Next Article in Journal
Full-Length Computational Model of the SARS-CoV-2 Spike Protein and Its Implications for a Viral Membrane Fusion Mechanism
Next Article in Special Issue
Are Humanized Mouse Models Useful for Basic Research of Hepatocarcinogenesis through Chronic Hepatitis B Virus Infection?
Previous Article in Journal
HIV-1 Drug Resistance Genotyping in Resource Limited Settings: Current and Future Perspectives in Sequencing Technologies
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Envelope Proteins of Hepatitis B Virus: Molecular Biology and Involvement in Carcinogenesis

Division of Gastroenterology, Tohoku University Graduate School of Medicine, 1-1 Seiryo-machi, Aoba-ku, Sendai 980-8574, Japan
*
Author to whom correspondence should be addressed.
Viruses 2021, 13(6), 1124; https://doi.org/10.3390/v13061124
Submission received: 26 May 2021 / Revised: 8 June 2021 / Accepted: 9 June 2021 / Published: 11 June 2021

Abstract

:
The envelope of hepatitis B virus (HBV), which is required for the entry to hepatocytes, consists of a lipid bilayer derived from hepatocyte and HBV envelope proteins, large/middle/small hepatitis B surface antigen (L/M/SHBs). The mechanisms and host factors for the envelope formation in the hepatocytes are being revealed. HBV-infected hepatocytes release a large amount of subviral particles (SVPs) containing L/M/SHBs that facilitate escape from the immune system. Recently, novel drugs inhibiting the functions of the viral envelope and those inhibiting the release of SVPs have been reported. LHBs that accumulate in ER is considered to promote carcinogenesis and, especially, deletion mutants in the preS1/S2 domain have been reported to be associated with the development of hepatocellular carcinoma (HCC). In this review, we summarize recent reports on the findings regarding the biological characteristics of HBV envelope proteins, their involvement in HCC development and new agents targeting the envelope.

1. Introduction

Hepatitis B virus (HBV) infects 292 million people chronically in the world despite the availability of an effective vaccine [1], and they are at risk of liver cirrhosis and hepatocellular carcinoma (HCC) [2]. Although antiviral therapies such as nucleos(t)ide analogues (NAs) and interferon have been developed, complete eradication of HBV is still difficult because covalently closed circular DNA (cccDNA), which is a highly stable and active episomal genome, is formed in the nucleus of HBV-infected hepatocytes [3]. Even if patients have received NA therapy and HBV replication is inhibited, they still have the potential to develop liver cancer [4,5]. Therefore, novel antiviral therapies are required clinically and several lines of clinical trials for HBV chronic infection are ongoing [6,7].
The mature infectious HBV particle with a diameter of 42 nm, which is also called a “Dane particle”, consists of an envelope and nucleocapsid containing HBV DNA and polymerase (Figure 1). A large amount of hepatitis B surface antigen (HBsAg) subviral particles (SVPs), which do not include nucleocapsid, is released from HBV-infected hepatocytes [8] and HBsAg assays detect both Dane particles and SVPs. Additionally, the infected hepatocytes release several kinds of other immature particles [9]. Hepatitis B e antigen (HBeAg) is also secreted from the infected hepatocytes and is detected in the serum of patients. HBeAg seroconversion, defined as loss of HBeAg and appearance of anti-HBe antibody, occurs both in the natural course and under treatment. HBeAg-defective precore mutants are found in most inactive HBV carriers with undetectable HBeAg [10,11]. SVPs and HBeAg are not essential for the infection, but they are considered to help the establishment of persistent infection of HBV via effects on the immune system [8]. The HBV envelope, which is considered to be essential for the infection [12], is composed of a cellular lipid bilayer and three HBV surface proteins: large/middle/small hepatitis B surface proteins (LHBs/MHBs/SHBs). The ratio of LHBs, MHBs and SHBs in the envelope of Dane particles was reported to be approximately 1:1:4 [13]. Moreover, hepatitis D virus (HDV) requires the HBV envelope for the packaging. Therefore, the HBV envelope is indispensable for the life cycle of HBV and HDV and could be an antiviral target against these viruses that are tough to beat.
As for the carcinogenesis of HBV, hepatitis B x protein (HBx) and LHBs are considered to play major roles in the infected hepatocytes. HBx is included often in the HBV DNA that is integrated into the host DNA of HCC tissue and has been reported to have multifunctional roles in the modulation of the expression/activities of many genes [14]. LHBs accumulation in the endoplasmic reticulum (ER) has been documented to associate with ER stress responses, which lead to pro-oncogenic effects [15]. In this review, we summarize the biological characteristics of HBV envelope and SVPs and the roles of envelope proteins in the hepatocarcinogenesis.

2. Biological Characteristics of HBV Envelope

The HBV envelope proteins, LHBs, MHBs and SHBs, are translated from 2.4/2.1 kb mRNAs, which are transcribed from cccDNA. Additionally, these mRNAs are transcribed from integrated viral genome. The C-terminus of LHBs, MHBs and SHBs is common, and the length is 400 (389), 281 and 226 amino acids, respectively (Figure 2A). The LHBs consists of preS1, preS2 and S domains, and MHBs consists of preS2 and S domains. There are 4 transmembrane domains in SHBs and “a” determinant that is exposed on the outside of the envelope (Figure 2B). The “a” determinant is an antigenic domain that is a major target of neutralizing antibodies [16]. The N-terminus is myristoylated at glycine 2, which is considered to be essential for the infection [17]. Myristoylation is a posttranslational modification and myristic acid is attached by the ubiquitous enzyme N-myristoyltransferase (NMT). Because the myristic acid is a hydrophobic moiety, the myristoylated part is inserted into hydrophobic regions in the lipid bilayer [18]. Using the myristic acid as an anchor, the preS1 and preS2 in LHBs are considered to be able to transit between the cytoplasmic side and the luminal side (Figure 2B) [19]. The preS1 on the cytoplasmic side is required for the interaction between the LHBs and capsid at the step of envelope formation [20], and the preS1 on the luminal side, which is exposed on the Dane particle, is essential for the infection by interacting with the HBV receptor, sodium-taurocholate cotransporting polypeptide (NTCP) [21].
The three HBV envelope proteins are glycosylated at an N-linked glycosylation site (N-X-S/T) (Figure 2A). All three proteins have a potential N-glycosylation site at N146 of the S domain, but LHBs and SHBs are also present in non-glycosylated forms. MHBs has an additional N-glycosylation site at N4 of the preS2 domain, which is not glycosylated in LHBs, probably because of the altered conformation of preS1/preS2 domains [22]. As a result, LHBs exists in 2 forms (p39 and gp42), MHBs in 2 forms (gp33 and gp36 [a diglycosylated form]) and SHBs in 2 forms (p24 and gp27). Additionally, HBV genotypes C and D have an O-linked glycosylation site at T34 of the preS2 domain [23]. The preS2 glycosylation is central to folding, assembly and secretion [24]. Additionally, N-glycosylation in envelope proteins has dual roles in the interaction with immune systems including dendritic-cell-specific ICAM-3-grabbing nonintegrin (DC-SIGN) [25] and in the occlusion of neutralizing epitopes [26].
A compartment of the late endocytic pathway, the multivesicular body (MVB), which generates exosomes as intraluminal vesicles, is considered to participate in the formation of the HBV envelope, based on the findings that endosomal sorting complex required for transport (ESCRT) is necessary for the release of Dane particles [27,28,29]. The nucleocapsids bud inwards into invaginations of the MVB membrane [30] (Figure 3). Direct interaction between the nucleocapsid and envelope proteins was demonstrated in cell-free binding assays [31,32] and cell-based assays [20], and several host factors have been identified to work in the process of envelope formation. Translated LHBs, MHBs and SHBs are inserted into the ER membrane interacting with chaperones such as calnexin [33], Hsc70 and BiP [34]. LHBs interacts with Hsc and BiP and MHBs interacts with calnexin, but SHBs does not interact with these chaperones. These different interactions might be one of the causes of their different destinies. Moreover, γ2-adaptin interacts with LHBs [35] and is considered to assist the transport of LHBs to MVB. Recently, our group reported that the small GTPase Rab5B is required for the transportation of LHBs from ER to MVBs [36]. On the MVB membrane, α-taxilin, which interacts with LHBs and ESCRT I component TSG101, was shown to be essential for the release of Dane particles [37]. Additionally, an exosome marker, CD63, was shown to work in the incorporation of LHBs in the HBV envelope by us [38]. Nedd4 binds to the late domain-like PPAY motif of core protein and it is considered that ubiquitinated Nedd4 binds to γ2-adaptin resulting in a linkage between LHBs and nucleocapsid [30,39].
The Dane particles are thought to be released from hepatocytes via exocytosis in the same manner as exosomes, but the details have yet to be determined. It was demonstrated that Dane particles are released in a form with preS1 hidden in the interior and that they are converted spontaneously into a form with preS1 on the surface [40]. Because the preS1 domain binds with heparan sulfate proteoglycans on the cellular membrane of various cell types, the dynamic topology switch is thought to be utilized to avoid non-productive attachment before the Dane particles reach the liver of a naïve host. A part of MVBs are considered to be fused with lysosome, resulting degradation of cargos. The small GTPase Rab7, which was activated by HBV infection, was shown to enhance the fusion and to regulate the HBV secretion [41].

3. Subviral Particles of HBV

SVPs were found in HBV-infected patients in 2 forms: sphere and filament (Figure 1). The sphere is 22 nm in a diameter, and the filament is variable in length. The secretory pathways of these SVPs were considered to be distinct from Dane particles. A recent report suggested that coat protein complex II (COPII) is required for the transport of SHBs from ER to ER-Golgi intermediate compartment (ERGIC) [42], where filamentous structures are unpacked and relaxed to form SVP spheres [43]. These are considered to be secreted via Golgi. Also, SVP filaments that contain a larger amount of LHBs than spheres were reported to be ESCRT-dependently released via MVBs [44].

4. HBV Envelope Proteins as Clinical Diagnostic Tools

As described above, a large amount of particles containing HBV envelope proteins are present in serum of HBV-infected patients and are detected as HBsAg. The serum HBsAg assays using techniques such as chemiluminescence immunoassay (CLIA) or chemiluminescence enzyme immunoassay (CLEIA) are easy methods to distinguish the HBV-infected patients, but the sensitivity is lower than serum HBV DNA in general, and sometimes underdiagnose patients who are at an early stage of infection or reactivation of HBV. To improve the sensitivity, immune complex transfer (ICT)-CLEIA, which includes a sample pretreatment to dissociate HBsAg-antibody immune complexes, has been developed [45], and its application for the detection of HBV reactivation following chemotherapies for hematological malignancies was reported [46,47]. In the late stage of infection, HBsAg is detectable for a long time even in patients whose HBV DNA became undetectable.
HBsAg quantification is considered to be a surrogate marker of viral suppression in patients under NA treatments whose serum HBV DNA is undetectable. HBsAg loss is regarded as the optimal treatment endpoint, termed as “functional cure” [2]. The level of HBsAg, as well as hepatitis B core-related antigen (HBcrAg), is associated with relapse of hepatitis after discontinuation of NA [48]. Moreover, high levels of HBsAg were reported to be a risk factor for HCC in patients with low HBV DNA levels [49].
As a novel evaluation method of HBV envelope proteins, quantification of LHBs and MHBs has been reported to predict clinical outcome [50,51]. Inactive HBV carriers had lower proportion of LHBs and MHBs [50], and during antiviral therapies, the decline of LHBs and MHBs was found before subsequent HBsAg loss [51].

5. Genetic Variations of HBV Envelope

Several mutations/deletions are found in HBV-infected patients and associations with specific disease outcomes have been reported. Vaccine-escape mutations such as G145R and D144A/E are variations that were found in the “a” determinant of HBV from patients with a vaccine-induced antibody response [16]. Moreover, these mutations were found in HBV-reactivated patients who were treated with anti-cancer chemotherapies or immunosuppression therapies [52]. Interestingly, several mutations in the “a” determinant that generate novel N-glycosylation sites were reported including M133T, which restores the secretion impairment of virion with G145R [53].
Many previous studies reported that preS1/preS2 deletions were found in patients with chronic HBV infection and were associated with the liver disease progression. Deletion in the preS1 C-terminal half was more frequently found in patients with chronic hepatitis and liver cirrhosis than in asymptomatic carriers [54], and deletion in the preS2 region was frequently found in HCC patients [55]. Especially, preS2-defective mutations including mutations at the level of preS2 start codon and deletions at the 5′-terminal half of the preS2 region were frequently found [56]. The frequency of deletions in these regions was low in patients with acute HBV infection, indicating that these deletions may emerge during the long-term persistence of the viral genome [57]. As described later, preS1/preS2 deletion mutants are reported to be retained in ER and subsequently induce ER stress [58].
A deletion of 11 amino acids at the N-terminus of preS1 domain is found almost specifically in genotype D HBV, and two recent reports showed that this deletion enhances the binding and infectivity to hepatocytes [59,60]. This deletion was found also in HBV strains of nonhuman primates, suggesting that the deletion may have advantages in the HBV life cycle beyond the host species [60].

6. HBV Envelope Protein as a Cause of HCC

The preS mutations including deletions in the preS1 and preS2 regions were found frequently in HCC patients [61] and many studies using in vitro and in vivo models showed an association between preS mutations and HCC. An association of preS mutations with HCC development was reported, including patients with low HBV DNA and alanine aminotransferase levels [62]. Because the deletion site of preS2 is known to include an epitope of the T-cell response and B-cell neutralization [63], HBV clones with such deletions might be selected as immune escape mutants [58]. These regions overlap the polymerase gene but correspond to the spacer region, which is not required for the enzyme activity [64]. LHBs with preS1 or preS2 deletion accumulates in ER and LHBs-accumulated hepatocytes are observed as ground-glass hepatocytes [58], which have been recognized as the histological hallmark of chronic HBV infection [65]. The misfolded protein of the preS1/preS2 mutants induces strong and sustained ER stress [15]. ER stress increases intracellular Ca2+ level that leads to mitochondrial dysfunction, hepatocyte apoptosis and liver fibrosis, resulting in the promotion of carcinogenesis [62]. The integrated HBV DNA can supply LHBs including its mutants, like as a hepatoma cell line PLC/PRF/5 [66].
The preS2 mutant initiates a mammalian target of rapamycin (mTOR)-dependent glycolytic signal cascade and might promote tumorigenesis by aerobic glycolysis [67]. Moreover, transgenic mouse livers harboring preS2 mutant showed lipid accumulation through the activation of sterol regulatory element binding transcription factor 1 (SREBF1) by mTOR signaling [68]. These changes in metabolic pathways might be linked to the HCC development. Additionally, negative feedback from the regulation of HBsAg synthesis by activation of the mTOR signal pathway was reported [69], indicating that a decreased level of HBsAg and HBV DNA in the serum may not necessarily represent a good sign during the clinical course.
Importantly, antiviral therapy with NAs has been reported to reduce HCC development [70,71], including patients infected with preS-mutant HBV [62]. Because the suppression effects of NAs on serum HBsAg [4] and intracellular cccDNA [72] are not ideal, NAs might not reduce accumulation of LHBs enough. Therefore, novel drugs that can inhibit the accumulation of LHBs in ER might reduce HCC further, which is still found in some patients who are treated with NAs [4,5].

7. HBV Envelope as a Target of Antiviral Therapy

7.1. Inhibition of HBV Entry

Recently, many studies have reported agents that inhibit the interaction between HBV preS1 and NTCP, including bulevertide (formerly Myrcludex B), a peptide that comprises N-terminal 47 preS1 amino acids with an N-terminal myristoyl moiety, inhibits the HBV and HDV entry [73,74] and was approved for chronic HDV infection. Moreover, cyclosporine A and cyclosporine derivatives [75,76,77], vanitaracin A [78], Ro41-5253 [79] and evans blue [80] were reported to bind to NTCP and inhibit its interaction with preS1. Some of these interfere with the primary NTCP function and inhibit bile acid intake, and selective inhibitors of HBV/HDV infection could be candidates for clinical application. Additionally, preS1-specific neutralizing antibodies induced by DNA vaccine containing preS1 and the large HDV antigen were reported to inhibit HBV entry in an animal model [81]. The HBV entry inhibitors are not expected to deplete cccDNA completely, but de novo infection via NTCP is reported to be necessary to maintain the cccDNA pool [82], and the benefit of entry inhibition might be more significant than expected [7,83].

7.2. Inhibition of HBV Envelope Protein Release

Because HBsAg SVPs are considered to induce exhaustion of the adaptive immune responses against HBV [8], the inhibition of HBsAg release can be a target for the HBV suppression. Nucleic acid polymers (NAPs) are single-stranded phosphorothioated oligonucleotides and interact with exposed amphipathic alpha helices in proteins in a sequence-independent manner [84], and some NAPs such as REP 2139 and REP 2165 were shown to inhibit HBsAg release; clinical trials are ongoing [83]. Especially, its combination with pegylated interferon (PEG-IFN) plus tenofovir disoproxil fumarate was reported to show a good response rate for HBsAg loss and HBsAg seroconversion during/after the therapy [85].
NAPs selectively inhibit the secretion of spherical SVPs without affecting Dane particles. Additionally, they reduced also intracellular HBsAg in an in vitro model probably via proteasomal and lysosomal degradation pathways [86].

7.3. Inhibition of HBV Enveloping of the HBV Nucleocapsids

As described before, the interaction between preS1 and nucleocapsids is necessary for the formation of infectious HBV particles. Only one report showed agents that inhibited the interaction between the envelope and the nucleocapsid [87], and further investigation for this step is required.

8. Conclusions

The envelope of HBV is indispensable for the life cycle of HBV and could be a promising target for anti-viral therapies. Because the accumulation of envelope proteins including the deletion mutants in ER cause ER stress leading to carcinogenesis, suppression of the envelope protein expression and/or reduction of its mutants might be a therapeutic target to decrease HCC patients. Further studies to clarify the biology of the HBV envelope and the development of agents targeting the envelope are required for the establishment of novel antiviral therapy.

Author Contributions

J.I., K.S. and M.N. compiled the complete draft and modification of the manuscript. A.M. provided critiques on the work and finalized the article prior to submission. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Japan Society for the Promotion of Science (19K08385) and Japan Agency for Medical Research and Development (JP18fk0310101h0002) to J.I.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Polaris Observatory, C. Global prevalence, treatment, and prevention of hepatitis B virus infection in 2016: A modelling study. Lancet Gastroenterol. Hepatol. 2018, 3, 383–403. [Google Scholar]
  2. Terrault, N.A.; Lok, A.S.F.; McMahon, B.J.; Chang, K.M.; Hwang, J.P.; Jonas, M.M.; Brown, R.S., Jr.; Bzowej, N.H.; Wong, J.B. Update on prevention, diagnosis, and treatment of chronic hepatitis B: AASLD 2018 hepatitis B guidance. Hepatology 2018, 67, 1560–1599. [Google Scholar] [CrossRef] [PubMed]
  3. Marchetti, A.L.; Guo, H. New Insights on Molecular Mechanism of Hepatitis B Virus Covalently Closed Circular DNA Formation. Cells 2020, 9, 2430. [Google Scholar] [CrossRef] [PubMed]
  4. Inoue, J.; Akahane, T.; Nakayama, H.; Kimura, O.; Kobayashi, T.; Kisara, N.; Sato, T.; Morosawa, T.; Izuma, M.; Kakazu, E.; et al. Comparison of hepatitis B virus genotypes B and C among chronically hepatitis B virus-infected patients who received nucleos(t)ide analogs: A multicenter retrospective study. Hepatol. Res. 2019, 49, 1263–1274. [Google Scholar] [CrossRef] [PubMed]
  5. Tseng, C.H.; Hsu, Y.C.; Chen, T.H.; Ji, F.; Chen, I.S.; Tsai, Y.N.; Hai, H.; Thuy, L.T.T.; Hosaka, T.; Sezaki, H.; et al. Hepatocellular carcinoma incidence with tenofovir versus entecavir in chronic hepatitis B: A systematic review and meta-analysis. Lancet Gastroenterol. Hepatol. 2020, 5, 1039–1052. [Google Scholar] [CrossRef]
  6. Xia, Y.; Liang, T.J. Development of Direct-acting Antiviral and Host-targeting Agents for Treatment of Hepatitis B Virus Infection. Gastroenterology 2019, 156, 311–324. [Google Scholar] [CrossRef]
  7. Martinez, M.G.; Villeret, F.; Testoni, B.; Zoulim, F. Can we cure hepatitis B virus with novel direct-acting antivirals? Liver Int. 2020, 40 (Suppl. 1), 27–34. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Vaillant, A. HBsAg, Subviral Particles, and Their Clearance in Establishing a Functional Cure of Chronic Hepatitis B Virus Infection. ACS Infect. Dis. 2020. [Google Scholar] [CrossRef]
  9. Tsukuda, S.; Watashi, K. Hepatitis B virus biology and life cycle. Antivir. Res. 2020, 182, 104925. [Google Scholar] [CrossRef]
  10. Carman, W.F.; Jacyna, M.R.; Hadziyannis, S.; Karayiannis, P.; McGarvey, M.J.; Makris, A.; Thomas, H.C. Mutation preventing formation of hepatitis B e antigen in patients with chronic hepatitis B infection. Lancet 1989, 2, 588–591. [Google Scholar] [CrossRef]
  11. Okamoto, H.; Yotsumoto, S.; Akahane, Y.; Yamanaka, T.; Miyazaki, Y.; Sugai, Y.; Tsuda, F.; Tanaka, T.; Miyakawa, Y.; Mayumi, M. Hepatitis B viruses with precore region defects prevail in persistently infected hosts along with seroconversion to the antibody against e antigen. J. Virol. 1990, 64, 1298–1303. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Herrscher, C.; Roingeard, P.; Blanchard, E. Hepatitis B Virus Entry into Cells. Cells 2020, 9, 1486. [Google Scholar] [CrossRef] [PubMed]
  13. Heermann, K.H.; Goldmann, U.; Schwartz, W.; Seyffarth, T.; Baumgarten, H.; Gerlich, W.H. Large surface proteins of hepatitis B virus containing the pre-s sequence. J. Virol. 1984, 52, 396–402. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Liu, S.; Koh, S.S.; Lee, C.G. Hepatitis B Virus X Protein and Hepatocarcinogenesis. Int. J. Mol. Sci. 2016, 17, 940. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Lin, W.L.; Hung, J.H.; Huang, W. Association of the Hepatitis B Virus Large Surface Protein with Viral Infectivity and Endoplasmic Reticulum Stress-mediated Liver Carcinogenesis. Cells 2020, 9, 2052. [Google Scholar] [CrossRef]
  16. Carman, W.F.; Zanetti, A.R.; Karayiannis, P.; Waters, J.; Manzillo, G.; Tanzi, E.; Zuckerman, A.J.; Thomas, H.C. Vaccine-induced escape mutant of hepatitis B virus. Lancet 1990, 336, 325–329. [Google Scholar] [CrossRef]
  17. Gripon, P.; Le Seyec, J.; Rumin, S.; Guguen-Guillouzo, C. Myristylation of the hepatitis B virus large surface protein is essential for viral infectivity. Virology 1995, 213, 292–299. [Google Scholar] [CrossRef]
  18. Udenwobele, D.I.; Su, R.C.; Good, S.V.; Ball, T.B.; Varma Shrivastav, S.; Shrivastav, A. Myristoylation: An Important Protein Modification in the Immune Response. Front. Immunol. 2017, 8, 751. [Google Scholar] [CrossRef] [Green Version]
  19. Prange, R.; Streeck, R.E. Novel transmembrane topology of the hepatitis B virus envelope proteins. EMBO J. 1995, 14, 247–256. [Google Scholar] [CrossRef]
  20. Pastor, F.; Herrscher, C.; Patient, R.; Eymieux, S.; Moreau, A.; Burlaud-Gaillard, J.; Seigneuret, F.; de Rocquigny, H.; Roingeard, P.; Hourioux, C. Direct interaction between the hepatitis B virus core and envelope proteins analyzed in a cellular context. Sci. Rep. 2019, 9, 16178. [Google Scholar] [CrossRef] [Green Version]
  21. Yan, H.; Zhong, G.; Xu, G.; He, W.; Jing, Z.; Gao, Z.; Huang, Y.; Qi, Y.; Peng, B.; Wang, H.; et al. Sodium taurocholate cotransporting polypeptide is a functional receptor for human hepatitis B and D virus. eLife 2012, 1, e00049. [Google Scholar] [CrossRef]
  22. Dobrica, M.O.; Lazar, C.; Branza-Nichita, N. N-Glycosylation and N-Glycan Processing in HBV Biology and Pathogenesis. Cells 2020, 9, 1404. [Google Scholar] [CrossRef]
  23. Schmitt, S.; Glebe, D.; Tolle, T.K.; Lochnit, G.; Linder, D.; Geyer, R.; Gerlich, W.H. Structure of pre-S2 N- and O-linked glycans in surface proteins from different genotypes of hepatitis B virus. J. Gen. Virol. 2004, 85, 2045–2053. [Google Scholar] [CrossRef] [PubMed]
  24. Mehta, A.; Lu, X.; Block, T.M.; Blumberg, B.S.; Dwek, R.A. Hepatitis B virus (HBV) envelope glycoproteins vary drastically in their sensitivity to glycan processing: Evidence that alteration of a single N-linked glycosylation site can regulate HBV secretion. Proc. Natl. Acad. Sci. USA 1997, 94, 1822–1827. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Wang, M.; Zou, X.; Tian, D.; Hu, S.; Jiang, L. Role of Dendritic Cell-Specific ICAM-3-Grabbing Nonintegrin on Dendritic Cells in the Recognition of Hepatitis B Virus. Viral Immunol. 2015, 28, 331–338. [Google Scholar] [CrossRef]
  26. Yu, D.M.; Li, X.H.; Mom, V.; Lu, Z.H.; Liao, X.W.; Han, Y.; Pichoud, C.; Gong, Q.M.; Zhang, D.H.; Zhang, Y.; et al. N-glycosylation mutations within hepatitis B virus surface major hydrophilic region contribute mostly to immune escape. J. Hepatol. 2014, 60, 515–522. [Google Scholar] [CrossRef]
  27. Kian Chua, P.; Lin, M.H.; Shih, C. Potent inhibition of human Hepatitis B virus replication by a host factor Vps4. Virology 2006, 354, 1–6. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Lambert, C.; Doring, T.; Prange, R. Hepatitis B virus maturation is sensitive to functional inhibition of ESCRT-III, Vps4, and gamma 2-adaptin. J. Virol. 2007, 81, 9050–9060. [Google Scholar] [CrossRef] [Green Version]
  29. Watanabe, T.; Sorensen, E.M.; Naito, A.; Schott, M.; Kim, S.; Ahlquist, P. Involvement of host cellular multivesicular body functions in hepatitis B virus budding. Proc. Natl. Acad. Sci. USA 2007, 104, 10205–10210. [Google Scholar] [CrossRef] [Green Version]
  30. Prange, R. Host factors involved in hepatitis B virus maturation, assembly, and egress. Med. Microbiol. Immunol. 2012, 201, 449–461. [Google Scholar] [CrossRef]
  31. Poisson, F.; Severac, A.; Hourioux, C.; Goudeau, A.; Roingeard, P. Both pre-S1 and S domains of hepatitis B virus envelope proteins interact with the core particle. Virology 1997, 228, 115–120. [Google Scholar] [CrossRef] [Green Version]
  32. Hourioux, C.; Touze, A.; Coursaget, P.; Roingeard, P. DNA-containing and empty hepatitis B virus core particles bind similarly to envelope protein domains. J. Gen. Virol. 2000, 81, 1099–1101. [Google Scholar] [CrossRef]
  33. Prange, R.; Werr, M.; Loffler-Mary, H. Chaperones involved in hepatitis B virus morphogenesis. Biol. Chem. 1999, 380, 305–314. [Google Scholar] [CrossRef] [PubMed]
  34. Lambert, C.; Prange, R. Chaperone action in the posttranslational topological reorientation of the hepatitis B virus large envelope protein: Implications for translocational regulation. Proc. Natl. Acad. Sci. USA 2003, 100, 5199–5204. [Google Scholar] [CrossRef] [Green Version]
  35. Hartmann-Stuhler, C.; Prange, R. Hepatitis B virus large envelope protein interacts with gamma2-adaptin, a clathrin adaptor-related protein. J. Virol. 2001, 75, 5343–5351. [Google Scholar] [CrossRef] [Green Version]
  36. Inoue, J.; Ninomiya, M.; Umetsu, T.; Nakamura, T.; Kogure, T.; Kakazu, E.; Iwata, T.; Takai, S.; Sano, A.; Fukuda, M.; et al. Small Interfering RNA Screening for the Small GTPase Rab Proteins Identifies Rab5B as a Major Regulator of Hepatitis B Virus Production. J. Virol. 2019, 93, e00621-e19. [Google Scholar] [CrossRef] [Green Version]
  37. Hoffmann, J.; Boehm, C.; Himmelsbach, K.; Donnerhak, C.; Roettger, H.; Weiss, T.S.; Ploen, D.; Hildt, E. Identification of alpha-taxilin as an essential factor for the life cycle of hepatitis B virus. J. Hepatol. 2013, 59, 934–941. [Google Scholar] [CrossRef] [PubMed]
  38. Ninomiya, M.; Inoue, J.; Krueger, E.W.; Chen, J.; Cao, H.; Masamune, A.; McNiven, M.A. The Exosome-Associated Tetraspanin CD63 Contributes to the Efficient Assembly and Infectivity of the Hepatitis B Virus. Hepatol. Commun. 2021. [Google Scholar] [CrossRef]
  39. Rost, M.; Mann, S.; Lambert, C.; Doring, T.; Thome, N.; Prange, R. Gamma-adaptin, a novel ubiquitin-interacting adaptor, and Nedd4 ubiquitin ligase control hepatitis B virus maturation. J. Biol. Chem. 2006, 281, 29297–29308. [Google Scholar] [CrossRef] [Green Version]
  40. Seitz, S.; Iancu, C.; Volz, T.; Mier, W.; Dandri, M.; Urban, S.; Bartenschlager, R. A Slow Maturation Process Renders Hepatitis B Virus Infectious. Cell Host Microbe 2016, 20, 25–35. [Google Scholar] [CrossRef] [Green Version]
  41. Inoue, J.; Krueger, E.W.; Chen, J.; Cao, H.; Ninomiya, M.; McNiven, M.A. HBV secretion is regulated through the activation of endocytic and autophagic compartments mediated by Rab7 stimulation. J. Cell Sci. 2015, 128, 1696–1706. [Google Scholar] [CrossRef] [Green Version]
  42. Zeyen, L.; Doring, T.; Stieler, J.T.; Prange, R. Hepatitis B subviral envelope particles use the COPII machinery for intracellular transport via selective exploitation of Sec24A and Sec23B. Cell Microbiol. 2020, 22, e13181. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Patient, R.; Hourioux, C.; Roingeard, P. Morphogenesis of hepatitis B virus and its subviral envelope particles. Cell. Microbiol. 2009, 11, 1561–1570. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Jiang, B.; Himmelsbach, K.; Ren, H.; Boller, K.; Hildt, E. Subviral Hepatitis B Virus Filaments, like Infectious Viral Particles, Are Released via Multivesicular Bodies. J. Virol. 2016, 90, 3330–3341. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Takeda, K.; Maruki, M.; Yamagaito, T.; Muramatsu, M.; Sakai, Y.; Tobimatsu, H.; Kobayashi, H.; Mizuno, Y.; Hamaguchi, Y. Highly sensitive detection of hepatitis B virus surface antigen by use of a semiautomated immune complex transfer chemiluminescence enzyme immunoassay. J. Clin. Microbiol. 2013, 51, 2238–2244. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Kusumoto, S.; Tanaka, Y.; Suzuki, R.; Watanabe, T.; Nakata, M.; Sakai, R.; Fukushima, N.; Fukushima, T.; Moriuchi, Y.; Itoh, K.; et al. Ultra-high sensitivity HBsAg assay can diagnose HBV reactivation following rituximab-based therapy in patients with lymphoma. J. Hepatol. 2020, 73, 285–293. [Google Scholar] [CrossRef] [PubMed]
  47. Shinkai, N.; Kusumoto, S.; Murakami, S.; Ogawa, S.; Ri, M.; Matsui, T.; Tamori, A.; Toyoda, H.; Ishida, T.; Iida, S.; et al. Novel monitoring of hepatitis B reactivation based on ultra-high sensitive hepatitis B surface antigen assay. Liver Int. 2017, 37, 1138–1147. [Google Scholar] [CrossRef]
  48. Matsumoto, A.; Tanaka, E.; Suzuki, Y.; Kobayashi, M.; Tanaka, Y.; Shinkai, N.; Hige, S.; Yatsuhashi, H.; Nagaoka, S.; Chayama, K.; et al. Combination of hepatitis B viral antigens and DNA for prediction of relapse after discontinuation of nucleos(t)ide analogs in patients with chronic hepatitis B. Hepatol Res. 2012, 42, 139–149. [Google Scholar] [CrossRef]
  49. Tseng, T.C.; Liu, C.J.; Yang, H.C.; Su, T.H.; Wang, C.C.; Chen, C.L.; Kuo, S.F.; Liu, C.H.; Chen, P.J.; Chen, D.S.; et al. High levels of hepatitis B surface antigen increase risk of hepatocellular carcinoma in patients with low HBV load. Gastroenterology 2012, 142, 1140–1149.e3, quiz 1113–1144. [Google Scholar] [CrossRef] [Green Version]
  50. Pfefferkorn, M.; Bohm, S.; Schott, T.; Deichsel, D.; Bremer, C.M.; Schroder, K.; Gerlich, W.H.; Glebe, D.; Berg, T.; van Bommel, F. Quantification of large and middle proteins of hepatitis B virus surface antigen (HBsAg) as a novel tool for the identification of inactive HBV carriers. Gut 2018, 67, 2045–2053. [Google Scholar] [CrossRef]
  51. Pfefferkorn, M.; Schott, T.; Bohm, S.; Deichsel, D.; Felkel, C.; Gerlich, W.H.; Glebe, D.; Wat, C.; Pavlovic, V.; Heyne, R.; et al. Composition of HBsAg is predictive of HBsAg loss during treatment in patients with HBeAg-positive chronic hepatitis B. J. Hepatol. 2021, 74, 283–292. [Google Scholar] [CrossRef] [PubMed]
  52. Inoue, J.; Nakamura, T.; Masamune, A. Roles of Hepatitis B Virus Mutations in the Viral Reactivation after Immunosuppression Therapies. Viruses 2019, 11, 457. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Ito, K.; Qin, Y.; Guarnieri, M.; Garcia, T.; Kwei, K.; Mizokami, M.; Zhang, J.; Li, J.; Wands, J.R.; Tong, S. Impairment of hepatitis B virus virion secretion by single-amino-acid substitutions in the small envelope protein and rescue by a novel glycosylation site. J. Virol. 2010, 84, 12850–12861. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Lin, Y.M.; Chen, B.F. A putative hepatitis B virus splice variant associated with chronic hepatitis and liver cirrhosis. Virology 2017, 510, 224–233. [Google Scholar] [CrossRef] [PubMed]
  55. Yeung, P.; Wong, D.K.; Lai, C.L.; Fung, J.; Seto, W.K.; Yuen, M.F. Association of hepatitis B virus pre-S deletions with the development of hepatocellular carcinoma in chronic hepatitis B. J. Infect. Dis. 2011, 203, 646–654. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Pollicino, T.; Saitta, C.; Raimondo, G. Hepatocellular carcinoma: The point of view of the hepatitis B virus. Carcinogenesis 2011, 32, 1122–1132. [Google Scholar] [CrossRef] [PubMed]
  57. Shen, F.C.; Su, I.J.; Wu, H.C.; Hsieh, Y.H.; Yao, W.J.; Young, K.C.; Chang, T.C.; Hsieh, H.C.; Tsai, H.N.; Huang, W. A pre-S gene chip to detect pre-S deletions in hepatitis B virus large surface antigen as a predictive marker for hepatoma risk in chronic hepatitis B virus carriers. J. Biomed. Sci. 2009, 16, 84. [Google Scholar] [CrossRef] [Green Version]
  58. Wang, H.C.; Huang, W.; Lai, M.D.; Su, I.J. Hepatitis B virus pre-S mutants, endoplasmic reticulum stress and hepatocarcinogenesis. Cancer Sci. 2006, 97, 683–688. [Google Scholar] [CrossRef]
  59. Li, J.; Li, J.; Chen, S.; Yuan, Q.; Zhang, J.; Wu, J.; Jiang, Q.; Wang, Q.; Xia, N.S.; Zhang, J.; et al. Naturally occurring 5′ preS1 deletions markedly enhance replication and infectivity of HBV genotype B and genotype C. Gut 2021, 70, 575–584. [Google Scholar] [CrossRef]
  60. Murayama, A.; Yamada, N.; Osaki, Y.; Shiina, M.; Aly, H.H.; Iwamoto, M.; Tsukuda, S.; Watashi, K.; Matsuda, M.; Suzuki, R.; et al. N-Terminal PreS1 Sequence Regulates Efficient Infection of Cell-Culture-Generated Hepatitis B Virus. Hepatology 2021, 73, 520–532. [Google Scholar] [CrossRef]
  61. Pollicino, T.; Cacciola, I.; Saffioti, F.; Raimondo, G. Hepatitis B virus PreS/S gene variants: Pathobiology and clinical implications. J. Hepatol. 2014, 61, 408–417. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Liang, Y.J.; Teng, W.; Chen, C.L.; Sun, C.P.; Teng, R.D.; Huang, Y.H.; Liang, K.H.; Chen, Y.W.; Lin, C.C.; Su, C.W.; et al. Clinical implications on HBV preS/S mutations and the effects of preS2 deletion on mitochondria, liver fibrosis, and cancer development. Hepatology 2021. [Google Scholar] [CrossRef]
  63. Chisari, F.V.; Ferrari, C. Hepatitis B virus immunopathogenesis. Annu. Rev. Immunol. 1995, 13, 29–60. [Google Scholar] [CrossRef] [PubMed]
  64. Locarnini, S.; Zoulim, F. Molecular genetics of HBV infection. Antivir. Ther. 2010, 15 (Suppl. 3), 3–14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Hadziyannis, S.; Gerber, M.A.; Vissoulis, C.; Popper, H. Cytoplasmic hepatitis B antigen in “ground-glass” hepatocytes of carriers. Arch. Pathol. 1973, 96, 327–330. [Google Scholar] [PubMed]
  66. Koch, S.; Freytag von Loringhoven, A.; Kahmann, R.; Hofschneider, P.H.; Koshy, R. The genetic organization of integrated hepatitis B virus DNA in the human hepatoma cell line PLC/PRF/5. Nucleic Acids Res. 1984, 12, 6871–6886. [Google Scholar] [CrossRef] [Green Version]
  67. Teng, C.F.; Hsieh, W.C.; Wu, H.C.; Lin, Y.J.; Tsai, H.W.; Huang, W.; Su, I.J. Hepatitis B Virus Pre-S2 Mutant Induces Aerobic Glycolysis through Mammalian Target of Rapamycin Signal Cascade. PLoS ONE 2015, 10, e0122373. [Google Scholar] [CrossRef] [Green Version]
  68. Teng, C.F.; Wu, H.C.; Hsieh, W.C.; Tsai, H.W.; Su, I.J. Activation of ATP citrate lyase by mTOR signal induces disturbed lipid metabolism in hepatitis B virus pre-S2 mutant tumorigenesis. J. Virol. 2015, 89, 605–614. [Google Scholar] [CrossRef] [Green Version]
  69. Teng, C.F.; Wu, H.C.; Tsai, H.W.; Shiah, H.S.; Huang, W.; Su, I.J. Novel feedback inhibition of surface antigen synthesis by mammalian target of rapamycin (mTOR) signal and its implication for hepatitis B virus tumorigenesis and therapy. Hepatology 2011, 54, 1199–1207. [Google Scholar] [CrossRef]
  70. Hosaka, T.; Suzuki, F.; Kobayashi, M.; Seko, Y.; Kawamura, Y.; Sezaki, H.; Akuta, N.; Suzuki, Y.; Saitoh, S.; Arase, Y.; et al. Long-term entecavir treatment reduces hepatocellular carcinoma incidence in patients with hepatitis B virus infection. Hepatology 2013, 58, 98–107. [Google Scholar] [CrossRef]
  71. Wu, C.Y.; Lin, J.T.; Ho, H.J.; Su, C.W.; Lee, T.Y.; Wang, S.Y.; Wu, C.; Wu, J.C. Association of nucleos(t)ide analogue therapy with reduced risk of hepatocellular carcinoma in patients with chronic hepatitis B: A nationwide cohort study. Gastroenterology 2014, 147, 143–151.e5. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Boyd, A.; Lacombe, K.; Lavocat, F.; Maylin, S.; Miailhes, P.; Lascoux-Combe, C.; Delaugerre, C.; Girard, P.M.; Zoulim, F. Decay of ccc-DNA marks persistence of intrahepatic viral DNA synthesis under tenofovir in HIV-HBV co-infected patients. J. Hepatol. 2016, 65, 683–691. [Google Scholar] [CrossRef] [PubMed]
  73. Petersen, J.; Dandri, M.; Mier, W.; Lutgehetmann, M.; Volz, T.; von Weizsacker, F.; Haberkorn, U.; Fischer, L.; Pollok, J.M.; Erbes, B.; et al. Prevention of hepatitis B virus infection in vivo by entry inhibitors derived from the large envelope protein. Nat. Biotechnol. 2008, 26, 335–341. [Google Scholar] [CrossRef] [PubMed]
  74. Lutgehetmann, M.; Mancke, L.V.; Volz, T.; Helbig, M.; Allweiss, L.; Bornscheuer, T.; Pollok, J.M.; Lohse, A.W.; Petersen, J.; Urban, S.; et al. Humanized chimeric uPA mouse model for the study of hepatitis B and D virus interactions and preclinical drug evaluation. Hepatology 2012, 55, 685–694. [Google Scholar] [CrossRef]
  75. Nkongolo, S.; Ni, Y.; Lempp, F.A.; Kaufman, C.; Lindner, T.; Esser-Nobis, K.; Lohmann, V.; Mier, W.; Mehrle, S.; Urban, S. Cyclosporin A inhibits hepatitis B and hepatitis D virus entry by cyclophilin-independent interference with the NTCP receptor. J. Hepatol. 2014, 60, 723–731. [Google Scholar] [CrossRef]
  76. Watashi, K.; Sluder, A.; Daito, T.; Matsunaga, S.; Ryo, A.; Nagamori, S.; Iwamoto, M.; Nakajima, S.; Tsukuda, S.; Borroto-Esoda, K.; et al. Cyclosporin A and its analogs inhibit hepatitis B virus entry into cultured hepatocytes through targeting a membrane transporter, sodium taurocholate cotransporting polypeptide (NTCP). Hepatology 2014, 59, 1726–1737. [Google Scholar] [CrossRef]
  77. Shimura, S.; Watashi, K.; Fukano, K.; Peel, M.; Sluder, A.; Kawai, F.; Iwamoto, M.; Tsukuda, S.; Takeuchi, J.S.; Miyake, T.; et al. Cyclosporin derivatives inhibit hepatitis B virus entry without interfering with NTCP transporter activity. J. Hepatol. 2017, 66, 685–692. [Google Scholar] [CrossRef]
  78. Kaneko, M.; Watashi, K.; Kamisuki, S.; Matsunaga, H.; Iwamoto, M.; Kawai, F.; Ohashi, H.; Tsukuda, S.; Shimura, S.; Suzuki, R.; et al. A Novel Tricyclic Polyketide, Vanitaracin A, Specifically Inhibits the Entry of Hepatitis B and D Viruses by Targeting Sodium Taurocholate Cotransporting Polypeptide. J. Virol. 2015, 89, 11945–11953. [Google Scholar] [CrossRef] [Green Version]
  79. Tsukuda, S.; Watashi, K.; Iwamoto, M.; Suzuki, R.; Aizaki, H.; Okada, M.; Sugiyama, M.; Kojima, S.; Tanaka, Y.; Mizokami, M.; et al. Dysregulation of retinoic acid receptor diminishes hepatocyte permissiveness to hepatitis B virus infection through modulation of sodium taurocholate cotransporting polypeptide (NTCP) expression. J. Biol. Chem. 2015, 290, 5673–5684. [Google Scholar] [CrossRef] [Green Version]
  80. Xiao, Y.; Liu, C.; Tang, W.; Zhang, H.; Chen, X. Evans Blue Inhibits HBV Replication Through a Dual Antiviral Mechanism by Targeting Virus Binding and Capsid Assembly. Front. Microbiol. 2019, 10, 2638. [Google Scholar] [CrossRef]
  81. Maravelia, P.; Frelin, L.; Ni, Y.; Caro Perez, N.; Ahlen, G.; Jagya, N.; Verch, G.; Verhoye, L.; Pater, L.; Johansson, M.; et al. Blocking Entry of Hepatitis B and D Viruses to Hepatocytes as a Novel Immunotherapy for Treating Chronic Infections. J. Infect. Dis. 2021, 223, 128–138. [Google Scholar] [CrossRef] [Green Version]
  82. Allweiss, L.; Volz, T.; Giersch, K.; Kah, J.; Raffa, G.; Petersen, J.; Lohse, A.W.; Beninati, C.; Pollicino, T.; Urban, S.; et al. Proliferation of primary human hepatocytes and prevention of hepatitis B virus reinfection efficiently deplete nuclear cccDNA in vivo. Gut 2018, 67, 542–552. [Google Scholar] [CrossRef]
  83. Soriano, V.; Barreiro, P.; Cachay, E.; Kottilil, S.; Fernandez-Montero, J.V.; de Mendoza, C. Advances in hepatitis B therapeutics. Ther. Adv. Infect. Dis. 2020, 7, 2049936120965027. [Google Scholar] [CrossRef] [PubMed]
  84. Vaillant, A. Nucleic acid polymers: Broad spectrum antiviral activity, antiviral mechanisms and optimization for the treatment of hepatitis B and hepatitis D infection. Antivir. Res. 2016, 133, 32–40. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Bazinet, M.; Pantea, V.; Placinta, G.; Moscalu, I.; Cebotarescu, V.; Cojuhari, L.; Jimbei, P.; Iarovoi, L.; Smesnoi, V.; Musteata, T.; et al. Safety and Efficacy of 48 Weeks REP 2139 or REP 2165, Tenofovir Disoproxil, and Pegylated Interferon Alfa-2a in Patients With Chronic HBV Infection Naive to Nucleos(t)ide Therapy. Gastroenterology 2020, 158, 2180–2194. [Google Scholar] [CrossRef] [PubMed]
  86. Boulon, R.; Blanchet, M.; Lemasson, M.; Vaillant, A.; Labonte, P. Characterization of the antiviral effects of REP 2139 on the HBV lifecycle in vitro. Antivir. Res. 2020, 183, 104853. [Google Scholar] [CrossRef] [PubMed]
  87. Asif-Ullah, M.; Choi, K.J.; Choi, K.I.; Jeong, Y.J.; Yu, Y.G. Identification of compounds that inhibit the interaction between core and surface protein of hepatitis B virus. Antivir. Res. 2006, 70, 85–90. [Google Scholar] [CrossRef]
Figure 1. Schema of HBV-associated particles. Dane particles are 42 nm in diameter, and their envelope consists of large/middle/small hepatitis B surface proteins (LHBs/MHBs/SHBs) and a nucleocapsid containing viral genome and polymerase. There are 2 forms of subviral particles, filament and sphere. The former consists of LHBs, MHBs and SHBs, and the latter consists of MHBs and SHBs.
Figure 1. Schema of HBV-associated particles. Dane particles are 42 nm in diameter, and their envelope consists of large/middle/small hepatitis B surface proteins (LHBs/MHBs/SHBs) and a nucleocapsid containing viral genome and polymerase. There are 2 forms of subviral particles, filament and sphere. The former consists of LHBs, MHBs and SHBs, and the latter consists of MHBs and SHBs.
Viruses 13 01124 g001
Figure 2. The structure of HBV envelope proteins. (A) LHBs contains preS1, preS2 and S domains. Typically, the length is 400 amino acids for most genotypes and 389 amino acids for genotype D. Myristic acid, which is a hydrophobic moiety, is attached at the N-terminus. MHBs contains preS2 and S domains, and SHBs contains only the S domain. There are 4 transmembrane domains (I–IV) in the S domain. Two N-glycosylation sites are indicated as “Y” and the O-glycosylation site for genotypes C and D is indicated as “Y*”. “(Y)” indicates only the part of the proteins that is glycosylated. (B) Two-dimensional conformation of LHBs. The myristic acid acts as an anchor and the preS1/preS2 domains in LHBs move between the cytoplasmic side and the luminal side.
Figure 2. The structure of HBV envelope proteins. (A) LHBs contains preS1, preS2 and S domains. Typically, the length is 400 amino acids for most genotypes and 389 amino acids for genotype D. Myristic acid, which is a hydrophobic moiety, is attached at the N-terminus. MHBs contains preS2 and S domains, and SHBs contains only the S domain. There are 4 transmembrane domains (I–IV) in the S domain. Two N-glycosylation sites are indicated as “Y” and the O-glycosylation site for genotypes C and D is indicated as “Y*”. “(Y)” indicates only the part of the proteins that is glycosylated. (B) Two-dimensional conformation of LHBs. The myristic acid acts as an anchor and the preS1/preS2 domains in LHBs move between the cytoplasmic side and the luminal side.
Viruses 13 01124 g002
Figure 3. Model depicting the steps of HBV envelope formation. Molecules/organelles in the cells that are utilized for the envelope formation are shown. Nucleocapsid is considered to interact with preS1/preS2 of LHBs on the cytoplasmic side on the multivesicular body (MVB) membrane and buds into the MVB with the action of endosomal sorting complex required for transport (ESCRT). After the release from cells, probably via exocytosis, the position of the preS1/preS2 domains changes to the luminal side to attach to the surface of hepatocytes. LHBs accumulation in the endoplasmic reticulum (ER), which is accelerated by preS mutation, leads to carcinogenesis.
Figure 3. Model depicting the steps of HBV envelope formation. Molecules/organelles in the cells that are utilized for the envelope formation are shown. Nucleocapsid is considered to interact with preS1/preS2 of LHBs on the cytoplasmic side on the multivesicular body (MVB) membrane and buds into the MVB with the action of endosomal sorting complex required for transport (ESCRT). After the release from cells, probably via exocytosis, the position of the preS1/preS2 domains changes to the luminal side to attach to the surface of hepatocytes. LHBs accumulation in the endoplasmic reticulum (ER), which is accelerated by preS mutation, leads to carcinogenesis.
Viruses 13 01124 g003
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Inoue, J.; Sato, K.; Ninomiya, M.; Masamune, A. Envelope Proteins of Hepatitis B Virus: Molecular Biology and Involvement in Carcinogenesis. Viruses 2021, 13, 1124. https://doi.org/10.3390/v13061124

AMA Style

Inoue J, Sato K, Ninomiya M, Masamune A. Envelope Proteins of Hepatitis B Virus: Molecular Biology and Involvement in Carcinogenesis. Viruses. 2021; 13(6):1124. https://doi.org/10.3390/v13061124

Chicago/Turabian Style

Inoue, Jun, Kosuke Sato, Masashi Ninomiya, and Atsushi Masamune. 2021. "Envelope Proteins of Hepatitis B Virus: Molecular Biology and Involvement in Carcinogenesis" Viruses 13, no. 6: 1124. https://doi.org/10.3390/v13061124

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

Article Metrics

Back to TopTop