Abstract
The core (capsid) protein of hepatitis B virus (HBV) is the building block of nucleocapsids where viral DNA reverse transcriptional replication takes place and mediates virus-host cell interaction important for the persistence of HBV infection. The pleiotropic role of core protein (Cp) in HBV replication makes it an attractive target for antiviral therapies of chronic hepatitis B, a disease that affects more than 257 million people worldwide without a cure. Recent clinical studies indicate that core protein allosteric modulators (CpAMs) have a great promise as a key component of hepatitis B curative therapies. Particularly, it has been demonstrated that modulation of Cp dimer-dimer interactions by several chemical series of CpAMs not only inhibit nucleocapsid assembly and viral DNA replication, but also induce the disassembly of double-stranded DNA-containing nucleocapsids to prevent the synthesis of cccDNA. Moreover, the different chemotypes of CpAMs modulate Cp assembly by interaction with distinct amino acid residues at the HAP pocket between Cp dimer-dimer interfaces, which results in the assembly of Cp dimers into either non-capsid Cp polymers (type I CpAMs) or empty capsids with distinct physical property (type II CpAMs). The different CpAMs also differentially modulate Cp metabolism and subcellular distribution, which may impact cccDNA metabolism and host antiviral immune responses, the critical factors for the cure of chronic HBV infection. This review article highlights the recent research progress on the structure and function of core protein in HBV replication cycle, the mode of action of CpAMs, as well as the current status and perspectives on the discovery and development of core protein-targeting antivirals. This article forms part of a symposium in Antiviral Research on “Wide-ranging immune and direct-acting antiviral approaches to curing HBV and HDV infections.”
Keywords: Hepatitis B virus, capsid, nucleocapsid, core protein allosteric modulators, cccDNA
1. Introduction
Hepatitis B virus (HBV) chronically infects 257 million people worldwide and causes approximately 880 thousand deaths annually due to cirrhosis, hepatocellular carcinoma (HCC) and liver failure (Polaris Observatory, 2018). The current standard of care for chronic HBV infection, that includes six nucleos(t)ide analog viral DNA polymerase inhibitors (NUCs) and pegylated interferon-alpha (pgIFN-α), can significantly reduce viral load and prevent liver disease progression, but rarely induce the loss of HBV surface antigen (HBsAg) or seroconvertion to anti-HBs, the indication of durable immune control of HBV infection or a functional cure (Block et al., 2013; Dienstag, 2009; Lok et al., 2017; Perrillo, 2009). Once started, termination of the NUC-based antiviral therapy usually results in the rebound of viral replication and flares of liver diseases (Alter et al., 2018). The failure to achieve a functional cure by the NUC therapy is attributed to the lack of direct effects on viral covalently closed circular (ccc) DNA in infected hepatocytes and inability to restore a functional antiviral immune response (Chang et al., 2014). Therefore, novel antiviral agents that can induce the loss or inactivation of cccDNA and activate a functional antiviral immune response with a finite therapeutic duration are urgently needed (Liang et al., 2015). Recently, HBV core protein (Cp), the building block of nucleocapsids where viral DNA replication takes place, has emerged as a promising target for the development of novel antiviral therapeutics (Tang et al., 2017; Zlotnick et al., 2015). Several core protein allosteric modulators (CpAMs) are currently under clinical development and have demonstrated a great promise as a key component of curative therapeutics for chronic hepatitis B (Cole, 2016; Yang et al., 2019b). This review highlights the recent research progress in our understanding of the structure and function of core protein in the HBV replication cycle, the mode of CpAM action, the current status and our perspectives on the discovery and development of core protein-targeting antivirals.
2. Structure and function of core protein in HBV life cycle
2.1. Cp is the building block of capsids and plays an essential role during multiple stages of viral replication in the cytoplasm
HBV is a para-retrovirus with a 3.2 kb, partially double-stranded, relaxed circular (rc) DNA genome that encode seven viral proteins, including the small (S), middle (M) and large (L) envelope proteins, core (capsid) protein (Cp), DNA polymerase (Pol), pre-core protein that is proteolytically processed and secreted as soluble e antigen (HBeAg), and X protein (HBx) (Hu et al., 2019). HBV replicates its DNA genome via the reverse transcription of a RNA pregenome in the cytoplasmic nucleocapsids of infected hepatocytes (Mason et al., 1982; Summers and Mason, 1982). As the building block of HBV nucleocapsid, Cp plays distinct roles at multiple steps of the HBV replication cycle. As illustrated in Fig. 1, by binding to its specific receptor, sodium taurocholate cotransporting polypeptide (NTCP), on the plasma membrane, the HBV virion enters hepatocyte by endocytosis. The nucleocapsid is delivered into the cytoplasm after the fusion of the viral envelope with the membrane of the endocytic vesicle (Yan et al., 2012). The Cp on the shell of nucleocapsid binds importin-β and directs the transport of the nucleocapsid to the nuclear pore complex (NPC), where the nucleocapsid disassembles and releases the rcDNA genome into the nucleus for the synthesis of cccDNA (Guo et al., 2010; Guo and Guo, 2015; Schmitz et al., 2010). The cccDNA serves as a template for the transcription of four major species of viral RNA. The 3.5 kb pre-genomic (pg) RNA translates Cp and Pol. Binding of Pol to the stem-loop structure (ε) at the 5’ terminal region of pgRNA initiates the packaging of the Pol-pgRNA complex by 120 copies of Cp dimers to form the nucleocapsid (Bartenschlager and Schaller, 1992; Tavis et al., 1994; Wang and Seeger, 1993). Inside the nucleocapsid, Pol reverse transcribes the pgRNA first to single-stranded (ss) DNA and then to rcDNA (Mason et al., 1982; Summers and Mason, 1982). The deficiency of second-strand DNA synthesis in the nucleocapsids assembled with the W124A mutant or a C-terminally truncated Cp strongly indicates that Cp does not builds an inert container to provide space for viral DNA synthesis, but actively participates in viral DNA synthesis, probably via interaction with Pol, pgRNA/DNA or encapsidated host cellular factors (Nassal, 1992; Tan et al., 2015). Through interaction with envelope proteins studded on the membrane of multi-vesicle bodies (MVBs), Cp directs the envelopment of the rcDNA-containing “mature” nucleocapsids for the assembly and secretion of virion particles (Hu and Liu, 2017). Alternatively, Cp can also mediate the transport of mature nucleocapsids to the nuclear pore complexes (NPCs) for the import of rcDNA into the nuclei to amplify the cccDNA pool (Guo et al., 2007; Guo et al., 2010; Tuttleman et al., 1986). Due to the strong tendency of Cp self-assembly into empty capsids, the nucleocapsids only constitute less than 10% of total capsids in an HBV infected hepatocyte (Ning et al., 2011). Similar to the mature nucleocapsids, the less matured nucleocapsids-containing reverse transcriptional intermediates as well as the empty capsids can also be enveloped and secreted as RNA-containing virions or “empty” virions, respectively. While the RNA-containing virions are 100-fold less abundant than the DNA-containing virions, the empty virions are approximately 100-fold more abundant than the DNA-containing virions in the blood of HBV carriers (Hu and Liu, 2017). It is thus apparent that by assembly of empty capsids and nucleocapsids, Cp involves in almost every steps of viral replication in the cytoplasm, which will be explained in detail below.
Fig. 1. HBV life cycle and steps of HBV replication disrupted by the standard of care medicines (NUCs and IFN-α) and CpAMs.
See text for detailed description on HBV life cycle. NUCs inhibit viral DNA synthesis in the cytoplasmic nucleocapsids. IFN-α inhibits cccDNA transcription and pgRNA encapsidation, in addition to its immunomodulatory activity. Hyperphosphorylated and hypophosphorylated core proteins are illustrated in purple and green, respectively. The positive strand and negative strand of HBV DNA are illustrated in black and red, respectively. DP-rcDNA, deproteinized relaxed circular DNA. MVB, multi-vesicle body.
2.2. Core protein and capsid structures
HBV Cp is a 183-amino acid (aa) polypeptide containing a N-terminal assembly domain (NTD, aa 1–140) and an arginine-rich C-terminal domain (CTD, aa150–183), that are connected by a linker of 9 amino acid residues (Venkatakrishnan and Zlotnick, 2016; Zlotnick et al., 2015) (Fig. 2A). As illustrated in Fig. 2C, the assembly domain has five α helices connected by loops, i.e., the short helix α1 (aa 13–17), longer helices α2 (aa 27–43) and α5 (aa 112–127)), helix bundles α3 (upward) and α4 (downward), an irregular proline-rich loop (aa 128–136) and an extended strand (aa 137–149). The hydrophobic interaction between α3 and α4 helices of two Cp molecules drives the formation of a four-helix bundle at their interface and results in the dimerization of Cp. The dimer is the basic structural and functional unit of Cp.
Fig. 2. Structural features of HBV core protein and HBeAg.
(A) Domain structure of Cp and structural motifs and phosphorylation sites at C-terminal domain are highlighted. (B) Pre-core protein processing and biogenesis of HBeAg and their structure domains are illustrated. (C and D) the secondary structure and dimers of Cp from pdb code 1QGT and HBeAg from pdb code 6CVK are presented with all the five alpha helices color coded to see how the structures differ even though they have the same secondary structures
The structure of HBV capsids assembled in E. coli expressing C-terminally truncated or full-length Cp or HBV infected hepatocytes had been determined by cryo-electron microscopy (Cryo-EM) and X-ray diffraction technologies (Bottcher and Nassal, 2018; Bourne et al., 2006; Wynne et al., 1999; Yu et al., 2013) (Table 1). The predominant capsid species is assembled from 120 copies of Cp dimers arranged in T=4 icosahedral symmetry (Fig. 3). The capsid is approximately 340 Å in diameter with an inner radius of 130 Å and thickness of 20 Å (pdb codes 1QGT, 2G33, 3J2V and 6HTX). The spikes with 25 Å in length and 20 Å in width located on the surface of the capsid are the four-helix bundles of Cp dimers. Cys61 located in the middle of the four-helix bundle can form a disulfide bond to crosslink the Cp dimer. However, such a disulfide bond is not required for capsid assembly (Bourne et al., 2006). In fact, the assembly of capsids is driven by the hydrophobic interaction between the dimer-dimer interfaces. Although each Cp, or subunit, has the same aa sequence, the 240 Cp molecules of a capsid take one of four quasi-equivalent conformations, A, B, C or D (Fig. 3).AB and CD dimers have a nearly two-fold symmetry. Five A subunits surround the five-fold axes and AB dimers extend from the five-fold to quasi-six-fold vertices. CD dimers connect the quasi-six-fold vertices and each quasi-six-fold vertex is surrounded by B–C–D–B–C–D subunits. There are pores around the five-fold, quasi-six-fold, three-fold and local three-fold axes of the capsid (Fig 3). The five-fold and quasi-six-fold axes primarily involve the residues from the irregular proline rich loop (aa 128–136) and the extended strand (aa 137–149). Especially the N137 forms a ring with P134 and P135 providing the structural clamp to hold the ring in place. The three-fold and local three-fold channels involve short helix α1 (aa 13–17) and longer helix α2 (aa 27–43), which are primarily polar residues (D2, K7, R39, E40, E43 and E46).
Table 1.
Structures of HBV capsids and Y132A Cp hexamer with or without CpAMs
| No. | PDB ID | Res. (Å) | Structure | Ligand | Method | Year |
|---|---|---|---|---|---|---|
| 1 | 1QGT | 3.3 | WT (T=4) | NA | X-ray Diffraction | 1999 |
| 2 | 2G33 | 4.0 | WT (T=4) | NA | X-ray Diffraction | 2006 |
| 3 | 2QIJ | 8.9 | WT (T=4) | NA | Electron Microscopy | 2007 |
| 4 | 3J2V | 3.5 | WT (T=4) Cp183 | NA | Electron Microscopy | 2013 |
| 5 | 6HTX | 2.7 | WT (T=4) Cp183 | NA | Electron Microscopy | 2018 |
| 6 | 2G34 | 5.1 | WT (T=4) | HAP1 | X-ray Diffraction | 2006 |
| 7 | 4G93 | 4.2 | WT (T=4) | AT130 | X-ray Diffraction | 2013 |
| 8 | 5D7Y | 3.9 | WT (T=4) | HAP18 | X-ray Diffraction | 2016 |
| 9 | 6BVN | 4.0 | WT (T=3) | HAP-TAMRA | Electron Microscopy | 2018 |
| 10 | 6BVF | 4.0 | WT (T=4) | HAP-TAMRA | Electron Microscopy | 2018 |
| 11 | 6CWT | 3.2 | WT (T=4) | Fab e21 | X-ray Diffraction | 2018 |
| 12 | 6CWD | 3.3 | WT (T=4) | scFv e13 | X-ray Diffraction | 2018 |
| 13 | 3KXS | 2.3 | Y132A | NA | X-ray Diffraction | 2010 |
| 14 | 4BMG | 3.0 | Y132A | NA | X-ray Diffraction | 2013 |
| 15 | 5WTW | 2.6 | Y132A | NA | X-ray Diffraction | 2017 |
| 16 | 5E0I | 2.0 | Y132A | NVR10-001E2 | X-ray Diffraction | 2016 |
| 17 | 5GMZ | 1.7 | Y132A | 4-CH3-HAP | X-ray Diffraction | 2016 |
| 18 | 5WRE | 1.9 | Y132A | HAP_R01 | X-ray Diffraction | 2017 |
| 19 | 5T2P | 1.7 | Y132A | SBA_R01 | X-ray Diffraction | 2017 |
| 20 | 6J10 | 2.3 | Y132A | Ciclopirox | X-ray Diffraction | 2019 |
| 21 | 6HU4 | 2.6 | F97L (T=4) Cp183 | NA | Electron Microscopy | 2018 |
| 22 | 6HU7 | 2.8 | F96L-phos (T=4) Cp183 | NA | Electron Microscopy | 2018 |
Fig. 3. Structure of HBV capsids.
Global icosahedral structure of T=4 HBV capsid from pdb code 1QGTwith highlighted sections showing five-fold, three-fold, quasi-six-fold and local three-fold axis. The A, B, C and D monomers are colored green, yellow, red and blue respectively.
2.3. Physical principles of capsid assembly learned from in vitro assembly system
HBV capsids can be assembled in vitro with the CTD-deleted Cp, Cp149. The in vitro assembly reaction occurs as a function of Cp149 dimer concentration, temperature and ionic strength, and results in capsids that are morphologically indistinguishable from capsids purified from HBV expressing cells (Wingfield et al., 1995; Zlotnick et al., 1997). Study of the in vitro Cp assembly reaction kinetics and characterization of the assembly intermediates and products by light scattering, size exclusion chromatography (SEC), charge detection mass spectrometry (CDMS), small angle X-ray scattering (SAXS) technologies has revealed very interesting physical principles of HBV capsid assembly (Perlmutter and Hagan, 2015).
Kinetically, the assembly of capsids in the in vitro condition is sigmoidal. The initial step is the nucleation of Cp dimers, i.e., the formation of the trimers of Cp dimers, which is slow or the lag phase (Zlotnick, 2005). The assembly incompetent Cp with a Y132A substitution gives a glimpse of intermediate structures - namely the kinetically trapped trimers of dimers namely AB, CD and EF for the capsid formation (Alexander et al., 2013; Packianathan et al., 2010; Zhou et al., 2017). Cp149_Y132A crystallizes in three forms – open triangle (3KXS), closed triangle (4BMG) and spike-contact packing (5WTW). The assembly becomes fast once more nuclei form, which is followed by an asymptotic phase leading to equilibrium as the free Cp dimer depletes (Zlotnick, 2005). Interestingly, analyses of assembled capsids by CDMS revealed that the in vitro assembly reaction initially yields defective and overgrown particles, which slowly anneal and relax back to form icosahedrons (Lutomski et al., 2017). Moreover, a recent time-resolved small angle X-ray scattering analysis of full-length Cp assembly indicates that HBV capsid is assembled through the formation of two transient, on-path intermediates, a trimer of dimers and a partially closed shell consisting of around 40 dimers (Oliver et al., 2020).
Based on the kinetically-limiting model established by Zlotnick and colleagues, thermodynamic parameters such as nucleus size, nucleation rate, ΔGcontact and elongation rate could be extracted. ΔGcontact, the association energy of dimers, is found to be weak (~3.5 kcal/mol), but it is enough to form stable capsids since each subunit must interact with at least three other subunits (~9–12 kcal/mol as association energies are additive) (Ceres and Zlotnick, 2002). The weak association energy allows thermodynamic editing which helps remove incorrect and weakly associated dimers. If ΔGcontact is too high (or, equivalently, temperature is too low), the system becomes kinetically trapped in intermediates which cannot rearrange anymore due to the strong binding. If ΔGcontact is too low (or, equivalently, temperature is too high), the target structure is not sufficiently stable. For example, a V124W substitution in Cp increases buried hydrophobic surface and in turn increases the rate of capsid assembly while substitutions such as Y132A and W102A decreases buried surface and prevents capsid assembly (Packianathan et al., 2010).
2.4. Capsid assembly in hepatocytes and its regulation by viral and host cellular factors
Assembly of capsids in hepatocytes should follow the general physical principles as observed in the in vitro assembly reaction. However, due to the complex intracellular environment, differences do exist. For instance, capsids can be assembled in hepatocytes expressing Cp142, but not Cp149, two forms of Cp with C-terminal truncation after residues 142 and 149, respectively, suggesting that cellular proteins may interact with the linker region to regulate capsid assembly (Ludgate et al., 2016; Wu et al., 2017). As shown in Fig. 2A, the CTD contains four arginine (R)-rich clusters with nucleic acid binding activity (Birnbaum and Nassal, 1990; Su et al., 2016; Yang et al., 1994) and seven serine (S)/ threonine (T) residues that can be phosphorylated or dephosphorylated during the viral replication cycle (Basagoudanavar et al., 2007; Ning et al., 2017; Perlman et al., 2005; Pugh et al., 1989). Expression of the wild-type Cp or Cp with substitutions at the seven phosphor-acceptor residues in the CTD with aspartic acid to mimic hyperphosphorylated Cp, results in assembled capsids without detectable packaged RNA. In contrast, the expression of Cp with substitution of the seven phosphor-acceptor residues with alanine (A) to mimic the dephosphorylated Cp results in assembled capsids that non-specifically package viral or cellular RNA (Heger-Stevic et al., 2018; Hu and Liu, 2017; Ludgate et al., 2016; Ning et al., 2011; Su et al., 2016). This is because the arginine-rich clusters in the CTD can non-specifically bind RNA through charge interaction. The RNA binding activity of the CTD can be neutralized by the introduction of negatively charged moieties through phosphorylation (Chua et al., 2010).
The selective packaging of Pol-pgRNA complex into a nucleocapsid is a key step of HBV replication and subjected to regulation by many viral and host cellular factors (Hu et al., 2019). Despite its lack of a role in the assembly of empty capsids, the CTD is absolutely required for pgRNA packaging. Sequential C-terminal truncation analyses indicate that deletion of the CTD beyond aa 164 impairs pgRNA encapsidation, whereas residues between aa 164 to 173 are required for primer translocation and positive strand DNA synthesis (Nassal, 1992). The function of the four arginine clusters in the CTD was analyzed by substitution of the individual arginine clusters with either alanine or lysine residues (Lewellyn and Loeb, 2011a). As expected, all the clusters were not required for capsid assembly. However, the alanine (but not lysine) substitution of cluster I to III, but not IV, significantly reduced the amount of encapsidated pgRNA. Moreover, it appears that all four clusters involved in viral DNA synthesis, which is, occasionally, not charge-dependent (Lewellyn and Loeb, 2011a). These results imply that by interaction with pgRNA and viral DNA, the arginine-rich CTD play an essential role in multiple steps of viral genome replication (Su et al., 2016).
HBV core proteins were discovered as phosphoproteins in the late 1980s by using metabolic labeling and phosphatase treatment/immunoblotting assays (Roossinck and Siddiqui, 1987; Yeh and Ou, 1991). Phosphomimic mutagenesis studies indicated that CTD phosphorylation, particularly at S162 and S170, was required for pgRNA packaging (Gazina et al., 2000; Lan et al., 1999; Lewellyn and Loeb, 2011b; Melegari et al., 2005). While several cellular kinases, such as SRPK1, SRPK2 (Daub et al., 2002; Zheng et al., 2005), PKA (Okabe et al., 2006), GAPDH-PK(Duclos-Vallee et al., 1998), PKC (Kann and Gerlich, 1994; Wittkop et al., 2010), CDK2 (Ludgate et al., 2012) and PLK1 (Diab et al., 2017), have been reported to phosphorylate one or multiple phospho-acceptor sites of CTD in vitro, their biological relevance in HBV replication in hepatocytes has not been established. However, our recent studies show that Cp is hyperphosphorylated in free dimers and empty capsids, but hypophosphorylated in pgRNA and DNA-containing nucleocapsids, suggesting that Cp dephosphorylation occurs during pgRNA encapsidation (Zhao et al., 2018). Indeed, further investigation demonstrated that cellular protein phosphatase 1 (PP1) catalyzes Cp dephosphorylation and is required for pgRNA packaging. Interestingly, both PP1 catalytic subunits α and β are co-packaged with Pol-pgRNA complex into nucleocapsids, but not empty capsids (Hu et al., 2020). We believe that the hyperphosphorylation of Cp dimers prevents its non-specific binding of cellular RNA and ensures the specific packaging of pgRNA through recognition of the Pol-pgRNA complex, most likely via Cp-Pol interaction. The sequential dephosphorylation of Cp during nucleocapsid assembly allows the CTD to interact with pgRNA, which may facilitate pgRNA packaging by organizing its folding to ensure its complete packaging. This hypothesis is supported by several independent studies showing that C-terminal truncation or mutation of the last two arginine-rich clusters resulted in failed encapsidation of the 3’ terminal region of pgRNA (Lewellyn and Loeb, 2011a). In addition, the involvement of the CTD in viral DNA synthesis, including template switching and primer translocation as well as DNA chain elongation, suggests that the CTD may be involved in interactions with viral and cellular factors in the viral replication environment (Lewellyn and Loeb, 2011a; Nassal, 1992). One possibility is that the CTD may also work as a nucleic acid chaperone to catalyze the breaking and reforming of base pair interactions. This hypothesis is supported by the fact that the CTD can catalyze nucleic acid strand exchange and hammerhead ribozyme activity in vitro (Chu et al., 2014). The other host cellular factors required for or have a role in regulating nucleocapsid assembly had been extensively reviewed elsewhere (Hu et al., 2019).
2.5. The morphogenesis of complete virions and empty virions depends on distinct structure features of capsids
As mentioned above, both the partially double stranded DNA-containing nucleocapsids and empty capsids can be efficiently enveloped and secreted as complete and empty virion particles, respectively (Blondot et al., 2016; Hu and Liu, 2017). Mutagenesis analyses identified amino acid residues of Cp on the shell of nucleocapsids that are essential for the envelopment of DNA-containing nucleocapsids (Koschel et al., 2000; Orabi et al., 2015; Pairan and Bruss, 2009; Ponsel and Bruss, 2003). While the tip of the spike of capsid appeared not to be directly involved in the morphogenesis of virions, alanine substitution of amino acid residues L60, L95, K96 or I126 abolished virion secretion (Ponsel and Bruss, 2003). These amino acid residues are located on the outer surface of capsid in the depressions between the spikes and possibly bind to envelope proteins and initiate budding of virion particles. Interestingly, F97L mutant Cp causes premature secretion of nucleocapsids containing less mature forms of viral DNA (Yuan et al., 1999), which is most likely due to the amino acid substitution leading to the increased size of an adjacent hydrophobic pocket in the center of the spike (Ceres et al., 2004). In marked contrast, Jianming Hu and colleagues recently demonstrated that the Cp residues essential for the production of complete virions as well as the entire CTD are not required for the envelopment of empty capsids (Ning et al., 2018). However, substitution or deletion mutations of the Cp linker region examined thus far completely blocks the production of empty virions, suggesting a critical role of the linker region in empty virion morphogenesis (Liu et al., 2018).
2.6. Nucleocapsid disassembly and its regulation
In addition to supporting viral DNA replication, another important function of HBV nucleocapsids is to serves as a vehicle to transmit viral genome to susceptible cells of the same or other individuals. On one hand, this latter function requires that the nucleocapsids be stable enough to protect the viral genome from degradation by environmental and host factors, while on the other hand, it also requires that the nucleocapsids disassemble in the proper subcellular compartment of infected cells to release the viral genome for replication. This is particularly a challenge for HBV and related viruses, because progeny mature nucleocapsids can recycle the newly replicated viral DNA back to the nucleus for amplification of the cccDNA pool (Guo and Guo, 2015). Because the uncontrolled cccDNA amplification loop results in hepatocyte death (Summers et al., 1990), the disassembly of mature progeny nucleocapsids must be strictly controlled by viral and host cellular factors (Cui et al., 2015a; Sheraz et al., 2019; Tang et al., 2019). In addition, timely- and spatially-controlled disassembly of nucleocapsids from infecting virions during de novo infection of hepatocytes is also critical for the successful establishment of cccDNA (Guo et al., 2017). Thus far, it is not known where the incoming virion-derived nucleocapsid disassembly occurs. However, it had been identified that several factors regulate the disassembly of intracellular progeny mature nucleocapsids and subsequently, cccDNA synthesis (Chen et al., 2016; Gallucci and Kann, 2017; Guo et al., 2010; Osseman et al., 2017).
The nucleocapsid disassembly is a multistep event that can be revealed as distinct species with different electrophoresis mobility in a native agarose gel electrophoresis (Cui et al., 2013). The physical force driving the nucleocapsid disassembly is most likely the elongation of double-stranded DNA, which is structurally rigid and pushes the nucleocapsid from inside (Dhason et al., 2012; Guo et al., 2007). In addition, interaction with NPC components, such as Nup153, is also important for the uncoating of viral rcDNA (Schmitz et al., 2010). Interestingly, Cp sequence motifs regulate nucleocapsid disassembly are identical or overlapped with that regulating virion particle envelopment (Cui et al., 2015b). This finding implies that the nucleocapsid disassembly (or genome uncoating) and envelopment are mechanistically two mutually excluding events. In addition, several mutagenesis studies indicated that phosphorylation of Cp at the CTD and assembly domain also modulates capsid stability and possibly, nucleocapsid disassembly (Barrasa et al., 2001; Selzer et al., 2015). Particularly, a phosphomimetic mutagenesis analysis of two putative phosphorylation sites in Cp assembly domain, S44 and S49, demonstrated that phosphorylation of the two residues suppressed pgRNA packaging, promoted nucleocapsid disassembly and subsequent cccDNA synthesis. On the contrary, dephosphorylation of the two residues did not affects pgRNA packaging and reverse transcriptional DNA synthesis, but reduced cccDNA formation (Luo et al., 2020). Because residues S44 and S49 locate at the internal surface of capsid, it is possible that phosphorylation of these two residues by encapsidated cellular kinases, such as CDK2, triggers mature nucleocapsid disassembly and cccDNA synthesis. In support of this notion, it was shown that treatment of cells with CDK2 inhibitors reduced cccDNA synthesis (Luo et al., 2020).
3. Core protein: Possible roles in cccDNA function and virus-host cell interaction
In addition to the capsid-related functions, Cp may also regulate cccDNA function as a component of cccDNA minichromosomes and mediate virus-host interaction by interacting with a variety of host cellular proteins. For instance, it was reported that Cp associates with cccDNA and regulates cccDNA transcription (Bock et al., 2001; Chong et al., 2017; Guo et al., 2011). Moreover, it was also demonstrated that Cp recruits IFN-induced cellular proteins APOBEC 3A or 3B to cccDNA minichromosomes for induction of cccDNA cytidine deamination and its subsequent degradation (Lucifora et al., 2014; Xia et al., 2014; Xia et al., 2016). Interestingly, although infection of primary duck hepatocytes with wild-type duck hepatitis B virus (DHBV) or a mutant DHBV deficient in expression of Cp results in the similar amounts of de novo cccDNA synthesis, the cccDNA in mutant DHBV infected cells transcribed significantly less amounts of viral RNA, suggesting an important role of Cp in DHBV cccDNA transcription regulation and/or RNA stability (Schultz et al., 1999). In marked contrast, HepG2-NTCP cells infected with wild-type and Cp-deficient HBV produced similar amounts of cccDNA, viral RNA and HBsAg during a three-week observation period (Qi et al., 2016). These contradictive findings on the role of Cp in DHBV and HBV cccDNA transcription regulation may reflect the distinct structure and unique biological function of the avian and mammalian hepadnaviral Cp in viral replication. Particularly, DHBV Cp is approximately 80 amino acid residues longer than HBV Cp. Because DHBV lacks X protein, it had been speculated that DHBV Cp may have a HBx-like function, which is, at least in part, involved in regulation of cccDNA transcription (Decorsiere et al., 2016; Lucifora et al., 2011).
Interestingly, it was reported recently that the phosphorylation status of residues S44 and S49 at Cp assembly domain affects the production and conversion of minus-strand covalently closed rcDNA, an intermediate of cccDNA synthesis (Luo et al., 2017; Tang et al., 2019; Zhao and Guo, 2020), to cccDNA, strongly suggesting a possible role of Cp in cccDNA synthesis in the nucleus (Luo et al., 2020). In addition, a recent proteomic analysis identified 60 Cp-interacting proteins in the nuclei of human hepatocytes. The majority of Cp-interacting host factors are RNA binding proteins that are involved in various aspects of mRNA metabolism. Particularly, SRSF10, a RNA binding protein that regulates RNA splicing, appears to modulate HBV RNA abundance, suggesting that Cp may regulate viral and cellular RNA metabolism via interaction with SRSF10 and other host cellular factors (Helene et al., 2020).
Because CpAM treatment usually induces the nuclear or cytoplasmic redistribution of Cp (Corcuera et al., 2018; Huber et al., 2018), it can be anticipated that CpAMs may modulate Cp interaction with its cytoplasmic and/or nuclear partners and consequentially modify HBV-hepatocyte interaction. Concerning the effects of CpAMs on the nuclear function of Cp, only one published study showed that prolonged treatment of HBV-infected HepaRG cells with JNJ-827 and JNJ-890 can reduce HBV RNA and HBsAg production, presumably by inhibition of cccDNA transcription or promotion of viral RNA decay (Lahlali et al., 2018). The effect of CpAM on Cp recruitment of APOBEC 3A and 3B for cccDNA decay as well as viral and cellular RNA metabolism remain to be investigated.
4. Core protein allosteric modulators: Antiviral activities and mode of action
4.1. Multiple chemotypes of CpAMs and their antiviral activity
Since the late 1990s, several chemotypes of compounds had been discovered to inhibit HBV replication via reduction of the amounts of capsids, such as heteroaryldihydropyrimidine (HAP) derivatives (Deres et al., 2003), or pgRNA-containing nucleocapsids, such as phenylpropenamides (PPA) and sulfamoylbenzamides (SBA) derivatives (Campagna et al., 2013b; King et al., 1998a). Mechanistic studies in the in vitro capsid assembly system indicated that those compounds accelerated, but not inhibited, Cp assembly into aberrant or normal capsids and were thus designated as core protein allosteric modulators (CpAMs), but not capsid assembly inhibitors (Zlotnick et al., 2015). As illustrated in Fig. 4, many structural distinct CpAMs had been discovered (Nijampatnam and Liotta, 2019). Based on the structures and metabolism of the induced Cp assembly products in vitro and in hepatocytes, CpAMs can be categorized into two distinct phenotypes. While type I CpAMs, including HAPs and possibly also Ciclopirox (Kang et al., 2019), induce the assembly of aberrant capsids or non-capsid Cp polymers that form aggregates and are subsequently degraded in hepatocytes (Bourne et al., 2006; Kang et al., 2019; Stray et al., 2005a; Stray and Zlotnick, 2006; Wu et al., 2013), type II CpAMs, such as PPA, SBA, etc., induce the assembly of morphologically “normal” capsids devoid of pgRNA and Pol (Amblard et al., 2020; Campagna et al., 2013a; Corcuera et al., 2018; King et al., 1998b; Wang et al., 2015; Yang et al., 2016). As will be discussed in detail below, all the CpAMs examined so far disrupt Cp assembly by binding to a hydrophobic pocket between the interface of Cp dimers to enhance dimer-dimer interaction. In agreement with the pleiotropic function of Cp in the HBV life cycle, CpAMs have also been demonstrated to disrupt multiple steps of HBV replication.
Fig. 4.
Chemical Structures of representative CpAMs.
4.2. Biophysics of CpAM disruption of capsid assembly
Altering the conformational state of Cp by changing the thermodynamics (pair-wise contact energies) or kinetics (rate of nucleation) can result in the misassembly of capsids and potentially disrupt the pregenome encapsidation process (Zlotnick and Mukhopadhyay, 2011). HAP compounds enhance the rate and extent of Cp assembly and act as allosteric effectors to induce an assembly-active state. While HAP1 misdirects assembly and leads to aberrant particles (Bourne et al., 2006; Stray et al., 2005b), HAP 12 leads to irregular non-capsid like polymers by increasing the association energy (~2 kcal/mol) and the rate of assembly (Venkatakrishnan et al., 2016). Type II CpAMs, such as PPAs, act as assembly accelerators by exclusively changing the reaction kinetics (Katen et al., 2010). Unlike HAPs, they do not misdirect assembly, but just speed it up and result in the production of ‘empty’ capsids due to the imbalance created between the rate of assembly of capsids and packing of Pol-pgRNA complex.
CpAMs differentially strengthen the association between Cp dimers and accelerate the kinetics of assembly. Thermodynamics and kinetics play a crucial role in binding energy and mode of CpAM action as well as the acquisition of CpAM resistance (Tan et al., 2013). While the thermodynamics and kinetics of in vitro Cp assembly have a strong and predictable effects on the Cp assembly product morphology, only the kinetics of in vitro assembly had a strong correlation with the activity of CpAM inhibition of HBV replication in human hepatocytes (Bourne et al., 2008). In addition, study of the wild-type and the V124 mutant Cp assembly kinetics revealed a nice correlation between the area of hydrophobic surface at the dimer-dimer interface and ΔGcontact (Tan et al., 2015). The resistance of T109 mutant Cp to GLS4, a HAP derivative, is also correlated with ΔGcontact. The most stable inter-dimer interface has the highest resistance to the CpAMs (Ruan et al., 2018).
4.3. Structure biology of CpAM-misdirected Cp assembly
The structures of Cp149 capsids incubated with CpAMs have been investigated using X-ray crystallography (pdb code 2G34, 4G93 and 5D7Y) (Table 1). All the assembly modulating compounds examined thus far bind to the “HAP” pocket, which is at the inter-dimer interface and consists of a concave interface made up of residues from one chain - F23, P25 of loop 2, D29, L30, T33, L37 from helix2, W102, I105, S106, T109, F110 of helix4, Y118 and F122 of helix5, I139, L140, S141 of loop6) and cap from the other chain – V124, W125, R127 from helix5 and T128, P129, Y132, R133, P134 of loop6 as shown in Fig. 2C.
CpAMs do not have a uniform allosteric effect on capsid structure. A small difference in structure leads to a large (mostly unpredictable) difference in the T=4 capsid after binding of the CpAM. HAP 1 (pdb code 2G34) causes the five-fold vertices to protrude, three-fold vertices to open, and quasi-six-fold vertices to flatten. It changes the quaternary but not the tertiary structure (Bourne et al., 2006). AT-130 (pdb code 4G93), a PPA derivative, causes quaternary changes like HAPs but also induces compensatory tertiary structural changes that lead to the formation of normal capsids (Katen et al., 2013). The quaternary changes might result in assembly misdirection. Binding of Cp by HAP1 and AT-130 expanded the capsid diameter by up to 10Å and caused the capsids to crystallize with different unit cell parameters from apo capsids. HAP18 (pdb code 5D7Y) caused only minor changes in quaternary structure and decreased the capsid diameter by 3Å (Venkatakrishnan et al., 2016).
The Y132 residue near the C-terminal end of the assembly domain plays an important role in interdimer interaction, which facilitates capsid assembly. The Y132A substitution causes the protein to become assembly deficient (Bourne et al., 2009; Packianathan et al., 2010). Four structures with CpAMs in the assembly inactive Y132 Cp hexamer are available till now (pdb codes 5E0I (Klumpp et al., 2015), 5GMZ (Qiu et al., 2016), 5WRE and 5T2P (Zhou et al., 2017)). These structures are at higher resolution compared to the X-ray structures of whole capsid and hence we can visualize water molecules and see some water-mediated hydrogen bonding of the ligand to the dimer interface. Fig. 5 shows the ligand-interaction diagram of the compounds HAP18, Ciclopirox, AT-130 and SBA_R01, respectively, with the core protein observed in the X-ray crystal structures. We can see that although the chemical structures are varied, there are some interactions such as hydrogen bond with W102 and aromatic ring (often with halogens on them) on the hydrophobic pocket formed by L30, L37, I105 and V124 has been observed in most instances.
Fig. 5. Illustration of the interaction of representative CpAMs with Cp at the dimer-dimer interface of the capsid and Y132A Cp hexamer X-ray crystal structures.
The ligand interaction diagram with negatively charged residues in red, positively charged residues in blue, hydrophobic residues in green and polar residues in blue. The purple line shows hydrogen bonding and orange line represents halogen bond interaction with the compounds. The green line denotes a stacking interaction. See text for detailed description.
4.4. Effects of CpAMs on assembled capsids and nucleocapsids
The dual effects of capsid-targeting antiviral agents on capsid assembly and disassembly have been reported in dengue virus and human immunodeficiency virus (Klumpp and Crepin, 2014; Scaturro et al., 2014). It is thus very interesting to note that CpAMs also disrupt the integrity of assembled HBV capsids. In fact, it was reported a decade ago that treatment with higher concentrations of HAPs can induce the disassembly of in vitro assembled capsids (Stray and Zlotnick, 2006). Using an engineered addressable thiol (mutating L30, I105, S106 and V124 to cysteine) in the HAP pocket, it was shown recently that ligands bound to the pocket triggers capsid disassembly in a dose-dependent manner (Qazi et al., 2018). Using a fluorophore [tetramethylrhodamine (TAMRA)]-labeled HAP, it was further demonstrated that binding of HAP compound induces capsid-wide asymmetric deformations (outbreaks, flat regions and sharp angles) (Schlicksup et al., 2018). These in vitro studies imply that the HAP pocket may transiently adopt destabilizing conformations and play a key role in both capsid assembly and disassembly (Hadden et al., 2018). However, CpAM treatment apparently does not induce the disassembly of capsids purified from the cytoplasm of hepatocytes supporting HBV replication. Instead, some HAP derivatives, such as HAP18 and GLS4, can induce the electrophoresis mobility shift of assembled empty capsids and DNA-containing nucleocapsids in native agarose gel electrophoresis (Ko et al., 2019), indicating the induction of global structural alteration of capsids (Schlicksup et al., 2018; Wu et al., 2018). Interestingly, we found recently that CpAMs can preferentially induce the disassembly of mature nucleocapsids and differentially modulate cccDNA synthesis from de novo and intracellular amplification pathways (Guo et al., 2017). Apparently, the mode of CpAM action on capsid structure alteration, mature nucleocapsid disassembly and de novo cccDNA synthesis remains to be further investigated.
Pharmacologically, inhibition of de novo cccDNA synthesis by CpAMs can only be achieved at the concentrations approximately 10 to 34 fold higher than that required to inhibit pgRNA assembly and viral DNA replication in primary human hepatocytes (Berke et al., 2020; Ko et al., 2019). While the finding is consistent with the observation that that disruption of capsid integrity requires the coordinative binding of CpAMs at many HAP pockets of a capsid (Stray and Zlotnick, 2006), it also argues whether the effective CpAM concentrations can be achieved in vivo in humans to inhibit de novo cccDNA synthesis. However, potent suppression of HBV pgRNA encapsidation will result in significant reduction of infectious virion production and consequentially reduce de novo infection and new cccDNA synthesis. Interestingly, a recent study showed that NUC therapy not only efficiently reduced HBV load, the residual virions produced by NUC-treated hepatocytes are not infectious, due to the incorporation of NUC, that irreversibly terminates viral DNA chain elongation (Liu et al., 2020). Therefore, it is conceivable that combination therapy of CpAM and NUC can achieve a more profound inhibition of HBV replication by targeting both pgRNA encapsidation and viral DNA synthesis, which should further reduce the production of infectious virion and thus prevent de novo cccDNA synthesis. In agreement with this notion, a recently reported clinical trial demonstrated that in the majority of patients, 6 months of combination therapy using entecavir and ABI-H0731 abolished residual HBV DNA replication and significantly reduced the level of HBsAg, indicative of a significant reduction of cccDNA levels in the liver (Mark S. Sulkowski, 2019).
5. CpAMs also disrupt the metabolism of pre-core protein
Pre-core ORF encodes a 25 kDa polypeptide (p25) that contains 29 amino acid residues extension at the N-terminus of core protein. As illustrated in Fig. 2B, p25 is co-translationally targeted by the signal peptide located in the N-terminal leader sequence to the endoplasmic reticulum (ER), where the first 19 residues of the leader sequence are cleaved by cellular signalase, producing a 22-kDa protein (p22). The p22 form is further processed by a furin-like protease in a post-ER compartment to remove C-terminal 29 residues to yield a 17 kD protein (p17), which is subsequently secreted out of the hepatocyte as HBeAg (Ito et al., 2009). Therefore, except for the 10 remaining leader residues at its N-terminal, HBeAg (p17) shares the entire assembly domain and linker region with Cp. Due to the formation of an intramolecular disulfide bond between the leader peptide C(−7) and C61 in the middle of assembly domain, HBeAg dimer has an entirely different quaternary structure in comparison with Cp dimer and is incompetent to assemble capsid (DiMattia et al., 2013). However, in a reducing condition, disruption of the intramolecular disulfide bond of p22 or HBeAg confers their ability to assemble capsids with reduced stability, due to the existence of the N-terminal leader peptide. Interestingly, a small fraction of p22 can be retro-transported from the ER lumen into the cytoplasm, where p22 may form heterodimers with Cp and assemble chimeric capsids and nucleocapsids. The observed inhibition of HBV replication by pre-core gene products in hepatocytes in vitro and in transgenic mice in vivo suggests an important regulatory role of p22 and/or HBeAg in HBV replication via assembly of unstable chimeric nucleocapsids with Cp (Guidotti et al., 1996). It had also reported that p22 attenuates interferon response by inhibiting STAT1 nuclear translocation (Mitra et al., 2019).
In addition to the intracellular effects of p22 on HBV replication and innate antiviral immune response, the secreted pre-core gene product HBeAg had been speculated to suppress host antiviral immune response and favor the establishment of persistent HBV infection (Milich, 2019). Pre-core gene is preserved in all the hepadnaviruses. While studies with duck hepatitis B virus (DHBV) and woodchuck hepatitis virus (WHV) showed that the deficiency of pre-core gene expression does not significantly impair the replication fitness of the viruses in ducks (Zhang and Summers, 1999) and woodchucks (Chen et al., 1992), infection of neonate woodchucks with WHV deficient in pre-core protein expression results in persistent infection, but wild-type WHV infection is transient (Chen et al., 1992). A recent study in mouse models revealed that the maternal-derived HBeAg potentiates persistent HBV infection by altering the function of intrahepatic macrophages in the offspring (Tian et al., 2016). A clinical study indicates that HBeAg induces the expansion of monocytic myeloid-derived suppressor cells to inhibit T cell function (Yang et al., 2019a).
Recently, a study showed that treatment of HBV replicating hepatocytes with HAP_R10, but not AT-130, efficiently inhibited HBeAg secretion. Mechanistic studies support a notion that HAP_10 induces p22 assembly into non-capsid aggregates, which is subsequently degraded (Yan et al., 2019). Interestingly, HAP_10 failed to inhibit HBeAg secretion from hepatocytes supporting the replication of a HBV variant encoding HAP-resistant core protein, suggesting that HAP_R01-mediated HBeAg and core protein reductions are mediated via the same mechanism. Considering the function of p22 in suppression of IFN response in hepatocytes as well as the potential role of HBeAg in regulation of host antiviral immune response, it is reasonable to speculate that HAP derivatives may be able to restore the host antiviral immune response via reducing intracellular and secreted pre-core gene products. It will be interesting to investigate whether the CpAMs that can inhibit HBeAg secretion, such as HAP_10 and JNJ-827 (Lahlali et al., 2018; Yan et al., 2019), are more efficacious in achieving the functional cure of chronic HBV infection than the CpAMs that cannot inhibit HBeAg secretion in clinical trials.
6. CpAMs in clinical development
A number of structurally diverse CpAMs, representing both type I and type II CpAMs, are in clinical trials and constitute one of the most actively investigated mechanisms for chronic hepatitis B (CHB) cure. NVR 3–778 (type II) was the first CpAM to demonstrate clinical proof of concept (POC), showing HBV DNA and RNA reductions in patients when administered as a monotherapy for 28 days and in combination with PegIFNα−2a (Yuen et al., 2019b). Since then, newer generations of CpAMs have demonstrated improved efficacy in POC studies, as monotherapy and in combinations with standard of care (SOC) drugs and investigational agents (Table 2).
Table 2.
Most advanced CpAMs in clinical trials for the treatment of CHB patients
| CpAM (MoA Type) | Chemical Class | CpAM Dose and dosing frequency | Reduction from BL (mean or median log10 IU/mL or copies/mL) in Serum | Clinical Status | Sponsor | Ongoing or Completed Clinical Trial^ | |||||
|---|---|---|---|---|---|---|---|---|---|---|---|
| Clinical Phase | HBV DNA | HBV RNA | HBeAg# | HBsAg | HBc rAg | ||||||
| NVR 3-778 (Type II) | SBA | 100 mg QD 200 mg QD 400 mg QD 600 mg BID 1000 mg BID 600 mg BID + PegIFNα2a (SC)/PegIFNα2a alone |
Ph 1b (28d/EOT) |
0.19 0.20 0.49 1.43 1.35 1.97/1.06 |
nd nd nd 1.42 1.27 2.10/0.89 |
0.02 0.04 0.04 0.09 0.07 0.21/0.29 |
0.03 0.03 0.03 0.02 1.00 0.08/0.02 |
nd nd nd 0.01 0.13 0.10/0.28 |
D/C | Novira/ JNJ |
NCT02112799; NCT02401737; |
| JNJ-6379 (Type II) | SPA | 25 mg QD* 75 mg QD 150 mg QD 250 mg QD |
Ph 1b (28d/EOT) |
2.16 2.89 2.70 2.70 |
2.30 1.85 1.83 1.43 |
NS NS NS NS |
NS NS NS NS |
≥0.5 in a subset of patients mainly with ALT>1x and <2.5x ULN | Ph 2 | Janssen |
NCT02662712, NCT04208399, NCT03864601, NCT03945539, NCT03111511, NCT02933580, NCT03982186, NCT04288310, NCT04129554, NCT03361956 |
| JNJ-0440 (Type II) | SBA | 750 mg QD 750 mg BID |
Ph 1 (28d/EOT) |
3.2 3.3 |
2.65 2.94 |
0.2 0.2 |
NS NS |
0.85 0.62 |
Ph 1 | NCT03439488 | |
| ABI-H0731 (Type II) | DBT | 100 mg QD 200 mg QD 300 mg QD 400 mg QD |
Ph 1b (28d/EOT) |
1.3/2.2** 1.9/2.4** 2.9/2.5**$ NR/3.9** |
1.2/NR** 1.7/NR** 2.3/NR**$ NA/NR** |
NS NS NS NS |
NS NS NS NS |
NS NS NS NS |
Ph 2 | Assembly |
NCT02908191, NCT03780543, NCT03577171, NCT03576066, NCT03109730 |
| 300 mg QD + ETV/ETV alone (HBeAg+ Rx naïve) 300 mg QD + NRTI/NRTI alone (HBeAg+/- NUC suppressed) |
Ph 2a Study 202 (24 week) Phase 2a Study 201 (24 week) | 5.30/4.19 TND in 22 of 27 pts/ 0 out of 12 pts |
2.34/0.61& 1.74/0.09 |
- | - | - | |||||
| ABI-H2158 (Type II) | 100 mg QD | Ph 1b (14d/EOT) |
2.3 | 2.1& | NR | NR | NR | Ph 1 |
NCT03714152, NCT04083716, NCT04142762 |
||
| AB-506 (Type II) | AI | 160 mg QD 400 mg QD |
Ph 1b (28d/EOT) |
2.10 2.80 |
2.37 2.40 |
NR NR |
0.021 0.113 |
NR NR |
D/C | Arbutus | |
| GLS4JHS (Type I) | HAP | 120 mg QD (+ RTV) | Ph 1b (28d/EOT) |
1.42 | NR | 0.25 | 0.06 | NR | Ph 2 | HEC Pharma |
NCT04147208, NCT03638076, NCT03662568 |
| 240 mg QD (+ RTV) | 2.13 | NR | 0.30 | 0.14 | NR | ||||||
| 120 mg BID (+ RTV) | Ph 2 (48-week study/TW24 data | 3.28¥ | ~3.0¥ | 0.57¥ | 0.20¥ | ~1.9¥ | |||||
| 120 mg TID (+RTV) | reported) | 4.40¥ | ~3.0¥ | 1.06¥ | 0.44¥ | ~1.9¥ | |||||
| RO9389 (Type I) | HAP | 200 mg BID 400 mg BID 200 mg QD 600 mg QD 1000 mg QD |
Ph 1b (28d/EOT) | 2.66@ 3.20@ 3.00@ 2.92@ 3.17@ |
2.09@ 2.55@ 2.55@ 2.54@ 2.43@ |
NS NS NS NS NS |
NS NS NS NS NS |
NR NR NR NR NR |
Ph 2 | Roche |
NCT03570658, NCT03717064, NCT02952924, NCT04225715 |
Notes: D/C = discontinued; SC = subcutaneous; QD = once daily; BID = twice daily; NR = not reported;
100 mg loading dose followed by 25 mg QD dosing; BL = baseline; EOT = end of treatment; NS = no significant changes; SBA = sulfamoylbenzamide; SPA = Sulphamoylpyrrolamide; AI = aminoindane; DBT -dibenzothiazepinecarboxamide; HAP = heteroaryldihydropyrimidine;
HBeAg positive/HBeAg Negative CHB patients; ETV = entecavir;
= pgRNA; TND = DNA target not detected in a semiquantitative PCR assay; JNJ-6379 = JNJ-56136379; JNJ-0440 = JNJ-64530440; RO9389 = RO7049389; RTV = ritonavir;
median decline log IU/mL; NRTI = approved nucleoside reverse transcriptase inhibitor therapy;
only trials registered at US clinicaltrials.gov shown; Ph = phase;
= HBV DNA and RNA data of one patient with Cp T109M variant in the 300 mg QD cohort was excluded from analyses.
= applies to HBeAg-positive patients; TW = treatment week.
= mean max log10 IU/ml reduction from BL.
In a 28-day POC study in treatment-naïve CHB patients, JNJ-6379 (type II) administration showed reductions in HBV DNA and RNA from baseline (BL) without significant reductions in HBeAg or HBsAg levels, although a subset of patients showed serum HBV core-related antigen (HBcrAg) reduction presumably due to direct inhibition of core protein release (Lenz et al., 2019; Zoulim et al., 2018). The antiviral activity plateaued at ~3 log reduction in HBV DNA, at doses exceeding 75 mg QD (Zoulim et al., 2018). Interestingly, JNJ-6379 showed a long plasma half-life of 120–148 h (Yogaratnam et al., 2019). JNJ-6379 is being evaluated in a triple combination study with an RNAi inhibitor and a NUC where interim results have shown robust reductions in HBsAg, HBV DNA and RNA after 3 months on therapy (Yuen et al., 2019c). JNJ-0440, a SBA derivative (Table 2), has also demonstrated clinical POC in a 28-day study, with 750 mg QD and BID dosing showing equivalent HBV DNA reductions in plasma and modest effects on HBcrAg levels (Gane et al., 2019a). In contrast to JNJ-6379, JNJ-0440 has a shorter plasma half-life of 9–16 h. The BID dosing of JNJ-0440 resulted in higher plasma trough concentrations (Ctau:plasma-adjusted EC90 ratio) in comparison to QD dosing, presumably increasing the likelihood of engaging the secondary mechanism of CpAMs (Gane et al., 2019a).
ABI-H0731 (type II CpAM) has also demonstrated antiviral activity in CHB patients in a 28-day POC study (Yuen et al., 2018; Yuen et al., 2020). One patient in the 300 mg cohort who exhibited a reduced level of HBV DNA decline (1 log vs mean 2.9 log10 IU/mL for the cohort), showed a 70% to 100% enrichment of Cp T109M BL variant over the course of treatment. ABI-H0731+NUC combination is currently being evaluated in NUC-suppressed HBeAg positive and HBeAg negative patients, as well as NUC-naïve HBeAg positive subjects (Sulkowski et al., 2019). In the NUC-naïve group, ABI-H0731 and entecavir combination demonstrated faster and greater declines in serum HBV DNA and pgRNA from baseline versus entecavir alone, following 24 weeks of administration. These patients enrolled for an additional 52 weeks of combination therapy showed continuous decrease in serum HBV DNA to week 60. Similarly, in NUC-suppressed HBeAg positive and HBeAg negative patients, ABI-0731 and NUC combination showed a greater reduction in HBV pgRNA levels from baseline in comparison to entecavir alone, with 22/27 patients exhibiting undetectable HBV DNA vs 0/12 patients in the NUC alone arm using an inhouse higher sensitivity assay. Interestingly, HBV pgRNA declines also correlated with reductions in HBeAg, HBcrAg and HBsAg levels in patients that had undergone treatment for 16–60 weeks with the combination. A second-generation CpAM, ABI-H2158, is also undergoing POC study in a 14-day monotherapy trial, with a 100 mg QD dose demonstrating reductions in serum HBV DNA and pgRNA in HBeAg positive patients. Higher doses of ABI-H2158 are under evaluation (Yuen et al., 2019a).
AB-506 (type II) has also demonstrated POC with a 400 mg QD dose resulting in mean HBV DNA and RNA reductions of 2.8 and 2.4 log10 IU/mL by day 29 (Sims et al., 2019). One subject (160 mg dose) showed non-response, attributed to the presence of Cp I105T mutation at baseline, which is known to reduce in vitro susceptibility to AB-506 by ~19 fold.
Two type I CpAMs, GLS4 and RO9389, are currently in Phase 2 clinical trials. Phase 1 studies showed that administration of GLS4 did not result in sufficient plasma trough levels, which was attributed to metabolic instability. A co-dosing strategy with the known CYP450 3A4 inhibitor ritonavir (RTV) was used to boost GLS4 exposures (Ding et al., 2017; Zhang et al., 2019; Zhao et al., 2019). RO9389 treatment has also demonstrated robust declines in serum HBV DNA (Gane et al., 2019b).
ALT elevations have been reported in some patients with CpAMs (Gane et al., 2019b; Lenz et al., 2019; Sims et al., 2019; Sulkowski et al., 2019; Vandenbossche et al., 2019). In the case of AB-506 a 28-day healthy volunteer study directly attributed ALT flares to AB-506, leading to its discontinuation. Whether ALT elevations observed with other CpAMs are related to an immunological reawakening of an anti-HBV response or liver toxicity require further studies.
It is clear that both types I and II CpAMs could potentially become part of future treatment regimens for CHB patients. While the above examples are limited to the most advanced CpAMs, a number of others in preclinical/Phase 1a studies are poised to shape a highly competitive landscape.
7. Perspectives of Cp-targeting antiviral therapy
CpAMs are an emerging new mechanistic class of anti-HBV agents for chronic hepatitis B patients. The clinical experience to date has shown that a combination of CpAM and NUC is a viable therapeutic regimen that leads to more pronounced viral suppression than a NUC alone and confirms the hypothesis that NUCs are inherently “leaky” in inhibiting viral replication. However, important questions still remain such as: What mechanistic combination(s) will ultimately increase the rate of functional cure? Is there a role of the secondary mechanism of CpAMs, such as inhibition of cccDNA synthesis and/or transcription and suppression of HBeAg secretion, on antiviral efficacy or rates of functional cure? Do type I vs type II CpAMs have differential effects on efficacy endpoints and/or safety? What is the implication of baseline Cp variants on clinical efficacy? What is the significance of ALT elevations observed with some CpAMs; are they good (i.e. due to immune reactivation) or bad (toxicity) flares? As more CpAM combinations with other investigational agents enter clinical testing, the field is set for an exciting future that will hopefully address these important questions.
Highlights.
Core protein is a building block of nucleocapsids and involved in multiple steps of the HBV replication.
CpAMs bind to a hydrophobic pocket between core protein dimers to misdirect nucleocapsid assembly and disassembly.
CpAMs not only block viral DNA replication, but also inhibit de novo cccDNA formation.
Targeting the multi-functional core protein is an effective antiviral approach against HBV infection.
Acknowledgments
Funding information
HHS | NIH | National Institute of Allergy and Infectious Diseases (NIAID) provided funding to Ju-Tao Guo under grant number R01 AI113267.
This work was partially supported by the Office of the Assistant Secretary of Defense for Health Affairs, through the Peer Reviewed Medical Research Program under Award No. W81XWH-17-1-0600. Opinions, interpretations, conclusions and recommendations are those of the author and are not necessarily endorsed by the Department of Defense.
Footnotes
Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
References
- Alexander CG, Jurgens MC, Shepherd DA, Freund SM, Ashcroft AE, Ferguson N, 2013. Thermodynamic origins of protein folding, allostery, and capsid formation in the human hepatitis B virus core protein. Proc Natl Acad Sci U S A 110, E2782–2791. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Alter H, Block T, Brown N, Brownstein A, Brosgart C, Chang KM, Chen PJ, Chisari FV, Cohen C, El-Serag H, Feld J, Gish R, Glenn J, Greten T, Guo H, Guo JT, Hoshida Y, Hu J, Kowdley KV, Li W, Liang J, Locarnini S, Lok AS, Mason W, McMahon B, Mehta A, Perrillo R, Revill P, Rice CM, Rinaudo J, Schinazi R, Seeger C, Shetty K, Tavis J, Zoulim F, 2018. A research agenda for curing chronic hepatitis B virus infection. Hepatology 67, 1127–1131. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Amblard F, Boucle S, Bassit L, Cox B, Sari O, Tao S, Chen Z, Ozturk T, Verma K, Russell O, Rat V, de Rocquigny H, Fiquet O, Boussand M, Di Santo J, Strick-Marchand H, Schinazi RF, 2020. Novel Hepatitis B Virus Capsid Assembly Modulator Induces Potent Antiviral Responses In Vitro and in Humanized Mice. Antimicrob Agents Chemother 64. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Barrasa MI, Guo JT, Saputelli J, Mason WS, Seeger C, 2001. Does a cdc2 kinase-like recognition motif on the core protein of hepadnaviruses regulate assembly and disintegration of capsids? J Virol 75, 2024–2028. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bartenschlager R, Schaller H, 1992. Hepadnaviral assembly is initiated by polymerase binding to the encapsidation signal in the viral RNA genome. EMBO J 11, 3413–3420. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Basagoudanavar SH, Perlman DH, Hu J, 2007. Regulation of hepadnavirus reverse transcription by dynamic nucleocapsid phosphorylation. J Virol 81, 1641–1649. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Berke JM, Dehertogh P, Vergauwen K, Mostmans W, Vandyck K, Raboisson P, Pauwels F, 2020. Antiviral Properties and Mechanism of Action Studies of the Hepatitis B Virus Capsid Assembly Modulator JNJ-56136379. Antimicrob Agents Chemother 64. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Birnbaum F, Nassal M, 1990. Hepatitis B virus nucleocapsid assembly: primary structure requirements in the core protein. J Virol 64, 3319–3330. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Block TM, Gish R, Guo H, Mehta A, Cuconati A, Thomas London W, Guo JT, 2013. Chronic hepatitis B: what should be the goal for new therapies? Antiviral Res 98, 27–34. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Blondot ML, Bruss V, Kann M, 2016. Intracellular transport and egress of hepatitis B virus. J Hepatol 64, S49–S59. [DOI] [PubMed] [Google Scholar]
- Bock CT, Schwinn S, Locarnini S, Fyfe J, Manns MP, Trautwein C, Zentgraf H, 2001. Structural organization of the hepatitis B virus minichromosome. J Mol Biol 307, 183–196. [DOI] [PubMed] [Google Scholar]
- Bottcher B, Nassal M, 2018. Structure of Mutant Hepatitis B Core Protein Capsids with Premature Secretion Phenotype. J Mol Biol 430, 4941–4954. [DOI] [PubMed] [Google Scholar]
- Bourne C, Lee S, Venkataiah B, Lee A, Korba B, Finn MG, Zlotnick A, 2008. Small-molecule effectors of hepatitis B virus capsid assembly give insight into virus life cycle. J Virol 82, 10262–10270. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bourne CR, Finn MG, Zlotnick A, 2006. Global structural changes in hepatitis B virus capsids induced by the assembly effector HAP1. J Virol 80, 11055–11061. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bourne CR, Katen SP, Fulz MR, Packianathan C, Zlotnick A, 2009. A mutant hepatitis B virus core protein mimics inhibitors of icosahedral capsid self-assembly. Biochemistry 48, 1736–1742. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Campagna MR, Liu F, Mao R, Mills C, Cai D, Guo F, Zhao X, Ye H, Cuconati A, Guo H, 2013a. Sulfamoylbenzamide derivatives inhibit the assembly of hepatitis B virus nucleocapsids. Journal of virology 87, 6931–6942. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Campagna MR, Liu F, Mao R, Mills C, Cai D, Guo F, Zhao X, Ye H, Cuconati A, Guo H, Chang J, Xu X, Block TM, Guo JT, 2013b. Sulfamoylbenzamide derivatives inhibit the assembly of hepatitis B virus nucleocapsids. J Virol 87, 6931–6942. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ceres P, Stray SJ, Zlotnick A, 2004. Hepatitis B virus capsid assembly is enhanced by naturally occurring mutation F97L. J Virol 78, 9538–9543. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ceres P, Zlotnick A, 2002. Weak protein− protein interacDons are sufficient to drive assembly of hepatitis B virus capsids. Biochemistry 41, 11525–11531. [DOI] [PubMed] [Google Scholar]
- Chang J, Guo F, Zhao X, Guo JT, 2014. Therapeutic strategies for a functional cure of chronic hepatitis B virus infection. Acta pharmaceutica Sinica. B 4, 248–257. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chen C, Wang JC, Pierson EE, Keifer DZ, Delaleau M, Gallucci L, Cazenave C, Kann M, Jarrold MF, Zlotnick A, 2016. Importin beta Can Bind Hepatitis B Virus Core Protein and Empty Core-Like Particles and Induce Structural Changes. PLoS Pathog 12, e1005802. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chen HS, Kew MC, Hornbuckle WE, Tennant BC, Cote PJ, Gerin JL, Purcell RH, Miller RH, 1992. The precore gene of the woodchuck hepatitis virus genome is not essential for viral replication in the natural host. J Virol 66, 5682–5684. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chong CK, Cheng CYS, Tsoi SYJ, Huang FY, Liu F, Seto WK, Lai CL, Yuen MF, Wong DK, 2017. Role of hepatitis B core protein in HBV transcription and recruitment of histone acetyltransferases to cccDNA minichromosome. Antiviral Res 144, 1–7. [DOI] [PubMed] [Google Scholar]
- Chu TH, Liou AT, Su PY, Wu HN, Shih C, 2014. Nucleic acid chaperone activity associated with the arginine-rich domain of human hepatitis B virus core protein. J Virol 88, 2530–2543. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chua PK, Tang FM, Huang JY, Suen CS, Shih C, 2010. Testing the balanced electrostatic interaction hypothesis of hepatitis B virus DNA synthesis by using an in vivo charge rebalance approach. J Virol 84, 2340–2351. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cole AG, 2016. Modulators of HBV capsid assembly as an approach to treating hepatitis B virus infection. Curr Opin Pharmacol 30, 131–137. [DOI] [PubMed] [Google Scholar]
- Corcuera A, Stolle K, Hillmer S, Seitz S, Lee JY, Bartenschlager R, Birkmann A, Urban A, 2018. Novel non-heteroarylpyrimidine (HAP) capsid assembly modifiers have a different mode of action from HAPs in vitro. Antiviral Res 158, 135–142. [DOI] [PubMed] [Google Scholar]
- Cui X, Guo JT, Hu J, 2015a. Hepatitis B Virus Covalently Closed Circular DNA Formation in Immortalized Mouse Hepatocytes Associated with Nucleocapsid Destabilization. J Virol 89, 9021–9028. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cui X, Luckenbaugh L, Bruss V, Hu J, 2015b. Alteration of Mature Nucleocapsid and Enhancement of Covalently Closed Circular DNA Formation by Hepatitis B Virus Core Mutants Defective in Complete-Virion Formation. J Virol 89, 10064–10072. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cui X, Ludgate L, Ning X, Hu J, 2013. Maturation-associated destabilization of hepatitis B virus nucleocapsid. J Virol 87, 11494–11503. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Daub H, Blencke S, Habenberger P, Kurtenbach A, Dennenmoser J, Wissing J, Ullrich A, Cotten M, 2002. Identification of SRPK1 and SRPK2 as the major cellular protein kinases phosphorylating hepatitis B virus core protein. J Virol 76, 8124–8137. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Decorsiere A, Mueller H, van Breugel PC, Abdul F, Gerossier L, Beran RK, Livingston CM, Niu C, Fletcher SP, Hantz O, Strubin M, 2016. Hepatitis B virus X protein identifies the Smc5/6 complex as a host restriction factor. Nature 531, 386–389. [DOI] [PubMed] [Google Scholar]
- Deres K, Schroder CH, Paessens A, Goldmann S, Hacker HJ, Weber O, Kramer T, Niewohner U, Pleiss U, Stoltefuss J, Graef E, Koletzki D, Masantschek RN, Reimann A, Jaeger R, Gross R, Beckermann B, Schlemmer KH, Haebich D, Rubsamen-Waigmann H, 2003. Inhibition of hepatitis B virus replication by drug-induced depletion of nucleocapsids. Science 299, 893–896. [DOI] [PubMed] [Google Scholar]
- Dhason MS, Wang JC, Hagan MF, Zlotnick A, 2012. Differential assembly of Hepatitis B Virus core protein on single- and double-stranded nucleic acid suggest the dsDNA-filled core is spring-loaded. Virology 430, 20–29. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Diab A, Foca A, Fusil F, Lahlali T, Jalaguier P, Amirache F, N’Guyen L, Isorce N, Cosset FL, Zoulim F, Andrisani O, Durantel D, 2017. Polo-like-kinase 1 is a proviral host factor for hepatitis B virus replication. Hepatology 66, 1750–1765. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dienstag JL, 2009. Benefits and risks of nucleoside analog therapy for hepatitis B. Hepatology 49, S112–121. [DOI] [PubMed] [Google Scholar]
- DiMattia MA, Watts NR, Stahl SJ, Grimes JM, Steven AC, Stuart DI, Wingfield PT, 2013. Antigenic switching of hepatitis B virus by alternative dimerization of the capsid protein. Structure 21, 133–142. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ding Y, Niu J, Zhang H, Chen H, Li C, Li X, Liu J, Wu M, Zhu X, Chen G, Mai J, Hu Y, 2017. Multiple Dose Study of GLS4JHS, Interfering With The Assembly Of HBV core particles, in Patients Infected with Hepatitis B Virus. Hepatology 66(1) SUPPL 491A. [Google Scholar]
- Duclos-Vallee JC, Capel F, Mabit H, Petit MA, 1998. Phosphorylation of the hepatitis B virus core protein by glyceraldehyde-3-phosphate dehydrogenase protein kinase activity. J Gen Virol 79 ( Pt 7), 1665–1670. [DOI] [PubMed] [Google Scholar]
- Gallucci L, Kann M, 2017. Nuclear Import of Hepatitis B Virus Capsids and Genome. Viruses 9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gane E, Schwabe C, Lenz O, Verbinnen T, Talloen W, Kakuda TN, Westland C, Patel M, Yogaratnam J, Dragone L, Remoortere PV, 2019a. JNJ-64530440 (JNJ-0440), A Novel, Class N Capsid Assembly Modulator (CAM-N): Safety, Tolerability, Pharmacokinetics and Antiviral Activity of Multiple Ascending Doses in Patients with Chronic Hepatitis B. Hepatology 70(1) SUPPL 61A. [Google Scholar]
- Gane E, Yuen MF, Bo Q, Schwabe C, Tanwandee T, Das S, Jin Y, Gao L, Zhou X, Wang Y, Feng S, Meinel D, Zhu M, 2019b. RO7049389, a core protein allosteric modulator, demonstrates robust decline in HBV DNA and HBV RNA in chronic HBV infected patients. Journal of Hepatology 70, e491. [Google Scholar]
- Gazina EV, Fielding JE, Lin B, Anderson DA, 2000. Core protein phosphorylation modulates pregenomic RNA encapsidation to different extents in human and duck hepatitis B viruses. J Virol 74, 4721–4728. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Guidotti LG, Matzke B, Pasquinelli C, Schoenberger JM, Rogler CE, Chisari FV, 1996. The hepatitis B virus (HBV) precore protein inhibits HBV replication in transgenic mice. J Virol 70, 7056–7061. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Guo F, Zhao Q, Sheraz M, Cheng J, Qi Y, Su Q, Cuconati A, Wei L, Du Y, Li W, Chang J, Guo JT, 2017. HBV core protein allosteric modulators differentially alter cccDNA biosynthesis from de novo infection and intracellular amplification pathways. PLoS Pathog 13, e1006658. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Guo H, Jiang D, Zhou T, Cuconati A, Block TM, Guo JT, 2007. Characterization of the intracellular deproteinized relaxed circular DNA of hepatitis B virus: an intermediate of covalently closed circular DNA formation. J Virol 81, 12472–12484. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Guo H, Mao R, Block TM, Guo JT, 2010. Production and function of the cytoplasmic deproteinized relaxed circular DNA of hepadnaviruses. J Virol 84, 387–396. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Guo JT, Guo H, 2015. Metabolism and function of hepatitis B virus cccDNA: Implications for the development of cccDNA-targeting antiviral therapeutics. Antiviral Res 122, 91–100. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Guo YH, Li YN, Zhao JR, Zhang J, Yan Z, 2011. HBc binds to the CpG islands of HBV cccDNA and promotes an epigenetic permissive state. Epigenetics 6, 720–726. [DOI] [PubMed] [Google Scholar]
- Hadden JA, Perilla JR, Schlicksup CJ, Venkatakrishnan B, Zlotnick A, Schulten K, 2018. All-atom molecular dynamics of the HBV capsid reveals insights into biological function and cryo-EM resolution limits. eLife 7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Heger-Stevic J, Zimmermann P, Lecoq L, Bottcher B, Nassal M, 2018. Hepatitis B virus core protein phosphorylation: Identification of the SRPK1 target sites and impact of their occupancy on RNA binding and capsid structure. PLoS Pathog 14, e1007488. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Helene C, Auclair H, Vegna S, Lahlali T, Pons C, Michelet M, Coute Y, Belmudes L, Kim Y, Bernardo A, Jalaguier P, Rivoire M, Arnold L, Lopatin U, Combet C, Zoulim F, Grierson D, Chabot B, Lucifora J, Salvetti A, 2020. Hepatitis B virus Core protein nuclear interactome identifies SRSF10 as a host RNA-binding protein restricting HBV RNA production. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hu J, Cheng J, Tang L, Hu Z, Luo Y, Li Y, Zhou T, Chang J, Guo JT, 2019. Virological Basis for the Cure of Chronic Hepatitis B. ACS infectious diseases 5, 659–674. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hu J, Liu K, 2017. Complete and Incomplete Hepatitis B Virus Particles: Formation, Function, and Application. Viruses 9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hu Z, Ban H, Zheng H, Liu M, Chang J, Guo JT, 2020. Protein phosphatase 1 catalyzes HBV core protein dephosphorylation and is co-packaged with viral pregenomic RNA into nucleocapsids. PLoS Pathog 16, e1008669. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Huber AD, Wolf JJ, Liu D, Gres AT, Tang J, Boschert KN, Puray-Chavez MN, Pineda DL, Laughlin TG, Coonrod EM, Yang Q, Ji J, Kirby KA, Wang Z, Sarafianos SG, 2018. The Heteroaryldihydropyrimidine Bay 38–7690 Induces Hepatitis B Virus Core Protein Aggregates Associated with Promyelocytic Leukemia Nuclear Bodies in Infected Cells. mSphere 3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ito K, Kim KH, Lok AS, Tong S, 2009. Characterization of genotype-specific carboxyl-terminal cleavage sites of hepatitis B virus e antigen precursor and identification of furin as the candidate enzyme. J Virol 83, 3507–3517. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kang JA, Kim S, Park M, Park HJ, Kim JH, Park S, Hwang JR, Kim YC, Jun Kim Y, Cho Y, Sun Jin M, Park SG, 2019. Ciclopirox inhibits Hepatitis B Virus secretion by blocking capsid assembly. Nature communications 10, 2184. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kann M, Gerlich WH, 1994. Effect of core protein phosphorylation by protein kinase C on encapsidation of RNA within core particles of hepatitis B virus. J Virol 68, 7993–8000. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Katen SP, Chirapu SR, Finn MG, Zlotnick A, 2010. Trapping of hepatitis B virus capsid assembly intermediates by phenylpropenamide assembly accelerators. ACS Chem Biol 5, 1125–1136. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Katen SP, Tan Z, Chirapu SR, Finn MG, Zlotnick A, 2013. Assembly-directed antivirals differentially bind quasiequivalent pockets to modify hepatitis B virus capsid tertiary and quaternary structure. Structure 21, 1406–1416. [DOI] [PMC free article] [PubMed] [Google Scholar]
- King RW, Ladner SK, Miller TJ, Zaifert K, Perni RB, Conway SC, Otto MJ, 1998a. Inhibition of human hepatitis B virus replication by AT-61, a phenylpropenamide derivative, alone and in combination with (−)beta-L- 2’,3’-dideoxy-3’-thiacytidine [published erratum appears in Antimicrob Agents Chemother 1999 Mar;43(3):726]. Antimicrob Agents Chemother 42, 3179–3186. [DOI] [PMC free article] [PubMed] [Google Scholar]
- King RW, Ladner SK, Miller TJ, Zaifert K, Perni RB, Conway SC, Otto MJ, 1998b. Inhibition of human hepatitis B virus replication by AT-61, a phenylpropenamide derivative, alone and in combination with (−) β-l-2′, 3′-dideoxy-3′-thiacytidine. Antimicrobial agents and chemotherapy 42, 3179–3186. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Klumpp K, Crepin T, 2014. Capsid proteins of enveloped viruses as antiviral drug targets. Curr Opin Virol 5, 63–71. [DOI] [PubMed] [Google Scholar]
- Klumpp K, Lam AM, Lukacs C, Vogel R, Ren S, Espiritu C, Baydo R, Atkins K, Abendroth J, Liao G, Efimov A, Hartman G, Flores OA, 2015. High-resolution crystal structure of a hepatitis B virus replication inhibitor bound to the viral core protein. Proc Natl Acad Sci U S A 112, 15196–15201. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ko C, Bester R, Zhou X, Xu Z, Blossey C, Sacherl J, Vondran FWR, Gao L, Protzer U, 2019. A New Role for Capsid Assembly Modulators To Target Mature Hepatitis B Virus Capsids and Prevent Virus Infection. Antimicrob Agents Chemother 64. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Koschel M, Oed D, Gerelsaikhan T, Thomssen R, Bruss V, 2000. Hepatitis B virus core gene mutations which block nucleocapsid envelopment. J Virol 74, 1–7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lahlali T, Berke JM, Vergauwen K, Foca A, Vandyck K, Pauwels F, Zoulim F, Durantel D, 2018. Novel potent capsid assembly modulators regulate multiple steps of the Hepatitis B virus life-cycle. Antimicrob Agents Chemother. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lan YT, Li J, Liao W, Ou J, 1999. Roles of the three major phosphorylation sites of hepatitis B virus core protein in viral replication. Virology 259, 342–348. [DOI] [PubMed] [Google Scholar]
- Lenz O, Verbinnen T, Hodari M, Talloen W, Vandenbossche JJ, Shukla U, Berby F, Scholtes C, Blue D, Yogaratnam J, Meyer SD, Zoulim F, 2019. HBcrAg Decline in JNJ-56136379-treated Chronic Hepatitis B Patients: Results of a Phase 1 Study. Journal of Hepatology 70(1) SUPPL, e473–e474. [Google Scholar]
- Lewellyn EB, Loeb DD, 2011a. The arginine clusters of the carboxy-terminal domain of the core protein of hepatitis B virus make pleiotropic contributions to genome replication. J Virol 85, 1298–1309. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lewellyn EB, Loeb DD, 2011b. Serine phosphoacceptor sites within the core protein of hepatitis B virus contribute to genome replication pleiotropically. PLoS One 6, e17202. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liang TJ, Block TM, McMahon BJ, Ghany MG, Urban S, Guo JT, Locarnini S, Zoulim F, Chang KM, Lok AS, 2015. Present and future therapies of hepatitis B: From discovery to cure. Hepatology 62, 1893–1908. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liu K, Luckenbaugh L, Ning X, Xi J, Hu J, 2018. Multiple roles of core protein linker in hepatitis B virus replication. PLoS Pathog 14, e1007085. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liu Y, Liu H, Hu Z, Ding Y, Pan XB, Zou J, Xi J, Yu G, Huang H, Luo MT, Guo F, Liu S, Sheng Q, Jia J, Zheng YT, Wang J, Chen X, Guo JT, Wei L, Lu F, 2020. Hepatitis B Virus Virions Produced Under Nucleos(t)ide Analogue Treatment Are Mainly Not Infectious Because of Irreversible DNA Chain Termination. Hepatology 71, 463–476. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lok AS, Zoulim F, Dusheiko G, Ghany MG, 2017. Hepatitis B cure: From discovery to regulatory approval. J Hepatol 67, 847–861. [DOI] [PubMed] [Google Scholar]
- Lucifora J, Arzberger S, Durantel D, Belloni L, Strubin M, Levrero M, Zoulim F, Hantz O, Protzer U, 2011. Hepatitis B virus X protein is essential to initiate and maintain virus replication after infection. J Hepatol 55, 996–1003. [DOI] [PubMed] [Google Scholar]
- 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, Huser N, Durantel D, Liang TJ, Munk C, Heim MH, Browning JL, Dejardin E, Dandri M, Schindler M, Heikenwalder M, Protzer U, 2014. Specific and nonhepatotoxic degradation of nuclear hepatitis B virus cccDNA. Science 343, 1221–1228. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ludgate L, Liu K, Luckenbaugh L, Streck N, Eng S, Voitenleitner C, Delaney W.E.t., Hu J, 2016. Cell-Free Hepatitis B Virus Capsid Assembly Dependent on the Core Protein C-Terminal Domain and Regulated by Phosphorylation. J Virol 90, 5830–5844. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ludgate L, Ning X, Nguyen DH, Adams C, Mentzer L, Hu J, 2012. Cyclin-dependent kinase 2 phosphorylates s/t-p sites in the hepadnavirus core protein C-terminal domain and is incorporated into viral capsids. J Virol 86, 12237–12250. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Luo J, Cui X, Gao L, Hu J, 2017. Identification of Intermediate in Hepatitis B Virus CCC DNA Formation and Sensitive and Selective CCC DNA Detection. J Virol. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Luo J, Xi J, Gao L, Hu J, 2020. Role of Hepatitis B virus capsid phosphorylation in nucleocapsid disassembly and covalently closed circular DNA formation. PLoS Pathog 16, e1008459. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lutomski CA, Lyktey NA, Zhao Z, Pierson EE, Zlotnick A, Jarrold MF, 2017. Hepatitis B Virus Capsid Completion Occurs through Error Correction. J Am Chem Soc 139, 16932–16938. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sulkowski Mark S., Fung K.A., Scott, Yuen Man-Fung, Ma Xiaoli, Lalezari Jacob, Nguyen Tuan T., Bae Ho, Schiff Eugene R., Hassanein Tarek, Hann Hie-Won, Elkhashab Magdy, Dieterich Douglas, Kwo Paul, Nahass Ronald, Ramji Alnoor, Park James, Ravendhran Natarajan, Chan Sing, Weilert Frank, Han Steven-Huy, Ayoub Walid, Gane Edward, Jacobson Ira, Bennett Michael, Bonacini Maurizio, Zayed Hany, Alves Katia, Huang Qi, Ruby Eric, Qiang Dongmei, Knox Steven J., and Colonno Richard 2019. Continued Therapy with ABI-H0731 + NrtI Results in Sequential Reduction/Loss of HBV DNA, HBV RNA, HBeAg, HBcrAg and HBsAg in HBeAg-Positive Patients, American Association for the Study of Liver Diseases (AASLD) Annual Meeting (The Liver Meeting). Conference Reports for NATAP, Boston, MA. [Google Scholar]
- Mason WS, Aldrich C, Summers J, Taylor JM, 1982. Asymmetric replication of duck hepatitis B virus DNA in liver cells: Free minus-strand DNA. Proc Natl Acad Sci U S A 79, 3997–4001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Melegari M, Wolf SK, Schneider RJ, 2005. Hepatitis B virus DNA replication is coordinated by core protein serine phosphorylation and HBx expression. J Virol 79, 9810–9820. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Milich DR, 2019. Is the function of the HBeAg really unknown? Hum Vaccin Immunother 15, 2187–2191. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mitra B, Wang J, Kim ES, Mao R, Dong M, Liu Y, Zhang J, Guo H, 2019. Hepatitis B Virus Precore Protein p22 Inhibits Alpha Interferon Signaling by Blocking STAT Nuclear Translocation. J Virol 93. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nassal M, 1992. The arginine-rich domain of the hepatitis B virus core protein is required for pregenome encapsidation and productive viral positive-strand DNA synthesis but not for virus assembly. J Virol 66, 4107–4116. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nijampatnam B, Liotta DC, 2019. Recent advances in the development of HBV capsid assembly modulators. Current opinion in chemical biology 50, 73–79. [DOI] [PubMed] [Google Scholar]
- Ning X, Basagoudanavar SH, Liu K, Luckenbaugh L, Wei D, Wang C, Wei B, Zhao Y, Yan T, Delaney W, Hu J, 2017. Capsid Phosphorylation State and Hepadnavirus Virion Secretion. J Virol 91. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ning X, Luckenbaugh L, Liu K, Bruss V, Sureau C, Hu J, 2018. Common and Distinct Capsid and Surface Protein Requirements for Secretion of Complete and Genome-Free Hepatitis B Virions. J Virol 92. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ning X, Nguyen D, Mentzer L, Adams C, Lee H, Ashley R, Hafenstein S, Hu J, 2011. Secretion of genome-free hepatitis B virus--single strand blocking model for virion morphogenesis of para-retrovirus. PLoS Pathog 7, e1002255. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Okabe M, Enomoto M, Maeda H, Kuroki K, Ohtsuki K, 2006. Biochemical characterization of suramin as a selective inhibitor for the PKA-mediated phosphorylation of HBV core protein in vitro. Biol Pharm Bull 29, 1810–1814. [DOI] [PubMed] [Google Scholar]
- Oliver RC, Potrzebowski W, Najibi SM, Nors Pedersen M, Arleth L, Mahmoudi N, Andre I, 2020. Assembly of Capsids from Hepatitis B Virus Core Protein Progresses through Highly Populated Intermediates in Presence and Absence of RNA. ACS Nano. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Orabi A, Bieringer M, Geerlof A, Bruss V, 2015. An Aptamer against the Matrix Binding Domain on the Hepatitis B Virus Capsid Impairs Virion Formation. J Virol 89, 9281–9287. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Osseman Q, Gallucci L, Au S, Cazenave C, Berdance E, Blondot ML, Cassany A, Begu D, Ragues J, Aknin C, Sominskaya I, Dishlers A, Rabe B, Anderson F, Pante N, Kann M, 2017. The chaperone dynein LL1 mediates cytoplasmic transport of empty and mature hepatitis B virus capsids. J Hepatol. [DOI] [PubMed] [Google Scholar]
- Packianathan C, Katen SP, Dann CE 3rd, Zlotnick A, 2010. Conformational changes in the hepatitis B virus core protein are consistent with a role for allostery in virus assembly. J Virol 84, 1607–1615. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pairan A, Bruss V, 2009. Functional surfaces of the hepatitis B virus capsid. J Virol 83, 11616–11623. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Perlman DH, Berg EA, O’Connor P B, Costello CE, Hu J, 2005. Reverse transcription-associated dephosphorylation of hepadnavirus nucleocapsids. Proc Natl Acad Sci U S A 102, 9020–9025. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Perlmutter JD, Hagan MF, 2015. Mechanisms of virus assembly. Annu Rev Phys Chem 66, 217–239. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Perrillo R, 2009. Benefits and risks of interferon therapy for hepatitis B. Hepatology 49, S103–111. [DOI] [PubMed] [Google Scholar]
- Polaris Observatory C, 2018. Global prevalence, treatment, and prevention of hepatitis B virus infection in 2016: a modelling study. Lancet Gastroenterol Hepatol 3, 383–403. [DOI] [PubMed] [Google Scholar]
- Ponsel D, Bruss V, 2003. Mapping of amino acid side chains on the surface of hepatitis B virus capsids required for envelopment and virion formation. J Virol 77, 416–422. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pugh J, Zweidler A, Summers J, 1989. Characterization of the major duck hepatitis B virus core particle protein. Journal of Virology 63, 1371–1376. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Qazi S, Schlicksup CJ, Rittichier J, VanNieuwenhze MS, Zlotnick A, 2018. An Assembly-Activating Site in the Hepatitis B Virus Capsid Protein Can Also Trigger Disassembly. ACS Chem Biol 13, 2114–2120. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Qi Y, Gao Z, Xu G, Peng B, Liu C, Yan H, Yao Q, Sun G, Liu Y, Tang D, Song Z, He W, Sun Y, Guo JT, Li W, 2016. DNA Polymerase kappa Is a Key Cellular Factor for the Formation of Covalently Closed Circular DNA of Hepatitis B Virus. PLoS Pathog 12, e1005893. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Qiu Z, Lin X, Zhou M, Liu Y, Zhu W, Chen W, Zhang W, Guo L, Liu H, Wu G, Huang M, Jiang M, Xu Z, Zhou Z, Qin N, Ren S, Qiu H, Zhong S, Zhang Y, Zhang Y, Wu X, Shi L, Shen F, Mao Y, Zhou X, Yang W, Wu JZ, Yang G, Mayweg AV, Shen HC, Tang G, 2016. Design and Synthesis of Orally Bioavailable 4-Methyl Heteroaryldihydropyrimidine Based Hepatitis B Virus (HBV) Capsid Inhibitors. J Med Chem 59, 7651–7666. [DOI] [PubMed] [Google Scholar]
- Roossinck MJ, Siddiqui A, 1987. In vivo phosphorylation and protein analysis of hepatitis B virus core antigen. J Virol 61, 955–961. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ruan L, Hadden JA, Zlotnick A, 2018. Assembly properties of Hepatitis B Virus core protein mutants correlate with their resistance to assembly-directed antivirals. J Virol. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Scaturro P, Trist IM, Paul D, Kumar A, Acosta EG, Byrd CM, Jordan R, Brancale A, Bartenschlager R, 2014. Characterization of the mode of action of a potent dengue virus capsid inhibitor. J Virol 88, 11540–11555. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Schlicksup CJ, Wang JC, Francis S, Venkatakrishnan B, Turner WW, VanNieuwenhze M, Zlotnick A, 2018. Hepatitis B virus core protein allosteric modulators can distort and disrupt intact capsids. eLife 7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Schmitz A, Schwarz A, Foss M, Zhou L, Rabe B, Hoellenriegel J, Stoeber M, Pante N, Kann M, 2010. Nucleoporin 153 arrests the nuclear import of hepatitis B virus capsids in the nuclear basket. PLoS Pathog 6, e1000741. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Schultz U, Summers J, Staeheli P, Chisari FV, 1999. Elimination of duck hepatitis B virus RNA-containing capsids in duck interferon-alpha-treated hepatocytes. J Virol 73, 5459–5465. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Selzer L, Kant R, Wang JC, Bothner B, Zlotnick A, 2015. Hepatitis B Virus Core Protein Phosphorylation Sites Affect Capsid Stability and Transient Exposure of the C-terminal Domain. J Biol Chem 290, 28584–28593. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sheraz M, Cheng J, Tang L, Chang J, Guo JT, 2019. Cellular DNA topoisomerases are required for the synthesis of hepatitis B virus covalently closed circular DNA. J Virol. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sims KD, Gane E, Yuen MF, Berliba E, Sukeepaisarnjaroen W, Ahn SH, Tanwandee T, Lim YS, Kim Y, Poovorawan K, Tangkijvanich P, Chan HLY, Brow J, Moore C, Mani N, Rijnbrand R, Cole A, Sofia M, Thi E, Kim J, Eley T, Lee ACH, Picchio G, 2019. Key findings leading to the discontinuation of a Capsid Inhibitor (CI), AB-506, in Healthy Subjects (HS) and Chronic Hepatitis B (CHB) Subjects, HepDART. Virology Education, Kauii, HI. [Google Scholar]
- Stray SJ, Bourne CR, Punna S, Lewis WG, Finn M, Zlotnick A, 2005a. A heteroaryldihydropyrimidine activates and can misdirect hepatitis B virus capsid assembly. Proceedings of the National Academy of Sciences 102, 8138–8143. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Stray SJ, Bourne CR, Punna S, Lewis WG, Finn MG, Zlotnick A, 2005b. A heteroaryldihydropyrimidine activates and can misdirect hepatitis B virus capsid assembly. Proc Natl Acad Sci U S A 102, 8138–8143. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Stray SJ, Zlotnick A, 2006. BAY 41–4109 has multiple effects on Hepatitis B virus capsid assembly. J Mol Recognit 19, 542–548. [DOI] [PubMed] [Google Scholar]
- Su PY, Yang CJ, Chu TH, Chang CH, Chiang C, Tang FM, Lee CY, Shih C, 2016. HBV maintains electrostatic homeostasis by modulating negative charges from phosphoserine and encapsidated nucleic acids. Sci Rep 6, 38959. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sulkowski MS, Agarwal K, Fung S, Yuen MF, Ma X, Lalezari J, Nguyen TT, Bae H, Schiff ER, Hassanein T, Hann HW, Elkhashab M, Dieterich D, Kwo P, Ramji NRA, Park J, Ravendhran N, Chan S, Weilert F, Han SH, Ayoub W, Gane E, Jacobson I, Bennett M, Bonacini M, Zayed H, Alves K, Huang Q, Ruby E, Qiang D, Knox SJ, Colonno R, 2019. Continued Therapy with ABI-H0731 + NrtI Results in Sequential Reduction/Loss of HBV DNA, HBV RNA, HBeAg, HBcrAg and HBsAg in HBeAg-Positive Patients. Hepatology 70(6) SUPPL 1487A. [Google Scholar]
- Summers J, Mason WS, 1982. Replication of the genome of a hepatitis B-like virus by reverse transcription of an RNA intermediate. Cell 29, 403–415. [DOI] [PubMed] [Google Scholar]
- Summers J, Smith PM, Horwich AL, 1990. Hepadnavirus envelope proteins regulate covalently closed circular DNA amplification. Journal of Virology 64, 2819–2824. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tan Z, Maguire ML, Loeb DD, Zlotnick A, 2013. Genetically altering the thermodynamics and kinetics of hepatitis B virus capsid assembly has profound effects on virus replication in cell culture. J Virol 87, 3208–3216. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tan Z, Pionek K, Unchwaniwala N, Maguire ML, Loeb DD, Zlotnick A, 2015. The interface between hepatitis B virus capsid proteins affects self-assembly, pregenomic RNA packaging, and reverse transcription. J Virol 89, 3275–3284. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tang L, Sheraz M, McGrane M, Chang J, Guo JT, 2019. DNA Polymerase alpha is essential for intracellular amplification of hepatitis B virus covalently closed circular DNA. PLoS Pathog 15, e1007742. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tang L, Zhao Q, Wu S, Cheng J, Chang J, Guo JT, 2017. The current status and future directions of hepatitis B antiviral drug discovery. Expert opinion on drug discovery 12, 5–15. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tavis JE, Perri S, Ganem D, 1994. Hepadnavirus reverse transcription initiates within the stem-loop of the RNA packaging signal and employs a novel strand transfer. Journal of Virology 68, 3536–3543. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tian Y, Kuo CF, Akbari O, Ou JH, 2016. Maternal-Derived Hepatitis B Virus e Antigen Alters Macrophage Function in Offspring to Drive Viral Persistence after Vertical Transmission. Immunity 44, 1204–1214. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tuttleman JS, Pourcel C, Summers J, 1986. Formation of the pool of covalently closed circular viral DNA in hepadnavirus-infected cells. Cell 47, 451–460. [DOI] [PubMed] [Google Scholar]
- Vandenbossche J, Jessner W, van den Boer M, Biewenga J, Berke JM, Talloen W, De Zwart L, Snoeys J, Yogaratnam J, 2019. Pharmacokinetics, Safety and Tolerability of JNJ-56136379, a Novel Hepatitis B Virus Capsid Assembly Modulator, in Healthy Subjects. Adv Ther 36, 2450–2462. [DOI] [PubMed] [Google Scholar]
- Venkatakrishnan B, Katen SP, Francis S, Chirapu S, Finn MG, Zlotnick A, 2016. Hepatitis B Virus Capsids Have Diverse Structural Responses to Small-Molecule Ligands Bound to the Heteroaryldihydropyrimidine Pocket. J Virol 90, 3994–4004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Venkatakrishnan B, Zlotnick A, 2016. The Structural Biology of Hepatitis B Virus: Form and Function. Annual review of virology 3, 429–451. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang GH, Seeger C, 1993. Novel mechanism for reverse transcription in hepatitis B viruses. J Virol 67, 6507–6512. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang Y-J, Lu D, Xu Y-B, Xing W-Q, Tong X-K, Wang G-F, Feng C-L, He P-L, Yang L, Tang W, 2015. A novel pyridazinone derivative inhibits hepatitis B virus replication by inducing genome-free capsid formation. Antimicrobial agents and chemotherapy 59, 7061–7072. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wingfield PT, Stahl SJ, Williams RW, Steven AC, 1995. Hepatitis core antigen produced in Escherichia coli: subunit composition, conformation analysis, and in vitro capsid assembly. Biochemistry 34, 4919–4932. [DOI] [PubMed] [Google Scholar]
- Wittkop L, Schwarz A, Cassany A, Grun-Bernhard S, Delaleau M, Rabe B, Cazenave C, Gerlich W, Glebe D, Kann M, 2010. Inhibition of protein kinase C phosphorylation of hepatitis B virus capsids inhibits virion formation and causes intracellular capsid accumulation. Cell Microbiol 12, 962–975. [DOI] [PubMed] [Google Scholar]
- Wu G, Liu B, Zhang Y, Li J, Arzumanyan A, Clayton MM, Schinazi RF, Wang Z, Goldmann S, Ren Q, 2013. Preclinical characterization of GLS4, an inhibitor of hepatitis B virus core particle assembly. Antimicrobial agents and chemotherapy 57, 5344–5354. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wu S, Luo Y, Viswanathan U, Kulp J, Cheng J, Hu Z, Xu Q, Zhou Y, Gong GZ, Chang J, Li Y, Guo JT, 2018. CpAMs induce assembly of HBV capsids with altered electrophoresis mobility: Implications for mechanism of inhibiting pgRNA packaging. Antiviral Res 159, 1–12. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wu S, Zhao Q, Zhang P, Kulp J, Hu L, Hwang N, Zhang J, Block TM, Xu X, Du Y, Chang J, Guo JT, 2017. Discovery and mechanistic study of benzamide derivatives that modulate hepatitis B virus capsid assembly. J Virol 91, e0051917. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wynne S, Crowther R, Leslie A, 1999. The crystal structure of the human hepatitis B virus capsid. Molecular cell 3, 771–780. [DOI] [PubMed] [Google Scholar]
- Xia Y, Lucifora J, Reisinger F, Heikenwalder M, Protzer U, 2014. Virology. Response to Comment on “Specific and nonhepatotoxic degradation of nuclear hepatitis B virus cccDNA”. Science 344, 1237. [DOI] [PubMed] [Google Scholar]
- Xia Y, Stadler D, Lucifora J, Reisinger F, Webb D, Hosel M, Michler T, Wisskirchen K, Cheng X, Zhang K, Chou WM, Wettengel JM, Malo A, Bohne F, Hoffmann D, Eyer F, Thimme R, Falk CS, Thasler WE, Heikenwalder M, Protzer U, 2016. Interferon-gamma and Tumor Necrosis Factor-alpha Produced by T Cells Reduce the HBV Persistence Form, cccDNA, Without Cytolysis. Gastroenterology 150, 194–205. [DOI] [PubMed] [Google Scholar]
- Yan H, Zhong G, Xu G, He W, Jing Z, Gao Z, Huang Y, Qi Y, Peng B, Wang H, Fu L, Song M, Chen P, Gao W, Ren B, Sun Y, Cai T, Feng X, Sui J, Li W, 2012. Sodium taurocholate cotransporting polypeptide is a functional receptor for human hepatitis B and D virus. eLife 1, e00049. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yan Z, Wu D, Hu H, Zeng J, Yu X, Xu Z, Zhou Z, Zhou X, Yang G, Young JAT, Gao L, 2019. Direct Inhibition of Hepatitis B e Antigen by Core Protein Allosteric Modulator. Hepatology 70, 11–24. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yang F, Yu X, Zhou C, Mao R, Zhu M, Zhu H, Ma Z, Mitra B, Zhao G, Huang Y, Guo H, Wang B, Zhang J, 2019a. Hepatitis B e antigen induces the expansion of monocytic myeloid-derived suppressor cells to dampen T-cell function in chronic hepatitis B virus infection. PLoS Pathog 15, e1007690. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yang L, Liu F, Tong X, Hoffmann D, Zuo J, Lu M, 2019b. Treatment of Chronic Hepatitis B Virus Infection Using Small Molecule Modulators of Nucleocapsid Assembly: Recent Advances and Perspectives. ACS infectious diseases 5, 713–724. [DOI] [PubMed] [Google Scholar]
- Yang L, Wang Y-J, Chen H-J, Shi L-P, Tong X-K, Zhang Y-M, Wang G-F, Wang W-L, Feng C-L, He P-L, 2016. Effect of a hepatitis B virus inhibitor, NZ-4, on capsid formation. Antiviral research 125, 25–33. [DOI] [PubMed] [Google Scholar]
- Yang W, Guo J, Ying Z, Hua S, Dong W, Chen H, 1994. Capsid assembly and involved function analysis of twelve core protein mutants of duck hepatitis B virus. J Virol 68, 338–345. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yeh CT, Ou JH, 1991. Phosphorylation of hepatitis B virus precore and core proteins. J Virol 65, 2327–2331. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yogaratnam J, Zoulim F, Vandenbossche J, Lenz O, Talloen W, Moscalu L, Mohamed R, Streinu-Cercel A, Chuang WL, Bourgeois S, Yang S-S, Buti M, Crespo J, Chen YC, Pascasio JM, Sarrazin C, Vanwolleghem T, Shukla U, Fry J, 2019. Safety, antiviral activity, and pharmacokinetics of a novel hepatitis B virus capsid assembly modulator, JNJ-56136379, in Asian and non-Asian patients with chronic hepatitis B. Journal of Hepatology 70(1) SUPPL, e489–e490. [Google Scholar]
- Yu X, Jin L, Jih J, Shih C, Zhou ZH, 2013. 3.5A cryoEM structure of hepatitis B virus core assembled from full-length core protein. PLoS One 8, e69729. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yuan TT, Sahu GK, Whitehead WE, Greenberg R, Shih C, 1999. The mechanism of an immature secretion phenotype of a highly frequent naturally occurring missense mutation at codon 97 of human hepatitis B virus core antigen. J Virol 73, 5731–5740. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yuen MF, Agarwal K, Gane E, Nguyen TT, Hassanein TI, Kim DJ, Alves K, Zayed H, Qiang D, Ruby E, Evanchik M, Huang Q, Knox SJ, Colonno R, 2019a. The Second-Generation Hepatitis B Virus (HBV) Core Inhibitor (CI) ABI-H2158 is Associated with Potent Antiviral Activity in a 14-Day Monotherapy Study in HBeAg-positive Patients with Chronic Hepatitis B (CHB). Hepatology 70(6) SUPPL 1497A. [Google Scholar]
- Yuen MF, Agarwal K, Gane E, Schwabe C, Ahn SH, Kim DJ, Lim YS, Cheng W, Sievert W, Visvanathan K, Ruby E, Liaw S, Yan R, Connelly E, Cai D, Huang Q, Colonno R, Lopatin U, 2018. Final Results of a Phase 1b 28-day Study of ABI-H0731, A Novel Core Inhibitor, In Non-cirrhotic Viremic Subjects With Chronic HBV. Hepatology 68(1) SUPPL 46A. [Google Scholar]
- Yuen MF, Agarwal K, Gane EJ, Schwabe C, Ahn SH, Kim DJ, Lim YS, Cheng W, Sievert W, Visvanathan K, Ruby E, Liaw S, Yan R, Huang Q, Colonno R, Lopatin U, 2020. Safety, pharmacokinetics, and antiviral effects of ABI-H0731, a hepatitis B virus core inhibitor: a randomised, placebo-controlled phase 1 trial. Lancet Gastroenterol Hepatol 5, 152–166. [DOI] [PubMed] [Google Scholar]
- Yuen MF, Gane EJ, Kim DJ, Weilert F, Yuen Chan HL, Lalezari J, Hwang SG, Nguyen T, Flores O, Hartman G, Liaw S, Lenz O, Kakuda TN, Talloen W, Schwabe C, Klumpp K, Brown N, 2019b. Antiviral Activity, Safety, and Pharmacokinetics of Capsid Assembly Modulator NVR 3–778 in Patients with Chronic HBV Infection. Gastroenterology 156, 1392–1403 e1397. [DOI] [PubMed] [Google Scholar]
- Yuen MF, Locarnini S, Given B, Schluep T, Hamilton J, Biermer M, Kalmeijer R, Beumont-Mauviel M, Lenz O, Cloherty G, Jackson K, Ferrari C, Lai CL, Liu KSH, Mak LY, Wong DKH, To WP, Ko KL, Gish RG, 2019c. First Clinical Experience with RNA Interference-based Triple Combination Therapy in Chronic Hepatitis B: JNJ-3989, JNJ-6379 and a Nucleos(t)ide Analogue. Hepatology 70(6) SUPPL, 1489A. [Google Scholar]
- Zhang H, Niu J, Ding HY, Zhang Y, Luo L, Li C, Wu M, Hu Y, Li X, Chen H, Zhu X, Liu Y, He Q, Chen Y, Ren Q, Gu B, Jing L, 2019. Analysis of Factors Influencing the Efficacy and Safety of GLS4 Treatment of Patients with Chronic Hepatitis B. Hepatology 70(1) SUPPL 429A. [Google Scholar]
- Zhang YY, Summers J, 1999. Enrichment of a precore-minus mutant of duck hepatitis B virus in experimental mixed infections. J Virol 73, 3616–3622. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhao N, Jia B, Zhao H, Xu J, Sheng X, Luo L, Huang Z, Wang X, Ren Q, Zhang Y, Zhao X, Cui Y, 2019. A First-in-Human Trial of GLS4, a Novel Inhibitor of Hepatitis B Virus Capsid Assembly, following Single- and Multiple-Ascending-Oral-Dose Studies with or without Ritonavir in Healthy Adult Volunteers. Antimicrob Agents Chemother 64, e01686–01619. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhao Q, Guo JT, 2020. Have the starting lineup of five for HBV cccDNA synthesis been identified? Hepatology. [DOI] [PubMed] [Google Scholar]
- Zhao Q, Hu Z, Cheng J, Wu S, Luo Y, Chang J, Hu J, Guo JT, 2018. Hepatitis B Virus Core Protein Dephosphorylation Occurs during Pregenomic RNA Encapsidation. J Virol 92. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zheng Y, Fu XD, Ou JH, 2005. Suppression of hepatitis B virus replication by SRPK1 and SRPK2 via a pathway independent of the phosphorylation of the viral core protein. Virology 342, 150–158. [DOI] [PubMed] [Google Scholar]
- Zhou Z, Hu T, Zhou X, Wildum S, Garcia-Alcalde F, Xu Z, Wu D, Mao Y, Tian X, Zhou Y, Shen F, Zhang Z, Tang G, Najera I, Yang G, Shen HC, Young JA, Qin N, 2017. Heteroaryldihydropyrimidine (HAP) and Sulfamoylbenzamide (SBA) Inhibit Hepatitis B Virus Replication by Different Molecular Mechanisms. Sci Rep 7, 42374. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zlotnick A, 2005. Theoretical aspects of virus capsid assembly. J Mol Recognit 18, 479–490. [DOI] [PubMed] [Google Scholar]
- Zlotnick A, Cheng N, Stahl S, Conway J, Steven A, Wingfield P, 1997. Localization of the C terminus of the assembly domain of hepatitis B virus capsid protein: implications for morphogenesis and organization of encapsidated RNA. Proceedings of the National Academy of Sciences 94, 9556–9561. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zlotnick A, Mukhopadhyay S, 2011. Virus assembly, allostery and antivirals. Trends Microbiol 19, 14–23. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zlotnick A, Venkatakrishnan B, Tan Z, Lewellyn E, Turner W, Francis S, 2015. Core protein: A pleiotropic keystone in the HBV lifecycle. Antiviral Res 121, 82–93. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zoulim F, Yogaratnam J, Vandenbossche JJ, Moscalu I, Streinu-Cercel A, Lenz O, Bourgeois S, Buti M, Talloen W, Crespo J, Pascasio JM, Sarrazin C, Vanwolleghem T, Shukla U, Fry J, 2018. Safety, Pharmacokinetics and Antiviral Activity of a Novel Hepatitis B Virus (HBV) Capsid Assembly Modulator, JNJ-56136379, in Patients with Chronic Hepatitis B (CHB), Hepatology, pp. 47A–48A. [Google Scholar]





