Hepatitis B virus X protein accelerates the development of hepatoma =================================================================== * Xiao-Dong Zhang * Yuan Wang * Li-Hong Ye ## Abstract The chronic infection of hepatitis B virus (HBV) is closely related to the occurrence and development of hepatocellular carcinoma (HCC). Accumulated evidence has shown that HBV X protein (HBx protein) is a multifunctional regulator with a crucial role in hepatocarcinogenesis. However, information on the mechanism by which HBV induces HCC is lacking. This review focuses on the pathological functions of HBx in HBV-induced hepatocarcinogenesis. As a transactivator, HBx can modulate nuclear factor kappa-light-chain-enhancer of activated B cells (NF-κB) and transcription factor AP-2. Moreover, HBx can affect regulatory non-coding RNAs (ncRNAs) including microRNAs and long ncRNAs (lncRNAs), such as miRNA-205 and highly upregulated in liver cancer (HULC), respectively. HBx is also involved in epigenetic modification, including methylation and acetylation. HBx interacts with various signal-transduction pathways, such as protein kinase B/Akt, Wnt/β-catenin, signal transducer and activator of transcription, and NF-κB pathways. Moreover, HBx affects cellular fate by shifting the balance toward cell survival. HBx may lead to the loss of apoptotic functions or directly contributes to oncogenesis by achieving transforming functions, which induce hepatocarcinogenesis. Additionally, HBx can modulate apoptosis and immune response by direct or indirect interaction with host factors. We conclude that HBx hastens the development of hepatoma. * Hepatocellular carcinoma (HCC) * hepatitis B virus (HBV) * HBV X protein (HBx protein) * hepatocarcinogenesis ## Introduction Hepatocellular carcinoma (HCC) is one of the most common malignant cancers worldwide and ranks the third in annual cancer mortality rates. The chronic infection of hepatitis B virus (HBV) has been identified as a major risk factor in the development of HCC, accounting for 55% of cases worldwide; 80% or more of such cases are found in the eastern Pacific region and sub-Saharan Africa, which are the areas with the highest tumor incidence1. The mechanisms underlying HBV-induced malignant transformation remain ambiguous, but evidence has suggested that HBV X (HBx) protein serves a crucial function in the pathogenesis of HCC2. In this review, we focused on the molecular mechanisms of HBx in HCC development. ## *HBV X* gene and HBx protein The HBV genome is circular with partly double-stranded DNA. The long (minus) strand is approximately 3,200 bases in length and contains four open reading frames (ORFs), namely, S, C, P, and X. *HBx* is the smallest gene of the four partially overlapping ORFs of HBV. The genome comprises 452 nucleotides, which encode a 154-amino acid regulatory protein with a molecular mass of 17 kDa3. HBx is the designated name of the gene and protein because the amino acid sequences failed to show homology with known proteins. HBx protein is highly conserved among the different subtypes of the virus and is common to all mammalian members of the *Hepadnaviridae* but not to avian viruses4. The HBx protein is present in the cytoplasm and, to a lesser extent, the nucleus of hepatocytes. The enigmatic HBx protein is a transactivator that can activate various viral and cellular promoters and enhancers. HBx protein fails to bind directly to DNA, and its transcriptional activity is mediated via protein-protein interaction. The transactivation function of HBx may be exerted in the cytoplasm via signaling pathways and in the nucleus via DNA-binding proteins. HBx transcriptional activity is necessary for viral replication5. A high number of transcriptional factors [such as nuclear factor kappa-light-chain-enhancer of activated B cells (NF-κB) and AP-2], which directly interact with HBx, were recently identified6. The clinical relevance of HBx begins with the integration of HBV DNA into the genome of the hepatocytes of chronic HBV carriers. The *HBx* gene is usually conserved in the integrant and is often found in the malignant hepatocytes of chronic HBV carriers. Our groups identified the *HBx* gene inserted into the upstream of a 20-bp *Alu* core sequence with a secondary deletion of 382 to 465 bp at its 3' end, followed by subtelomeric DNA; *HBx* gene-integration causes the recombination of HBx/*Alu* core sequence/subtelomeric DNA in the host genome7. HBx transcripts accumulate in the hepatocytes; the HBx induction in the hepatocytes seems to alter the expression of many genes involved in signal transduction pathways, cell cycle control, metastasis, transcriptional regulation, immune response, and metabolism8. Recently, Qiu *et al*.9 analyzed the differentially expressed genes between the LO2–HBx and LO2-vector control cells using microarray analysis, and the results were validated using quantitative real-time PCR. HBx upregulates 456 genes and downregulates 843 genes. These initial cellular targets of HBx provide a starting point in mapping the HBx mechanisms of action during the progressive development of HBx-mediated hepatocarcinogenesis. ## Mechanisms of HBx-induced hepatocarcinogenesis ### HBx and regulatory non-coding RNAs (ncRNAs) The perspective on gene regulation in biology has mostly focused on the protein-coding genes via the central dogma of DNA→mRNA→protein. However, with the development of new techniques, such as high-resolution microarray and massively parallel sequencing, a minimum of 90% of the human genome has been actively transcribed into ncRNAs, and less than 2% of the genome sequences encode proteins9. According to their size, ncRNAs are divided into two groups, as follows: small ncRNAs (<200 nt) and long ncRNAs (lncRNAs). The interaction between HBx and ncRNAs is evident in HCC development. #### HBx and small ncRNAs Small ncRNAs include microRNAs (miRNAs), small-interfering RNAs (siRNAs), or PIWI-interacting RNAs. Among these small ncRNAs, miRNAs are the most detailed. MiRNAs are evolutionarily conserved noncoding RNAs (21 to 25 nt in length). These RNAs serve critical functions in gene expression regulation and multiple cellular processes, including proliferation, development, differentiation, and tumorigenesis. The alterations in miRNA expression have been observed in HCCs and are linked to the molecular pathogenesis of HCC because of their effect on the expression of crucial mRNAs. Depending on the target genes, miRNAs can function as tumor suppressor genes or oncogenes through mRNA cleavage or translational repression. HBx suppresses the p53-mediated activation of miR-148a, resulting in the upregulation of AKT and ERK and subsequent activation of mTOR to promote cancer growth and metastasis in a mouse model of HCC10. Recent studies reported that HBx suppresses the expression of miR-15b, which directly targets fucosyltransferase 2 and increases the levels of Globo H to enhance HCC cell proliferation11. miR-205 was downregulated in the liver of HBx-transgenic mice. HBx can abrogate the effect of miR-205 on tumor suppression in carcinogenesis12. The downregulation of miR-205 mediated by HBx remarkably increases the expression levels of acyl-CoA synthetase long-chain family member 1 (ACSL1), and its metabolite triglyceride levels are remarkably increased in HBx-induced liver cancer tissues, as shown in an HBx transgenic mice model13. HBx-transgenic mice have higher miR-224 expression than normal mice. Moreover, miR-224 promotes hepatoma cell migration and tumor formation by silencing its target gene *Smad4*14. HBx-mediated miRNA regulation serves a crucial function in hepatocarcinogenesis. #### HBx and lncRNAs LncRNAs are broadly defined as endogenous cellular RNA molecules that are longer than 200 nt. LncRNAs are poorly conserved and can regulate gene expression at various levels, such as during chromatin modification, transcription, and post-transcriptional processing. An increasing amount of evidence showed that lncRNAs serve critical functions in the development and progression of HBV-related HCC. For example, highly upregulated in liver cancer (HULC) is dysregulated in HCC. HULC expression level is positively correlated with HBx expression in clinical HCC tissues. Moreover, HBx can upregulate the HULC expression in human immortalized normal liver L-O2 cells and hepatoma HepG2 cells. The upregulation of HULC by HBx can promote the proliferation of hepatoma cells by p18 suppression15. Some studies have shown that HBx-transgenic mice show a specific profile of liver lncRNAs compared with wild-type mice; such profiles include the downregulated expression by HBx (Dreh; lncRNA–Dreh), which can inhibit HCC growth and metastasis *in vitro* and *in vivo* and functions as a tumor suppressor in the development of HBV-related HCC16. The lncRNA dysregulation, which is mediated by HBx, is closely related to tumorigenesis, metastasis, prognosis, or diagnosis of HCC; thus, lncRNA dysregulation serves an important function in hepatocarcinogenesis17. ### HBx and modification of epigenetics Early research on the pathogenesis of HBx-induced HCC and on other etiological forms of the tumor focused on the changes at genetic level. Recent studies showed that epigenetic changes, which inactivate tumor suppressor genes or chromosomal instability, serve an increasingly important function in hepatocarcinogenesis18. Epigenetic changes refer to the changes in gene expression mediated by mechanisms other than the alterations in the nucleotide sequence of that gene. Such alterations include aberrations in DNA methylation and histone modifications. #### HBx and methylation Reversible DNA methylation occurs at the sites of cytosine-guanine dinucleotides (CpGs) and can be generated by adding a methyl group to the carbon-5 position of cytosine. DNA methylation patterns are established by a family of DNA methyltransferases (DNMTs). The HBx-induced upregulation of DNMT activity is implicated in the aberrant methylation of CpG islands in some host genes that are involved in tumor suppression. Moreover, these DNMT-mediated methylated genes serve a function in cell cycle regulation, cell growth, dedifferentiation, invasiveness, apoptosis, immune escape, and xenobiotic metabolism, thereby contributing to HBV-related HCC onset and/or progression19. Recent studies showed that HBx can upregulate the expression of DNMT1 and DNMT3A by HBx transcriptional transactivation activity20. HBx induces the promoter hypermethylation of the genes with tumor-suppressing activity by upregulating the expression of DNMTs, as shown in the *p16INK4A* gene that negatively regulates cell cycle21. In addition, HBx significantly elevates the expression of genes by inducing the targeted promoter hypomethylation; these genes include several tumor promotion-related genes, such as retinal dehydrogenase 1, plasma retinol-binding protein precursor, and cellular retinol-binding protein I22. #### HBx and acetylation HCC usually occurs with aberrant acetylation, particularly histone acetylation. Such abnormal histone modifications perturb cellular gene expression, thereby disrupting normal cellular activities. Acetylation is controlled by histone acetyltransferases and histone deacetylases. HBx can promote HBV-induced HCC pathogenesis by inducing the histone acetylation of some tumor promotion-related genes. For example, HBx can strongly induce the expression of HDAC1 at transcription level, thereby enhancing the stability of hypoxia-inducible factor-1α, which may serve a critical function in the angiogenesis and metastasis of HBV-associated HCCs23. HBx can also promote the development of HBV-induced HCC by inducing the histone deacetylations of specific tumor suppression-related genes. Recent studies demonstrated that the HBx-mediated suppression of *cadherin-1* (*CDH1*) genes involves the recruitment of the transcription factor mSin3A/HDAC1 complex to the promoters of the Snail-binding sites in human *CDH1*24. HBx recruits HDAC1 to form a complex with Sp1 in a p53-independent manner and with deacetylated Sp1; subsequently, the binding of Sp1 to the targeted DNA during transcriptional repression is diminished25. NAD-dependent deacetylase sirtuin-1 (SIRT1), a member of the class III histone deacetylases, is conserved during evolution of different species, from yeasts to humans. SIRT1 is implicated in several biological processes, including metabolism, cell division, differentiation, survival, and senescence26. In particular, SIRT1 mediates adaptation process to enhance the survival of an organism under cellular stress conditions, such as nutrient deprivation and oxidation. The presence of HBx attenuates the interaction between SIRT1 and β-catenin, thereby protecting β-catenin from the inhibitory action of SIRT1, which leads to anticancer drug treatment resistance27. The upregulation of SIRT1 expression or activity exposes Hep3B cells, which stably express HBx, to oxidation-induced apoptosis via the downregulation of c-Jun N-Terminal protein kinase activity28. These results suggest that SIRT1 serves an important function in protecting HBx-expressing HCC cells from oxidative stress, thereby promoting hepatocarcinogenesis. ### HBx and signaling pathway #### HBx and Akt signaling pathway As one of the most versatile kinase families, Akt (also known as protein kinase B) serine–threonine kinases critically regulate cell survival, proliferation, metabolism, and migration. The dysregulation of Akt kinases is frequently associated with human diseases, such as cancer29. Previous studies focused on the functional interaction between HBx and Akt; HBx can be a substrate of Akt kinase30. The molecular cooperation between HBx and Akt serves a key function in cell survival and oncogenic transformation. Researchers recently isolated two HBx isoforms from an HBV carrier. The first HBx isoform contains an Akt phosphorylation site at Ser31 and functions as an anti-apoptotic protein (designated as HBx-S31), whereas the other isoform lacks an Akt phosphorylation site and functions as an apoptotic protein (designated as HBx-L31). HBx-S31 can activate Akt, whereas HBx-L31 lacks this ability; the former enhances tumor growth, whereas the latter suppresses tumorigenesis. This finding shows that HBx can serve two functions (pro-apoptotic and anti-apoptotic) through different isoforms; HBx with Ser31 induces Akt signaling31. HBx-activated Akt phosphorylates the downstream target glycogen synthase kinase 3β (GSK-3β), thereby stabilizing β-catenin and enhancing promoter recruitment and expression of cyclin D1 and Bcl-XL32. In addition, HBx-activated Akt promotes the IKKα nuclear translocation via phosphorylating its threonine-23. IKKα ubiquitination, which is enhanced by HBx and Akt, contributes to IKKα accumulation in the nucleus, thereby indicating the involvement of ubiquitination in the Akt-increased IKKα nuclear transport in response to HBx32. The proliferation and colony formation of HCC cells are suppressed by miR-132-mediated inhibition of the Akt-signaling pathway in miR-132-transfected cells, as revealed by data from recent studies. Such data also revealed that HBx can repress the expression of miR-132 by DNA methylation33. The oncogenic cooperation between HBx and Akt may be important for cell proliferation, abrogation of apoptosis, and tumorigenic transformation of cells. #### HBx and Wnt signaling pathway Wnt pathway, an essential regulator of early embryonic development, contributes to somatic stem cell homeostasis. The canonical Wnt signal is initiated by the binding of Wnt ligands to the receptor Frizzled to activate Dishevelled (Dvl) proteins. Dvl transduces Wnt signal by suppressing the degradation of β-catenin via the adenomatosis polyposis coli (APC)/Axin/GSK3β destruction complex. The accumulation of β-catenin in the cytosol leads to its translocation into the nucleus, where the transcription of Wnt target genes is initiated34. The dysregulation of the Wnt pathway produces an improperly decayed β-catenin and an elevated β-catenin level, thereby hyper activating the transcription of cyclin D, VEGF, and c-MYC to subsequently promote the development and progression of HCC35. The increased expression of β-catenin in the cytoplasm or nucleus occurs in 50% to 70% of patients with HCC. Currently available studies showed that HBx competitively binds to APC to displace β-catenin from its degradation complex, resulting in β-catenin upregulation in the nucleus and the activation of Wnt signaling; thus, malignant transformation is induced35. In addition, the ectopic expression of HBx with Wnt-1 can activate Wnt/β-catenin signaling in Huh7 cells by stabilizing cytoplasmic β-catenin. Furthermore, the stabilization of β-catenin by HBx can be achieved by suppressing GSK-3 activity via the activation of Src kinase36. HBV-related HCC predominantly occurs in men; HBx is involved in this disparity between sexes by enhancing the transcriptional activity of the androgen receptor (AR) gene in a ligand-dependent manner. HBx can also enhance the AR transcriptional activity via c-Src and GSK-3β37. Thus, HBx is important in the activation of Wnt/β-catenin signaling in hepatocarcinogenesis. #### HBx and signal transducer and activator of transcription (STAT) STAT pathway serves a crucial function in the regulation of many key cellular processes, such as survival, proliferation, and differentiation. In the liver, the activation of STAT proteins, including STAT3 and STAT5, is critical for anti-viral defense against hepatitis viral infection and for controlling injury, repair, inflammation, and tumorigenesis38. In particular, HBx specifically upregulates the expression of tyrosine phosphorylation and increases the kinase activity of Jak1 through protein–protein interaction with Jak1. HBx induces the Jak–STAT signaling pathway39. The expression levels of p-STAT3 are positively correlated with those of the HBx expression. HBx can activate the Twist promoter by activating STAT3, which promotes the EMT occurrence in liver cells40. Previous studies revealed atypical activity. These observations emphasize the active role of HBx in hepatocarcinogenesis, particularly in the dysregulation of STAT signaling. #### HBx and NF-κB HBx is an activator of the transcription factor NF-κB, which is the first HBx-responsive motif to be identified6. HBx-mediated regulation of the heat shock protein gp96 expression requires an NF-κB cis-regulatory element in the promoter of gp96. HBx promotes the binding of NF-κB to the gp96 promoter41. Moreover, HBx directly regulates Notch1 signaling, which cross-talked with the NF-κB pathway [p65, p50, and IκB kinase α (IKKα)]. The downregulation of Notch1 decreases the binding of p65 to its target gene promoter, reduces NF-κB expression, and enhances IKKα expression42. Tumor necrosis factor (TNF)-α, which is an NF-κB signaling activator, induces the accumulation of HBx in cells by increasing protein stability, which is mainly caused by the reduction of proteasomal degradation. The effects of TNF-α on HBx protein stability are mediated via the activation of NF-κB effector kinases, such as IKKα, IKKβ, and p65, thereby indicating a positive feedback loop between HBx and NF-κB signaling pathways43. The ability of NF-κB, to mediate apoptosis is inhibited by HBx-expressing cells. The regulation of NF-κB mediated by HBx serves a central function in HBV-related HCC. ## Functions of HBx in hepatocarcinogenesis ### HBx enhances cell transformation The hepatocarcinogenic effect of HBx on HBx-transgenic mice has been investigated. *HBx* gene expression can disrupt the normal cell growth in transgenic mice44. Studies suggested that HBx-transgenic mouse cells are characterized by malignant transformation45. Moreover, HBx protein exerts a strong growth arrest on hepatocytes and an imbalanced cell cycle progression, resulting in abnormal cell death; this condition is accompanied by severe fat accumulation and impaired glycogen storage in the HBx-transgenic livers46. HBx protein also blocks the G1/S transition of the hepatocyte cell cycle progression and causes liver functionality failure and cell death in the regenerating liver of HBx-transgenic mice46. The stable HBx transfection results in a malignant phenotype in the engineered cells *in vivo* and *in vitro*. HBx can increase the transcription of NF-κB, AP-1, and survivin to upregulate the expression levels of c-Myc and survivin. Abnormal centrosome duplication and activated human telomerase reverse transcriptase are responsible for the transformation47. The aberrant high expression of the *HBx* gene in normal cells strongly affects cell cycle progression and growth advantage, which promotes cell transformation47. HBx affects cellular fate by shifting the balance toward cell survival, probably causing loss of apoptotic functions or directly contributing to oncogenesis by gain of transforming functions. Therefore, HBx contributes to oncogenesis by stimulating cell proliferation. These studies indicate that HBx may function as an activator of hepatocarcinogenesis. ### HBx and apoptosis Programmed cell death (apoptosis) is necessary for the elimination of redundant, damaged, and virally infected cells. Generally, proteins regulate apoptosis in three ways, as follows: as effectors and initiators of apoptosis, as inducers and suppressors of apoptosis, and as intermediate proteins2. HBx can regulate apoptosis through its action on caspases (cysteine–aspartate specific proteases), the mitochondria, and SIRT. ### HBx and cysteine aspartate specific proteases (caspases) Caspases are a group of cysteine-containing proteases responsible for cellular maturation and reconstruction, number and quality of cell regulation, and apoptosis initiation of old cells or those that cannot perform their normal roles48. The apoptosis induced by the ectopic expression of HBx is associated with the alterations in mitochondrial membrane and caspases. The inhibition of Notch1 markedly promotes the apoptosis of LO2–HBx cells (stably transfected with HBx) via the caspase-9–caspase-3 pathway, whereas HBx can significantly improve the activity of Notch149. Recent research provided *in vitro* and *in vivo* evidence that HBx can enhance the susceptibility of hepatocytes toward oxidative stress-induced apoptotic killing by accelerating the loss of Mcl-1 protein, which is mainly caspase-3 dependent50. This effect is an important factor in controlling HBx-related apoptosis. #### HBx and the mitochondria HBV infection alters mitochondrial metabolism. The selective removal of damaged mitochondria is important for maintaining mitochondrial and cellular homeostasis. Raf serine/threonine kinases 1 (Raf-1) can translocate to the mitochondria, thereby protecting the cells from stress-mediated apoptosis. Recent studies showed that HBx stimulates the mitochondrial translocation of Raf-1, thereby contributing to the anti-apoptotic effect51. In addition, HBV induces the perinuclear clustering of the mitochondria and triggers the mitochondrial translocation of dynamin-related protein by stimulating its phosphorylation at Ser616, leading to mitochondrial fission. HBx protein serves a central function in promoting aberrant mitochondrial dynamics when expressed alone or in the context of the viral genome52. Previous evidence indicated that HBx-expressing cells have an elevated mitochondrial calcium uptake. The increased mitochondrial calcium uptake mediated by HBx can sustain higher cytosolic calcium and stimulate HBV replication to accelerate the development of HCC53. We speculate that the altered mitochondrial dynamics associated with HBx contributes to mitochondrial injury and HCC pathogenesis. #### HBx and programmed cell death 4 (PDCD4) PDCD4 was originally identified as a tumor-related gene in humans. PDCD4 is expressed ubiquitously in normal tissues, and the highest level is found in the liver. PDCD4 is a proapoptotic molecule involved in the TGF-β1-induced apoptosis in human HCC cells and a possible tumor suppressor in hepatocarcinogenesis54. PDCD4 is downregulated in clinical HCC specimens, and such downregulation is correlated with HBx9. The mechanisms underlying PDCD4 downregulation by HBx are partly through miR-2155. These data elucidate the anti-apoptotic function of the HBx viral protein and its contribution to HCC. ### HBx and immune system To successfully detect and eliminate invading pathogens by discriminating self from non-self, the mammalian immune system develops mechanisms, which comprise two distinct components, namely, the innate immunity and the adaptive immunity. In most multicellular organisms, the highly conserved innate immune system provides the first line of defense to limit infection by detecting pathogens using germline-encoded proteins56. The adaptive immunity, which is present only in vertebrates, detects non-self by recognizing peptide antigens through the receptors expressed on the surface of B and T cells57. HBx has different strategies for affecting immune system to promote the initiation and progression of HCC. #### HBx and innate immune The innate immune system is characterized by the production of type I interferon-α/β (IFN-α/β) cytokines and the activation of NK cells; this immune system has important roles in detecting and eliminating invading pathogens. Previous studies reported that HBx interacts with mitochondrial antiviral signaling (MAVS) and promotes the degradation of MAVS through Lys136 ubiquitin in MAVS protein to prevent the induction of IFN-β58. Further analysis on clinical samples revealed that MAVS protein is downregulated in the HCCs of HBV origin; this downregulation is correlated with the increased sensitivities of the primary murine hepatocytes isolated from HBx knock-in transgenic mice upon vesicular stomatitis virus (VSV) infections58. In addition, experimental evidence revealed that IFN induction is blocked by HBV replication in HepG2.2.15 cells. This effect is partially due to HBx, which impairs the IFN-β promoter activation using Sendai virus and the components implicated in signaling by viral sensors59. Moreover, HBx–siRNAs significantly promote PKR activation, leading to the higher production of type I IFN60. All these studies suggest that HBx can attenuate the antiviral response of the innate immune system and then promotes hepatocarcinogenesis. These findings delineated some functions of HBx-induced immune dysfunction in the development of HBV-associated HCC. #### HBx and adaptive immunity CD8+ T cells contribute to the clearance of HBV infection, and the lack of CD8+ T cells may be among the major factors that cause chronic HBV infection and HCC development. HBx expression induces the low production of interferon-γ and the apoptosis of CD8+ T cells without affecting CD8+ T cell proliferation61. As critical component in the tumor microenvironment, immunocytes can affect tumor prognosis by affecting adaptive immunity62. Interleukin-12 (IL-12) is a disulfide-linked heterodimeric cytokine with potent immunostimulatory activity. Recent data revealed that co-expression of HBx and IL-12 can induce a massive accumulation of CD8+ T cells, which inhibits stromal cell growth, such as in vascular endothelial cells62. Other studies showed the apparent infiltration of CD8+ T lymphocytes in the tumor of an HBx-treated group. Moreover, flow cytometric analysis showed that the component of CD8+ T lymphocytes after HBx vaccination is larger than that of the control groups. Moreover, the depletion of CD8+ T lymphocytes showed a complete abrogation of the antitumor immune activity *in vivo*63. These findings strongly confirm that HBx-related adaptive immunity is very important in the development of HCC. Overall, we summarized the mechanism of HBx in hepatocarcinogenesis. HBV-associated hepatocarcinogenesis can be viewed as a multi-factorial process, which includes direct and indirect mechanisms that may act cooperatively (Table 1). However, considering the numerous evidence implicating many different molecules and pathways in the development of HCC, the exact mechanisms by which HBx induces hepatocarcinogenesis remain ambiguous. Further research is required. The above mentioned data provide insights into the cell-transforming potential of HBx. HBx can interact with many transcription factors, regulatory RNAs, and various signaling transduction pathways. HBx is also involved in the modification of epigenetics. Additionally, HBx induces HBV-related HCC and can modulate apoptosis and the immune system. HBx serves a major function in cell survival and hepatocellular transformation during HBV infection. Biologically elucidating the significant activity of HBx may provide further insight into the role of HBx and may ultimately lead to the development of novel therapeutic strategies for managing HBV-related HCC and for the clinical therapy of HCC. The definition of HBx function *in vivo* should be the focus of future studies. View this table: [Table 1](http://www.cancerbiomed.org/content/11/3/182/T1) Table 1 Summary of HBx in hepatocarcinogenesis ## Footnotes * **Conflict of interest statement** No potential conflicts of interest are disclosed. * Received April 18, 2014. * Accepted June 2, 2014. * Copyright: © 2014, Cancer Biology & Medicine [https://creativecommons.org/licenses/by/4.0/](https://creativecommons.org/licenses/by/4.0/) This is an open access article distributed under the terms of the [Creative Commons Attribution License (CC BY) 4.0](https://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution and reproduction in any medium, provided the original author and source are credited. ## References 1. 1. Kew MC. Epidemiology of chronic hepatitis B virus infection, hepatocellular carcinoma, and hepatitis B virus-induced hepatocellular carcinoma. Pathol Biol (Paris) 2010;58:273–277. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1016/j.patbio.2010.01.005&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=20378277&link_type=MED&atom=%2Fcbm%2F11%2F3%2F182.atom) 2. 2. Zhang X, Zhang H, Ye L. Effects of hepatitis B virus X protein on the development of liver cancer. J Lab Clin Med 2006;147:58–66. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1016/j.lab.2005.10.003&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=16459163&link_type=MED&atom=%2Fcbm%2F11%2F3%2F182.atom) [Web of Science](http://www.cancerbiomed.org/lookup/external-ref?access_num=000235505700002&link_type=ISI) 3. 3. Seeger C, Mason WS. Hepatitis B virus biology. Microbiol Mol Biol Rev 2000;64:51–68. [Abstract/FREE Full Text](http://www.cancerbiomed.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6NDoibW1iciI7czo1OiJyZXNpZCI7czo3OiI2NC8xLzUxIjtzOjQ6ImF0b20iO3M6MTg6Ii9jYm0vMTEvMy8xODIuYXRvbSI7fXM6ODoiZnJhZ21lbnQiO3M6MDoiIjt9) 4. 4. Kodama K, Ogasawara N, Yoshikawa H, Murakami S. Nucleotide sequence of a cloned woodchuck hepatitis virus genome: evolutional relationship between hepadnaviruses. J Virol 1985;56:978–986. [Abstract/FREE Full Text](http://www.cancerbiomed.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6MzoianZpIjtzOjU6InJlc2lkIjtzOjg6IjU2LzMvOTc4IjtzOjQ6ImF0b20iO3M6MTg6Ii9jYm0vMTEvMy8xODIuYXRvbSI7fXM6ODoiZnJhZ21lbnQiO3M6MDoiIjt9) 5. 5. Nakatake H, Chisaka O, Yamamoto S, Matsubara K, Koshy R. Effect of X protein on transactivation of hepatitis B virus promoters and on viral replication. Virology 1993;195:305–314. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1006/viro.1993.1381&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=8337816&link_type=MED&atom=%2Fcbm%2F11%2F3%2F182.atom) 6. 6. Seto E, Mitchell PJ, Yen TS. Transactivation by the hepatitis B virus X protein depends on AP-2 and other transcription factors. Nature 1990;344:72–74. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1038/344072a0&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=2154703&link_type=MED&atom=%2Fcbm%2F11%2F3%2F182.atom) 7. 7. Zhang X, You X, Li N, Zhang W, Gagos S, Wang Q, et al. Involvement of hepatitis B virus X gene (HBx) integration in hepatocarcinogenesis via a recombination of HBx/Alu core sequence/subtelomeric DNA. FEBS Lett 2012;586:3215–3221. [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=22771476&link_type=MED&atom=%2Fcbm%2F11%2F3%2F182.atom) 8. 8. Ng RK, Lau CY, Lee SM, Tsui SK, Fung KP, Waye MM. cDNA microarray analysis of early gene expression profiles associated with hepatitis B virus X protein-mediated hepatocarcinogenesis. Biochem Biophys Res Commun 2004;322:827–835. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1016/j.bbrc.2004.07.188&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=15336538&link_type=MED&atom=%2Fcbm%2F11%2F3%2F182.atom) [Web of Science](http://www.cancerbiomed.org/lookup/external-ref?access_num=000223710500018&link_type=ISI) 9. 9. Qiu X, Dong S, Qiao F, Lu S, Song Y, Lao Y, et al. HBx-mediated miR-21 upregulation represses tumor-suppressor function of PDCD4 in hepatocellular carcinoma. Oncogene 2013;32:3296–3305. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1038/onc.2013.150&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=23604124&link_type=MED&atom=%2Fcbm%2F11%2F3%2F182.atom) 10. 10. Esteller M. Non-coding RNAs in human disease. Nat Rev Genet 2011;12:861–874. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1038/nrg3074&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=22094949&link_type=MED&atom=%2Fcbm%2F11%2F3%2F182.atom) 11. 11. Xu X, Fan Z, Kang L, Han J, Jiang C, Zheng X, et al. Hepatitis B virus X protein represses miRNA-148a to enhance tumorigenesis. J Clin Invest 2013;123:630–645. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1172/JCI64265&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=23321675&link_type=MED&atom=%2Fcbm%2F11%2F3%2F182.atom) [Web of Science](http://www.cancerbiomed.org/lookup/external-ref?access_num=000314553600020&link_type=ISI) 12. 12. Wu CS, Yen CJ, Chou RH, Chen JN, Huang WC, Wu CY, et al. Downregulation of microRNA-15b by hepatitis B virus X enhances hepatocellular carcinoma proliferation via fucosyltransferase 2-induced Globo H expression. Int J Cancer 2014;134:1638–1647. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1002/ijc.28501&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=24122375&link_type=MED&atom=%2Fcbm%2F11%2F3%2F182.atom) [Web of Science](http://www.cancerbiomed.org/lookup/external-ref?access_num=000329579500012&link_type=ISI) 13. 13. Zhang T, Zhang J, Cui M, Liu F, You X, Du Y, 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. 14. 14. 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. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1016/j.bbrc.2014.01.051&link_type=DOI) 15. 15. Lan SH, Wu SY, Zuchini R, Lin XZ, Su IJ, Tsai TF, et al. Autophagy suppresses tumorigenesis of hepatitis B virus-associated hepatocellular carcinoma through degradation of microRNA-224. Hepatology 2014;59:505–517. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1002/hep.26659&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=23913306&link_type=MED&atom=%2Fcbm%2F11%2F3%2F182.atom) 16. 16. Du Y, Kong G, You X, Zhang S, Zhang T, Gao Y, et al. 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. [Abstract/FREE Full Text](http://www.cancerbiomed.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6MzoiamJjIjtzOjU6InJlc2lkIjtzOjEyOiIyODcvMzEvMjYzMDIiO3M6NDoiYXRvbSI7czoxODoiL2NibS8xMS8zLzE4Mi5hdG9tIjt9czo4OiJmcmFnbWVudCI7czowOiIiO30=) 17. 17. Huang JF, Guo YJ, Zhao CX, Yuan SX, Wang Y, Tang GN, et al. 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. Hepatology 2013;57:1882–1892. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1002/hep.26195&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=23239537&link_type=MED&atom=%2Fcbm%2F11%2F3%2F182.atom) [Web of Science](http://www.cancerbiomed.org/lookup/external-ref?access_num=000318162200024&link_type=ISI) 18. 18. He Y, Meng XM, Huang C, Wu BM, Zhang L, Lv XW, et al. Long noncoding RNAs: Novel insights into hepatocelluar carcinoma. Cancer Lett 2014;344:20–27. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1016/j.canlet.2013.10.021&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=24183851&link_type=MED&atom=%2Fcbm%2F11%2F3%2F182.atom) 19. 19. Kew MC. Hepatitis B virus x protein in the pathogenesis of hepatitis B virus-induced hepatocellular carcinoma. J Gastroenterol Hepatol 2011;26 Suppl 1:144–152. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1111/j.1440-1746.2010.06546.x&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=21199526&link_type=MED&atom=%2Fcbm%2F11%2F3%2F182.atom) [Web of Science](http://www.cancerbiomed.org/lookup/external-ref?access_num=000285880400019&link_type=ISI) 20. 20. 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. [Abstract/FREE Full Text](http://www.cancerbiomed.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6MzoibWNiIjtzOjU6InJlc2lkIjtzOjEwOiIzMy8xNS8yODEwIjtzOjQ6ImF0b20iO3M6MTg6Ii9jYm0vMTEvMy8xODIuYXRvbSI7fXM6ODoiZnJhZ21lbnQiO3M6MDoiIjt9) 21. 21. Zhu YZ, Zhu R, Fan J, Pan Q, Li H, Chen Q, et al. Hepatitis B virus X protein induces hypermethylation of p16(INK4A) promoter via DNA methyltransferases in the early stage of HBV-associated hepatocarcinogenesis. J Viral Hepat 2010;17:98–107. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1111/j.1365-2893.2009.01156.x&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=19732323&link_type=MED&atom=%2Fcbm%2F11%2F3%2F182.atom) 22. 22. Park SH, Jung JK, Lim JS, Tiwari I, Jang KL. Hepatitis B virus X protein overcomes all-trans retinoic acid-induced cellular senescence by downregulating levels of p16 and p21 via DNA methylation. J Gen Virol 2011;92:1309–1317. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1099/vir.0.029512-0&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=21325480&link_type=MED&atom=%2Fcbm%2F11%2F3%2F182.atom) [Web of Science](http://www.cancerbiomed.org/lookup/external-ref?access_num=000291370100007&link_type=ISI) 23. 23. Tong A, Gou L, Lau QC, Chen B, Zhao X, Li J, et al. Proteomic profiling identifies aberrant epigenetic modifications induced by hepatitis B virus X protein. J Proteome Res 2009;8:1037–1046. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1021/pr8008622&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=19117405&link_type=MED&atom=%2Fcbm%2F11%2F3%2F182.atom) 24. 24. Yoo YG, Na TY, Seo HW, Seong JK, Park CK, Shin YK, et al. Hepatitis B virus X protein induces the expression of MTA1 and HDAC1, which enhances hypoxia signaling in hepatocellular carcinoma cells. Oncogene 2008;27:3405–3413. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1038/sj.onc.1211000&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=18264140&link_type=MED&atom=%2Fcbm%2F11%2F3%2F182.atom) [Web of Science](http://www.cancerbiomed.org/lookup/external-ref?access_num=000256309900005&link_type=ISI) 25. 25. Arzumanyan A, Friedman T, Kotei E, Ng IO, Lian Z, Feitelson MA. Epigenetic repression of E-cadherin expression by hepatitis B virus x antigen in liver cancer. Oncogene 2012;31:563–572. [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=21706058&link_type=MED&atom=%2Fcbm%2F11%2F3%2F182.atom) [Web of Science](http://www.cancerbiomed.org/lookup/external-ref?access_num=000300221800003&link_type=ISI) 26. 26. Shon JK, Shon BH, Park IY, Lee SU, Fa L, Chang KY, et al. 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. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1016/j.virusres.2008.09.006&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=18948152&link_type=MED&atom=%2Fcbm%2F11%2F3%2F182.atom) [Web of Science](http://www.cancerbiomed.org/lookup/external-ref?access_num=000262606700003&link_type=ISI) 27. 27. Brooks CL, Gu W. How does SIRT1 affect metabolism, senescence and cancer? Nat Rev Cancer 2009;9:123–128. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1038/nrc2562&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=19132007&link_type=MED&atom=%2Fcbm%2F11%2F3%2F182.atom) [Web of Science](http://www.cancerbiomed.org/lookup/external-ref?access_num=000263180900013&link_type=ISI) 28. 28. Srisuttee R, Koh SS, Kim SJ, Malilas W, Boonying W, Cho IR, 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. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.3892/or.2012.1798&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=22562294&link_type=MED&atom=%2Fcbm%2F11%2F3%2F182.atom) 29. 29. Srisuttee R, Koh SS, Malilas W, Moon J, Cho IR, Jhun BH, et al. SIRT1 sensitizes hepatocellular carcinoma cells expressing hepatitis B virus X protein to oxidative stress-induced apoptosis. Biochem Biophys Res Commun 2012;429:45–50. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1016/j.bbrc.2012.10.102&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=23137540&link_type=MED&atom=%2Fcbm%2F11%2F3%2F182.atom) 30. 30. Vivanco I, Sawyers CL. The phosphatidylinositol 3-Kinase AKT pathway in human cancer. Nat Rev Cancer 2002;2:489–501. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1038/nrc839&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=12094235&link_type=MED&atom=%2Fcbm%2F11%2F3%2F182.atom) [Web of Science](http://www.cancerbiomed.org/lookup/external-ref?access_num=000180447000010&link_type=ISI) 31. 31. Khattar E, Mukherji A, Kumar V. Akt augments the oncogenic potential of the HBx protein of hepatitis B virus by phosphorylation. FEBS J 2012;279:1220–1230. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1111/j.1742-4658.2012.08514.x&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=22309289&link_type=MED&atom=%2Fcbm%2F11%2F3%2F182.atom) 32. 32. Lee WP, Lan KH, Li CP, Chao Y, Lin HC, Lee SD. Pro-apoptotic or anti-apoptotic property of X protein of hepatitis B virus is determined by phosphorylation at Ser31 by Akt. Arch Biochem Biophys 2012;528:156–162. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1016/j.abb.2012.08.008&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=22982405&link_type=MED&atom=%2Fcbm%2F11%2F3%2F182.atom) 33. 33. Huang WC, Chen WS, Chen YJ, Wang LY, Hsu SC, Chen CC, et al. Hepatitis B virus X protein induces IKKα nuclear translocation via Akt-dependent phosphorylation to promote the motility of hepatocarcinoma cells. J Cell Physiol 2012;227:1446–1454. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1002/jcp.22860&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=21618535&link_type=MED&atom=%2Fcbm%2F11%2F3%2F182.atom) 34. 34. Wei X, Tan C, Tang C, Ren G, Xiang T, Qiu Z, et al. Epigenetic repression of miR-132 expression by the hepatitis B virus x protein in hepatitis B virus-related hepatocellular carcinoma. Cell Signal 2013;25:1037–1043. 35. 35. Grigoryan T, Wend P, Klaus A, Birchmeier W. Deciphering the function of canonical Wnt signals in development and disease: conditional loss- and gain-of-function mutations of beta-catenin in mice. Genes Dev 2008;22:2308–2341. [Abstract/FREE Full Text](http://www.cancerbiomed.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6ODoiZ2VuZXNkZXYiO3M6NToicmVzaWQiO3M6MTA6IjIyLzE3LzIzMDgiO3M6NDoiYXRvbSI7czoxODoiL2NibS8xMS8zLzE4Mi5hdG9tIjt9czo4OiJmcmFnbWVudCI7czowOiIiO30=) 36. 36. Hsieh A, Kim HS, Lim SO, Yu DY, Jung G. Hepatitis B viral X protein interacts with tumor suppressor adenomatous polyposis coli to activate Wnt/β-catenin signaling. Cancer Lett 2011;300:162–172. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1016/j.canlet.2010.09.018&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=20971552&link_type=MED&atom=%2Fcbm%2F11%2F3%2F182.atom) [Web of Science](http://www.cancerbiomed.org/lookup/external-ref?access_num=000285813900007&link_type=ISI) 37. 37. Cha MY, Kim CM, Park YM, Ryu WS. Hepatitis B virus X protein is essential for the activation of Wnt/beta-catenin signaling in hepatoma cells. Hepatology 2004;39:1683–1693. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1002/hep.20245&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=15185310&link_type=MED&atom=%2Fcbm%2F11%2F3%2F182.atom) [Web of Science](http://www.cancerbiomed.org/lookup/external-ref?access_num=000221849000026&link_type=ISI) 38. 38. Gao B, Wang H, Lafdil F, Feng D. STAT proteins - key regulators of anti-viral responses, inflammation, and tumorigenesis in the liver. J Hepatol 2012;57:430–441. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1016/j.jhep.2012.01.029&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=22504331&link_type=MED&atom=%2Fcbm%2F11%2F3%2F182.atom) 39. 39. Waris G, Huh KW, Siddiqui A. Mitochondrially associated hepatitis B virus X protein constitutively activates transcription factors STAT-3 and NF-kappa B via oxidative stress. Mol Cell Biol 2001;21:7721–7730. [Abstract/FREE Full Text](http://www.cancerbiomed.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6MzoibWNiIjtzOjU6InJlc2lkIjtzOjEwOiIyMS8yMi83NzIxIjtzOjQ6ImF0b20iO3M6MTg6Ii9jYm0vMTEvMy8xODIuYXRvbSI7fXM6ODoiZnJhZ21lbnQiO3M6MDoiIjt9) 40. 40. Teng J, Wang X, Xu Z, Tang N. HBx-dependent activation of Twist mediates STAT3 control of epithelium-mesenchymal transition of liver cells. J Cell Biochem 2013;114:1097–1104. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1002/jcb.24450&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=23161342&link_type=MED&atom=%2Fcbm%2F11%2F3%2F182.atom) 41. 41. Fan H, Yan X, Zhang Y, Zhang X, Gao Y, Xu Y, et al. Increased expression of Gp96 by HBx-induced NF-κB activation feedback enhances hepatitis B virus production. PLoS One 2013;8:e65588. 42. 42. Luo J, Zhou H, Wang F, Xia X, Sun Q, Wang R, et al. The hepatitis B virus X protein downregulates NF-κB signaling pathways through decreasing the Notch signaling pathway in HBx-transformed L02 cells. Int J Oncol 2013;42:1636–1643. [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=23450368&link_type=MED&atom=%2Fcbm%2F11%2F3%2F182.atom) 43. 43. Shukla R, Yue J, Siouda M, Gheit T, Hantz O, Merle P, et al. Proinflammatory cytokine TNF-α increases the stability of hepatitis B virus X protein through NF-κB signaling. Carcinogenesis 2011;32:978–985. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1093/carcin/bgr057&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=21459755&link_type=MED&atom=%2Fcbm%2F11%2F3%2F182.atom) [Web of Science](http://www.cancerbiomed.org/lookup/external-ref?access_num=000292337800005&link_type=ISI) 44. 44. Wang Y, Cui F, Lv Y, Li C, Xu X, Deng C, et al. HBsAg and HBx knocked into the p21 locus causes hepatocellular carcinoma in mice. Hepatology 2004;39:318–324. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1002/hep.20076&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=14767984&link_type=MED&atom=%2Fcbm%2F11%2F3%2F182.atom) [Web of Science](http://www.cancerbiomed.org/lookup/external-ref?access_num=000220375600010&link_type=ISI) 45. 45. Kim CM, 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. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1038/351317a0&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=2034275&link_type=MED&atom=%2Fcbm%2F11%2F3%2F182.atom) [Web of Science](http://www.cancerbiomed.org/lookup/external-ref?access_num=A1991FM97600061&link_type=ISI) 46. 46. Wu BK, Li CC, Chen HJ, Chang JL, Jeng KS, Chou CK, et al. Blocking of G1/S transition and cell death in the regenerating liver of Hepatitis B virus X protein transgenic mice. Biochem Biophys Res Commun 2006;340:916–928. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1016/j.bbrc.2005.12.089&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=16403455&link_type=MED&atom=%2Fcbm%2F11%2F3%2F182.atom) [Web of Science](http://www.cancerbiomed.org/lookup/external-ref?access_num=000235092000026&link_type=ISI) 47. 47. Zhang WY, Cai N, Ye LH, Zhang XD. Transformation of human liver L-O2 cells mediated by stable HBx transfection. Acta Pharmacol Sin 2009;30:1153–1161. [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=19578387&link_type=MED&atom=%2Fcbm%2F11%2F3%2F182.atom) 48. 48. Hengartner MO. The biochemistry of apoptosis. Nature 2000;407:770–776. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1038/35037710&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=11048727&link_type=MED&atom=%2Fcbm%2F11%2F3%2F182.atom) [Web of Science](http://www.cancerbiomed.org/lookup/external-ref?access_num=000089773900051&link_type=ISI) 49. 49. Sun Q, Wang R, Wang Y, Luo J, Wang P, Cheng B. Notch1 is a potential therapeutic target for the treatment of human hepatitis B virus X protein-associated hepatocellular carcinoma. Oncol Rep 2014;31:933–939. 50. 50. Hu L, Chen L, Yang G, Li L, Sun H, Chang Y, et al. HBx sensitizes cells to oxidative stress-induced apoptosis by accelerating the loss of Mcl-1 protein via caspase-3 cascade. Mol Cancer 2011;10:43. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1186/1476-4598-10-43&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=21504623&link_type=MED&atom=%2Fcbm%2F11%2F3%2F182.atom) 51. 51. Chen J, Siddiqui A. Hepatitis B virus X protein stimulates the mitochondrial translocation of Raf-1 via oxidative stress. J Virol 2007;81:6757–6760. [Abstract/FREE Full Text](http://www.cancerbiomed.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6MzoianZpIjtzOjU6InJlc2lkIjtzOjEwOiI4MS8xMi82NzU3IjtzOjQ6ImF0b20iO3M6MTg6Ii9jYm0vMTEvMy8xODIuYXRvbSI7fXM6ODoiZnJhZ21lbnQiO3M6MDoiIjt9) 52. 52. Kim SJ, Khan M, Quan J, Till A, Subramani S, Siddiqui A. Hepatitis B virus disrupts mitochondrial dynamics: induces fission and mitophagy to attenuate apoptosis. PLoS Pathog 2013;9:e1003722. 53. 53. Yang B, Bouchard MJ. The hepatitis B virus X protein elevates cytosolic calcium signals by modulating mitochondrial calcium uptake. J Virol 2012;86:313–327. [Abstract/FREE Full Text](http://www.cancerbiomed.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6MzoianZpIjtzOjU6InJlc2lkIjtzOjg6Ijg2LzEvMzEzIjtzOjQ6ImF0b20iO3M6MTg6Ii9jYm0vMTEvMy8xODIuYXRvbSI7fXM6ODoiZnJhZ21lbnQiO3M6MDoiIjt9) 54. 54. Zhang H, Ozaki I, Mizuta T, Hamajima H, Yasutake T, Eguchi Y, et al. Involvement of programmed cell death 4 in transforming growth factor-beta1-induced apoptosis in human hepatocellular carcinoma. Oncogene 2006;25:6101–6112. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1038/sj.onc.1209634&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=16682950&link_type=MED&atom=%2Fcbm%2F11%2F3%2F182.atom) [Web of Science](http://www.cancerbiomed.org/lookup/external-ref?access_num=000241101200007&link_type=ISI) 55. 55. Damania P, Sen B, Dar SB, Kumar S, Kumari A, Gupta E, et al. 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. 56. 56. Hoffmann JA, Kafatos FC, Janeway CA, Ezekowitz RA. Phylogenetic perspectives in innate immunity. Science 1999;284:1313–1318. [Abstract/FREE Full Text](http://www.cancerbiomed.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6Mzoic2NpIjtzOjU6InJlc2lkIjtzOjEzOiIyODQvNTQxOC8xMzEzIjtzOjQ6ImF0b20iO3M6MTg6Ii9jYm0vMTEvMy8xODIuYXRvbSI7fXM6ODoiZnJhZ21lbnQiO3M6MDoiIjt9) 57. 57. Akira S, Takeda K. Toll-like receptor signalling. Nat Rev Immunol 2004;4:499–511. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1038/nri1391&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=15229469&link_type=MED&atom=%2Fcbm%2F11%2F3%2F182.atom) [Web of Science](http://www.cancerbiomed.org/lookup/external-ref?access_num=000222435700012&link_type=ISI) 58. 58. Wei C, Ni C, Song T, Liu Y, Yang X, Zheng Z, et al. The hepatitis B virus X protein disrupts innate immunity by downregulating mitochondrial antiviral signaling protein. J Immunol 2010;185:1158–1168. [Abstract/FREE Full Text](http://www.cancerbiomed.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6ODoiamltbXVub2wiO3M6NToicmVzaWQiO3M6MTA6IjE4NS8yLzExNTgiO3M6NDoiYXRvbSI7czoxODoiL2NibS8xMS8zLzE4Mi5hdG9tIjt9czo4OiJmcmFnbWVudCI7czowOiIiO30=) 59. 59. Jiang J, Tang H. Mechanism of inhibiting type I interferon induction by hepatitis B virus X protein. Protein Cell 2010;1:1106–1117. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1007/s13238-010-0141-8&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=21213104&link_type=MED&atom=%2Fcbm%2F11%2F3%2F182.atom) [Web of Science](http://www.cancerbiomed.org/lookup/external-ref?access_num=000208511900008&link_type=ISI) 60. 60. Han Q, Zhang C, Zhang J, Tian Z. Involvement of activation of PKR in HBx-siRNA-mediated innate immune effects on HBV inhibition. PLoS One 2011;6:e27931. 61. 61. Lee MJ, Jin YH, Kim K, Choi Y, Kim HC, Park S. Expression of hepatitis B virus x protein in hepatocytes suppresses CD8 T cell activity. Immune Netw 2010;10:126–134. [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=20844737&link_type=MED&atom=%2Fcbm%2F11%2F3%2F182.atom) 62. 62. He H, Fan P, Yin T, Chen Q, Shi H, Liu S, et al. Local delivery of recombinant adenovirus expressing hepatitis B virus X protein and interleukin-12 results in antitumor effects via inhibition of hepatoma cell growth and intervention of tumor microenvironment. Int J Mol Med 2012;30:599–605. [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=22710287&link_type=MED&atom=%2Fcbm%2F11%2F3%2F182.atom) 63. 63. Li Y, Cheng P, Wen Y, Chen P, Yang L, Zhao X, et al. T lymphocyte responses against hepatitis B virus-related hepatocellular carcinoma induced by adenovirus vaccine encoding HBx. Int J Mol Med 2010;26:869–876. [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=21042781&link_type=MED&atom=%2Fcbm%2F11%2F3%2F182.atom) 64. 64. Yang WJ, Chang CJ, Yeh SH, Lin WH, Wang SH, Tsai TF, et al. Hepatitis B virus X protein enhances the transcriptional activity of the androgen receptor through c-Src and glycogen synthase kinase-3beta kinase pathways. Hepatology 2009;49:1515–1524. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1002/hep.22833&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=19205031&link_type=MED&atom=%2Fcbm%2F11%2F3%2F182.atom)