Next Article in Journal
Meta-Analysis of Differential Connectivity in Gene Co-Expression Networks in Multiple Sclerosis
Next Article in Special Issue
Dexmedetomidine Protects Rat Liver against Ischemia-Reperfusion Injury Partly by the α2A-Adrenoceptor Subtype and the Mechanism Is Associated with the TLR4/NF-κB Pathway
Previous Article in Journal
Mitochondria Synthesize Melatonin to Ameliorate Its Function and Improve Mice Oocyte’s Quality under in Vitro Conditions
Previous Article in Special Issue
Persistence of HCV in Acutely-Infected Patients Depletes C24-Ceramide and Upregulates Sphingosine and Sphinganine Serum Levels
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Hepatitis B Virus X Protein and Hepatocarcinogenesis

by
Shuaichen Liu
1,2,
Samantha S. Y. Koh
3 and
Caroline G. L. Lee
1,3,4,5,*
1
Department of Biochemistry, Yong Loo Lin School of Medicine, National University of Singapore, 117597 Singapore, Singapore
2
Department of Hepatobiliary & Pancreas Surgery, The First Hospital, Jilin University, Changchun 130021, China
3
Division of Medical Sciences, Humphrey Oei Institute of Cancer Research, National Cancer Centre Singapore, 169610 Singapore, Singapore
4
NUS Graduate School for Integrative Sciences and Engineering, National University of Singapore, 117456 Singapore, Singapore
5
Duke-NUS Graduate Medical School, 169857 Singapore, Singapore
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2016, 17(6), 940; https://doi.org/10.3390/ijms17060940
Submission received: 3 May 2016 / Revised: 18 May 2016 / Accepted: 24 May 2016 / Published: 14 June 2016
(This article belongs to the Special Issue Molecular Mechanisms of Human Liver Diseases)

Abstract

:
Chronic hepatitis B virus (HBV) infection is one of the most associated factors in hepatocarcinogenesis. HBV is able to integrate into the host genome and encode the multi-functional hepatitis B virus x protein (HBx). Although the mechanism between HBx and carcinogenesis is still elusive, recent studies have shown that HBx was able to influence various signaling pathways, as well as epigenetic and genetic processes. This review will examine and summarize recent literature about HBx’s role in these various processes.

1. Introduction

Hepatocellular carcinoma (HCC) is one of the most prevalent malignancies in the world, accounting for 70%–90% of primary liver cancer [1]. By 2012, it became the fifth most common cause of cancer in men and ninth in women, with approximately 554,400 and 228,100 new cases, respectively. Liver cancer ranked second (521,000) in men, and sixth (224,500) in women, in estimated mortality [2]. A recently published report from the Liver Cancer Study Group of Japan (LCSGJ), which included over 20,000 HCC patients from 1996–2007, showed that with improved HCC surveillance and development of new treatment strategies, the three-, five-, and 10-year cumulative survival rates for HCC patients were 62.1%, 44.3%, and 20.5%, respectively [3].
Risk factors that have been associated with the development of HCC are excessive alcohol consumption, hepatitis B/C virus infection, exposure to aflatoxin, and patient factors such as non-alcoholic fatty liver disease (NAFLD), type 2 diabetes, and gender. About 54.4% of HCC cases are attributed to hepatitis B virus (HBV) infection [4], hence drawing attention to research investigating the underlying mechanism.
HBV, with a length of 3.2 kb, is a member of the hepadnaviridae family. It is an enveloped, partially double-stranded DNA virus that replicates its genome by reverse transcription. There are four overlapping open reading frames (ORFs) which encode seven proteins (pre-S1, pre-S2, S, pre-C, C, viral polymerase, HBx protein) and four regulatory elements (enhancer II/basal core promoter, preS1 promoter, preS2/S promoter, and enhancer I/X promoter) (Figure 1) [5,6].
Hepatitis B virus x (HBx) is a highly-conserved 154-amino acid protein with a molecular mass of approximately 17 kDa. HBx is the designated name for both the gene and protein because the amino acid sequence is not homologous to any known protein [7]. It is mostly localized in the cytoplasm and, to a lesser extent, the nucleus of hepatocytes [8]. As HBV DNA can integrate into the host genome, the HBx gene can be maintained and transcribed in human HCC cells even in the absence of complete HBV replication [9,10]. HBx protein does not directly bind to DNA but activates various viral and cellular promoters and enhancers [8]. In this review, we will discuss the mechanism of HBx protein in hepatocarcinogenesis in detail.

2. Mechanism of HBx Protein

HBx has been reported to modulate the expression and activities of numerous genes, as well as epigenetic molecules (e.g., miRNAs and lncRNAs) and events (e.g., methylation and Acetylation), leading to the deregulation of various pathways and function (Figure 2 and Table 1) as discussed below.

3. Pathways

3.1. Signal Pathways

Signaling pathways are widely engaged in carcinogenesis. As an important oncoprotein, HBx interacts with several signaling pathways during HCC development. Early studies mainly focus on the expression of HBx relative to key molecules in different pathways. With the advancements in research, the mechanisms behind these expression changes have been further illuminated.
The Notch signaling pathway is highly conserved in evolution and can affect cell proliferation, differentiation, and apoptosis. Previous studies have demonstrated that HBx can activate the Notch pathway to induce hepatocarcinogenesis, but did not specify which of the four transmembrane receptors, Notch1 to Notch4, were involved. One study aimed to establish the link between HBx and Notch pathway by observing the effect γ-secretase inhibitor had in HBx infected cells. However, all four receptors in the Notch pathway were blocked so it was impossible to ascertain which receptors interacted with HBx. To find out if HBx acts on a single receptor to activate the Notch pathway, another study used shRNA to inhibit Notch1 and found significant suppression in the growth of L02 cells [58]. In HepG2X cells, HBx is found to upregulate the expression of both Notch1 and 4, which indicates Notch4 is also involved in HBx induced Notch signaling activation [11].
The PI3K/mTOR pathway is also vital in hepatocarcinogenesis. The activation of this pathway upregulates Ras, Scr, and CXCR4 expression resulting in the promotion of progression, invasion, and metastasis of cancer cells. A recent study revealed that upregulated α-fetoprotein (AFP) can bind to phosphatase and tensin homologue (PTEN) and attenuate its inhibition on PI3K/mTOR pathway. This upregulation has been shown to be induced by HBx [12].
Secreted frizzled-related proteins (SFRPs) are a family of extracellular glycoproteins that can bind to Wnt ligands and antagonize the Wnt/β-catenin signaling pathway—a pathway involved in cell proliferation, differentiation and metastasis. The reduction in expression of SFRP1 and SFRP5 in hepatoma cells induced by HBx, correlates with the recruitment of DNA methyltransferase 1 (DNMT1) and DNMT3 to their promoters and subsequent hypermethylation. This regression of SFRPs reduces the antagonistic effect on Wnt pathway, resulting in epithelial mesenchymal transition (EMT) and contributing to hepatocarcinogenesis [13].

3.2. DNA Repair

DNA damage is one of the key mechanisms driving HCC carcinogenesis [59]. HBx can hamper DNA repair to promote HCC carcinogenesis through its interaction with transcription factor IIH (TFIIH), a multiprotein complex of 10 polypeptides, which is an integral component of the DNA repair pathway. HBx interferes with TFIIH’s function, increasing ultraviolet (UV) sensitivity and reducing DNA repair capacity [14]. There is a striking similarity between the predicted structure of the HBx protein and the central domain of human thymine DNA glycosylase (TDG), a key enzyme in the base excision repair (BER) pathway [15]. This similarity enables HBx to inhibit the BER pathway, possibly by interfering with a downstream substance [60]. Human 8-oxoguanine DNA glycosylase 1 (hOGG1) and DNA glycosylase a (hMYHa) are two DNA repair enzymes for oxidative DNA damage. There was a slight, but insignificant, reduction of both hOGG1 and hMYHa at the transcription level and a significant reduction of hMYHa at an mRNA level in HBx-infected cells, indicating that HBx can hinder the DNA repair process in hepatic cells [16].

3.3. Oxidative Stress

The intracellular redox state and the generation of reactive oxygen species (ROS) can lead to a series of cell damage, including dysfunction, mutations, and toxicity [61]. NAD(P) H:quinone oxidoreductase 1 (NQO1) is a phase II enzyme that participates in the detoxification of dopamine-derived quinone molecules and ROS. HBx can induce mitochondrial injury and oxidative stress by transcriptional downregulation of NQO1 [17]. HBx induced ROS can also cause DNA damage that is more severe in mitochondrial DNA (mtDNA) than in nucleus DNA, thought to bring about mutation and carcinogenensis. This effect is related to the C-terminal truncation of HBx [18]. Conversely, HBx is also reported to exert anti-oxidative stress functions through the upregulation of Forkhead box class O 4 (Foxo4), an activator of pyruvate dehydrogenase kinase 4 and of Mn superoxide dismutase (SOD) [19].

3.4. Immune System

The evasion of cellular innate immunity is a protective factor for HBV survival in infected hepatocytes. Toll-like receptor 3 (TLR3) is an important pattern recognition receptor (PRR) in innate immunity while TIR-domain-containing adaptor inducing interferon-β (TRIF) is its adaptor. HBx can reduce TRIF expression in a dose and proteasome dependent manner, thus attenuating the TLR3 signaling pathway. This will promote HBV replication and enable the virus to evade innate immunity [20]. In early HBV infection, there is no activation of IFN-β, an innate immunity activation indicator. β Interferon promoter stimulator 1 (IPS-1) is another PRR adaptor, which can recruit kinase, required for IFN-β production. HBx can bind to IPS-1 and diminish IFN-β signaling. This is another possible way for HBV to evade the immune system [21]. However, this evasion does not occur in all infected hepatocytes. In acute infection or in the early stages of liver disease, HBV-specific cytotoxic T lymphocytes (CTLs) can still eliminate HBV. The accumulation of chemokine attractive immune cells, like T cells, precursor B cells, monocytes, and neutrophils, have been shown to damage the liver. One such chemokine involved is stromal cell-derived factor-1 (SDF-1), which is induced by HBx via endoplasmic reticulum stress. The recruited immune cells in the liver will engender a more severe chronic hepatitis [22].

3.5. Apoptosis

HBx-induced apoptosis contributes to HCC carcinogenesis in two contradictory ways, depending on the factors HBx interacts with. Under certain conditions, multiple mechanisms are involved in HBx-induced apoptosis.

3.5.1. Pro-Apoptosis

TNF-α related apoptosis inducing ligand (TRAIL) is a member of the tumor necrosis factor (TNF) superfamily. The function of TRAIL is to induce cell apoptosis, mainly in tumor cells. HBx can induce the expression of TRAIL-R2 (DR5), one of two death receptors, and promote TRAIL-induced apoptosis. However, the NF-κB pathway is also responsible for the upregulation of DR5 [23]. A20 is an E3 ubiquitin ligase, which can modulate caspase-8 ubiquitination and inhibit its activity. Caspase-8 is a well-known TRAIL signaling mediator, the suppression of which will in turn attenuate TRAIL induced apoptosis. HBx can inhibit the expression of A20, through the upregulation of miR-125a, which would decrease the inhibition of caspase-8, thus sensitizing hepatocytes to TRAIL-induced apoptosis [24].

3.5.2. Anti-Apoptosis

B-cell lymphoma 2 (Bcl-2) and myeloid cell leukemia-1 (Mcl-1) are two apoptosis inhibition genes. HBx is found to increase Bcl-2 and Mcl-1 expression and downregulate their counterpart, Bcl-2-associated X protein (Bax), in hepatic progenitor cells. A decrease in caspase-9 and caspase-3, and an increase in β-catenin, were also observed. This shows that HBx inhibits apoptosis in hepatic progenitor cells. Considering all its effects, HBx has both an apoptosis and an anti-apoptosis function [25].

4. Epigenetics

Epigenetic changes, which include DNA methylation, histone acetylation, and non-coding RNA (ncRNA) expression, have been extensively studied in recent years [62]. These changes can affect coding and non-coding gene expression, due to chromosome remodeling [63]. Epigenetic changes can be induced by many factors, such as pathogens, chemicals, UV light, etc. In HBV-induced HCC, HBx is regarded as the most important factor [64].

4.1. Methylation

The overexpression of insulin-like growth factor-2 (IGF-2) is observed in HBV-infected HCC patients but not in those who are uninfected [65]. HBx can directly interact with methyl-CpG binding domain protein 2 (MBD2), CREB-binding protein (CBP), and its paralog p300, to form the MBD2-HBx-CBP/p300 complex. This complex recruits to IGF-2 P3 and P4 promoters, which induces hypomethylation and transcriptional activation of the P3 and P4 promoters [26,66]. HBx can inhibit the expression of CD82, a tumor suppressor gene that prevents metastasis, through hypermethylation of CD82 promoter. This effect can be reversed by 5‑aza‑CdR, a methyl enzyme inhibitor [27]. Similar results are observed in protocadherin 10 (PCDH10), caveolin-1, and metastasis-associated protein 1 (MTA1) [28,29,30]. Some normally-hypermethylated intragenic CpG islands (mCGIs) have been observed to be hypomethylation before the development of HCC in HBx transgenic mouse liver. Most of these mCGIs are located in the 5′ end of the transcripts and play important roles in tumorigenesis. HBx causes this hypomethylation by binding to the promoters of Dnmt3A and Dnmt3L and recruiting HDAC1 to downregulate their expressions, thus, influencing cell differentiation [31]. Another study conducted in the HBx stably-expressed cell line showed the opposite result that HBx upregulated Dnmt3A/3B expression at both mRNA and protein levels. This upregulation led to promoter CpG island methylation of the suppressors of cytokine signaling‑1 (SOCS‑1). SOCS-1 is a negative regulator of the Janus kinase (JAK)/signal transducer and activator of transcription (STAT) signaling pathway and functions as a tumor suppressor. The promoter CpG island methylation of SOCS-1 led to SOCS-1 expression decrease, which might be a possible mechanism of HBx-induced carcinogenesis [32]. HBx can also recruit DNMT1 to the promoter regions of Ras association domain family 1 isoform A (RASSF1A), a tumor suppressor gene, and then induce methylation of SP1, a specific transcription factor binding site, to suppress RASSF1A expression [33].

4.2. Acetylation

Insulin-like growth factor binding protein 3 (IGFBP-3) gene can inhibit the mitogenic and anti-apoptosis processes in cells by binding to IGF-1. HBx has been proven to recruit HDAC1 to IGFBP-3 promoter to form a SP1/HDAC1 complex. This complex can cause the deacetylation of SP1 and reduce IGFBP-3 transcription [34]. E-cadherin gene (CDH1) is a metastasis promoting gene, while snail-1 is its repressor. HBx can inhibit CDH1 expression via recruitment of mSin3A/HDAC complex to the snail-1 binding site in the CDH1 promoter region and cause deacetylation. This exerts the same effect as direct snail-1 binding [35]. HBx can attenuate the interaction between NAD-dependent deacetylase sirtuin-1 (SIRT1)—a member of the class III histone deacetylases, which has controversial roles in tumor formation—and β-catenin, thus protecting β-catenin from degradation [36].

4.3. Non-Coding RNAs

Not all RNAs transcribed from the genome are coding RNAs. Recent RNA-seq data showed that more than 90% of the human genome is transcribed into ncRNAs. MicroRNAs (miRNAs) and long non-coding RNAs (lncRNAs) are two important members of the ncRNA family and exert their functions on the transcriptional and post-transcriptional levels. Both of them play crucial roles in human cancer [67]. In this section, we summarized recent studies about HBx interaction with miRNAs and lncRNAs.

4.3.1. miRNAs

HBx is able to regulate miRNA expression and influence miRNA-related targets. HBx can downregulate miR-145, which releases Cullin-5 (CUL5) from the inhibitory effect of miR-145, influencing the cell cycle [37]. Mammary serine protease inhibitor (maspin) can reduce integrin expression and, thus, increase cellular adherence to fibronectin, which results in less invasion. HBx increases the levels of miR-7, -21, and -107, which can directly target and inhibit maspin expression, resulting in the promotion of HCC migration, invasion, anoikis resistance, and chemoresistance [38]. HBx can also increase miR-146a expression, which targets and causes the downregulation of complement factor H (CFH), a protein which negatively regulates the alternative pathway of complement activation. These are key interactions in the development of HBV-induced hepatitis [39]. HBx is able to inhibit miR-216b transcription by reducing the p53 recruitment to miR-216b promoter in vitro and in vivo. Since miR-216b inhibits cell proliferation through regulating insulin-like growth factor 2 mRNA-binding protein 2 (IGF2BP2), the inhibition of miR-216b enables IGF2BP2 to activate insulin-like growth factor 2 (IGF2), resulting in the activation of the IGFBP2-IGF2-MAKP/PI3K pathway, which plays a role in tumorigenesis. The HBx-induced p53 recruitment to miR-216b promoter is also associated with the hypomethylation of IGF2 promoter [40]. Another mRNA affected by HBx is MiR-221. MiR-221 is upregulated by HBx and can directly target estrogen receptor-α (ERα), which is thought to be a protective factor against HCC. The downregulation of ERα will promote cell proliferation [41]. HBx can downregulate miR-375 and miR-136, which target astrocyte elevated gene-1 (AEG-1)—an oncoprotein that is overexpressed in HCC—and, thus, promotes cell migration [42]. HBx can also upregulate miR-21, which inhibits programmed cell death protein-4 (PDCD4) and PTEN, which finally results in increased cell proliferation [43]. Lipid metabolic disorder may lead to HCC [68]. Acyl-CoA synthetase long-chain family member (ACSL) catalyzes lipid metabolism and has five isoforms. Two of them, ACSL1 and ACSL4, were proven to be the target of miR-205. HBx can downregulate miR-205, resulting in upregulation of ACSL1/4. The acceleration of lipogenesis by ACSL1/4 might be a new insight in hepatocarcinogenesis [44,45]. Another study attempted to establish the relationship between HBx and miRNA expression from a broader perspective. Among the genes that are important in miRNA processing, DiGeorge critical region 8 (DGCR8) was found to be downregulated by HBx, through the HBx-induced inhibition of transcription factor Yin Yang 1 (YY1), which resulted in the inhibition of DGCR8 promoter activity. These findings indicate that HBx may influence the upstream of miRNA expression. In addition to HBx, HBs is also found to induce the same effect [46].

4.3.2. LncRNAs

The function of lncRNA is slightly more complicated than that of miRNA. Nevertheless, HBx can still interact with some lncRNAs and affect hepatocarcinogenesis. At present, there are a few lncRNAs that are known to play a role in hepatocarcinogenesis. DBH-AS1 is a ~2 kb lncRNA that can promote cell proliferation in vitro and in vivo. It has been shown to accelerate G1/S and G2/M transition. HBx can markedly induce DBH-AS1 expression and activate the ERK/p38/JNK MAPK signaling pathway [47]. Highly upregulated in liver cancer (HULC), as its name suggests, is a dramatically upregulated lncRNA in HCC patients. This HBx- induced upregulation promotes cell proliferation through the downregulation of tumor suppression gene p18 [48]. HBx downregulated lncRNA (Dreh) was identified and found to be downregulated in the livers of HBx transgenic mice. Its human ortholog gene, hDREH, was also non-coding and downregulated in HCC tissue. Vimentin is a protein upregulated in cells experiencing EMT. Dreh can inhibit HCC metastasis by combining and inhibiting vimentin expression which, in the end, changes the normal cytoskeleton structure [49]. A recent study showed that an RNA transcribed by HBx fusion with long interspersed nuclear elements (LINE) can function as an lncRNA. LINEs are retrotranspons that make up 17% of the human genome. Researchers have detected a co-transcribed LINE sequence, identified as LINE1, from chr.8p11 with HBx insertion in four HCC cell lines. This transcript induces EMT β-catenin nuclear translocation, not by encoding protein, but by the hybrid RNA which functions as an lncRNA. These findings present a promising therapeutic target, but more detailed studies are required [50].

5. Genetic

5.1. Integration

In HCC patients, HBV is often observed to be integrated into the host genome. A very interesting observation is that the 3’ end of the HBx gene, which is the site for viral replication/transcription initiation, was also found to often be the site of integration of HBV into the human genome [69]. Notably, upon integration, the 3’ end of the HBx protein is often deleted and fused with human sequences [69]. These chimeric transcripts have been shown to be capable of being translated into HBx-human fusion protein [69]. It would, thus, be important to elucidate the functional significance of these HBx-human fusion proteins in tumorigenesis as these are more often observed in HCC patients than even the truncated HBx protein which has been characterized.

5.2. HBx Gene Mutation

In HCC patients, the HBV genome is highly mutated with significant structural alterations, including duplications, deletions, and inversions, which differ between the tumor and adjacent non-tumorous tissues [69], as well as nucleotide mutations that resulted in amino acid substitutions [69]. Significant non-conservative substitutions were observed in the HBx gene suggesting that these mutations are likely to change the property/structure of the HBx protein, hence, influencing the tumorigenic potential or facilitating escape of the HBx protein from host immune surveillance [69]. Truncated HBx proteins were also commonly observed [70,71]. The role of the various mutated forms of HBx, as well as the chimeric HBx-human proteins in hepatocarcinogenesis, still remains unclear.
Overgrowth of tumour cells has been seen in mouse models with ectopic expression of truncated HBx [52,72]. In humans, an examination of 50 HCC tumor tissue samples shows that 23 (46%) of them have C-terminal truncation. A stable expression of truncated HBx protein (with 24 amino acid deleted in the C-terminal) increases the invasive ability of HepG2 cells, by activating matrix metalloproteinase 10 (MMP10) [51]. Another study, which used immunohistochemistry, found that 88 out of 111 (79.3%) HCC tissue samples have truncated HBx. Truncation leads to faster proliferation and lower apoptotic frequency in vitro and in vivo [52]. The degree of truncation is also important. One study indicates that 14 and 23 amino acids truncation of C-terminal will lower the steady levels of HBx. The NF-κB activity and HBV replication are also much lower than full length HBx in HepG2 cells [53]. Yet another study demonstrates that deletion within 14 amino acids of HBx will have no effect on its transactivation property [73]. The C-terminal is vital for the subcellular location of HBx. The 44 (111–154) amino acids of the C-terminal are sufficient for mitochondrial targeting. In the case of truncation, seven amino acids (111–117) are sufficient and have been known to be the mitochondrial targeting sequence (MTS) for HBx, with 115 cysteine being the key amino acid [54]. Residual cancer stem cells (CSC) are thought to be the cause of tumor relapse after chemotherapy typifies HCC. CD133 positive subset of CSC is able to initiate tumor and resist chemotherapy. C-terminal truncated HBx, at the breakpoints 140 and 119 amino acid, is able to enhance the stemness properties and drug resistant ability of the CD133 positive cell [55].
A10R-S144R double point mutation within the HBx gene was found between central and para-tumor isolates. This mutation arrests the cell cycle and has a lower p53 binding ability compared to wild type (WT)-HBx, which decreases p53 dependent apoptosis [74]. In contrast, K130M/V131I mutation can increase p53 binding ability of HBx. This dual mutation also strengthens the transcriptional activity of hypoxia-inducible factor-1 (HIF-1), which increases in a variety of human tumors. A more potent secondary structure brought by this dual mutation is thought to be the mechanism behind the interaction with HIF-1a [56]. A recent study aimed to link HBx point mutation with HCC survival in Chinese populations. They used the Cox proportional hazard model and found eight mutational sites as independent predictors. The nucleotide sites in HBV are 1383, 1461, 1485, 1544, 1613, 1653, 1719, and 1753. However, these results have yet to be tested in other populations or achieve experimental validation [57]. Sequencing of 44 HBV-associated HCC samples shows that there is a high mutation frequency (13 in 44) at L30F/S144A, but the role of this mutation in hepatocarcinogenesis is yet to be investigated [75].

6. Conclusions and Prospects

The majority of HCC cases in Asia are associated with HBV infection. Due to its high morbidity and mortality worldwide, there has been intense research over the years on HCC carcinogenesis to elucidate the molecular mechanisms to facilitate the design of better strategies to treat this cancer. HBx, encoded by HBV, is believed to play important roles in hepatocarcinogenesis through both direct and indirect mechanisms. With the development of next generation technologies, we may be able to probe deeper into how the various mutated forms of HBx observed in HCC patients can modulate hepatocarcinogenesis. This may then lead to better therapeutic strategies to manage HBV-associated HCC patients.

Acknowledgments

This work was supported by a grant from the National Medical Research Council (NMRC) (NMRC/CBRG/0095/2015 and NMRC/1238/2009), as well as block funding from National Cancer Center, Singapore and Duke-National University of Singapore (Duke-NUS) Graduate Medical School to A/P Caroline G. L. Lee.

Author Contributions

Conceived and designed the draft: Shuaichen Liu and Caroline G. L. Lee; edited manuscript: Samantha S. Y. Koh; Shuaichen Liu, Caroline G. L. Lee.

Conflicts of Interest

The authors declare no conflict of interest. The funding sponsors had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, and in the decision to publish the results.

References

  1. Torre, L.A.; Siegel, R.L.; Ward, E.M.; Jemal, A. Global cancer incidence and mortality rates and trends—An update. Cancer Epidemiol. Biomark. Prev. 2015. [Google Scholar] [CrossRef] [PubMed]
  2. Torre, L.A.; Bray, F.; Siegel, R.L.; Ferlay, J.; Lortet-Tieulent, J.; Jemal, A. Global cancer statistics, 2012. CA Cancer J. Clin. 2015, 65, 87–108. [Google Scholar] [CrossRef] [PubMed]
  3. Kudo, M.; Izumi, N.; Ichida, T.; Ku, Y.; Kokudo, N.; Sakamoto, M.; Takayama, T.; Nakashima, O.; Matsui, O.; Matsuyama, Y. Report of the 19th follow-up survey of primary liver cancer in Japan. Hepatol. Res. 2016, 46, 372–390. [Google Scholar] [CrossRef] [PubMed]
  4. Parkin, D.M. The global health burden of infection-associated cancers in the year 2002. Int. J. Cancer 2006, 118, 3030–3044. [Google Scholar] [CrossRef] [PubMed]
  5. Seeger, C.; Mason, W.S. Hepatitis B virus biology. Microbiol. Mol. Biol. Rev. 2000, 64, 51–68. [Google Scholar] [CrossRef] [PubMed]
  6. Minor, M.M.; Slagle, B.L. Hepatitis B virus HBx protein interactions with the ubiquitin proteasome system. Viruses 2014, 6, 4683–4702. [Google Scholar] [CrossRef] [PubMed]
  7. Ali, A.; Abdel-Hafiz, H.; Suhail, M.; Al-Mars, A.; Zakaria, M.K.; Fatima, K.; Ahmad, S.; Azhar, E.; Chaudhary, A.; Qadri, I. Hepatitis B virus, HBx mutants and their role in hepatocellular carcinoma. World J. Gastroenterol. 2014, 20, 10238–10248. [Google Scholar] [CrossRef] [PubMed]
  8. Zhang, X.D.; Wang, Y.; Ye, L.H. Hepatitis B virus X protein accelerates the development of hepatoma. Cancer Biol. Med. 2014, 11, 182–190. [Google Scholar] [PubMed]
  9. Peng, Z.; Zhang, Y.; Gu, W.; Wang, Z.; Li, D.; Zhang, F.; Qiu, G.; Xie, K. Integration of the hepatitis B virus X fragment in hepatocellular carcinoma and its effects on the expression of multiple molecules: A key to the cell cycle and apoptosis. Int. J. Oncol. 2005, 26, 467–473. [Google Scholar] [CrossRef] [PubMed]
  10. Hwang, G.Y.; Lin, C.Y.; Huang, L.M.; Wang, Y.H.; Wang, J.C.; Hsu, C.T.; Yang, S.S.; Wu, C.C. Detection of the hepatitis B virus X protein (HBx) antigen and anti-HBx antibodies in cases of human hepatocellular carcinoma. J. Clin. Microbiol. 2003, 41, 5598–5603. [Google Scholar] [CrossRef] [PubMed]
  11. Gao, J.; Xiong, Y.; Wang, Y.; Wang, Y.; Zheng, G.; Xu, H. Hepatitis B virus X protein activates notch signaling by its effects on Notch1 and Notch4 in human hepatocellular carcinoma. Int. J. Oncol. 2016, 48, 329–337. [Google Scholar] [CrossRef] [PubMed]
  12. Zhu, M.; Guo, J.; Li, W.; Lu, Y.; Fu, S.; Xie, X.; Xia, H.; Dong, X.; Chen, Y.; Quan, M.; et al. Hepatitis B virus X protein induces expression of α-fetoprotein and activates PI3K/mTOR signaling pathway in liver cells. Oncotarget 2015, 6, 12196–12208. [Google Scholar] [CrossRef] [PubMed]
  13. Xie, Q.; Chen, L.; Shan, X.; Shan, X.; Tang, J.; Zhou, F.; Chen, Q.; Quan, H.; Nie, D.; Zhang, W.; et al. Epigenetic silencing of SFRP1 and SFRP5 by hepatitis B virus X protein enhances hepatoma cell tumorigenicity through Wnt signaling pathway. Int. J. Cancer 2014, 135, 635–646. [Google Scholar] [CrossRef] [PubMed]
  14. Qadri, I.; Fatima, K.; Abde, L.H.H. Hepatitis B virus X protein impedes the DNA repair via its association with transcription factor, TFIIH. BMC Microbiol. 2011, 11, 48. [Google Scholar] [CrossRef] [PubMed]
  15. Van Hemert, F.J.; van de Klundert, M.A.; Lukashov, V.V.; Kootstra, N.A.; Berkhout, B.; Zaaijer, H.L. Protein x of hepatitis B virus: Origin and structure similarity with the central domain of DNA glycosylase. PLoS ONE 2011, 6, e23392. [Google Scholar] [CrossRef] [PubMed]
  16. Cheng, B.; Zheng, Y.; Guo, X.; Wang, Y.; Liu, C. Hepatitis B viral x protein alters the biological features and expressions of DNA repair enzymes in LO2 cells. Liver Int. Off. J. Int. Assoc. Study Liver 2010, 30, 319–326. [Google Scholar] [CrossRef] [PubMed]
  17. Wu, Y.L.; Wang, D.; Peng, X.E.; Chen, Y.L.; Zheng, D.L.; Chen, W.N.; Lin, X. Epigenetic silencing of NAD(P)H: Quinone oxidoreductase 1 by hepatitis B virus X protein increases mitochondrial injury and cellular susceptibility to oxidative stress in hepatoma cells. Free Radic. Biol. Med. 2013, 65, 632–644. [Google Scholar] [CrossRef] [PubMed]
  18. Jung, S.Y.; Kim, Y.J. C-terminal region of hbx is crucial for mitochondrial DNA damage. Cancer Lett. 2013, 331, 76–83. [Google Scholar] [CrossRef] [PubMed]
  19. Srisuttee, R.; Koh, S.S.; Park, E.H.; Cho, I.R.; Min, H.J.; Jhun, B.H.; Yu, D.Y.; Park, S.; Park do, Y.; Lee, M.O.; et al. Up-regulation of Foxo4 mediated by hepatitis B virus X protein confers resistance to oxidative stress-induced cell death. Int. J. Mol. Med. 2011, 28, 255–260. [Google Scholar] [PubMed]
  20. Hong, Y.; Zhou, L.; Xie, H.; Zheng, S. Innate immune evasion by hepatitis B virus-mediated downregulation of TRIF. Biochem. Biophys. Res. Commun. 2015, 463, 719–725. [Google Scholar] [CrossRef] [PubMed]
  21. Kumar, M.; Jung, S.Y.; Hodgson, A.J.; Madden, C.R.; Qin, J.; Slagle, B.L. Hepatitis B virus regulatory HBx protein binds to adaptor protein IPS-1 and inhibits the activation of β interferon. J. Virol. 2011, 85, 987–995. [Google Scholar] [CrossRef] [PubMed]
  22. Cho, H.K.; Kim, S.Y.; Seong, J.K.; Cheong, J. Hepatitis B virus X increases immune cell recruitment by induction of chemokine SDF-1. FEBS Lett. 2014, 588, 733–739. [Google Scholar] [CrossRef] [PubMed]
  23. Kong, F.; You, H.; Zhao, J.; Liu, W.; Hu, L.; Luo, W.; Hu, W.; Tang, R.; Zheng, K. The enhanced expression of death receptor 5 (DR5) mediated by HBV x protein through NF-κB pathway is associated with cell apoptosis induced by (TNF-α related apoptosis inducing ligand) trail in hepatoma cells. Virol. J. 2015, 12, 192. [Google Scholar] [CrossRef] [PubMed]
  24. Zhang, H.; Huang, C.; Wang, Y.; Lu, Z.; Zhuang, N.; Zhao, D.; He, J.; Shi, L. Hepatitis B virus X protein sensitizes trail-induced hepatocyte apoptosis by inhibiting the E3 ubiquitin ligase A20. PLoS ONE 2015, 10, e0127329. [Google Scholar] [CrossRef] [PubMed]
  25. Shen, L.; Zhang, X.; Hu, D.; Feng, T.; Li, H.; Lu, Y.; Huang, J. Hepatitis B virus X (HBx) play an anti-apoptosis role in hepatic progenitor cells by activating Wnt/β-catenin pathway. Mol. Cell. Biochem. 2013, 383, 213–222. [Google Scholar] [CrossRef] [PubMed]
  26. Liu, X.Y.; Tang, S.H.; Wu, S.L.; Luo, Y.H.; Cao, M.R.; Zhou, H.K.; Jiang, X.W.; Shu, J.C.; Bie, C.Q.; Huang, S.M.; et al. Epigenetic modulation of insulin-like growth factor-II overexpression by hepatitis B virus X protein in hepatocellular carcinoma. Am. J. Cancer Res. 2015, 5, 956–978. [Google Scholar] [PubMed]
  27. Yu, G.; Bing, Y.; Li, W.; Xia, L.; Liu, Z. Hepatitis B virus inhibits the expression of CD82 through hypermethylation of its promoter in hepatoma cells. Mol. Med. Rep. 2014, 10, 2580–2586. [Google Scholar] [PubMed]
  28. Fang, S.; Huang, S.F.; Cao, J.; Wen, Y.A.; Zhang, L.P.; Ren, G.S. Silencing of PCDH10 in hepatocellular carcinoma via de novo DNA methylation independent of HBV infection or HBx expression. Clin. Exp. Med. 2013, 13, 127–134. [Google Scholar] [CrossRef] [PubMed]
  29. Yan, J.; Lu, Q.; Dong, J.; Li, X.; Ma, K.; Cai, L. Hepatitis B virus X protein suppresses caveolin-1 expression in hepatocellular carcinoma by regulating DNA methylation. BMC Cancer 2012, 12, 353. [Google Scholar] [CrossRef] [PubMed]
  30. Lee, M.H.; Na, H.; Na, T.Y.; Shin, Y.K.; Seong, J.K.; Lee, M.O. Epigenetic control of metastasis-associated protein 1 gene expression by hepatitis B virus X protein during hepatocarcinogenesis. Oncogenesis 2012, 1, e25. [Google Scholar] [CrossRef] [PubMed]
  31. Lee, S.M.; Lee, Y.G.; Bae, J.B.; Choi, J.K.; Tayama, C.; Hata, K.; Yun, Y.; Seong, J.K.; Kim, Y.J. HBx induces hypomethylation of distal intragenic CPG islands required for active expression of developmental regulators. Proc. Natl. Acad. Sci. USA 2014, 111, 9555–9560. [Google Scholar] [CrossRef] [PubMed]
  32. Fu, X.; Song, X.; Li, Y.; Tan, D.; Liu, G. Hepatitis B virus X protein upregulates DNA methyltransferase 3a/3b and enhances SOCS-1CPG island methylation. Mol. Med. Rep. 2016, 13, 301–308. [Google Scholar] [PubMed]
  33. Qiu, X.; Zhang, L.; Lu, S.; Song, Y.; Lao, Y.; Hu, J.; Fan, H. Upregulation of DNMT1 mediated by HBx suppresses RASSF1A expression independent of DNA methylation. Oncol. Rep. 2014, 31, 202–208. [Google Scholar] [PubMed]
  34. Shon, J.K.; Shon, B.H.; Park, I.Y.; Lee, S.U.; Fa, L.; Chang, K.Y.; Shin, J.H.; Lee, Y.I. Hepatitis B virus-x protein recruits histone deacetylase 1 to repress insulin-like growth factor binding protein 3 transcription. Virus Res. 2009, 139, 14–21. [Google Scholar] [CrossRef] [PubMed]
  35. Arzumanyan, A.; Friedman, T.; Kotei, E.; Ng, I.O.; Lian, Z.; Feitelson, M.A. Epigenetic repression of E-cadherin expression by hepatitis B virus X antigen in liver cancer. Oncogene 2012, 31, 563–572. [Google Scholar] [CrossRef] [PubMed]
  36. Srisuttee, R.; Koh, S.S.; Kim, S.J.; Malilas, W.; Boonying, W.; Cho, I.R.; Jhun, B.H.; Ito, M.; Horio, Y.; Seto, E.; et al. Hepatitis B virus X (HBx) protein upregulates β-catenin in a human hepatic cell line by sequestering sirt1 deacetylase. Oncol. Rep. 2012, 28, 276–282. [Google Scholar] [PubMed]
  37. 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. Int. J. Exp. Cell Physiol. Biochem. Pharmacol. 2015, 37, 1547–1559. [Google Scholar] [CrossRef] [PubMed]
  38. Chen, W.S.; Yen, C.J.; Chen, Y.J.; Chen, J.Y.; Wang, L.Y.; Chiu, S.J.; Shih, W.L.; Ho, C.Y.; Wei, T.T.; Pan, H.L.; et al. MiRNA-7/21/107 contribute to HBx-induced hepatocellular carcinoma progression through suppression of maspin. Oncotarget 2015, 6, 25962–25974. [Google Scholar] [CrossRef] [PubMed]
  39. Li, J.F.; Dai, X.P.; Zhang, W.; Sun, S.H.; Zeng, Y.; Zhao, G.Y.; Kou, Z.H.; Guo, Y.; Yu, H.; Du, L.Y.; et al. Upregulation of microRNA-146a by hepatitis B virus X protein contributes to hepatitis development by downregulating complement factor H. mBio 2015, 6. [Google Scholar] [CrossRef] [PubMed]
  40. 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] [PubMed]
  41. Chen, J.J.; Tang, Y.S.; Huang, S.F.; Ai, J.G.; Wang, H.X.; Zhang, L.P. Hbx protein-induced upregulation of microRNA-221 promotes aberrant proliferation in HBv related hepatocellular carcinoma by targeting estrogen receptor-α. Oncol. Rep. 2015, 33, 792–798. [Google Scholar] [CrossRef] [PubMed]
  42. Zhao, J.; Wang, W.; Huang, Y.; Wu, J.; Chen, M.; Cui, P.; Zhang, W.; Zhang, Y. HBx elevates oncoprotein AEG-1 expression to promote cell migration by downregulating miR-375 and miR-136 in malignant hepatocytes. DNA Cell Biol. 2014, 33, 715–722. [Google Scholar] [CrossRef] [PubMed]
  43. Damania, P.; Sen, B.; Dar, S.B.; Kumar, S.; Kumari, A.; Gupta, E.; Sarin, S.K.; Venugopal, S.K. Hepatitis B virus induces cell proliferation via HBx-induced microRNA-21 in hepatocellular carcinoma by targeting programmed cell death protein4 (PDCD4) and phosphatase and tensin homologue (PTEN). PLoS ONE 2014, 9, e91745. [Google Scholar] [CrossRef] [PubMed]
  44. Cui, M.; Wang, Y.; Sun, B.; Xiao, Z.; Ye, L.; Zhang, X. MiR-205 modulates abnormal lipid metabolism of hepatoma cells via targeting acyl-CoA synthetase long-chain family member 1 (ACSL1) mRNA. Biochem. Biophys. Res. Commun. 2014, 444, 270–275. [Google Scholar] [CrossRef] [PubMed]
  45. Cui, M.; Xiao, Z.; Sun, B.; Wang, Y.; Zheng, M.; Ye, L.; Zhang, X. Involvement of cholesterol in hepatitis B virus X protein-induced abnormal lipid metabolism of hepatoma cells via up-regulating miR-205-targeted ACSL4. Biochem. Biophys. Res. Commun. 2014, 445, 651–655. [Google Scholar] [CrossRef] [PubMed]
  46. Shan, X.; Ren, M.; Chen, K.; Huang, A.; Tang, H. Regulation of the microRNA processor DGCR8 by hepatitis B virus proteins via the transcription factor YY1. Arch. Virol. 2015, 160, 795–803. [Google Scholar] [CrossRef] [PubMed]
  47. Huang, J.L.; Ren, T.Y.; Cao, S.W.; Zheng, S.H.; Hu, X.M.; Hu, Y.W.; Lin, L.; Chen, J.; Zheng, L.; Wang, Q. HBx-related long non-coding RNA DBH-AS1 promotes cell proliferation and survival by activating MAPK signaling in hepatocellular carcinoma. Oncotarget 2015, 6, 33791–33804. [Google Scholar] [PubMed]
  48. Du, Y.; Kong, G.; You, X.; Zhang, S.; Zhang, T.; Gao, Y.; Ye, L.; Zhang, X. Elevation of highly up-regulated in liver cancer (HULC) by hepatitis B virus X protein promotes hepatoma cell proliferation via down-regulating p18. J. Biol. Chem. 2012, 287, 26302–26311. [Google Scholar] [CrossRef] [PubMed]
  49. Huang, J.F.; Guo, Y.J.; Zhao, C.X.; Yuan, S.X.; Wang, Y.; Tang, G.N.; Zhou, W.P.; Sun, S.H. Hepatitis B virus X protein (HBx)-related long noncoding RNA (lncRNA) down-regulated expression by HBx (DREH) inhibits hepatocellular carcinoma metastasis by targeting the intermediate filament protein vimentin. Hepatol. 2013, 57, 1882–1892. [Google Scholar] [CrossRef] [PubMed]
  50. Lau, C.C.; Sun, T.; Ching, A.K.; He, M.; Li, J.W.; Wong, A.M.; Co, N.N.; Chan, A.W.; Li, P.S.; Lung, R.W.; et al. Viral-human chimeric transcript predisposes risk to liver cancer development and progression. Cancer Cell 2014, 25, 335–349. [Google Scholar] [CrossRef] [PubMed]
  51. Sze, K.M.; Chu, G.K.; Lee, J.M.; Ng, I.O. C-terminal truncated hepatitis B virus X protein is associated with metastasis and enhances invasiveness by c-Jun/matrix metalloproteinase protein 10 activation in hepatocellular carcinoma. Hepatology 2013, 57, 131–139. [Google Scholar] [CrossRef] [PubMed]
  52. Ma, N.F.; Lau, S.H.; Hu, L.; Xie, D.; Wu, J.; Yang, J.; Wang, Y.; Wu, M.C.; Fung, J.; Bai, X.; et al. Cooh-terminal truncated HBV X protein plays key role in hepatocarcinogenesis. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2008, 14, 5061–5068. [Google Scholar] [CrossRef] [PubMed]
  53. Lizzano, R.A.; Yang, B.; Clippinger, A.J.; Bouchard, M.J. The C-terminal region of the hepatitis B virus X protein is essential for its stability and function. Virus Res. 2011, 155, 231–239. [Google Scholar] [CrossRef] [PubMed]
  54. Li, S.K.; Ho, S.F.; Tsui, K.W.; Fung, K.P.; Waye, M.Y. Identification of functionally important amino acid residues in the mitochondria targeting sequence of hepatitis B virus X protein. Virology 2008, 381, 81–88. [Google Scholar] [CrossRef] [PubMed]
  55. Ng, K.Y.; Chai, S.; Tong, M.; Guan, X.Y.; Lin, C.H.; Ching, Y.P.; Xie, D.; Cheng, A.S.; Ma, S. C-terminal truncated hepatitis B virus X protein promotes hepatocellular carcinogenesis through induction of cancer and stem cell-like properties. Oncotarget 2016. [Google Scholar] [CrossRef] [PubMed]
  56. Liu, L.P.; Hu, B.G.; Ye, C.; Ho, R.L.; Chen, G.G.; Lai, P.B. Hbx mutants differentially affect the activation of hypoxia-inducible factor-1α in hepatocellular carcinoma. Br. J. Cancer 2014, 110, 1066–1073. [Google Scholar] [CrossRef] [PubMed]
  57. Xie, Y.; Liu, S.; Zhao, Y.; Guo, Z.; Xu, J. X protein mutations in hepatitis b virus DNA predict postoperative survival in hepatocellular carcinoma. Tumour Biol. J. Int. Soc. Oncodev. Biol. Med. 2014, 35, 10325–10331. [Google Scholar] [CrossRef] [PubMed]
  58. Wang, Q.; Wei, F.; Li, C.; Lv, G.; Wang, G.; Liu, T.; Bellail, A.C.; Hao, C. Combination of mTOR and EGFR kinase inhibitors blocks mTORC1 and mTORC2 kinase activity and suppresses the progression of colorectal carcinoma. PLoS ONE 2013, 8, e73175. [Google Scholar] [CrossRef] [PubMed]
  59. Nagahashi, M.; Matsuda, Y.; Moro, K.; Tsuchida, J.; Soma, D.; Hirose, Y.; Kobayashi, T.; Kosugi, S.I.; Takabe, K.; Komatsu, M.; et al. DNA damage response and sphingolipid signaling in liver diseases. Surg. Today 2015. [Google Scholar] [CrossRef] [PubMed]
  60. Van de Klundert, M.A.; van Hemert, F.J.; Zaaijer, H.L.; Kootstra, N.A. The hepatitis B virus X protein inhibits thymine DNA glycosylase initiated base excision repair. PLoS ONE 2012, 7, e48940. [Google Scholar] [CrossRef] [PubMed]
  61. Shih, P.H.; Yeh, C.T.; Yen, G.C. Anthocyanins induce the activation of phase ii enzymes through the antioxidant response element pathway against oxidative stress-induced apoptosis. J. Agric. Food Chem. 2007, 55, 9427–9435. [Google Scholar] [CrossRef] [PubMed]
  62. He, Y.; Meng, X.M.; Huang, C.; Wu, B.M.; Zhang, L.; Lv, X.W.; Li, J. Long noncoding RNAs: Novel insights into hepatocelluar carcinoma. Cancer Lett. 2014, 344, 20–27. [Google Scholar] [CrossRef] [PubMed]
  63. Li, B.; Carey, M.; Workman, J.L. The role of chromatin during transcription. Cell 2007, 128, 707–719. [Google Scholar] [CrossRef] [PubMed]
  64. Tian, Y.; Yang, W.; Song, J.; Wu, Y.; Ni, B. Hepatitis B virus X protein-induced aberrant epigenetic modifications contributing to human hepatocellular carcinoma pathogenesis. Mol. Cell. Biol. 2013, 33, 2810–2816. [Google Scholar] [CrossRef] [PubMed]
  65. Kew, M.C. Hepatitis B virus X protein in the pathogenesis of hepatitis B virus-induced hepatocellular carcinoma. J. Gastroenterol. Hepatol. 2011, 26, 144–152. [Google Scholar] [CrossRef] [PubMed]
  66. Tang, S.; Hu, W.; Hu, J.; Wu, S.; Li, J.; Luo, Y.; Cao, M.; Zhou, H.; Jiang, X. Hepatitis B virus X protein promotes p3 transcript expression of the insulin-like growth factor 2 gene via inducing hypomethylation of p3 promoter in hepatocellular carcinoma. Liver Int. 2015, 35, 608–619. [Google Scholar] [CrossRef] [PubMed]
  67. Liz, J.; Esteller, M. LncRNAs and microRNAs with a role in cancer development. Biochim. Biophys. Acta 2016, 1859, 169–176. [Google Scholar] [CrossRef] [PubMed]
  68. Bechmann, L.P.; Hannivoort, R.A.; Gerken, G.; Hotamisligil, G.S.; Trauner, M.; Canbay, A. The interaction of hepatic lipid and glucose metabolism in liver diseases. J. Hepatol. 2012, 56, 952–964. [Google Scholar] [CrossRef] [PubMed]
  69. Toh, S.T.; Jin, Y.; Liu, L.; Wang, J.; Babrzadeh, F.; Gharizadeh, B.; Ronaghi, M.; Toh, H.C.; Chow, P.K.; Chung, A.Y.; et al. Deep sequencing of the hepatitis B virus in hepatocellular carcinoma patients reveals enriched integration events, structural alterations and sequence variations. Carcinogenesis 2013, 34, 787–798. [Google Scholar] [CrossRef] [PubMed]
  70. Iavarone, M.; Trabut, J.B.; Delpuech, O.; Carnot, F.; Colombo, M.; Kremsdorf, D.; Brechot, C.; Thiers, V. Characterisation of hepatitis B virus X protein mutants in tumour and non-tumour liver cells using laser capture microdissection. J. Hepatol. 2003, 39, 253–261. [Google Scholar] [CrossRef]
  71. Wang, Y.; Lau, S.H.; Sham, J.S.; Wu, M.C.; Wang, T.; Guan, X.Y. Characterization of hbv integrants in 14 hepatocellular carcinomas: Association of truncated X gene and hepatocellular carcinogenesis. Oncogene 2004, 23, 142–148. [Google Scholar] [CrossRef] [PubMed]
  72. Tu, H.; Bonura, C.; Giannini, C.; Mouly, H.; Soussan, P.; Kew, M.; Paterlini-Brechot, P.; Brechot, C.; Kremsdorf, D. Biological impact of natural cooh-terminal deletions of hepatitis B virus X protein in hepatocellular carcinoma tissues. Cancer Res. 2001, 61, 7803–7810. [Google Scholar] [PubMed]
  73. Kumar, V.; Jayasuryan, N.; Kumar, R. A truncated mutant (residues 58–140) of the hepatitis B virus X protein retains transactivation function. Proc. Natl. Acad. Sci. USA 1996, 93, 5647–5652. [Google Scholar] [CrossRef] [PubMed]
  74. Shi, Y.; Wang, J.; Wang, Y.; Wang, A.; Guo, H.; Wei, F.; Mehta, S.R.; Espitia, S.; Smith, D.M.; Liu, L.; et al. A novel mutant 10ALA/ARG together with mutant 144SER/ARG of hepatitis B virus X protein involved in hepatitis B virus-related hepatocarcinogenesis in HEPG2 cell lines. Cancer Lett. 2016, 371, 285–291. [Google Scholar] [CrossRef] [PubMed]
  75. Tai, J.; Wang, G.; Liu, T.; Wang, L.; Lin, C.; Li, F. Effects of siRNA targeting c-Myc and VEGF on human colorectal cancer volo cells. J. Biochem. Mol. Toxicol. 2012, 26, 499–505. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Hepatitis B virus (HBV) genome map. The genome of HBV is a double-stranded DNA (3.2 kb), which contains four overlapping open reading frames (ORFs) coding for viral envelope (pre-S1/pre-S2/S) (blue arrow), core proteins (pre-C/C) (yellow arrow), viral polymerase (green arrow), and HBx protein (red arrow). The genome contains four promoters, two enhancer regions (Enh1, Enh2), and two direct repeats (DR1, DR2).
Figure 1. Hepatitis B virus (HBV) genome map. The genome of HBV is a double-stranded DNA (3.2 kb), which contains four overlapping open reading frames (ORFs) coding for viral envelope (pre-S1/pre-S2/S) (blue arrow), core proteins (pre-C/C) (yellow arrow), viral polymerase (green arrow), and HBx protein (red arrow). The genome contains four promoters, two enhancer regions (Enh1, Enh2), and two direct repeats (DR1, DR2).
Ijms 17 00940 g001
Figure 2. HBx and its mutifuntional roles in hepatocarcinogenesis. Each box represents processes that HBx may potentially play a role. The black arrows indicate up or downregulation of miRNAs. Abbreviations are as follow: DBH-AS1: DBH antisense RNA 1; HULC: highly upregulated in liver cancer; LINE: long interspersed nuclear element; Dreh: downregulated expression by HBx; IGF2: insulin-like growth factor-2; PCDH10: protocadherin 10; MTA1: metastasis-associated protein 1; mCGIs: hypermethylated intragenic CpG islands; IGFBP3: IGF binding protein 3; CDH1: E-cadherin gene; SIRT1: NAD-dependent deacetylase sirtuin-1; DR5 death receptor 5; Bcl-2: B-cell lymphoma 2; Mcl-1: myeloid cell leukemia-1; TFIIH: transcription factor IIH; BER: base excision repair; hOGG1: human 8-oxoguanine DNA glycosylase 1; hMYHa: human 8-oxoguanine DNA glycosylase 1(hOGG1) and DNA glycosylase a; NQO1: NAD(P)H: quinone oxidoreductase 1; ROS: reactive oxygen species; Foxo4: forkhead box class O 4.
Figure 2. HBx and its mutifuntional roles in hepatocarcinogenesis. Each box represents processes that HBx may potentially play a role. The black arrows indicate up or downregulation of miRNAs. Abbreviations are as follow: DBH-AS1: DBH antisense RNA 1; HULC: highly upregulated in liver cancer; LINE: long interspersed nuclear element; Dreh: downregulated expression by HBx; IGF2: insulin-like growth factor-2; PCDH10: protocadherin 10; MTA1: metastasis-associated protein 1; mCGIs: hypermethylated intragenic CpG islands; IGFBP3: IGF binding protein 3; CDH1: E-cadherin gene; SIRT1: NAD-dependent deacetylase sirtuin-1; DR5 death receptor 5; Bcl-2: B-cell lymphoma 2; Mcl-1: myeloid cell leukemia-1; TFIIH: transcription factor IIH; BER: base excision repair; hOGG1: human 8-oxoguanine DNA glycosylase 1; hMYHa: human 8-oxoguanine DNA glycosylase 1(hOGG1) and DNA glycosylase a; NQO1: NAD(P)H: quinone oxidoreductase 1; ROS: reactive oxygen species; Foxo4: forkhead box class O 4.
Ijms 17 00940 g002
Table 1. Reported mechanisms of HBx on the various pathways, epigenetic and genetic events.
Table 1. Reported mechanisms of HBx on the various pathways, epigenetic and genetic events.
GroupSub-GroupTargetMechanismReference
PathwaysSignaling pathwayNotch1,Notch41. Activate Notch pathway by receptor Notch1 and Notch4, inducing cell growth, cell cycle prograssion and anti-apoptosisGao et al. [11]
AFP2. Induce AFP expression to activate PI3K/mTOR pathway, resulting in the promotion of progression, invasion and metastasis of cancer cellsZhu et al. [12]
SFRP1, SFRP53. Inhibit Wnt/β-catenin pathway by reducing its two ligands, SFRP1 and SFRP5, resulting in EMT Xie et al. [13]
DNA repairTFIIH1. Reduce DNA repaire capacity by interfering TFIIHQadri et al. [14]
DNA glycosylases2. Has a similar structure of DNA glycosylases but doesn't have the capapility in DNA repair Hement et al. [15]
hOGG1, hMYHa3. Inhibit DNA repair by hindering DNA repair enzyme hOGG1 and hMYHaCheng et al. [16]
Oxidative stressNQO11. Induce mitochondria injury and oxidative stress by downregulating NQO1Wu et al. [17]
mitochondria DNA2. Induce ROS and damage mitochondria DNAJung et al. [18]
Foxo43. Enhance resistances to oxidative stress-induced cell death by upregulating Foxo4Srisuttee et al. [19]
ImmuneTRIF1. Enable HBV replication and evasion from innate immunity by reducing TRIF expressionHong et al. [20]
IPS-12. Bind to IPS-1 and diminish IFN-β signalingKumar et al. [21]
SDF-13. Recruit immune cells into liver by inducing SDF-1 via endoplasmic reticulum stress Cho et al. [22]
ApoptosisDR51. Promote TRAIL induced apoptosis by increasing DR5Kong et al. [23]
A202. Sensitize TRAIL induced apoptosis by inhibiting caspase-8 inhibitor A20Zhang et al. [24]
Bcl-2, Mcl-13. Inhibit apoptosis by increasing apoptosis inhibition gene Bcl-2 and Mcl-1Shen et al. [25]
EpegneticsMethylationIGF-21. Hypomethylation the promoter of IGF-2 promoterLiu et al. [26]
CD822. Hypermethylation the promoter of metastasis-inhibit gene CD82Yu et al. [27]
PCDH103. Hypermethylation the promoter of tumor suppressor gene PCDH10 Fang et al. [28]
Caveolin-14. Hypermethylation the promoter of tumor suppressor gene Caveolin-1Yan et al. [29]
MTA15. Hypermethylation the promoter of tumor suppressor gene MTA1Lee et al. [30]
mCGIs6. Hypomethylation of mCGIs then influence cell differenciationLee et al. [31]
SOCS-17. Hypermethylation the promoter of tumor supprssor gene SOCS-1Fu et al. [32]
RASSF1A8. Hypermethylation the promoter of tumor suppressor gene RASSF1AQiu et al. [33]
AcetylationSP11. Deacetylation of SP1 then romotes cell survival, transformation, and progression to cancerShon et al. [34]
CDH12. Deacetylation of CDH1 promoter then inhibit metastasisArzumanyan et al. [35]
SIRT13. Attenuate the interaction between SIRT1 and β-catenin then protecting β-catenin from degradation Srisuttee et al. [36]
miRNAsmiR-1451. Induce CUL5 by down regulation of miR-145 then promote cell growthGao et al. [37]
miR-7,21,1072. Induce maspin by down regulation of miR-7,21,107 then promote migration, ivasion and chemoresistanceChen et al. [38]
miR-146a3. Inhibit CFH by up regulation of miR-146a then enhance the alternative pathway of complement activationLi et al. [39]
miR-216b4. Induce IGFBP2 by down regulation of miR-216b then increase cell proliferationLiu et al. [40]
miR-2215. Inhibit ERα, which is a protective factor against HCC, by up regulation of miR-221Chen et al. [41]
miR-136,3756. Induce AEG-1 by down regulation of miR-136 and 375 then promote cell migration Zhao et al. [42]
miR-217. Inhibit PDCD4 and PTEN by up regulation of miR-21 then increase cell proliferationDamania et al. [43]
miR-2058. Upregulate ACSL1/4 by inhibit miR-205 then affect lipid metabolismCui et al. [44,45]
DGCR89. Inhibit miRNA processor DGCR8 and interfer miRNA productionShan et al. [46]
LncRNAsDBH-AS11. Induce DBH-AS1 expression which activates ERK/p38/JNK MAPK signalling pathwayHuang et al. [47]
HULC2. Up regulate HULC which promotes cell proliferationDu et al. [48]
Dreh3. Inhibit HCC metastasis by downregulation of DrehHuang et al. [49]
LINE14. HBx-LINE1 fusion exerts lncRNA function and indeces EMTLau et al. [50]
MutationC-terminal truncationMMP101. Increse invasion by activating MMP10Sze et al. [51]
HBx2. Promote proliferation and inhibit apoptotic frequencyMa et al. [52]
HBx3. Decrease HBx steady level and slow HBV replicationLizzano et al. [53]
mitochondrial4. Target mitichondrial then aggregate it at perinuclear spaceLi et al. [54]
CSC5. Enhance stemness of CSC and drug resistancy Ng et al. [55]
Point mutationp531. A10R-S144R arrests cell cycle and attenuate p53 bindingLiu et al. [56]
HIF-12. K130M/V131I strengthens the transcriptional activity of HIF-1Xie et al. [57]

Share and Cite

MDPI and ACS Style

Liu, S.; Koh, S.S.Y.; Lee, C.G.L. Hepatitis B Virus X Protein and Hepatocarcinogenesis. Int. J. Mol. Sci. 2016, 17, 940. https://doi.org/10.3390/ijms17060940

AMA Style

Liu S, Koh SSY, Lee CGL. Hepatitis B Virus X Protein and Hepatocarcinogenesis. International Journal of Molecular Sciences. 2016; 17(6):940. https://doi.org/10.3390/ijms17060940

Chicago/Turabian Style

Liu, Shuaichen, Samantha S. Y. Koh, and Caroline G. L. Lee. 2016. "Hepatitis B Virus X Protein and Hepatocarcinogenesis" International Journal of Molecular Sciences 17, no. 6: 940. https://doi.org/10.3390/ijms17060940

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