Open access peer-reviewed chapter

Pathogenesis of Hepatitis B Virus Associated Chronic Liver Disease

Written By

Mark A. Feitelson

Submitted: 27 March 2018 Reviewed: 25 June 2018 Published: 05 November 2018

DOI: 10.5772/intechopen.79746

From the Edited Volume

Liver Cancer

Edited by Ahmed Lasfar

Chapter metrics overview

2,411 Chapter Downloads

View Full Metrics

Abstract

Hepatitis B virus (HBV) infection is associated with chronic liver diseases (CLD), which progress from hepatitis to fibrosis, cirrhosis, and finally hepatocellular carcinoma (HCC) over 30–50 years. The pathogenesis of CLD is immune mediated, which is characterized by persistent immune responses against virus infected hepatocytes. During bouts of CLD, the virus gene encoding the hepatitis B x antigen (HBx) is increasingly found integrated at multiple sites within the human genome. Many of these integrated templates express HBx, which is a trans-regulatory protein that supports virus gene expression and replication on one hand, but also alters patterns of gene expression in the infected cell. HBx alters gene expression by constitutively activating signal transduction pathways in the cytoplasm and promoting epigenetic mediated changes in the expression of cellular genes. In doing so, HBx contributes to the persistence of virus infected cells and to the pathogenesis of CLD by triggering multiple hallmarks which are characteristic of cancer.

Keywords

  • hepatitis B virus
  • chronic liver disease
  • hepatocellular carcinoma
  • hepatitis B x
  • immune mediated liver disease
  • epigenetics
  • hallmarks of cancer

1. Introduction

Hepatitis B virus (HBV) is a blood-borne virus that infects the liver. Until the discovery of the virus in the 1960s [1], it was transmitted sexually and by transfusion of contaminated blood and blood fractions. Today, the virus has been virtually eliminated from the blood supply by a simple blood test while infection has been prevented by a highly efficacious vaccine [2, 3]. Prior to establishment of vaccination programs in various countries, infants born to infected mothers replicating virus often acquired the virus at birth by exposure to contaminated maternal blood. More than 90% of these children became HBV carriers, characterized by the persistence of virus or virus antigens in their blood for years to decades. These children were a high risk for the development of chronic liver disease (CLD), which progressed from hepatitis, to fibrosis, cirrhosis, and hepatocellular carcinoma (HCC) [4]. Fortunately, newborns in many countries receive the HBV vaccine at birth, which helps to prevent mother-to-infant transmission as well as protect from exposure later in life. Among unvaccinated adults engaging in unprotected sex, roughly 5–10% become carriers, and these individuals are also at high risk for the development of CLD and HCC. Although estimates vary, there are ~300–350 million carriers of HBV worldwide [5]. HCC is the sixth most prevalent cancer worldwide, with about 600,000 newly diagnosed cases annually, and the second leading cause of cancer deaths [6]. Interferon, and in more recent years, powerful nucleoside analogs, have successfully treated patients with chronic hepatitis B, but presently there is no cure [7, 8]. HCC is curable by surgical resection, but this is often accompanied by relapse. Dozens of drugs, alone or in combination, have been evaluated in clinical trials for patients with advanced HCC, but only the multi-kinase inhibitors, Sorafenib and regorafenib, and the immune checkpoint inhibitor nivolumab, have been useful in modestly extending the lifespan of such patients [9, 10]. Given that the carrier state and CLD are the major risk factors for HCC [11], there is strong rationale to better understand the host-virus relationship that contributes to the pathogenesis of chronic infection.

Advertisement

2. Variations in pathogenesis

A hallmark in the pathogenesis of HBV infection is its’ variability. Among acutely infected adults, up to 65% develop a subclinical infection characterized only by the appearance of one or more viral antibodies in the blood, while another 25% develop acute resolving infection, which may or may not include a bout of hepatitis. The remaining 10% of patients develop chronic infection (i.e., the persistence of virus and virus antigens in the blood for more than 6 months). In chimpanzees [12] and woodchucks [13], acute infections are characterized by the nearly complete clearance of virus from the blood and liver followed by seroconversion from surface antigen to corresponding antibody. In this case, virus is mostly cleared by non-cytolytic cytokines (e.g., interferon gamma [IFNɣ] and tumor necrosis factor alpha [TNFα]) prior to the appearance of T and other inflammatory cells in the liver, suggesting that most virus clearance occurs prior to the development of acute hepatitis. Further work showed that CD4+ and CD8+ T cells, natural killer (NK) cells, Fas, various IFNs and corresponding receptors, and the TNF receptor 1 participate in virus clearance, suggesting redundant pathways inhibit HBV replication in the liver [14]. The subsequent contribution of a T cell response appears to clear virus infected cells by cytolytic mechanisms involving Fas and granzymes. In this context, CD4+ T cells are required to prime CD8+ T cells to facilitate virus elimination in acute infection [14]. When this happens in acute, resolving infection, the T cell response to HBV is vigorous, polyclonal and multi-specific, while among those who go on to develop chronic infection, adaptive immunity is relatively weak and narrowly focused, suggesting that clearance of HBV is T cell dependent. When T cell responses are not adequate, CLD may develop and progress to cirrhosis and HCC. However, CLD may spontaneously resolve at any of these stages. While the origin of this variability is not completely characterized, it is clear that the ability of the host to mount adaptive immune responses is a key element to limiting virus spread.

Advertisement

3. Contributions of hepatitis B surface, core and e antigens to the pathogenesis of chronic infection

3.1. Hepatitis B surface antigen (HBsAg)

HBV is a small virus consisting of only four open reading frames (ORF) [15]. One ORF encodes a family of envelope polypeptides (Figure 1). The major envelope polypeptide, HBsAg, triggers neutralizing antibody which is central to virus clearance after acute exposure and is the major component of the HBV vaccine [2]. HBsAg polypeptides are transmembrane proteins and glycoproteins that are on the envelope of virus particles, and are also secreted as small, spherical and variably long filamentous subviral particles that lack the virus nucleocapsid and HBV DNA. It is thought that these subviral particles, which are produced at concentrations several logs above that of infectious virus particles, absorb neutralizing antibody and trigger immunological tolerance, both of which promote virus persistence in the blood. Moreover, in patients with CLD, there does not seem to be any correlation between intrahepatic HBsAg expression patterns and inflammatory infiltrates [16, 17], nor have HBsAg specific T cell clones been isolated from such patients [18]. In addition, T cell sensitization to HBsAg in acute and chronic HBV infection is usually undetectable [19], so while HBsAg clearance occurs in acute, resolving infections, it is not clear that it is an immunological target in established infections.

Figure 1.

Genetic organization of HBV showing the ORFs (in color). The positions of enhancer 1 (EN1) and 2 (EN2) are also shown. The direct repeat 1 (DR1) and 2 (DR2) sequences at the ends of the long and short DNA strands are also indicted. The pregenomic RNA (3.5 kb) is greater than genome length, while the 2.1 and 2.4 kb subgenomic mRNAs encode surface antigen polypeptides, and the 0.7 kb mRNA encodes the X protein. Reproduced from [20] with permission.

3.2. Hepatitis B core antigen (HBcAg)

The second ORF, or core gene, encodes the hepatitis B core antigen (HBcAg) or nucleocapsid protein that polymerizes as an icosahedron around the virus replication complex, the latter of which consists of the virus nucleic acid and HBV encoded polymerase [20]. The fact that the pregenomic RNA and the reverse transcribed viral DNA product are sequestered within a nucleocapsid means that they are not readily detected by pattern recognition receptors, (e.g., toll-like receptors, retinoic acid inducible gene 1 [RIG-1], and mitochondrial anti-viral signaling [MAVS]) that trigger innate immunity [21]. Moreover, innate immune responses do not develop in the liver of acutely infected chimpanzees [22], suggesting that HBV replication and spread may be conducted in “stealth” mode with virus nucleocapsids upon infection and again during virus replication. If so, then this may explain why up to 70% of acutely infected adults who become carriers do not develop CLD. However, carriers who develop CLD also have intrahepatic core antigen, suggesting that HBcAg may be an important immunological target in CLD [23]. Alternatively, patients with acute, resolving hepatitis show a vigorous peripheral blood mononuclear cell response to HBcAg that is temporally associated with the clearance of HBsAg, while in patients with chronic infection, T cell responsiveness to HBcAg is relatively weak, providing an opportunity for HBV to spread in the liver and establish a chronic infection [19].

3.3. Hepatitis B e antigen (HBeAg)

A proteolytic fragment of HBcAg, known as HBeAg, is secreted into the circulation and serves as a surrogate marker of virus replication. Seroconversion from HBeAg to anti-HBe is usually accompanied by a significant decrease in virus replication in both the liver and blood and resolution of CLD [24]. The detection of HBcAg specific cytotoxic T lymphocytes (CTL) is associated with the clearance of virus replication, often a transient exacerbation of CLD, and seroconversion to anti-HBe during the natural history of infection [24], suggesting that HBcAg is an important virus target in CLD. HBcAg specific T cells have been detected in the peripheral blood and liver [18, 25] of patients with CLD, suggesting that HBcAg is an immunological target in chronic hepatitis B. Interestingly, HBeAg in serum may attenuate immune responses against virus infected liver, because some patients who develop mutations in HBV that no longer express HBeAg, continue to support high levels of virus replication and ongoing, CLD [26, 27]. In fact, HBeAg appears to be a T cell tolerogen that down-regulates immune responses against HBcAg [28]. HBeAg may also stimulate the appearance of regulatory dendritic cells, which would also suppress virus specific immunity and promote virus persistence [29] by up-regulating the expression of suppressor of cytokine signaling 2 (SOCS2), which in turn represses IFN signaling, thereby blunting innate anti-viral responses and promoting virus persistence [30]. Thus, HBeAg polypeptides, like subviral HBsAg particles, promote chronicity by acting as tolerogens.

3.4. Hepatitis B polymerase

The HBV encoded polymerase, encoded by a third ORF, has DNA dependent and RNA dependent DNA polymerase (DNAp) activities, and RNase H activity. Upon infection, the partially double stranded viral DNA is made fully double stranded by the endogenous DNAp activity [20] (Figure 2). The HBV genome then appears as a supercoiled mini-chromosome in the nuclei of infected cells, and this acts as a template for the transcription of subgenomic RNAs and a greater than genome length pre-genomic RNA. The latter then migrates into the cytoplasm, where it is packaged with the virus polymerase into nascent (“immature”) core (or nucleocapsid) particles, where the pregenomic RNA is reverse transcribed into minus strand DNA, with the latter then being used as a template for partial plus strand synthesis just prior to the budding and secretion of progeny virus (Figure 2). Some immature core particles are recycled into the nucleus to replenish the pool of covalently closed circular (ccc) HBV DNA. Although the HBV polymerase triggers antibody responses [31], there is no evidence that immune responses against the polymerase directly impact pathogenesis or virus persistence. However, HBV polymerase inhibits RIG-1 and nuclear factor kappa B (NF-ĸB) induction of IFNβ, suggesting that the polymerase could block innate signaling [32, 33], thereby contributing to virus persistence.

Figure 2.

General scheme of HBV replication. See the text for additional details. Reproduced from [20] with permission.

Advertisement

4. Relationship between persistent virus replication, integration of HBV DNA, and the risk for the development of HCC

There is evidence to suggest that persistent, high levels of HBV replication correlate with the progression of CLD to HCC [34]. However, independent work showed an elevated risk for HCC among patients with CLD but low virus titers [35, 36]. Other observations have shown no correlation between HBV DNA levels in serum (>105 copies/ml) and histological grade or stage of liver disease in carriers [37, 38]. In addition, it is controversial as to whether long term nucleoside analog therapy resulted in a decreased risk for the development of HCC [6, 35]. Given that HBV is not directly cytopathic [39], that carriers with high levels of HBV DNA in serum are often asymptomatic, and that the pathogenesis of CLD is immune mediated [17, 40], a correlation between virus replication and CLD may contribute to, but not by itself, determine disease progression. Moreover, most carriers with CLD who develop cirrhosis and HCC have long since seroconverted from HBeAg (reflecting high levels of virus replication) to anti-HBe (reflecting low or undetectable virus replication), indicating that disease progression may occur at low virus titers [36]. Among patients with sustained high levels of HBV replication and successive bouts of CLD, there is a wave of liver regeneration following each episode of hepatitis to restore full liver function. At these times, fragments of HBV DNA, mostly encoding the HBx ORF (and sometimes the HBx plus preS/S ORFs), become integrated at multiple sites within host DNA [41, 42] (Figure 3). Over time, these integration events result in increased intrahepatic expression levels of HBx that alter patterns of host (and support virus) gene expression (Figure 3). HBV integrates early after infection, not only in permissive liver cell lines, but also in non-replicating primary human hepatocytes [43]. Many fragments of integrated HBV DNA encode HBx that is capable of trans-activation [44]. Although the relatively low levels of HBx made from the virus mini-chromosome support virus gene expression and replication, it is hypothesized that as intrahepatic levels of HBx increase [45] (Figure 3), it epigenetically alter the expression patterns of selected host genes [46] that contribute to both virus persistence and to malignant transformation. Thus, the changing intrahepatic levels of HBx promote virus persistence and ultimately, contribute to malignant transformation [47].

Figure 3.

Natural history of chronic hepatitis B featuring the progressive lesions that develop in CLD compared to increased number of integration events, many of which produce functional HBx (modified from [53] with permission).

4.1. Covalently closed circular HBV (ccc) DNA

Given that the current treatment of chronic hepatitis B with nucleoside analogs is not curative, there has been a major effort to eliminate ccc DNA [47], especially since ccc DNA is the template for all virus transcripts. Since nucleoside analogs do not eliminate integrated HBV templates or the HBV mini-chromosome, continued virus gene expression from these templates will drive pathogenesis toward HCC. Formation of ccc DNA is a complex process that involves a variety of host proteins, including several DNA polymerases [48] that could potentially be therapeutic targets, although this approach may be accompanied by toxicity. As outlined below, HBx regulates the formation, function and intracellular copy number of ccc DNA by several epigenetic mechanisms that involve altered expression of histone methyltransferases and histone deacetylases, by promoting degradation of the anti-viral restriction factor Smc5/6, and by increasing expression of DNA methyltransferases [48]. Anti-viral immune responses in which selected cytokines mediate non-cytolytic degradation of ccc DNA have also been documented in vitro [48, 49]. Among these, IFN alpha up-regulated expression of APOBEC3 nuclear deaminase resulted in a modest reduction in ccc DNA copy number via deamination [50]. Gene editing approaches, such as CRISPR/Cas9 have also been demonstrated to work in vitro and in vivo [51], but off-target effects, ability to access and act on all susceptible cells, and recognition of all HBV genotypes, remain to be addressed. In addition, the recent finding of ccc host DNA in both normal and tumor cells, as a mechanism whereby host cells regulate gene expression [52], implies that targeting ccc DNA may also have toxic effects on the treated cells whether or not they are virally infected. Thus, it is not clear whether this approach in a liver which is already damaged will exacerbate that damage and/or have an anti-tumor effect.

Advertisement

5. Contribution of HBx to pathogenesis of CLD by regulation of HBV replication

HBx, the trans-activation protein of HBV, trans-activates virus gene expression and replication in vitro [54, 55]. The contribution of this regulatory protein to virus persistence in the carrier state was shown in woodchucks experimentally infected with the HBV-like virus, woodchuck hepatitis virus (WHV). Wild type WHV readily establishes a chronic infection, characterized by persistent virus replication and CLD that progresses to HCC [56]. However, experimental infection with a mutant of WHV that does not encode woodchuck hepatitis x (WHx) antigen yielded no carrier state and no CLD [57, 58], suggesting that trans-activation of virus gene expression and replication is central to the establishment of the carrier state. Among infected woodchucks, there was co-staining between WHV core antigen (where virus replication takes place) and WHx [59, 60], while in human infection, HBx often co-existed with HBe in serum [61] and replication complexes (i.e., with HBcAg) in the liver [62]. Thus, HBx expression is associated with virus replication.

5.1. Mechanisms regulating HBV replication

At the molecular level, HBx regulates HBV replication by binding to various cellular proteins. For example, HBx binds to jumonji C-domain-containing 5 (JMJD5), a arginyl-hydroxylase, which promotes the expression of transcription factors (e.g., such as hepatocyte nuclear factors 3 gamma and 4 alpha [HNF3G and HNF4A] and CCAAT/enhancer-binding protein alpha) that facilitate hepatocyte differentiation [63]. Given that HBV replicates in differentiated hepatocytes, the binding of HBx to JMJD5 facilitates HBV replication via epigenetic alterations in host gene expression. In addition, HBx promotes the formation of ccc DNA by recruiting the transcriptional scaffold, p300; the cAMP response element binding protein CREB; the CREB transcription factor binding protein, CBP; the histone acetyltransferase p300/CBP-associated factor, as well as the histone deactylases HDAC1, Sirt1 [48] and Sirt2 [12]. Once ccc DNA is formed, HBx up-regulates HBV replication, in part, by binding to cullin4-damage specific DNA binding protein (CUL4-DDB1) ubiquitin ligase [64, 65], suggesting that HBx may function, at least in part, at the level of the proteasome. HBx modulates proteasome activity by direct binding to the 26S proteasomal subunit [66], which is responsible for degradation of HBx and several anti-viral proteins. One of the latter is Smc5/6, which is involved in the structural maintenance of chromosomes (i.e., genome stability) and DNA repair [67]. Smc5/6 and HBx bind to the HBV mini-chromosome [67, 68], resulting in epigenetic changes of virus gene expression. HBx binding to CUL4-DDB1 triggers altered enzymatic activity of the E3 ligase CRL4, which then stimulates the ubiquitination and subsequent proteasomal degradation of Smc5/6 [68, 69, 70], thereby promoting virus replication. Other anti-viral systems, such as IFN induced APOBEC3A [50], may also be similarly degraded. In this context, HBV is not very good in triggering innate immunity, which may underscore why there are hundreds of millions of carriers worldwide [71]. As mentioned above, sequestration and reverse transcription of pregenomic HBV RNA in immature nucleocapsids (Figure 2) may block the induction of innate immunity. In addition, although HBV replication is exquisitely sensitive to inhibition by IFNs, HBx appears to block IFN expression and signaling [72, 73, 74], suggesting that both innate and adaptive immunity could be compromised, thereby permitting virus persistence. Under these circumstances, CLD would continue to damage the liver while being unable to resolve the virus infection. HBx also regulates HBV replication by stimulating the expression of DNA methyl-transferases (DNMTs), which suppresses HBV transcription via DNA methylation [75]. DNMTs also methylate tumor suppressor genes, thereby down-regulating their expression, and permitting the accumulation of mutations and chromosomal instability that contribute importantly to HCC. Thus, HBx regulates the activity of ccc DNA in both positive and negative ways, and in doing so, impacts upon the pathogenesis of CLD. The reason why it is important to regulate the intrahepatic levels of ccc DNA is because when virus antigens are greatly overproduced, they could trigger cytopathic effects (CPE), thereby limiting virus replication. For example, mutations in the preS region of the S gene prevent secretion of surface antigen and complete virus particles, and eventually CPE. Pre-S mutations also promote recycling of viral DNA into the nucleus where it results in increased levels of viral ccc DNA, which potentially promotes virus persistence [76] (Figure 2). In transgenic mice overproducing HBsAg, CPE develops and eventually evolves into HCC [77]. Although the latter is not characteristic of HCC pathogenesis among human carriers, it does underscore that selected HBV mutants that may arise during chronic infection potentially contribute to pathogenesis via CPE.

5.2. Oxidative damage and inflammation

Although HBV is not cytopathic, HBx strongly activates NF-ĸB [78], which promotes the expression of many pro-inflammatory cytokines and chemokines that attenuate virus replication and contribute to the pathogenesis of CLD and HCC. For example, HBx stimulates the expression of IFN inducible proteins, such as the CXC chemokine IP-10 [79] which promotes leukocyte chemotaxis. HBx also stimulates production of interleukin-23 (IL-23) [79], which contributes to the maintenance and expansion of pro-inflammatory Th17 cells. Among others, IL-6 is up-regulated by HBx in a MyD88 manner [80], which indicates that HBx is activating a pro-inflammatory environment via innate immune pathways early on after infection. The repressive effect of IL-6 upon HBV replication is demonstrated by the fact that IL-6 treatment of infected cells results in the loss of HNF1a and HNF4a, both of which bind to ccc DNA. Il-6 also redistributes signal transducers and activators of transcription 3 (STAT3) signaling from ccc DNA to IL-6 target genes [49]. HBx targets up-regulation of IL-18, which up-regulates FasL [81], which in this case blocks the killing of infected cells by CTLs. HBx also up-regulates tumor TNFα [82], which was shown to suppress HBcAg expression [83], thereby inhibiting virus replication. In addition, the pro-inflammatory IL-32 was up-regulated by HBx in a NF-ĸB dependent manner [84]. This is not an exhaustive list. Many of these molecules are turned on as a result of HBx stimulating multiple signal transduction pathways in the cytoplasm (in addition to NF-ĸB), but the bigger question is trying to understand how a non-cytopathic virus is mediating these and other related changes in infected cells.

The fact that HBx plays a central role in HBV replication suggests that intracellular conditions that stimulate HBx activity would also promote the carrier state, which would be evolutionally selected for because it would provide a large window of time for virus to be transmitted to other hosts. In this context, the expression and activity of HBx is stimulated in an oxidative environment, since the addition of anti-oxidants to cells expressing HBx strongly diminish HBx trans-activation activity [85, 86]. An oxidative environment (accompanied by oxidative stress of cellular organelles) could be created in the infected cell by virtue of the association of HBx with mitochondria [87]. HBx interacts with the voltage dependent anion channel on the outer mitochondrial membrane, altering transmembrane potential [88], resulting in diminished electron transport, increased free radical accumulation, including elevated lipid peroxidation products [89], release of calcium into the cytosol [55], and under specific circumstances, cell death [90]. Release of calcium into the cytosol, resulted in the activation of the protein tyrosine kinase 2 and Src kinase families, leading to stimulation of ras, raf, mitogen activated protein kinase, and Jun, which stimulate HBV transcription and replication [55]. HBx also induces oxidative stress in the endoplasmic reticulum, which activates the unfolded protein response and expression of pro-inflammatory cyclooxygenase-2 through the activating transcription factor 4 pathway [91]. Free radicals are also characteristic of immune responses aimed at damaging and destroying infected cells that are replicating HBV. In addition, mitochondrial associated HBx induces oxidative stress, which activates selected transcription factors, such as NF-ĸB, STAT3 and activating protein 1 [86]. However, HBx is also known to block mitochondrial triggered cell death, not only by activation of survival [21, 92] and hepato-protective pathways such as NF-ĸB that over-ride apoptosis signaling, but also by blocking key caspases and promoting autophagy [93] and mitophagy [94]. The maintenance of mitochondrial and cellular homeostasis by mitophagy acts to attenuate virus induced apoptosis, so that on the one hand, autophagy and mitophagy promote cell survival and virus persistence, while simultaneous mitochondrial damage may contribute to CLD [94].

5.3. HBx and inflammation

In this chronic pro-inflammatory environment, one would expect to see a correlation between HBx staining and the intensity of CLD. In fact, WHx staining has been observed around inflammatory foci in chronically infected woodchuck livers [95], and among human carriers, relatively low levels of intrahepatic HBx staining was observed in patient biopsy samples from people with low grade hepatitis, while intense and widespread HBx staining was observed in patient biopsies from those with advanced fibrosis and cirrhosis [45, 96], suggesting a direct correlation between HBx staining and liver damage. Independent work also showed low levels of HBx mRNA in the livers of patients with mild CLD (e.g., mild hepatitis), and much higher levels among patients with severe lesions in the liver (advanced fibrosis and cirrhosis) [97]. The relationship of HBx expression to disease severity is also consistent with the idea that when the liver regenerates following each bout of hepatitis, fragments of HBV DNA encoding the HBx region (and sometimes part of the preS/S encoding gene as well) increasingly integrate into multiple regions of the host genome during normal host DNA replication, resulting in increasing accumulation of intrahepatic HBx as CLD progresses. In contrast, the copy number of ccc DNA per cell decreases with regeneration.

The relationship between HBx expression and CLD has been recapitulated in HBx transgenic mice, where the presence, frequency and distribution of HBx in the liver increase with age, as does liver pathology, which progressively develops from hepatitis and steatosis, to dysplasia and microscopic nodules of HCC, and finally to multi-nodular macroscopic HCC with age [98]. In this model, HBx is expressed from its own enhancer and promoter, which is not active until after birth when appropriate transcription factors in the liver begin to appear. HBx expression triggers immune responses in the absence of other HBV gene products, so it is likely that the pathogenesis observed is due to the impact of increasing levels of HBx upon host gene expression combined with immune responses directed against virus infected cells. There is no ccc DNA in this system, just as it is difficult to detect HBV replication among patients with advanced stages of CLD (i.e., cirrhosis). Thus, it is possible that early in chronic infection, and immune responses to virus antigens emanating from ccc DNA templates play an important role in triggering and sustaining immune mediated pathogenesis, but following bouts of CLD and liver regeneration, where the levels of virus replication decrease at the same time that integration of virus DNA fragments increase, pathogenesis appears to be increasingly driven by one or more antigens made from integrated HBV DNA. Although cis-acting mechanisms have been postulated to contribute importantly to the pathogenesis of HCC in selected cases, the broadly distributed integration events of the HBx ORF into most chromosomes [99], suggests that the HBx proteins encoded by most integration events promote CLD and HCC in trans [47]. In this model, integration of HBV sequences would accumulate in areas of euchromatin and fragile sites much more frequently that at or within specific genes [100].

The model above suggests that targeting ccc DNA in HBeAg carriers with CLD may be an important therapeutic goal to bring about a functional (but not sterilizing) cure due to the presence of integrated virus DNA that express one or more virus proteins. Among anti-HBe carriers with advanced CLD, targeting the much lower levels of ccc DNA may not be effective in preventing progression to cirrhosis and HCC, because at this stage, most of the HBx made probably comes from integrated templates. Under these circumstances, ccc DNA may persist in a transcriptionally inactive form, which is consistent with the absence of HBV DNA in the blood, even after treatment with direct acting anti-viral agents or therapy aimed at stimulating immune responses against virus infected cells [101, 102]. In fact, early work already pointed out that seroconversion to anti-HBe is sometimes associated with the progression of CLD [103, 104], even though later work showed that disease progression was associated with continued replication of HBV DNA carrying one or more mutations in the core gene that blocks production of HBeAg [27]. These mutations were probably selected for during the natural history of infection by immune responses targeting HBcAg [105]. Although these findings suggest that CLD progresses in the liver supporting replication of selected virus mutants, it has also recently been suggested that linear HBV DNA, and not ccc DNA, is the template for integration into host DNA [43], from which one or more virus gene products are made, and contribute to pathogenesis. Thus, persistent inflammation in a chronically damaged liver may result in the development of HCC despite low levels or undetectable levels of virus replication.

Advertisement

6. Conclusions

HBV encodes polypeptides from four ORFs that trigger corresponding immune responses during acute and chronic infections. When these responses are rapid, strong and multi-specific, acute, resolving infection can be achieved. When these immune responses are weak and of limited specificity (against few virus epitopes), the carrier state may develop. Although the pathogenesis of HBV is variable in different hosts, the virus encodes proteins that blunt innate immunity, and as a consequence, adaptive immunity is not triggered at all or to a limited extent. The latter causes liver damage over many years without eliminating the virus. Even though available treatments suppress virus replication, none are curative, and the persistence of viral ccc DNA sustains infection. Production of HBx regulates virus gene expression and replication, but over time, increased integration of HBV DNA fragments encoding HBx results in high levels of HBx expression that epigenetically alter the expression of numerous host genes that up- or down-regulate HBV replication and impact disease activity. For example, HBx activation of AKT decreased HBV replication, but this was accompanied by an inhibition of apoptosis, suggesting that HBx balances HBV replication and cell survival by stimulating signaling that enhance hepatocyte survival at the expense of higher levels of HBV replication [106]. The generation of free radicals by immune responses against virus infected cells, combined with HBx mediated alterations in mitochondrial function, promote HBx activity. These events result in the activation of signaling pathways (e.g., AP-1 and NF-ĸB) that over-ride apoptosis and/or directly block the activation of critical caspases, so that whether HBx stimulates or block apoptosis depends upon whether the liver is experiencing inflammation and oxidative stress. It also depends upon whether HBx is being expressed in normal hepatocytes, where apoptotic pathways could be triggered, or whether HBx is expressed at high levels in cells where apoptotic pathways are compromised. In addition to being pro-inflammatory, activated NF-ĸB protects infected cells against immune elimination. Thus, the dichotomy of HBx activity may be a reflection of the environment wherein HBx is expressed. Importantly, the epigenetic mechanisms whereby HBx regulates virus replication also have an impact on cell growth and survival, and many of these same alterations in host gene expression are also hallmarks of cancer [107], which may explain why there is such a high risk of HCC among carriers with CLD [11]. The common denominator is that many of the pathways and molecules that support HBV gene expression and replication also protect infected cells from elimination, and contribute centrally to malignant transformation.

Advertisement

Acknowledgments

The author wants to acknowledge Temple University for providing him with the opportunity to work on this project.

Advertisement

Conflict of interest

The author declares no conflict of interest.

References

  1. 1. Blumberg BS. Australia antigen and the biology of hepatitis B. Science. 1977;197:17-25
  2. 2. Szmuness W, Stevens CE, Zang EA, Harley EJ, Kellner A. A controlled clinical trial of the efficacy of the hepatitis B vaccine (Heptavax B): A final report. Hepatology. 1981;1:377-385
  3. 3. Senior JR, London WT, Sutnick AI. The Australia antigen and role of the late Philadelphia general hospital in reducing post-transfusion hepatitis and sequelae. Hepatology. 2011;54:753-756. DOI: 10.1002/hep.24593
  4. 4. Villeneuve JP. The natural history of chronic hepatitis B virus infection. Journal of Clinical Virology. 2005;34(Suppl 1):S139-S142
  5. 5. Nguyen VT, Law MG, Dore GJ. Hepatitis B-related hepatocellular carcinoma: Epidemiological characteristics and disease burden. Journal of Viral Hepatitis. 2009;16:453-463. DOI: 10.1111/j.1365-2893.2009.01117.x
  6. 6. Yoo J, Hann HW, Coben R, Conn M, DiMarino AJ. Update treatment for HBV infection and persistent risk for hepatocellular carcinoma: Prospect for an HBV cure. Diseases. 2018;6. pii: E27. DOI: 10.3390/diseases6020027
  7. 7. Clark DN, Hu J. Hepatitis B virus reverse transcriptase—Target of current antiviral therapy and future drug development. Antiviral Research. 2015;123:132-137. DOI: 10.1016/j.antiviral.2015.09.011
  8. 8. Sun D, Zhu L, Yao D, Chen L, Fu L, Ouyang L. Recent progress in potential anti-hepatitis B virus agents: Structural and pharmacological perspectives. European Journal of Medicinal Chemistry. 2018;147:205-217. DOI: 10.1016/j.ejmech.2018.02.001
  9. 9. Ikeda M, Morizane C, Ueno M, Okusaka T, Ishii H, Furuse J. Chemotherapy for hepatocellular carcinoma: Current status and future perspectives. Japanese Journal of Clinical Oncology. 2018;48:103-114. DOI: 10.1093/jjco/hyx180
  10. 10. Kudo M. Immune checkpoint inhibition in hepatocellular carcinoma: Basics and ongoing clinical trials. Oncology. 2017;92(Suppl 1):50-62. DOI: 10.1159/000451016
  11. 11. Beasley RP, Hwang LY, Lin CC, Chien CS. Hepatocellular carcinoma and hepatitis B virus. A prospective study of 22 707 men in Taiwan. Lancet. 1981;2:1129-1133
  12. 12. Guidotti LG, Rochford R, Chung J, Shapiro M, Purcell R, Chisari FV. Viral clearance without destruction of infected cells during acute HBV infection. Science. 1999;284:825-829
  13. 13. Wang Y, Jacob JR, Menne S, Bellezza CA, Tennant BC, Gerin JL, Cote PJ. Interferon-gamma-associated responses to woodchuck hepatitis virus infection in neonatal woodchucks and virus-infected hepatocytes. Journal of Viral Hepatitis. 2004;11:404-417
  14. 14. Yang PL, Althage A, Chung J, Maier H, Wieland S, Isogawa M, Chisari FV. Immune effectors required for hepatitis B virus clearance. Proceedings of the National Academy of Sciences of the United States of America. 2010;107:798-802. DOI: 10.1073/pnas.0913498107
  15. 15. Tiollais P, Pourcel C, Dejean A. The hepatitis B virus. Nature. 1985;317:489-495
  16. 16. Kasprzak A, Biczysko W, Zabel M, Wysocki J, Surdyk-Zasada J. Studies on tissue expression of HBV in children with chronic hepatitis type B using Immunomax technique. Polish Journal of Pathology. 1999;50:249-258
  17. 17. Feitelson MA. Hepatitis B virus gene products as immunological targets in chronic infection. Molecular Biology & Medicine. 1989;6:367-393
  18. 18. Ferrari C, Penna A, Giuberti T, Tong MJ, Ribera E, Fiaccadori F, Chisari FV. Intrahepatic, nucleocapsid antigen-specific T cells in chronic active hepatitis B. Journal of Immunology. 1987;139:2050-2058
  19. 19. Ferrari C, Penna A, Bertoletti A, Valli A, Antoni AD, Giuberti T, Cavalli A, Petit MA, Fiaccadori F. Cellular immune response to hepatitis B virus-encoded antigens in acute and chronic hepatitis B virus infection. Journal of Immunology. 1990;145:3442-3449
  20. 20. Tong S, Li J, Wands JR, Wen YM. Hepatitis B virus genetic variants: Biological properties and clinical implications. Emerging Microbes and Infections. 2013;2:e10. DOI: 10.1038/emi.2013.10
  21. 21. Faure-Dupuy S, Lucifora J, Durantel D. Interplay between the hepatitis B virus and innate immunity: From an understanding to the development of therapeutic concepts. Viruses. 2017;9. pii: E95. DOI: 10.3390/v9050095
  22. 22. Wieland S, Thimme R, Purcell RH, Chisari FV. Genomic analysis of the host response to hepatitis B virus infection. Proceedings of the National Academy of Sciences of the United States of America. 2004;101:6669-6674. DOI: 10.1073/pnas.0401771101
  23. 23. Kim CW, Yoon SK, Jung ES, Jung CK, Jang JW, Kim MS, Lee SY, Bae SH, Choi JY, Choi SW, Han NI, Lee CD. Correlation of hepatitis B core antigen and beta-catenin expression on hepatocytes in chronic hepatitis B virus infection: Relevance to the severity of liver damage and viral replication. Journal of Gastroenterology and Hepatology. 2007;22:1534-1542. DOI: 10.1111/j.1440-1746.2007.04849.x
  24. 24. Lin CL, Kao JH. Hepatitis B viral factors and clinical outcomes of chronic hepatitis B. Journal of Biomedical Science. 2008;15:137-145. DOI: 10.1007/s11373-007-9225-8
  25. 25. Vento S, Hegarty JE, Alberti A, O'Brien CJ, Alexander GJ, Eddleston AL, Williams R. T lymphocyte sensitization to HBcAg and T cell-mediated unresponsiveness to HBsAg in hepatitis B virus-related chronic liver disease. Hepatology. 1985;5:192-197
  26. 26. Alexopoulou A, Karayiannis P. HBeAg negative variants and their role in the natural history of chronic hepatitis B virus infection. World Journal of Gastroenterology. 2014;20:7644-7652. DOI: 10.3748/wjg.v20.i24.7644
  27. 27. Lin CL, Liao LY, Wang CS, Chen PJ, Lai MY, Chen DS, Kao JH. Basal core-promoter mutant of hepatitis B virus and progression of liver disease in hepatitis B e antigen-negative chronic hepatitis B. Liver International. 2005;25:564-570. DOI: 10.1111/j.1478-3231.2005.01041.x
  28. 28. Chen MT, Billaud JN, Sällberg M, Guidotti LG, Chisari FV, Jones J, Hughes J, Milich DR. A function of the hepatitis B virus precore protein is to regulate the immune response to the core antigen. Proceedings of the National Academy of Sciences of the United States of America. 2004;101:14913-14918. DOI: 10.1073/pnas.0406282101
  29. 29. Lan S, Wu L, Wang X, Wu J, Lin X, Wu W, Huang Z. Impact of HBeAg on the maturation and function of dendritic cells. International Journal of Infectious Diseases. 2016;46:42-48. DOI: 10.1016/j.ijid.2016.03.024
  30. 30. Yu Y, Wan P, Cao Y, Zhang W, Chen J, Tan L, Wang Y, Sun Z, Zhang Q, Wan Y, Zhu Y, Liu F, Wu K, Liu Y, Wu J. Hepatitis B virus e antigen activates the suppressor of cytokine signaling 2 to repress interferon action. Scientific Reports. 2017;7:1729. DOI: 10.1038/s41598-017-01773-6
  31. 31. Feitelson MA, Millman I, Duncan GD, Blumberg BS. Presence of antibodies to the polymerase gene product(s) of hepatitis B and woodchuck hepatitis virus in natural and experimental infections. Journal of Medical Virology. 1988;24:121-136
  32. 32. Liu Y, Li J, Chen J, Li Y, Wang W, Du X, Song W, Zhang W, Lin L, Yuan Z. Hepatitis B virus polymerase disrupts K63-linked ubiquitination of STING to block innate cytosolic DNA-sensing pathways. Journal of Virology. 2015;89:2287-2300. DOI: 10.1128/JVI.02760-14
  33. 33. Liu D, Wu A, Cui L, Hao R, Wang Y, He J, Guo D. Hepatitis B virus polymerase suppresses NF-κB signaling by inhibiting the activity of IKKs via interaction with Hsp90β. PLoS One. 2014;9:e91658. DOI: 10.1371/journal.pone.0091658
  34. 34. Chen CJ, Yang HI. Natural history of chronic hepatitis B REVEALed. Journal of Gastroenterology and Hepatology. 2011;26:628-638. DOI: 10.1111/j.1440-1746.2011.06695.x
  35. 35. Sinn DH, Lee J, Goo J, Kim K, Gwak GY, Paik YH, Choi MS, Lee JH, Koh KC, Yoo BC, Paik SW. Hepatocellular carcinoma risk in chronic hepatitis B virus-infected compensated cirrhosis patients with low viral load. Hepatology. 2015;62:694-701. DOI: 10.1002/hep.27889
  36. 36. Chen JD, Yang HI, Iloeje UH, You SL, Lu SN, Wang LY, Su J, Sun CA, Liaw YF, Chen CJ. Risk evaluation of viral load elevation and associated liver disease/cancer in HBV (REVEAL-HBV) study group. Carriers of inactive hepatitis B virus are still at risk for hepatocellular carcinoma and liver-related death. Gastroenterology. 2010;138:1747-1754. DOI: 10.1053/j.gastro.2010.01.042
  37. 37. Shao J, Wei L, Wang H, Sun Y, Zhang LF, Li J, Dong JQ. Relationship between hepatitis B virus DNA levels and liver histology in patients with chronic hepatitis B. World Journal of Gastroenterology. 2007;13:2104-2107. DOI: 10.3748/wjg.v13.i14.2104
  38. 38. Mendy ME, Welzel T, Lesi OA, Hainaut P, Hall AJ, Kuniholm MH, McConkey S, Goedert JJ, Kaye S, Rowland-Jones S, Whittle H, Kirk GD. Hepatitis B viral load and risk for liver cirrhosis and hepatocellular carcinoma in the Gambia, West Africa. Journal of Viral Hepatitis. 2010;17:115-122. DOI: 10.1111/j.1365-2893.2009.01168.x
  39. 39. Bertoletti A, Gehring A. Immune response and tolerance during chronic hepatitis B virus infection. Hepatology Research. 2007;37(Suppl 3):S331-S338. DOI: 10.1111/j.1872-034X.2007.00221.x
  40. 40. Dudley FJ, Fox RA, Sherlock S. Cellular immunity and hepatitis-associated, Australia antigen liver disease. Lancet. 1972;1:723-726
  41. 41. Matsubara K, Tokino T. Integration of hepatitis B virus DNA and its implications for hepatocarcinogenesis. Molecular Biology & Medicine. 1990;7:243-260
  42. 42. Feitelson MA, Lee J. Hepatitis B virus integration, fragile sites, and hepatocarcinogenesis. Cancer Letters. 2007;252:157-170. DOI: 10.1016/j.canlet.2006.11.010
  43. 43. Tu T, Budzinska MA, Vondran FWR, Shackel NA, Urban S. Hepatitis B virus DNA integration occurs early in the viral life cycle in an in vitro infection model via NTCP-dependent uptake of enveloped virus particles. Journal of Virology 2018;92:e02007-17. DOI: 10.1128/JVI.02007-17
  44. 44. Wollersheim M, Debelka U, Hofschneider PH. A transactivating function encoded in the hepatitis B virus X gene is conserved in the integrated state. Oncogene. 1988;3:545-552
  45. 45. Wang WL, London WT, Lega L, Feitelson MA. HBxAg in the liver from carrier patients with chronic hepatitis and cirrhosis. Hepatology. 1991;14:29-37
  46. 46. 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. Molecular and Cellular Biology. 2013;33:2810-2816. DOI: 10.1128/MCB.00205-13
  47. 47. Caselmann WH. Transactivation of cellular gene expression by hepatitis B viral proteins: A possible molecular mechanism of hepatocarcinogenesis. Journal of Hepatology. 1995;22(1 Suppl):34-37
  48. 48. Ji M, Hu K. Recent advances in the study of hepatitis B virus covalently closed circular DNA. Virologica Sinica. 2017;32:454-464. DOI: 10.1007/s12250-017-4009-4
  49. 49. Palumbo GA, Scisciani C, Pediconi N, Lupacchini L, Alfalate D, Guerrieri F, Calvo L, Salerno D, Di Cocco S, Levrero M, Belloni L. IL6 inhibits HBV transcription by targeting the epigenetic control of the nuclear ccc DNA minichromosome. PLoS One. 2015;10:e0142599. DOI: 10.1371/journal.pone.0142599. Erratum in: PLoS One. 2015;10(12):e0145555
  50. 50. Lucifora J, Xia Y, Reisinger F, Zhang K, Stadler D, Cheng X, Sprinzl MF, Koppensteiner H, Makowska Z, Volz T, Remouchamps C, Chou WM, Thasler WE, Hüser N, Durantel D, Liang TJ, Münk C, Heim MH, Browning JL, Dejardin E, Dandri M, Schindler M, Heikenwalder M, Protzer U. Specific and nonhepatotoxic degradation of nuclear hepatitis B virus ccc DNA. Science. 2014;343:1221-1228. DOI: 10.1126/science.1243462
  51. 51. Lin SR, Yang HC, Kuo YT, Liu CJ, Yang TY, Sung KC, Lin YY, Wang HY, Wang CC, Shen YC, Wu FY, Kao JH, Chen DS, Chen PJ. The CRISPR/Cas9 system facilitates clearance of the intrahepatic HBV templates in vivo. Molecular Therapy: Nucleic Acids. 2014;3:e186. DOI: 10.1038/mtna.2014.38
  52. 52. Møller HD, Mohiyuddin M, Prada-Luengo I, Sailani MR, Halling JF, Plomgaard P, Maretty L, Hansen AJ, Snyder MP, Pilegaard H, Lam HYK, Regenberg B. Circular DNA elements of chromosomal origin are common in healthy human somatic tissue. Nature Communications. 2018;9:1069. DOI: 10.1038/s41467-018-03369-8
  53. 53. Arzumanyan A, Reis HM, Feitelson MA. Pathogenic mechanisms in HBV- and HCV-associated hepatocellular carcinoma. Nature Reviews. Cancer. 2013;13:123-135. DOI: 10.1038/nrc3449
  54. 54. Tang H, Oishi N, Kaneko S, Murakami S. Molecular functions and biological roles of hepatitis B virus x protein. Cancer Science. 2006;97:977-983. DOI: 10.1111/j.1349-7006.2006.00299.x
  55. 55. Bouchard MJ, Schneider RJ. The enigmatic X gene of hepatitis B virus. Journal of Virology. 2004;78:12725-12734. DOI: 10.1128/JVI.78.23.12725-12734.2004
  56. 56. Cote PJ, Toshkov I, Bellezza C, Ascenzi M, Roneker C, Ann Graham L, Baldwin BH, Gaye K, Nakamura I, Korba BE, Tennant BC, Gerin JL. Temporal pathogenesis of experimental neonatal woodchuck hepatitis virus infection: Increased initial viral load and decreased severity of acute hepatitis during the development of chronic viral infection. Hepatology. 2000;32(4 Pt 1):807-817. DOI: 10.1053/jhep.2000.17681
  57. 57. Zoulim F, Saputelli J, Seeger C. Woodchuck hepatitis virus X protein is required for viral infection in vivo. Journal of Virology. 1994;68:2026-2030
  58. 58. Chen HS, Kaneko S, Girones R, Anderson RW, Hornbuckle WE, Tennant BC, Cote PJ, Gerin JL, Purcell RH, Miller RH. The woodchuck hepatitis virus X gene is important for establishment of virus infection in woodchucks. Journal of Virology. 1993;67:1218-1226
  59. 59. Dandri M, Schirmacher P, Rogler CE. Woodchuck hepatitis virus X protein is present in chronically infected woodchuck liver and woodchuck hepatocellular carcinomas which are permissive for viral replication. Journal of Virology. 1996;70:5246-5254
  60. 60. Jacob JR, Ascenzi MA, Roneker CA, Toshkov IA, Cote PJ, Gerin JL, Tennant BC. Hepatic expression of the woodchuck hepatitis virus X-antigen during acute and chronic infection and detection of a woodchuck hepatitis virus X-antigen antibody response. Hepatology. 1997;26:1607-1615. DOI: 10.1002/hep.510260632
  61. 61. Feitelson MA, Clayton MM. X antigen polypeptides in the sera of hepatitis B virus-infected patients. Virology. 1990;177:367-371
  62. 62. Feitelson MA, Clayton MM, Phimister B. Monoclonal antibodies raised to purified woodchuck hepatitis virus core antigen particles demonstrate X antigen reactivity. Virology. 1990;177:357-366
  63. 63. Kouwaki T, Okamoto T, Ito A, Sugiyama Y, Yamashita K, Suzuki T, Kusakabe S, Hirano J, Fukuhara T, Yamashita A, Saito K, Okuzaki D, Watashi K, Sugiyama M, Yoshio S, Standley DM, Kanto T, Mizokami M, Moriishi K, Matsuura Y. Hepatocyte factor JMJD5 regulates hepatitis B virus replication through interaction with HBx. Journal of Virology. 2016;90:3530-3542. DOI: 10.1128/JVI.02776-15
  64. 64. Guo L, Wang X, Ren L, Zeng M, Wang S, Weng Y, Tang Z, Wang X, Tang Y, Hu H, Li M, Zhang C, Liu C. HBx affects CUL4-DDB1 function in both positive and negative manners. Biochemical and Biophysical Research Communications. 2014;450:1492-1497. DOI: 10.1016/j.bbrc.2014.07.019
  65. 65. Leupin O, Bontron S, Schaeffer C, Strubin M. Hepatitis B virus X protein stimulates viral genome replication via a DDB1-dependent pathway distinct from that leading to cell death. Journal of Virology. 2005;79:4238-4245. DOI: 10.1128/JVI.79.7.4238-4245.2005
  66. 66. Hu Z, Zhang Z, Doo E, Coux O, Goldberg AL, Liang TJ. Hepatitis B virus X protein is both a substrate and a potential inhibitor of the proteasome complex. Journal of Virology. 1999;73:7231-7240
  67. 67. Decorsière A, Mueller H, Van Breugel PC, Abdul F, Gerossier L, Beran RK, Livingston CM, Niu C, Fletcher SP, Hantz O, Strubin M. Hepatitis B virus X protein identifies the Smc5/6 complex as a host restriction factor. Nature. 2016;531:386-389. DOI: 10.1038/nature17170
  68. 68. Belloni L, Pollicino T, De Nicola F, Guerrieri F, Raffa G, Fanciulli M, Raimondo G, Levrero M. Nuclear HBx binds the HBV minichromosome and modifies the epigenetic regulation of ccc DNA function. Proceedings of the National Academy of Sciences of the United States of America. 2009;106:19975-19979. DOI: 10.1073/pnas.0908365106
  69. 69. Liang TJ. Virology: The X-files of hepatitis B. Nature. 2016;531:313-314. DOI: 10.1038/531313a
  70. 70. Murphy CM, Xu Y, Li F, Nio K, Reszka-Blanco N, Li X, Wu Y, Yu Y, Xiong Y, Su L. Hepatitis B virus X protein promotes degradation of SMC5/6 to enhance HBV replication. Cell Reports. 2016;16:2846-2854. DOI: 10.1016/j.celrep.2016.08.026
  71. 71. Guidotti LG, Isogawa M, Chisari FV. Host-virus interactions in hepatitis B virus infection. Current Opinion in Immunology. 2015;36:61-66. DOI: 10.1016/j.coi.2015.06.016
  72. 72. Tsunematsu S, Suda G, Yamasaki K, Kimura M, Izumi T, Umemura M, Ito J, Sato F, Nakai M, Sho T, Morikawa K, Ogawa K, Tanaka Y, Watashi K, Wakita T, Sakamoto N. Hepatitis B virus X protein impairs α-interferon signaling via up-regulation of suppressor of cytokine signaling 3 and protein phosphatase 2A. Journal of Medical Virology. 2017;89:267-275. DOI: 10.1002/jmv.24643
  73. 73. Cho IR, Oh M, Koh SS, Malilas W, Srisuttee R, Jhun BH, Pellegrini S, Fuchs SY, Chung YH. Hepatitis B virus X protein inhibits extracellular IFN-α-mediated signal transduction by downregulation of type I IFN receptor. International Journal of Molecular Medicine. 2012;29:581-586. DOI: 10.3892/ijmm.2012.879
  74. 74. Kumar M, Jung SY, Hodgson AJ, Madden CR, Qin J, Slagle BL. Hepatitis B virus regulatory HBx protein binds to adaptor protein IPS-1 and inhibits the activation of beta interferon. Journal of Virology. 2011;85:987-995. DOI: 10.1128/JVI.01825-10
  75. 75. Vivekanandan P, Daniel HD, Kannangai R, Martinez-Murillo F, Torbenson M. Hepatitis B virus replication induces methylation of both host and viral DNA. Journal of Virology. 2010;84:4321-4329. DOI: 10.1128/JVI.02280-09
  76. 76. Zhang YY, Hu KQ. Rethinking the pathogenesis of hepatitis B virus (HBV) infection. Journal of Medical Virology. 2015;87:1989-1999. DOI: 10.1002/jmv.24270
  77. 77. Chisari FV, Klopchin K, Moriyama T, Pasquinelli C, Dunsford HA, Sell S, Pinkert CA, Brinster RL, Palmiter RD. Molecular pathogenesis of hepatocellular carcinoma in hepatitis B virus transgenic mice. Cell. 1989 Dec 22;59(6):1145-1156
  78. 78. Su F, Schneider RJ. Hepatitis B virus HBx protein activates transcription factor NF-kappaB by acting on multiple cytoplasmic inhibitors of rel-related proteins. Journal of Virology. 1996;70:4558-4566
  79. 79. Xia L, Tian D, Huang W, Zhu H, Wang J, Zhang Y, Hu H, Nie Y, Fan D, Wu K. Upregulation of IL-23 expression in patients with chronic hepatitis B is mediated by the HBx/ERK/NF-κB pathway. Journal of Immunology. 2012;188:753-764. DOI: 10.4049/jimmunol.1101652
  80. 80. Xiang WQ, Feng WF, Ke W, Sun Z, Chen Z, Liu W. Hepatitis B virus X protein stimulates IL-6 expression in hepatocytes via a MyD88-dependent pathway. Journal of Hepatology. 2011;54:26-33. DOI: 10.1016/j.jhep.2010.08.006
  81. 81. Lee MO, Choi YH, Shin EC, Kang HJ, Kim YM, Jeong SY, Seong JK, Yu DY, Cho H, Park JH, Kim SJ. Hepatitis B virus X protein induced expression of interleukin 18 (IL-18): A potential mechanism for liver injury caused by hepatitis B virus (HBV) infection. Journal of Hepatology. 2002;37:380-386. DOI: 10.1016/S0168-8278(02)00181-2
  82. 82. Lara-Pezzi E, Majano PL, Gómez-Gonzalo M, García-Monzón C, Moreno-Otero R, Levrero M, López-Cabrera M. The hepatitis B virus X protein up-regulates tumor necrosis factor alpha gene expression in hepatocytes. Hepatology. 1998;28:1013-1021. DOI: 10.1002/hep.510280416
  83. 83. Romero R, Lavine JE. Cytokine inhibition of the hepatitis B virus core promoter. Hepatology. 1996;23:17-23. DOI: 10.1002/hep.510230103
  84. 84. Pan X, Cao H, Lu J, Shu X, Xiong X, Hong X, Xu Q, Zhu H, Li G, Shen G. Interleukin-32 expression induced by hepatitis B virus protein X is mediated through activation of NF-κB. Molecular Immunology. 2011;48:1573-1577. DOI: 10.1016/j.molimm.2011.03.012
  85. 85. Meyer M, Caselmann WH, Schlüter V, Schreck R, Hofschneider PH, Baeuerle PA. Hepatitis B virus transactivator MHBst: Activation of NF-kappa B, selective inhibition by antioxidants and integral membrane localization. The EMBO Journal. 1992;11:2991-3001
  86. 86. 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. Molecular and Cellular Biology. 2001;21:7721-7730
  87. 87. Chen J, Siddiqui A. Hepatitis B virus X protein stimulates the mitochondrial translocation of Raf-1 via oxidative stress. Journal of Virology. 2007;81:6757-6760
  88. 88. Rahmani Z, Huh KW, Lasher R, Siddiqui A. Hepatitis B virus X protein co-localizes to mitochondria with a human voltage-dependent anion channel, HVDAC3, and alters its transmembrane potential. Journal of Virology. 2000;74:2840-2846
  89. 89. Lee YI, Hwang JM, Im JH, Lee YI, Kim NS, Kim DG, Yu DY, Moon HB, Park SK. Human hepatitis B virus-X protein alters mitochondrial function and physiology in human liver cells. The Journal of Biological Chemistry. 2004;279:15460-15471. DOI: 10.1074/jbc.M309280200
  90. 90. Shirakata Y, Koike K. Hepatitis B virus X protein induces cell death by causing loss of mitochondrial membrane potential. The Journal of Biological Chemistry. 2003;278:22071-22078. DOI: 10.1074/jbc.M301606200
  91. 91. Cho HK, Cheong KJ, Kim HY, Cheong J. Endoplasmic reticulum stress induced by hepatitis B virus X protein enhances cyclo-oxygenase 2 expression via activating transcription factor 4. The Biochemical Journal. 2011;435:431-439. DOI: 10.1042/BJ20102071
  92. 92. Minczuk M, Mroczek S, Pawlak SD, Stepien PP. Human ATP-dependent RNA/DNA helicase hSuv3p interacts with the cofactor of survivin HBXIP. The FEBS Journal. 2005;272:5008-5019. DOI: 10.1111/j.1742-4658.2005.04910.x
  93. 93. Mao Y, Da L, Tang H, Yang J, Lei Y, Tiollais P, Li T, Zhao M. Hepatitis B virus X protein reduces starvation-induced cell death through activation of autophagy and inhibition of mitochondrial apoptotic pathway. Biochemical and Biophysical Research Communications. 2011;415:68-74. DOI: 10.1016/j.bbrc.2011.10.013
  94. 94. 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 Pathogens. 2013;9:e1003722. DOI: 10.1371/journal.ppat.1003722
  95. 95. Feitelson MA, Lega L, Duan LX, Clayton M. Characteristics of woodchuck hepatitis X-antigen in the livers and sera from infected animals. Journal of Hepatology. 1993;17(Suppl 3):S24-S34
  96. 96. Wang WL, London WT, Feitelson MA. Hepatitis B x antigen in hepatitis B virus carrier patients with liver cancer. Cancer Research. 1991;51:4971-4977
  97. 97. Diamantis ID, McGandy CE, Chen TJ, Liaw YF, Gudat F, Bianchi L. Hepatitis B X-gene expression in hepatocellular carcinoma. Journal of Hepatology. 1992;15:400-403
  98. 98. Yu DY, Moon HB, Son JK, Jeong S, Yu SL, Yoon H, Han YM, Lee CS, Park JS, Lee CH, Hyun BH, Murakami S, Lee KK. Incidence of hepatocellular carcinoma in transgenic mice expressing the hepatitis B virus X-protein. Journal of Hepatology. 1999;31:123-132
  99. 99. Furuta M, Tanaka H, Shiraishi Y, Unida T, Imamura M, Fujimoto A, Fujita M, Sasaki-Oku A, Maejima K, Nakano K, Kawakami Y, Arihiro K, Aikata H, Ueno M, Hayami S, Ariizumi SI, Yamamoto M, Gotoh K, Ohdan H, Yamaue H, Miyano S, Chayama K, Nakagawa H. Characterization of HBV integration patterns and timing in liver cancer and HBV-infected livers. Oncotarget. 2018;9:25075-25088. DOI: 10.18632/oncotarget.25308
  100. 100. Robinson WS. Molecular events in the pathogenesis of hepadnavirus-associated hepatocellular carcinoma. Annual Review of Medicine. 1994;45:297-323. DOI: 10.1146/annu-rev.med.45.1.297
  101. 101. Lin CL, Kao JH. Review article: Novel therapies for hepatitis B virus cure - advances and perspectives. Alimentary Pharmacology & Therapeutics. 2016;44:213-222. DOI: 10.1111/apt.13694
  102. 102. Cao G-W. Cancer Evo-Dev, a novel hypothesis derived from studies on hepatitis B virus-induced carcinogenesis. Hepatoma Research. 2017;3:241-259. ISSN: 2454-2520. Available at: http://hrjournal.net/article/view/2267
  103. 103. Sheen IS, Liaw YF, Tai DI, Chu CM. Hepatic decompensation associated with hepatitis B e antigen clearance in chronic type B hepatitis. Gastroenterology. 1985;89:732-735
  104. 104. Bonino F, Rosina F, Rizzetto M, Rizzi R, Chiaberge E, Tardanico R, Callea F, Verme G. Chronic hepatitis in HBsAg carriers with serum HBV-DNA and anti-HBe. Gastroenterology. 1986;90(5 Pt 1):1268-1273
  105. 105. Lee YI, Hur GM, Suh DJ, Kim SH. Novel pre-C/C gene mutants of hepatitis B virus in chronic active hepatitis: Naturally occurring escape mutants. The Journal of General Virology. 1996;77(Pt 6):1129-1138
  106. 106. Rawat S, Bouchard MJ. The hepatitis B virus (HBV) HBx protein activates AKT to simultaneously regulate HBV replication and hepatocyte survival. Journal of Virology. 2015;89:999-1012. DOI: 10.1128/JVI.02440-14
  107. 107. Mesri EA, Feitelson MA, Munger K. Human viral oncogenesis: A cancer hallmarks analysis. Cell Host & Microbe. 2014;15:266-282. DOI: 10.1016/j.chom.2014.02.011

Written By

Mark A. Feitelson

Submitted: 27 March 2018 Reviewed: 25 June 2018 Published: 05 November 2018