima, T.; Matsuyama, K.; Koike, K. Multiple integration site of hepatitis B virus DNA in hepatocellular carcinoma and chronic active hepatitis tissues from children. J. Virol. 1987, 61, 1808–1813. [Google Scholar]
  160. Zhou, Y.Z.; Butel, J.S.; Li, P.J.; Finegold, M.J.; Melnick, J.L. Integrated state of subgenomic fragments of hepatitis B virus DNA in hepatocellular carcinoma from mainland China. J. Natl. Cancer Inst. 1987, 79, 223–231. [Google Scholar]
  161. Kim, C.M.; Koike, K.; Saito, I.; Miyamura, T.; Jay, G. HBx gene of hepatitis B virus induces liver cancer in transgenic mice. Nature 1991, 351, 317–320. [Google Scholar] [CrossRef]
  162. Wollersheim, M.; Debelka, U.; Hofschneider, P.H. A transactivating function encoded in the hepatitis B virus X gene is conserved in the integrated state. Oncogene 1988, 3, 545–552. [Google Scholar] [PubMed]
  163. Kekulé, A.S.; Lauer, U.; Weiss, L.; Luber, B.; Hofschneider, P.H. Hepatitis B virus transactivator HBx uses a tumour promoter signalling pathway. Nature 1993, 361, 742–745. [Google Scholar] [CrossRef]
  164. Choi, B.H.; Choi, M.; Jeon, H.Y.; Rho, H.M. Hepatitis B viral X protein overcomes inhibition of E2F1 activity by pRb on the human Rb gene promoter. DNA Cell Biol. 2001, 20, 75–80. [Google Scholar] [CrossRef]
  165. Staib, F.; Hussain, S.P.; Hofseth, L.J.; Wang, X.W.; Harris, C.C. TP53 and liver carcinogenesis. Hum. Mutat. 2003, 21, 201–216. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Cha, M.Y.; Kim, C.M.; Park, Y.M.; Ryu, W.S. Hepatitis B virus X protein is essential for the activation of Wnt/beta-catenin signaling in hepatoma cells. Hepatology 2004, 39, 1683–1693. [Google Scholar] [CrossRef]
  167. Longato, L.; de la Monte, S.; Kuzushita, N.; Horimoto, M.; Rogers, A.B.; Slagle, B.L.; Wands, J.R. Overexpression of insulin receptor substrate-1 and hepatitis Bx genes causes premalignant alterations in the liver. Hepatology 2009, 49, 1935–1943. [Google Scholar] [CrossRef] [PubMed]
  168. Keng, V.W.; Tschida, B.R.; Bell, J.B.; Largaespada, D.A. Modeling hepatitis B virus X-induced hepatocellular carcinoma in mice with the Sleeping Beauty transposon system. Hepatology 2011, 53, 781–790. [Google Scholar] [CrossRef]
  169. Wang, C.; Yang, W.; Yan, H.X.; Luo, T.; Zhang, J.; Tang, L.; Wu, F.Q.; Zhang, H.L.; Yu, L.X.; Zheng, L.Y.; et al. Hepatitis B virus X (HBx) induces tumorigenicity of hepatic progenitor cells in 3,5-diethoxycarbonyl-1,4-dihydrocollidine-treated HBx transgenic mice. Hepatology 2012, 55, 108–120. [Google Scholar] [CrossRef]
  170. Von Olshausen, G.; Quasdorff, M.; Bester, R.; Arzberger, S.; Ko, C.; van de Klundert, M.; Zhang, K.; Odenthal, M.; Ringelhan, M.; Niessen, C.M.; et al. Hepatitis B virus promotes β-catenin-signalling and disassembly of adherens junctions in a Src kinase dependent fashion. Oncotarget 2018, 9, 33947–33960. [Google Scholar] [CrossRef]
  171. Singh, A.K.; Swarnalatha, M.; Kumar, V. c-ETS1 facilitates G1/S-phase transition by up-regulating cyclin E and CDK2 genes and cooperates with hepatitis B virus X protein for their deregulation. J. Biol. Chem. 2011, 286, 21961–21970. [Google Scholar] [CrossRef]
  172. Luo, L.; Chen, S.; Gong, Q.; Luo, N.; Lei, Y.; Guo, J.; He, S. Hepatitis B virus X protein modulates remodelling of minichromosomes related to hepatitis B virus replication in HepG2 cells. Int. J. Mol. Med. 2013, 31, 197–204. [Google Scholar] [CrossRef] [PubMed]
  173. Saeed, U.; Kim, J.; Piracha, Z.Z.; Kwon, H.; Jung, J.; Chwae, Y.J.; Park, S.; Shin, H.J.; Kim, K. Parvulin 14 and parvulin 17 bind to HBx and cccDNA and upregulate HBV replication from cccDNA to virion in a HBx-dependent manner. J. Virol. 2019, 93, e01840-18. [Google Scholar] [CrossRef] [PubMed]
  174. Swarnalatha, M.; Singh, A.K.; Kumar, V. Promoter occupancy of MLL1 histone methyltransferase seems to specify the proliferative and apoptotic functions of E2F1 in a tumour microenvironment. J. Cell Sci. 2013, 126, 4636–4646. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Huang, Y.H.; Tseng, Y.H.; Lin, W.R.; Hung, G.; Chen, T.C.; Wang, T.H.; Lee, W.C.; Yeh, C.T. HBV polymerase overexpression due to large core gene deletion enhances hepatoma cell growth by binding inhibition of microRNA-100. Oncotarget 2016, 7, 9448–9461. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Chung, H.J.; Chen, X.; Yu, Y.; Lee, H.K.; Song, C.H.; Choe, H.; Lee, S.; Kim, H.J.; Hong, S.T. A critical role of hepatitis B virus polymerase in cirrhosis, hepatocellular carcinoma, and steatosis. FEBS Open Bio 2017, 8, 130–145. [Google Scholar] [CrossRef] [Green Version]
  177. Kakumu, S.; Okumura, A.; Ishikawa, T.; Yano, M.; Enomoto, A.; Nishimura, H.; Yoshioka, K.; Yoshika, Y. Serum levels of IL-10, IL-15 and soluble tumour necrosis factor-alpha (TNF-alpha) receptors in type C chronic liver disease. Clin. Exp. Immunol. 1997, 109, 458–463. [Google Scholar] [CrossRef]
  178. Kanda, T.; Yokosuka, O.; Omata, M. Hepatitis C virus and hepatocellular carcinoma. Biology 2013, 2, 304–316. [Google Scholar] [CrossRef]
  179. Braconi, C.; Valeri, N.; Gasparini, P.; Huang, N.; Taccioli, C.; Nuovo, G.; Suzuki, T.; Croce, C.M.; Patel, T. Hepatitis C virus proteins modulate microRNA expression and chemosensitivity in malignant hepatocytes. Clin. Cancer Res. 2010, 16, 957–966. [Google Scholar] [CrossRef] [Green Version]
  180. Zhang, Y.; Wei, W.; Cheng, N.; Wang, K.; Li, B.; Jiang, X.; Sun, S. Hepatitis C virus-induced up-regulation of microRNA-155 promotes hepatocarcinogenesis by activating Wnt signaling. Hepatology 2012, 56, 1631–1640. [Google Scholar] [CrossRef]
  181. Hsu, S.H.; Wang, B.; Kota, J.; Yu, J.; Costinean, S.; Kutay, H.; Yu, L.; Bai, S.; La Perle, K.; Chivukula, R.R.; et al. Essential metabolic, anti-inflammatory, and anti-tumorigenic functions of miR-122 in liver. J. Clin. Investig. 2012, 122, 2871–2883. [Google Scholar] [CrossRef] [Green Version]
  182. Zhao, L.; Li, F.; Taylor, E.W. Can tobacco use promote HCV-induced miR-122 hijacking and hepatocarcinogenesis? Med. Hypotheses 2013, 80, 131–133. [Google Scholar] [CrossRef] [PubMed]
  183. Salvi, A.; Abeni, E.; Portolani, N.; Barlati, S.; De Petro, G. Human hepatocellular carcinoma cell-specific miRNAs reveal the differential expression of miR-24 and miR-27a in cirrhotic/non-cirrhotic HCC. Int. J. Oncol. 2013, 42, 391–402. [Google Scholar] [CrossRef] [PubMed]
  184. Elfimova, N.; Sievers, E.; Eischeid, H.; Kwiecinski, M.; Noetel, A.; Hunt, H.; Becker, D.; Frommolt, P.; Quasdorff, M.; Steffen, H.M.; et al. Control of mitogenic and motogenic pathways by miR-198, diminishing hepatoma cell growth and migration. Biochim. Biophys. Acta 2013, 1833, 1190–1198. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Thomas, M.; Deiters, A. MicroRNA miR-122 as a therapeutic target for oligonucleotides and small molecules. Curr Med Chem. 2013, 20, 3629–3640. [Google Scholar] [CrossRef] [PubMed]
  186. Huang, S.; Xie, Y.; Yang, P.; Chen, P.; Zhang, L. HCV core protein-induced down-regulation of microRNA-152 promoted aberrant proliferation by regulating Wnt1 in HepG2 cells. PLoS ONE 2014, 9, e81730. [Google Scholar] [CrossRef] [PubMed]
  187. Mukherjee, A.; Shrivastava, S.; Bhanja Chowdhury, J.; Ray, R.; Ray, R.B. Transcriptional suppression of miR-181c by hepatitis C virus enhances homeobox A1 expression. J. Virol. 2014, 88, 7929–7940. [Google Scholar] [CrossRef]
  188. Xu, G.; Yang, F.; Ding, C.L.; Wang, J.; Zhao, P.; Wang, W.; Ren, H. MiR-221 accentuates IFN׳s anti-HCV effect by downregulating SOCS1 and SOCS3. Virology 2014, 462–463, 343–350. [Google Scholar] [CrossRef]
  189. Pan, L.; Ren, F.; Rong, M.; Dang, Y.; Luo, Y.; Luo, D.; Chen, G. Correlation between down-expression of miR-431 and clinicopathological significance in HCC tissues. Clin. Transl. Oncol. 2015, 17, 557–563. [Google Scholar] [CrossRef]
  190. El-Abd, N.E.; Fawzy, N.A.; El-Sheikh, S.M.; Soliman, M.E. Circulating miRNA-122, miRNA-199a, and miRNA-16 as Biomarkers for Early Detection of Hepatocellular Carcinoma in Egyptian Patients with Chronic Hepatitis C Virus Infection. Mol. Diagn. Ther. 2015, 19, 213–220. [Google Scholar] [CrossRef]
  191. Devhare, P.B.; Steele, R.; Di Bisceglie, A.M.; Kaplan, D.E.; Ray, R.B. Differential Expression of MicroRNAs in Hepatitis C Virus-Mediated Liver Disease Between African Americans and Caucasians: Implications for Racial Health Disparities. Gene Expr. 2017, 17, 89–98. [Google Scholar] [CrossRef]
  192. Shiu, T.Y.; Shih, Y.L.; Feng, A.C.; Lin, H.H.; Huang, S.M.; Huang, T.Y.; Hsieh, C.B.; Chang, W.K.; Hsieh, T.Y. HCV core inhibits hepatocellular carcinoma cell replicative senescence through downregulating microRNA-138 expression. J. Mol. Med. 2017, 95, 629–639. [Google Scholar] [CrossRef]
  193. Shaker, O.; Alhelf, M.; Morcos, G.; Elsharkawy, A. miRNA-101-1 and miRNA-221 expressions and their polymorphisms as biomarkers for early diagnosis of hepatocellular carcinoma. Infect. Genet. Evol. 2017, 51, 173–181. [Google Scholar] [CrossRef] [PubMed]
  194. Shehata, R.H.; Abdelmoneim, S.S.; Osman, O.A.; Hasanain, A.F.; Osama, A.; Abdelmoneim, S.S.; Toraih, E.A. Deregulation of miR-34a and Its Chaperon Hsp70 in Hepatitis C virus-Induced Liver Cirrhosis and Hepatocellular Carcinoma Patients. Asian Pac. J. Cancer Prev. 2017, 18, 2395–2401. [Google Scholar] [PubMed]
  195. Sur, S.; Sasaki, R.; Devhare, P.; Steele, R.; Ray, R.; Ray, R.B. Association between MicroRNA-373 and Long Noncoding RNA NORAD in Hepatitis C Virus-Infected Hepatocytes Impairs Wee1 Expression for Growth Promotion. J. Virol. 2018, 92, e01215-18. [Google Scholar] [CrossRef] [PubMed]
  196. Rashad, N.M.; El-Shal, A.S.; Shalaby, S.M.; Mohamed, S.Y. Serum miRNA-27a and miRNA-18b as potential predictive biomarkers of hepatitis C virus-associated hepatocellular carcinoma. Mol. Cell. Biochem. 2018, 447, 125–136. [Google Scholar] [CrossRef]
  197. Hou, W.; Bonkovsky, H.L. Non-coding RNAs in hepatitis C-induced hepatocellular carcinoma: Dysregulation and implications for early detection, diagnosis and therapy. World J. Gastroenterol. 2013, 19, 7836–7845. [Google Scholar] [CrossRef] [PubMed]
  198. Zhang, H.; Zhu, C.; Zhao, Y.; Li, M.; Wu, L.; Yang, X.; Wan, X.; Wang, A.; Zhang, M.Q.; Sang, X.; et al. Long non-coding RNA expression profiles of hepatitis C virus-related dysplasia and hepatocellular carcinoma. Oncotarget 2015, 6, 43770–43778. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  199. Kamel, M.M.; Matboli, M.; Sallam, M.; Montasser, I.F.; Saad, A.S.; El-Tawdi, A.H.F. Investigation of long noncoding RNAs expression profile as potential serum biomarkers in patients with hepatocellular carcinoma. Transl. Res. 2016, 168, 134–145. [Google Scholar] [CrossRef] [PubMed]
  200. Fu, N.; Niu, X.; Wang, Y.; Du, H.; Wang, B.; Du, J.; Li, Y.; Wang, R.; Zhang, Y.; Zhao, S.; et al. Role of LncRNA-activated by transforming growth factor beta in the progression of hepatitis C virus-related liver fibrosis. Discov. Med. 2016, 22, 29–42. [Google Scholar]
  201. Kanda, T.; Tada, M.; Imazeki, F.; Yokosuka, O.; Nagao, K.; Saisho, H. 5-aza-2’-deoxycytidine sensitizes hepatoma and pancreatic cancer cell lines. Oncol. Rep. 2005, 14, 975–979. [Google Scholar] [CrossRef]
  202. Ray, R.B.; Lagging, L.M.; Meyer, K.; Ray, R. Hepatitis C virus core protein cooperates with ras and transforms primary rat embryo fibroblasts to tumorigenic phenotype. J. Virol. 1996, 70, 4438–4443. [Google Scholar] [PubMed]
  203. Moriya, K.; Fujie, H.; Shintani, Y.; Yotsuyanagi, H.; Tsutsumi, T.; Ishibashi, K.; Matsuura, Y.; Kimura, S.; Miyamura, T.; Koike, K. The core protein of hepatitis C virus induces hepatocellular carcinoma in transgenic mice. Nat. Med. 1998, 4, 1065–1067. [Google Scholar] [CrossRef]
  204. Kanda, T.; Steele, R.; Ray, R.; Ray, R.B. Hepatitis C virus core protein augments androgen receptor-mediated signaling. J. Virol. 2008, 82, 11066–11072. [Google Scholar] [CrossRef] [PubMed]
  205. Ghosh, A.K.; Majumder, M.; Steele, R.; Meyer, K.; Ray, R.; Ray, R.B. Hepatitis C virus NS5A protein protects against TNF-alpha mediated apoptotic cell death. Virus Res. 2000, 67, 173–178. [Google Scholar] [CrossRef]
  206. Majumder, M.; Ghosh, A.K.; Steele, R.; Ray, R.; Ray, R.B. Hepatitis C virus NS5A physically associates with p53 and regulates p21/waf1 gene expression in a p53-dependent manner. J. Virol. 2001, 75, 1401–1407. [Google Scholar] [CrossRef] [PubMed]
  207. Imazeki, F.; Yokosuka, O.; Fukai, K.; Kanda, T.; Kojima, H.; Saisho, H. Prevalence of diabetes mellitus and insulin resistance in patients with chronic hepatitis C: Comparison with hepatitis B virus-infected and hepatitis C virus-cleared patients. Liver Int. 2008, 28, 355–362. [Google Scholar] [CrossRef] [PubMed]
  208. Neuschwander-Tetri, B.A.; Clark, J.M.; Bass, N.M.; Van Natta, M.L.; Unalp-Arida, A.; Tonascia, J.; Zein, C.O.; Brunt, E.M.; Kleiner, D.E.; McCullough, A.J.; et al. NASH Clinical Research Network. Clinical, laboratory and histological associations in adults with nonalcoholic fatty liver disease. Hepatology 2010, 52, 913–924. [Google Scholar] [CrossRef] [PubMed]
  209. Haga, Y.; Kanda, T.; Sasaki, R.; Nakamura, M.; Nakamoto, S.; Yokosuka, O. Nonalcoholic fatty liver disease and hepatic cirrhosis: Comparison with viral hepatitis-associated steatosis. World J. Gastroenterol. 2015, 21, 12989–12995. [Google Scholar] [CrossRef] [PubMed]
  210. Kawaguchi, T.; Yoshida, T.; Harada, M.; Hisamoto, T.; Nagao, Y.; Ide, T.; Taniguchi, E.; Kumemura, H.; Hanada, S.; Maeyama, M.; et al. Hepatitis C virus down-regulates insulin receptor substrates 1 and 2 through up-regulation of suppressor of cytokine signaling 3. Am. J. Pathol. 2004, 165, 1499–1508. [Google Scholar] [CrossRef]
  211. Miyamoto, H.; Moriishi, K.; Moriya, K.; Murata, S.; Tanaka, K.; Suzuki, T.; Miyamura, T.; Koike, K.; Matsuura, Y. Involvement of the PA28gamma-dependent pathway in insulin resistance induced by hepatitis C virus core protein. J. Virol. 2007, 81, 1727–1735. [Google Scholar] [CrossRef]
  212. Banerjee, S.; Saito, K.; Ait-Goughoulte, M.; Meyer, K.; Ray, R.B.; Ray, R. Hepatitis C virus core protein upregulates serine phosphorylation of insulin receptor substrate-1 and impairs the downstream akt/protein kinase B signaling pathway for insulin resistance. J. Virol. 2008, 82, 2606–2612. [Google Scholar] [CrossRef] [PubMed]
  213. Banerjee, A.; Meyer, K.; Mazumdar, B.; Ray, R.B.; Ray, R. Hepatitis C virus differentially modulates activation of forkhead transcription factors and insulin-induced metabolic gene expression. J. Virol. 2010, 84, 5936–5946. [Google Scholar] [CrossRef] [PubMed]
  214. Bose, S.K.; Shrivastava, S.; Meyer, K.; Ray, R.B.; Ray, R. Hepatitis C virus activates the mTOR/S6K1 signaling pathway in inhibiting IRS-1 function for insulin resistance. J. Virol. 2012, 86, 6315–6322. [Google Scholar] [CrossRef] [PubMed]
  215. Tanaka, S.; Wands, J.R. Insulin receptor substrate 1 overexpression in human hepatocellular carcinoma cells prevents transforming growth factor beta1-induced apoptosis. Cancer Res. 1996, 56, 3391–3394. [Google Scholar] [PubMed]
  216. Qadri, I.; Choudhury, M.; Rahman, S.M.; Knotts, T.A.; Janssen, R.C.; Schaack, J.; Iwahashi, M.; Puljak, L.; Simon, F.R.; Kilic, G.; et al. Increased phosphoenolpyruvate carboxykinase gene expression and steatosis during hepatitis C virus subgenome replication: Role of nonstructural component 5A and CCAAT/enhancer-binding protein β. J. Biol. Chem. 2012, 287, 37340–37351. [Google Scholar] [CrossRef]
  217. Parvaiz, F.; Manzoor, S.; Iqbal, J.; McRae, S.; Javed, F.; Ahmed, Q.L.; Waris, G. Hepatitis C virus nonstructural protein 5A favors upregulation of gluconeogenic and lipogenic gene expression leading towards insulin resistance: A metabolic syndrome. Arch. Virol. 2014, 159, 1017–1025. [Google Scholar] [CrossRef] [PubMed]
  218. Parvaiz, F.; Manzoor, S.; Iqbal, J.; Sarkar-Dutta, M.; Imran, M.; Waris, G. Hepatitis C virus NS5A promotes insulin resistance through IRS-1 serine phosphorylation and increased gluconeogenesis. World J. Gastroenterol. 2015, 21, 12361–12369. [Google Scholar] [CrossRef]
  219. Kwon, Y.C.; Bose, S.K.; Steele, R.; Meyer, K.; Di Bisceglie, A.M.; Ray, R.B.; Ray, R. Promotion of Cancer Stem-Like Cell Properties in Hepatitis C Virus-Infected Hepatocytes. J. Virol. 2015, 89, 11549–11556. [Google Scholar] [CrossRef] [Green Version]
  220. Kwon, Y.C.; Sasaki, R.; Meyer, K.; Ray, R. Hepatitis C Virus Core Protein Modulates Endoglin (CD105) Signaling Pathway for Liver Pathogenesis. J. Virol. 2017, 91, e01235-17. [Google Scholar] [CrossRef]
  221. Uthaya Kumar, D.B.; Chen, C.L.; Liu, J.C.; Feldman, D.E.; Sher, L.S.; French, S.; DiNorcia, J.; French, S.W.; Naini, B.V.; Junrungsee, S.; et al. TLR4 Signaling via NANOG Cooperates with STAT3 to Activate Twist1 and Promote Formation of Tumor-Initiating Stem-Like Cells in Livers of Mice. Gastroenterology 2016, 150, 707–719. [Google Scholar] [CrossRef]
  222. Kudo, M. Systemic Therapy for Hepatocellular Carcinoma: Latest Advances. Cancers 2018, 10, 412. [Google Scholar] [CrossRef] [PubMed]
  223. Okazaki, T.; Honjo, T. PD-1 and PD-1 ligands: From discovery to clinical application. Int. Immunol. 2007, 19, 813–824. [Google Scholar] [CrossRef] [PubMed]
  224. Iwai, Y.; Ishida, M.; Tanaka, Y.; Okazaki, T.; Honjo, T.; Minato, N. Involvement of PD-L1 on tumor cells in the escape from host immune system and tumor immunotherapy by PD-L1 blockade. Proc. Natl. Acad. Sci. USA 2002, 99, 12293–12297. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  225. Leach, D.R.; Krummel, M.F.; Allison, J.P. Enhancement of antitumor immunity by CTLA-4 blockade. Science 1996, 271, 1734–1736. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Occurrence of HCC in natural course of HBV and HCV infection.
Figure 1. Occurrence of HCC in natural course of HBV and HCV infection.
Ijms 20 01358 g001
Figure 2. The p53-RB pathway in HCC.
Figure 2. The p53-RB pathway in HCC.
Ijms 20 01358 g002
Figure 3. Molecular mechanisms of liver cirrhosis and its progression to HCC.
Figure 3. Molecular mechanisms of liver cirrhosis and its progression to HCC.
Ijms 20 01358 g003
Table 1. Representative driver-gene candidates in HCC.
Table 1. Representative driver-gene candidates in HCC.
Gene SymbolGene NamePathwaysReferences
TERTTelomerase reverse transcriptaseTERT[52,53,54,55,56]
CTNNB1Catenin β1β-catenin[52,53]
TP53Tumor protein p53p53–RB[52,53]
ARID2AT-rich interaction domain 2Chromatin remodeling[52]
AXIN1Axin 1β-catenin[52]
TSC2TSC complex subunit 2PI3K–mTOR[52]
RB1Retinoblastoma 1p53–RB[52]
ACVR2AActivin A receptor type 2ASMAD[52]
BRD7Bromodomain containing 7Chromatin remodeling[52]
CDKN1ACyclin-dependent kinase inhibitor 1Aβ-catenin[52]
MEN1Menin 1(MEN1 syndrome)[52]
GALN11Polypeptide N-acetylgalactosaminyltransferase 11NOTCH[52]
FGF19Fibroblast growth factor 19β-catenin[52]
CCND1Cyclin D1p53–RB[52]
ARID1AAT-rich interaction domain 1AChromatin remodeling[52,53,54,57,58]
CDKN2ACyclin-dependent kinase inhibitor 2Ap53–RB[52]
CDKN2BCyclin-dependent kinase inhibitor 2Bp53–RB[52]
RPS6KA3Ribosomal protein S6 kinase, 90kDa, polypeptide 3p53–RB[52]
NFE2L2Nuclear factor, erythroid 2-like 2NRF2–KEAP1[52]
NCOR1Nuclear receptor corepressor 1β-catenin/chromatin remodeling[52]
ADH1BAlcohol dehydrogenase 1B, β polypeptide [52]
SRCAPSnf2-related CREB binding protein activator proteinChromatin remodeling[52]
FCRL1Fc receptor like 1 [52]
PTENPhosphatase and tensin homologPI3K–mTOR[52]
HNRNPA2B1Heterogeneous nuclear ribonucleoprotein A2/B1MAPK[52]
CYP2E1Cytochrome P450 family 2 subfamily E member 1 [52]
MAP2K3Mitogen-activated protein kinase 3MAPK[52]
TSC1Tuberous sclerosis 1mTOR[52]
TMEM99Transmembrane protein 99 [52]
G6PCGlucose-6-phosphatase, catalytic subunitFoxO[52]
Table 2. Hepatic and serum miRs involved in HBV-associated HCC.
Table 2. Hepatic and serum miRs involved in HBV-associated HCC.
miRsUpregulation/DownregulationTarget GenesReferences
Hepatic miRs
miR-223UpregulationStathmin 1 (STMN1)[92]
miR-143UpregulationFibronectin type III domain containing 3B (FNDC3B)[93]
miR-602UpregulationRas-association domain family member 1 (RASSF1A)[94]
miR-224UpregulationN/A[98,110,116]
miR-22Upregulation in male HCC- adjacent tissueEstrogen receptor alpha (ERα)[99]
miR-96UpregulationN/A[112]
miR-183UpregulationN/A[112]
miR-196aUpregulationN/A[112]
miR-545/374aUpregulationEstrogen-related receptor gamma (ESRRG)[120]
miR-331-3pUpregulationInhibitor of growth family member 5 (ING5)[124]
miR-519aUpregulationForkhead box F2 (FOXF2)[124]
miR-106bUpregulationN/A[131]
miR-1269bUpregulationCell division cycle 40 homolog (CDC40)[132]
let-7aDownregulationSignal transducer and activator of transcription 3 (STAT3)[95]
miR-152DownregulationDNA methyltransferase 1 (DNMT1), TNFRF6B[96,115]
miR-145DownregulationCullin 5 (CUL5)[98,125]
miR-199bDownregulationN/A[98]
miR-29cDownregulationTNF-α-induced protein 3 (TNFAIP3)[100]
miR-92DownregulationN/A[102]
miR-338-3pDownregulation3′ UTR region of CCND1[104]
miR-34aDownregulationC-C motif chemokine ligand 22 (CCL22)[115]
miR-101DownregulationDNA methyltransferase (DNMT)3A[116]
miR-122DownregulationPituitary tumor-transforming gene 1 (PTTG1) binding factor (PBF)[117,134,139]
miR-148aDownregulationHematopoietic pre-B cell leukemia transcription factor-interacting protein (HPIP)[108]
miR-22DownregulationCDKN1A[109,134]
let-7cDownregulationN/A[112]
miR-138DownregulationN/A[112]
miR-205DownregulationN/A[113]
miR-101-3pDownregulationRAB GTPase 5A (RAB5A)[114]
miR-429DownregulationNOTCH1[121]
miR-216DownregulationInsulin-like growth factor 2 mRNA-binding protein 2 (IGF2BP2)[122]
miR-34cDownregulationTransforming growth factor-β-induced factor homeobox 2 (TGIF2)[127]
miR-18aDownregulationConnective tissue growth factor (CTGF)[133]
miR-30b-5pDownregulationDNMT3α, USP37[135]
miR-384DownregulationPleiotrophin (PTN)[136]
miR-125a-5pDownregulationerb-b2 receptor tyrosine kinase 3 (ERBB3)[137]
miR-26aDownregulationN/A[140]
miR-302c-3pDownregulationTNF receptor-associated factor 4 (TRAF4)[141]
miR-1271DownregulationCCNA1[142]
Serum miRs
miR-25UpregulationN/A[97]
miR-375UpregulationN/A[97]
let-7fUpregulationN/A[97]
miR-122UpregulationN/A[99]
miR-18aUpregulationN/A[103]
miR-101UpregulationN/A[111]
miR-24-3pUpregulationN/A[117]
miR-545/374aUpregulationERR gamma (ERRG)[120]
miR-96UpregulationN/A[129]
miR-21DownregulationN/A[118]
miR-222DownregulationN/A[118]
miR-29DownregulationN/A[119]
miR-150DownregulationN/A[123]
miR-126DownregulationN/A[128]
miR-125bDownregulationN/A[130]
miR-143/145DownregulationN/A[138]
Table 3. Hepatic and serum miRs involved in HCV-associated HCC.
Table 3. Hepatic and serum miRs involved in HCV-associated HCC.
miRsUpregulation/DownregulationTarget GenesReferences
Hepatic miRs
miR-193bUpregulationSTMN1[179]
miR-155UpregulationN/A[180]
miR-122UpregulationCyclin G1 [181,182,186]
miR-373UpregulationWee1[195]
miR-24DownregulationN/A[183]
miR-27aDownregulationN/A[183]
miR-198DownregulationN/A[184]
miR-152DownregulationN/A[186]
miR-181cDownregulationHomeobox A1 [187]
miR-431DownregulationN/A[189]
miR-138DownregulationTERT[192]
Serum miRs
miR-150UpregulationN/A[191]
miR-221UpregulationN/A[193]
miR-101-1UpregulationN/A[193]
miR-27aUpregulationN/A[194]
miR-221DownregulationSuppressor of cytokine signaling (SOCS)1 and SOCS3[188]
miR-16DownregulationN/A[190]
miR-34aDownregulationHeat-shock protein 70[194]

Share and Cite

MDPI and ACS Style

Kanda, T.; Goto, T.; Hirotsu, Y.; Moriyama, M.; Omata, M. Molecular Mechanisms Driving Progression of Liver Cirrhosis towards Hepatocellular Carcinoma in Chronic Hepatitis B and C Infections: A Review. Int. J. Mol. Sci. 2019, 20, 1358. https://doi.org/10.3390/ijms20061358

AMA Style

Kanda T, Goto T, Hirotsu Y, Moriyama M, Omata M. Molecular Mechanisms Driving Progression of Liver Cirrhosis towards Hepatocellular Carcinoma in Chronic Hepatitis B and C Infections: A Review. International Journal of Molecular Sciences. 2019; 20(6):1358. https://doi.org/10.3390/ijms20061358

Chicago/Turabian Style

Kanda, Tatsuo, Taichiro Goto, Yosuke Hirotsu, Mitsuhiko Moriyama, and Masao Omata. 2019. "Molecular Mechanisms Driving Progression of Liver Cirrhosis towards Hepatocellular Carcinoma in Chronic Hepatitis B and C Infections: A Review" International Journal of Molecular Sciences 20, no. 6: 1358. https://doi.org/10.3390/ijms20061358

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop