Open access

Subversion of RNA Processing Pathways by the Hepatitis delta Virus

Written By

Yasnee Beeharry and Martin Pelchat

Submitted: 12 November 2010 Published: 29 August 2011

DOI: 10.5772/21332

From the Edited Volume

RNA Processing

Edited by Paula Grabowski

Chapter metrics overview

2,559 Chapter Downloads

View Full Metrics

1. Introduction

Viruses are the most opportunistic pathogens, since they extensively use host cell components for their replication. Among the various host factors usurped by viruses, proteins involved in RNA processing pathways are frequently used to regulate their own replication and to affect host protein expression (Lai, 1998). Over the years, the study of viral replication led to discoveries on how cellular RNA processing pathways are complex and how each step is crucial and extremely regulated. In addition to provide a better understanding of viral pathogenesis, investigation of viruses continue to be used as tools to understand the normal cellular RNA processing pathways and to identify new actors or new functions of known components.

The hepatitis delta virus (HDV) is a fascinating model. This virus is the smallest known pathogenic RNA capable of infecting human cells, and has to rely more heavily than any other viruses on host proteins for its replication (for general reviews of HDV biology and replication, see (Lai, 2005; Taylor, 2006, 2009; Tseng & Lai, 2009; Pascarella & Negro, 2011)). In this chapter, we present the current knowledge of the use of host proteins involved in RNA processing pathways by HDV, from RNA synthesis to RNA maturation. In addition, we discuss the potential contribution of these proteins to the life cycle of HDV, and possible impacts these interactions might have on natural host processes during an HDV infection.

Advertisement

2. The hepatitis delta virus

HDV is a defective RNA virus that requires the hepatitis B virus (HBV) envelope proteins for encapsidation and dissemination (Rizzetto et al., 1977). Co- and super-infection of HBV carriers by HDV causes acute exacerbation of the disease and subsequent chronic hepatitis. People with HDV superinfection have a much greater risk of developing fulminant hepatitis, hepatocellular carcinoma, cirrhosis, and liver failure than people infected with HBV alone (Fattovich et al., 2004; Su et al., 2006; Taylor, 2006; Romeo et al., 2009).

As stated above, HDV is the smallest known human RNA pathogen. It consists of a circular negative single-stranded RNA of about 1,680 nucleotides. This RNA can fold on itself using 74% intra-molecular base-pairing to form an unbranched, rod-like structure that can be divided into two domains (Fig. 1; reviewed by (Chen et al., 1986; Lai, 2005; Taylor, 2006, 2009; Tseng & Lai, 2009)). The left-terminal domain (~360 nt) includes both genomic and antigenomic self-cleaving motifs (i.e. delta ribozymes or δRz). The right-terminal domain contains a single open reading frame (ORF) encoding two viral proteins produced alternatively: the small HDAg (HDAg-S) and the large HDAg (HDAg-L; Weiner et al., 1988). The difference between these two proteins is that HDAg-L contains 19 additional amino acids at its C-terminus. During replication, post-transcriptional editing of HDV antigenomic RNA results in transcriptional read-through from the HDAg-S gene (24 kDa; 195 amino acids) and to the production of HDAg-L (27 kDa; 214 amino acids; Casey et al., 1992; Casey & Gerin, 1995; Wong & Lazinski, 2002). Although they are mostly identical, each protein has a distinct function. HDAg-S is essential for HDV replication (Kuo et al., 1989; Yamaguchi et al., 2001), while the HDAg-L is necessary for virion assembly and is reported to be a dominant negative inhibitor of replication (Chao et al., 1990; Chang et al., 1991; Ryu et al., 1992; Sureau et al., 1992; Lee et al., 1995).

Figure 1.

The Hepatitis delta virus RNA genome. The HDV genome is depicted as a superposition of both genomic and antigenomic polarities. The delta ribozyme motifs (δRz) and their respective cleavage sites are indicated. The post-transcriptional editing performed by ADAR-1 is also indicated on the top stand. The arrows indicate the proposed HDAg mRNA initiation site. Numbering is in accordance with (Kuo et al., 1988).

HDAgs are RNA binding proteins that associate with HDV genomic RNA to form ribonucleoprotein (RNP) complexes in the HDV-containing HBV virion and in transfected cells (Chang et al., 1988; Niranjanakumari et al., 2002; Chang et al., 2008). This binding is likely facilitated by the rod-like structure adopted by the HDV RNA genome. In addition, HDAg-S can function as an RNA chaperone and modulate the ribozyme activity of HDV RNA (Wang et al., 2003), and is proposed to shuttle HDV RNA to the nucleus for replication, to the cytosol for packaging and export (Ryu et al., 1993; Lai, 2005; Taylor, 2006), and to modulate both HDV transcription and replication (Yamaguchi et al., 2001). Both HDAgs are post-translationally modified (i.e. serine and threonine phosphorylation (Chang et al., 1988; Mu et al., 1999; Mu et al., 2001), arginine methylation (Li et al., 2004), lysine acetylation (Mu et al., 2004), and cysteine farnesylation (Glenn et al., 1992; Otto & Casey, 1996)). These modifications regulate intermolecular interactions affecting subcellular localization of HDAgs and HDV RNA synthesis (Mu et al., 2001; Li et al., 2004; Mu et al., 2004).

Replication of HDV RNA is considered to take place in the nucleus of the infected cells using a symmetrical rolling cycle mechanism (Fig. 2; Taylor, 2009). Replication of the infectious circular RNA monomer (which is assigned genomic polarity by convention and accumulates to a greater intracellular abundance than the antigenomic species) produces linear, multimeric strands which are subsequently cleaved by endogenous ribozymes and ligated, yielding antigenomic circular RNA monomers (~30,000 copies). Using the latter RNAs as templates, the same three steps are repeated to generate the genomic RNA progeny (~300,000 copies). A third RNA also accumulates during replication, which corresponds to the unique HDV mRNA (~600 copies).

Figure 2.

Symmetrical rolling cycle mechanism used by HDV to replicate its RNA genome. Genomic RNA (blue) serves as a template for the synthesis of antigenomic RNA (red) multimers, which self-cleave and are ligated to form circular antigenomic RNA monomers. The antigenomic RNA monomers are used to synthetize circular genomic RNA monomers using the same steps. Genomic RNA also serve as a template for the transcription of HDAg mRNA.

Advertisement

3. The pathogenesis of hepatitis delta virus

Despite its apparent simplicity, HDV causes one of the most serious and rapidly progressive viral hepatitis. It provokes severe acute and chronic liver diseases, and it is associated with the development of hepatocellular carcinoma (Fattovich et al., 2004; Su et al., 2006; Taylor, 2006; Romeo et al., 2009). Interestingly, it was found that HBV replication is repressed in the presence of HDV (Krogsgaard et al., 1987; Farci et al., 1988), suggesting that the induced liver damage is likely caused by HDV replication rather than by HBV.

Since the genome of HDV can replicate in animal cells independently of HBV, many experimental systems using transfection of mammalian cells have examined the impact of the accumulation of HDAg and/or HDV RNA on the phenotype of cells in culture. It has been suggested that expression of HDAg-S or HDAg-L in mammalian cells might cause HDV pathogenesis (Cole et al., 1991; Cole et al., 1993). In contrast, some reports have shown that HDV RNA replication, rather than expression of the HDAgs, might be responsible for the cytopathic effects (Wang et al., 2001). At low concentrations of HDAg-S and without HDAg-L, low levels of HDV genome replication was reported to be able to proceed for at least 2 years without any deleterious effects on the host cell (Chang et al., 2005). When higher level of HDV RNA was induced, cell-cycle arrest in G1/G0 was observed followed by cell death (Chang et al., 2005). Using such systems, experiments have shown that when HDAg and/or HDV RNA are over-expressed, host proteins involved in regulation of cell metabolism and energy pathways, nucleic acid and protein metabolism, transport, signal transduction, apoptosis, and cell growth and maintenance are differentially produced (Mota et al., 2008; Mota et al., 2009). During viral replication, interaction of host factors with both HDAg and the HDV RNA genome might thus affect normal cellular activities.

Advertisement

4. Hijacking of cellular RNA processing machinery by HDV

Because HDV has an extremely limited protein coding capacity, it is completely dependent on its host cell for its replication. HDV has to interact and divert several proteins during its life cycle, including those in the nucleus involved in both viral replication and transcription and those in the cytoplasm that are required for the production of HDAgs. The high accumulation of HDV RNA in infected cells (Chen et al., 1986) suggests that the interaction of HDV RNA with host factors might interfere with several of the normal cellular functions, thereby eliciting the virus’ pathogenic effect. Over the last few years, studies have investigated cellular proteins that might interact with either HDAgs or HDV RNA. Several of these proteins are involved in RNA processing, including RNA polymerase subunits, heterogeneous ribonucleoproteins (hnRNPs), RNA helicases, RNA-binding proteins, and both transcription and splicing factors (for a review, see (Greco-Stewart & Pelchat, 2010)). However, the physiological significance of the interaction of most of these proteins remains enigmatic and the identification of additional host factors involved in viral replication is still required to gain a better understanding of both HDV replication and the associated pathogenicity.

4.1. HDV uses host RNA polymerase(s) for both its replication and transcription.

Because HDV does not encode its own replicase, host DNA-dependent RNA polymerases (RNAPs) are considered to be involved in the replication and transcription of HDV RNAs. Several approaches have been used to identify the host RNAP(s) involved in HDV replication/transcription and to study how this redirection from DNA to RNA templates occurs. However, it is largely unknown how the RNA-dependent RNA synthesis of HDV proceeds and there is still controversy on the identity of the RNAP(s) involved in HDV replication.

Many studies using cultured cells and cell extracts have provided evidence of the involvement of RNAP II in HDV replication based on the sensitivity of the accumulation of HDV mRNA and processed unit-length genomic HDV RNAs to low levels of α-amanitin, a mycotoxin that inhibits DNA-dependent RNA synthesis by RNAP II (MacNaughton et al., 1991; Fu & Taylor, 1993; Filipovska & Konarska, 2000; Moraleda & Taylor, 2001; Chang et al., 2008). These results were substantiated by experiments using cells containing an α-amanitin-resistant allele of the largest subunit of RNAP II, which partially relieved transcription inhibition by α-aminitin (Filipovska & Konarska, 2000). RNAP II is furthermore speculated to be involved in the transcription of the HDAg mRNA because, in vivo, this mRNA was shown to be post-transcriptionally processed with a 5'-cap and a 3'-poly(A) tail, which are typical features of transcripts generated by RNAP II (Gudima et al., 1999).

Synthesis of complementary strands was possible in transcription assays using nuclear extract (NE) from HeLa cells and RNA derived from the left terminal stem-loop domain of antigenomic HDV RNA (Filipovska & Konarska, 2000). Accumulation of this RNA product was highly sensitive to α-amanitin. This sensitivity was partially abrogated in experiments conducted in NE from cells containing an α-amanitin-resistant allele of the largest subunit of RNAP II, suggesting the involvement of RNAP II in this reaction. Interestingly, the transcription did not proceed by de novo initiation, but rather by cleavage of the RNA template followed by extension of the new 3’ end, generating a chimeric template/transcript RNA product. In addition, transcription stopped after the elongation of only 41 nucleotides, and addition of HDAg-S allowed the RNA synthesis to resume (Filipovska & Konarska, 2000). Although not believed to be required for the initiation of RNA synthesis, HDAg-S was suggested to be implicated in the regulation of the elongation reaction through direct binding to RNAP II and displacement of the elongation repressors NELF and DSIF (Yamaguchi et al., 2001; Yamaguchi et al., 2007). Interestingly, HDAg has been shown to affect transcription from numerous bacterial and eukaryotic promoters (Brazas & Ganem, 1996; Lo et al., 1998; Wei & Ganem, 1998). For example, transcription from the serum response element (SRE) pathway was reported to be stimulated by HDAg-L in a manner independent of the binding of the serum response factor (SRF) protein (Goto et al., 2000; Goto et al., 2003). Because HDAg directly binds RNAP II and stimulates transcription elongation (Yamaguchi et al., 2001), it is possible that this protein affects the rate of host mRNA transcription, and consequently some RNA processing events.

Recently, RNAP II was reported to interact with HDV-derived RNAs at sites located within the terminal stem-loop domains of both polarities of HDV RNA (Greco-Stewart et al., 2007; Abrahem & Pelchat, 2008; Chang et al., 2008). Mutagenesis near these tips of the rod affects both HDV accumulation in cells and RNAP II binding in vitro (Beard et al., 1996; Greco-Stewart et al., 2007; Abrahem & Pelchat, 2008). To obtain insights into the recognition of HDV RNA promoter by RNAP II, an RNA fragment derived from the right terminal stem-loop region of genomic HDV RNA was used. This RNA fragment includes the reported initiation site for HDAg mRNA transcription (i.e. position 1630; Gudima et al., 2000) and it has been used in several investigations as templates for in vitro transcription using NE (Beard et al., 1996; Abrahem & Pelchat, 2008). Inhibition of HDV transcription by an antibody raised against the C-terminal domain of RNAP II, and direct binding of the RNA polymerase II confirmed that this HDV-derived RNA acts as a promoter for RNAP II (Abrahem & Pelchat, 2008). Using RNA affinity chromatography, it was established that an active RNAP II pre-initiation complex forms on this RNA promoter, and that this complex contains the same RNAP II subunits as those found on a typical DNA promoter during promoter recognition (i.e. RNAP II, TFIIA, TFIIB, TFIID, TFIIE, TFIIF, TFIIH, and TFIIS; Abrahem & Pelchat, 2008). In addition, analogous to what is observed to occur on DNA promoters during transcription (Yudkovsky et al., 2000), it was found that a re-initiation complex (i.e. TFIIA, TFIID, and TFIIE) might remain on the RNA promoter following transcription initiation (Abrahem & Pelchat, 2008). Furthermore, the direct binding of the TATA-binding protein was demonstrated, and it was suggested that this protein might be required to nucleate the RNAP II pre-initiation complex on the RNA promoter (Abrahem & Pelchat, 2008). Finally, the tertiary structure of purified RNAP II engaged in transcription on an HDV-RNA promoter has been reported (Lehmann et al., 2007). Superposing the DNA and RNA template-product crystal structures on the RNAP II transcription active site showed that they both occupied the same site. This finding suggests that RNAP II is recognizing RNA and DNA templates in similar way.

In contrast to the genomic RNA and HDAg mRNA, the synthesis of the antigenomic species has been shown to be resistant to higher doses of α-amanitin and to take place in the nucleolus, suggesting the involvement of another yet unknown RNAP in the life cycle of HDV (Modahl et al., 2000; Macnaughton et al., 2002; Li et al., 2006). It has been hypothesized that RNAP I or an RNAP I-like polymerase might be involved in HDV replication, because in vitro antigenomic HDV RNA synthesis is significantly decreased upon immunodepletion of the NE with an α-SL1 antibody (Li et al., 2006), and both polarities of HDV RNA were shown to associate with RNAP I in cells replicating HDV RNA (Greco-Stewart et al., 2009). Similarly, a role for RNAP III was also suggested based on its association with the HDV RNA genome in cells replicating HDV (Greco-Stewart et al., 2009). However, whether or not these associations are relevant for HDV replication is uncertain.

4.2. Interaction of host proteins with HDAgs.

Several RNA helicases were found to interact with HDAgs, including nucleolin, nucleophosmin, DDX1 and DHX15 (Lee et al., 1998; Huang et al., 2001; Cao et al., 2009). Nucleolin is a multifunctional nuclear phosphoprotein with both DNA/RNA helicase and ATPase activities, and is involved in several processes, including synthesis and maturation of the ribosome, cell proliferation and growth, nuclear trafficking, cytokinesis, nucleogenesis, transcriptional repression, replication, signal transduction, and chromosome remodelling (Tuteja & Tuteja, 1998). The nucleolin-binding domain of HDAgs is conserved among HDV variants, and it was shown that this domain is required for nucleolar targeting and accumulation of HDV RNA. Like Nucleolin, nucleophosmin is another multifunctional nuclear phosphoprotein that has been shown to interact with both HDAg-S and HDAg-L (Huang et al., 2001). This protein is involved in cell growth and proliferation, nuclear shuttling, and ribosome biogenesis. Nucleophosmin was reported to be upregulated in cells replicating HDV and to co-localize with HDAg in the nucleolus (Okuwaki, 2008). Both DDX1 and DHX15 are ATP-dependent RNA helicases involved in cellular processes such as alteration of RNA secondary structure during translation initiation, splicing, and both ribosome and spliceosome assembly (de la Cruz et al., 1999; Wen et al., 2008). Although the function of all these RNA helicases in HDV replication is unknown, their levels correlates with the amount of HDV RNA present in cells (Huang et al., 2001; Cao et al., 2009), suggesting that they might enhance HDV replication by affecting the localization of HDV RNPs. Alternatively, they might assist to the refolding of the HDV RNA genome during viral replication (Ghisolfi et al., 1992; Tuteja & Tuteja, 1998). However, it is possible that their interaction with HDAgs might be indirect since all the interactions were also observed in the absence of HDV replication (Hiscox, 2002; Cao et al., 2009; Han et al., 2009).

HDAg-S was also found to interact with many proteins capable of associating with the Argonaute (Ago) proteins, and suspected to be involved in miRNA-mediated silencing (Haussecker et al., 2008; Cao et al., 2009). Using a combined proteomic-RNAi screen, several RNA-binding proteins (IMP2, IMP3, RALY, RBM14, ILF2, and ILF3), the polyadenylate-binding protein 4 (BABPC4), and the putative RNA helicase MOV10 were found to be required for efficient HDV replication (Haussecker et al., 2008; Cao et al., 2009). Interestingly, HDV accumulation was found to be diminished only when Ago4 expression wass prevented, but not in the absence of Ago1-3 (Haussecker et al., 2008). Although the precise function of all these proteins for HDV is unclear, it was suggested that some uncharacterized downstream effectors in the microRNA (miRNA) interference pathway might remodel HDV RNA and facilitate HDV replication/transcription (Haussecker et al., 2008). It is also possible that during HDV replication, specific miRNA pathways might be deregulated.

4.3. Interaction of host proteins with the HDV RNA genome.

An essential host protein for the progression of HDV infection is the adenosine deaminase that acts on RNA (ADAR). This enzyme is involved in site-selective RNA editing, by changing specific adenosine residues to inosine (Fig. 1; Casey et al., 1992; Casey & Gerin, 1995). The small isoform of adenosine deaminase acting on RNA (ADAR-1) performs the post-transcriptional RNA editing of the HDV antigenome leading to the production of HDAg-L (Wong & Lazinski, 2002). Specifically, during HDV replication, the UAG amber termination codon of HDAg-S is converted to UIG on the antigenomic RNA genome. When genomic RNA is generated from the modified antigenome, inosine is read as guanine yielding a codon for tryptophan instead of a stop codon. This modification is retained in subsequent rounds of replication and results in translational readthrough which produces HDAg-L. HDAg-S was shown to inhibit this editing event (Polson et al., 1998), thus delaying HDAg-L production and regulating HDV life cycle. As stated above, HDV RNA accumulating at high levels was shown to be pathogenic to the cell (Chang et al., 2005). Thus, it is possible that during a typical HDV infection, HDV RNA-editing by ADAR-1, with the appearance of HDAg-L and the resulting down-regulation of viral RNA replication might have been selected to insure cell survival.

Recently, several approaches were used to identify additional host proteins that might be involved in HDV replication using HDV-derived RNA fragments as baits (Sikora et al., 2009). These studies led to the identification of SC35, ASF/SF2, the heterogeneous ribonucleoproteins L (hnRNP-L), the polypyrimidine tract-binding (PTB) protein-associated splicing factor (PSF), p54nrb, the glyceraldehyde 3-phosphate dehydrogenase (GAPDH), and the eukaryotic translation elongation factor 1 alpha 1 (eEF1A1). Due to their specific interactions with both polarities of HDV RNA (Sikora et al., 2009), it is possible that these RNA processing proteins could play a role in HDV RNA biology.

SC35, ASF/SF2 and hnRNP-L are abundant nuclear proteins playing important roles in pre-mRNA splicing and various RNA metabolism pathways (Hastings & Krainer, 2001). Both ASF/SF2 and hnRNP-L have also been shown to enhance cytoplasmic accumulation of an intronless mRNA (Liu & Mertz, 1995; Huang & Steitz, 2001). For example, hnRNP-L was reported to enhance cytoplasmic export of the herpes simplex virus thymidine kinase intronless mRNA by binding to the pre-mRNA processing enhancer (PPE; Liu & Mertz, 1995). Although SC35, ASF/SF2 and hnRNP-L have the ability to bind to the HDV RNA genome (Sikora et al., 2009), the functions of these proteins in the HDV life cycle is unknown. One possibility is that they might have a role in HDV RNA cellular localization, since HDV RNA is intronless. Furthermore, because hnRNP-L can also interact with HDAg-S (Cao et al., 2009), a complex of these two proteins might modulate HDV RNA shuttling in the cell. HDV RNA was shown to transiently co-localize with SC35 in nuclear speckles, and SC35 was reported to be binding HDV RNA during active transcription by RNAP II (Bichko & Taylor, 1996; Abrahem & Pelchat, 2008). It is thus possible that SC35 might have a role in HDV RNA transcription. The interaction of HDV RNA with these proteins might also give clues to HDV pathogenesis. Dysregulation of the splicing pathways have been associated with cell transformation and tumorogenesis (Castiglioni et al., 2006; Lee et al., 2006; Pospisil et al., 2006). ASF/SF2 is also involved in the maintenance of genome integrity. It was reported that stable RNA-DNA hybrids (R-loops), that form between nascent RNA transcripts and template DNA, accumulate in ASF/SF2-depleted cells (Li & Manley, 2005; Li et al., 2005). Although not tested, it is possible that under high HDV RNA accumulation, ASF/SF2 might be diverted from protecting chromosomal DNA, which in turn could lead to accumulation of R-loops, ultimately leading to genomic instability, a characteristic of almost all human cancers.

Two additional host proteins known to be involved in splicing were also found to bind HDV RNA: PSF and p54nrb (Greco-Stewart et al., 2006; Sikora et al., 2009). PSF is a multifunctional protein involved in many processes such as splicing, polyadenylation, transcriptional regulation, retention of defective RNAs, nucleic acid unwinding and annealing, nuclear shuttling, and pH homeostasis (Shav-Tal & Zipori, 2002). P54nrb is related to PSF, and both proteins share homology within their C-terminal portions (Dong et al., 1993). PSF forms a heterotetramer with p54nrb, and together they were shown to interact with the C-terminal domain (CTD) of RNAP II during both transcription initiation and elongation (Emili et al., 2002). It is also suggested that PSF might be able to interact with RNA and the CTD of RNAP II simultaneously (Emili et al., 2002). Using co-immunoprecipitations and electrophoretic mobility shift assays of a series of HDV-derived RNAs with purified, recombinant hexahistidine-tagged PSF, it was determined that PSF binds directly to terminal stem-loop domains of both polarities of HDV RNA at locations corresponding to the regions bound by RNAP II (Greco-Stewart et al., 2006), including a section of the HDV RNA genome reported to have RNA promoter activity (Beard et al., 1996; Abrahem & Pelchat, 2008). Recent results indicate that RNAP II interaction with HDV RNA might require both the PSF-HDV RNA interaction and the RNAP II-PSF interaction (Greco-Stewart and Pelchat, unpublished data). This suggests that PSF might provide a direct physical link between HDV RNA and RNAP II to promote initial binding of RNAP II and subsequent transcription on an HDV RNA template. The interaction of PSF with HDV RNA might also provide a mechanism for HDV pathogenesis. PSF mutations have been identified in both cervical cancer and papillary renal carcinoma cell lineages, and PSF dysregulation is associated with induction of several oncogenes (Clark et al., 1997; Benn et al., 2000; Song et al., 2004; Song et al., 2005; Li et al., 2009; Wang et al., 2009). More importantly, it was shown that noncoding RNAs can bind PSF and reverse PSF-mediated repression of several proto-oncogenes, thus promoting cell proliferation and oncogenesis (Song et al., 2004; Song et al., 2005). Thus, it is possible that the interaction of PSF with HDV RNA can reverse PSF repression and promote liver tumor development.

GAPDH was found to bind to the extremities of both polarities of the HDV RNA genome (Lin et al., 2000; Sikora et al., 2009). GAPDH is an enzyme historically known to be involved in glycolysis. In addition to this metabolic role, GAPDH is a multifunctional protein involved in apoptotic cell death, DNA proofreading, nuclear fusion, telomere maintenance, nuclear translocation, cell entry into S-phase, and hyperglycemic stress (Sirover, 2005). GAPDH might also be involved in transcriptional regulation, as it associates with Schizosaccharomyces pombe RNAP II (Mitsuzawa et al., 2005), stimulates RNAP II transcription in Xenopus laevis oocyte (Morgenegg et al., 1986), and was reported to transcriptionally activate histone H2B (Zheng et al., 2003). Although the precise function of GAPDH in HDV biology is still unknown, this protein was reported to facilitate the shuttling of HDV RNA to the nucleus, and to act as a molecular chaperone enhancing delta ribozyme activity almost two-fold (Lin et al., 2000). Interestingly, GAPDH can also bind both the 5’-UTR (+) of the hepatitis A virus and the 3’-UTR (+/-) of the human parainfluenza virus-3 (De et al., 1996; Schultz et al., 1996). These interactions were suggested to have a role in RNA-dependent viral RNA synthesis of these two viruses. Because GAPDH plays essential roles in several cellular processes, it is possible that the high level of HDV RNA occurring during infection might affect one or more GAPDH activities and contribute to liver disease.

Recently, the eukaryotic translation elongation factor 1 alpha 1 (eEF1A1) was reported to interact with both polarities of the HDV RNA genome (Sikora et al., 2009). eEF1A1 is mostly known for its role in translation where it binds and delivers aminoacyl-tRNA to the ribosome (Brands et al., 1986). Interestingly, eEF1A1 is one of the protein repetitively reported to be involved in viral RNA synthesis. eEF1A1 associates and enhances the RNA-dependent RNA polymerase activities of poliovirus, vesicular stomatitis virus, turnip yellow mosaic virus, and West Nile virus (Joshi et al., 1986; Harris et al., 1994; Blackwell & Brinton, 1997; Das et al., 1998). eEF1A1 has also been reported to be involved in the recruitment of HIV RNA to RNAP II using the HIV TAR element (Wu-Baer et al., 1996). More importantly, eEF1A1 was recently shown to interact with the peach latent mosaic viroid (Dube et al., 2009). Viroids are RNA pathogens very similar to HDV, because they have no replicase and have to rely heavily on their hosts. They are a small, single-stranded, circular RNA pathogens able to infect plants and do not encode any proteins (Flores et al., 2011). Although the function for eEF1A1 interaction with HDV RNA is still unknown, based on its involvement in viral RNA synthesis, it is tempting to speculate that this protein might enhance the RNAP II-mediated replication of HDV RNA. Interaction of HDV RNA with eEF1A1 might also give clues to HDV pathology. In addition to its main role in cellular translation through the binding of tRNAs (Brands et al., 1986), eEF1A1 also binds other highly-structured single-stranded RNA, such as the non-coding RNA HSR1. The eEF1A1-HSR1 complex activates the heat shock transcription factor 1 (HSF1), which binds to DNA elements and induces the expression of heat shock proteins. Although not studied, it is possible that HDV RNA might compete with HSR1 for eEF1A1 binding and block a heat shock response. In agreement with this hypothesis, hsp105, which belongs to the heat shock protein 70 family which is regulated by HSF1, was found to be downregulated in Huh7 cells expressing HDV mRNA (Mota et al., 2008).

Advertisement

5. Conclusion

The mechanisms by which HDV infection contributes to clinical hepatitis are poorly understood. HDV is unique among human viral pathogens in that it has a very limited coding capacity and must rely heavily on host proteins for its life cycle. Over the last few years, several host proteins have been shown to interact with either HDAgs or HDV RNA. Similar to what is observed for other RNA viruses, normal cellular components associated with RNA-processing pathways appear to be exploited by HDV. However, it is likely that several of these proteins interact indirectly with the HDV components, either through host RNA or protein interaction.

Although the normal functions of these proteins are frequently linked to viral replication and/or transcription, their precise roles in both HDV biology and HDV-mediated pathogenesis still need to be clarified. However, their interaction with the HDV RNA genome is consistent with their known biological properties. Several of these proteins are very abundant and are often referred as housekeeping proteins. Nonetheless, it is possible that the high level of HDV RNA occurring during infection could divert some of these proteins from their normal cellular functions, and produce ill effects. Further investigation on these interactions is needed in order to improve our understanding of the mechanisms of both HDV replication and pathogenesis.

Advertisement

Acknowledgments

This work was funded by a Canadian Institutes of Health Research grant to M.P.

References

  1. 1. AbrahemA.PelchatM.2008Formation of an RNA polymerase II preinitiation complex on an RNA promoter derived from the hepatitis delta virus RNA genome, Nucleic Acids Res. 3652015211
  2. 2. BeardM. R.MacNaughton. T. B.GowansE. J.1996Identification and characterization of a hepatitis delta virus RNA transcriptional promoter, J. Virol. 7049864995
  3. 3. BennD. E.DwightT.RichardsonA. L.DelbridgeL.BambachC. P.StowasserM.GordonR. D.MarshD. J.RobinsonB. G.2000Sporadic and familial pheochromocytomas are associated with loss of at least two discrete intervals on chromosome 1p, Cancer Res. 6070487051
  4. 4. BichkoV. V.TaylorJ. M.1996Redistribution of the delta antigens in cells replicating the genome of hepatitis delta virus, J. Virol. 7080648070
  5. 5. BlackwellJ. L.BrintonM. A.1997Translation elongation factor-1 alpha interacts with the 3’ stem-loop region of West Nile virus genomic RNA, J. Virol. 7164336444
  6. 6. BrandsJ. H.MaassenJ. A.van HemertF. J.AmonsR.MollerW.1986The primary structure of the alpha subunit of human elongation factor 1. Structural aspects of guanine-nucleotide-binding sites,Eur. J. Biochem. 155167171
  7. 7. BrazasR.GanemD.1996A cellular homolog of hepatitis delta antigen: implications for viral replication and evolution, Science 2749094
  8. 8. CaoD.HausseckerD.HuangY.KayM. A.2009Combined proteomic-RNAi screen for host factors involved in human hepatitis delta virus replication, RNA 1519711979
  9. 9. CaseyJ. L.BergmannK. F.BrownT. L.GerinJ. L.1992Structural requirements for RNA editing in hepatitis delta virus: evidence for a uridine-to-cytidine editing mechanism, Proc. Natl. Acad. Sci. USA 8971497153
  10. 10. CaseyJ. L.GerinJ. L.1995Hepatitis D virus RNA editing: specific modification of adenosine in the antigenomic RNA, J. Virol. 6975937600
  11. 11. CastiglioniF.TagliabueE.CampiglioM.PupaS. M.BalsariA.MenardS.2006Role of exon-16-deleted HER2 in breast carcinomas, Endocr Relat Cancer 13221232
  12. 12. ChangF. L.ChenP. J.TuS. J.WangC. J.ChenD. S.1991The large form of hepatitis delta antigen is crucial for assembly of hepatitis delta virus, Proc. Natl. Acad. Sci. USA 8884908494
  13. 13. ChangJ.GudimaS. O.TarnC.NieX.TaylorJ. M.2005Development of a novel system to study hepatitis delta virus genome replication, J Virol 7981828188
  14. 14. ChangJ.GudimaS. O.TaylorJ. M.2005Evolution of hepatitis delta virus RNA genome following long-term replication in cell culture, J Virol 791331013316
  15. 15. ChangJ.NieX.ChangH. E.HanZ.TaylorJ.2008Transcription of hepatitis delta virus RNA by RNA polymerase II, J. Virol. 8211181127
  16. 16. ChangM. F.BakerS. C.SoeL. H.KamahoraT.KeckJ. G.MakinoS.GovindarajanS.LaiM. M.1988Human hepatitis delta antigen is a nuclear phosphoprotein with RNA-binding activity, J. Virol. 6224032410
  17. 17. ChaoM.HsiehS. Y.TaylorJ.1990Role of two forms of hepatitis delta virus antigen: evidence for a mechanism of self-limiting genome replication, J. Virol. 6450665069
  18. 18. ChenP. J.KalpanaG.GoldbergJ.MasonW.WernerB.GerinJ.TaylorJ.1986Structure and replication of the genome of the hepatitis delta virus, Proc. Natl. Acad. Sci. USA 8387748778
  19. 19. ClarkJ.LuY. J.SidharS. K.ParkerC.GillS.SmedleyD.HamoudiR.LinehanW. M.ShipleyJ.CooperC. S.1997Fusion of splicing factor genes PSF and NonO (54nrbto the TFE3 gene in papillary renal cell carcinoma, Oncogene 15, 2233-2239
  20. 20. ColeS. M.GowansE. J.MacnaughtonT. B.HallP. D.BurrellC. J.1991Direct evidence for cytotoxicity associated with expression of hepatitis delta virus antigen, Hepatology 13845851
  21. 21. ColeS. M.MacnaughtonT. B.GowansE. J.1993Differential roles for HDAg-24and-p27 in HDV pathogenesis, Prog Clin Biol Res 382, 131-138
  22. 22. DasT.MathurM.GuptaA. K.JanssenG. M.BanerjeeA. K.1998RNA polymerase of vesicular stomatitis virus specifically associates with translation elongation factor-1 alphabetagamma for its activity, Proc. Natl. Acad. Sci. USA 9514491454
  23. 23. De B. P.GuptaS.ZhaoH.DrazbaJ. A.BanerjeeA. K.1996Specific interaction in vitro and in vivo of glyceraldehyde-3-phosphate dehydrogenase and LA protein with cis-acting RNAs of human parainfluenza virus type 3, J Biol Chem 2712472824735
  24. 24. de la CruzJ.KresslerD.LinderP.1999Unwinding RNA in Saccharomyces cerevisiae: DEAD-box proteins and related families, Trends Biochem Sci 24192198
  25. 25. DongB.HorowitzD. S.KobayashiR.KrainerA. R.1993Purification and cDNA cloning of HeLa cell 54nrba nuclear protein with two RNA recognition motifs and extensive homology to human splicing factor PSF and Drosophila NONA/BJ6, Nucleic Acids Res. 21, 4085-4092
  26. 26. DubeA.BisaillonM.PerreaultJ. P.2009Identification of Proteins from Prunus persica That Interact with Peach Latent Mosaic Viroid, Journal of Virology 831205712067
  27. 27. EmiliA.ShalesM.Mc CrackenS.XieW.TuckerP. W.KobayashiR.BlencoweB. J.InglesC. J.2002Splicing and transcription-associated proteins PSF and 54nrbnonO bind to the RNA polymerase II CTD, RNA 8, 1102-1111
  28. 28. FarciP.KarayiannisP.LaiM. E.MarongiuF.OrgianaG.BalestrieriA.ThomasH. C.1988Acute and chronic hepatitis delta virus infection: direct or indirect effect on hepatitis B virus replication?, J Med Virol 26279288
  29. 29. FattovichG.StroffoliniT.ZagniI.DonatoF.2004Hepatocellular carcinoma in cirrhosis: incidence and risk factors, Gastroenterology 127, S3550
  30. 30. FilipovskaJ.KonarskaM. M.2000Specific HDV RNA-templated transcription by pol II in vitro, RNA 64154
  31. 31. FloresR.GrubbD.ElleuchA.NohalesM. A.DelgadoS.GagoS.2011Rolling-circle replication of viroids, viroid-like satellite RNAs and hepatitis delta virus: Variations on a theme, RNA Biol 8,
  32. 32. FuT. B.TaylorJ.1993The RNAs of hepatitis delta virus are copied by RNA polymerase II in nuclear homogenates, J. Virol. 6769656972
  33. 33. GhisolfiL.JosephG.AmalricF.ErardM.1992The glycine-rich domain of nucleolin has an unusual supersecondary structure responsible for its RNA-helix-destabilizing properties, J. Biol. Chem. 26729552959
  34. 34. GlennJ. S.WatsonJ. A.HavelC. M.WhiteJ. M.1992Identification of a prenylation site in delta virus large antigen, Science 25613311333
  35. 35. GotoT.KatoN.Ono-NitaS. K.YoshidaH.OtsukaM.ShiratoriY.OmataM.2000Large isoform of hepatitis delta antigen activates serum response factor-associated transcription, J. Biol. Chem. 2753731137316
  36. 36. GotoT.KatoN.YoshidaH.OtsukaM.MoriyamaM.ShiratoriY.KoikeK.MatsumuraM.OmataM.2003Synergistic activation of the serum response element-dependent pathway by hepatitis B virus x protein and large-isoform hepatitis delta antigen, J. Infect. Dis. 187820828
  37. 37. Greco-StewartV. S.MironP.AbrahemA.PelchatM.2007The human RNA polymerase II interacts with the terminal stem-loop regions of the hepatitis delta virus RNA genome, Virology 3576878
  38. 38. Greco-StewartV. S.PelchatM.2010Interaction of Host Cellular Proteins with Components of the Hepatitis Delta Virus, Viruses 2189212
  39. 39. Greco-StewartV. S.SchisselE.PelchatM.2009The hepatitis delta virus RNA genome interacts with the human RNA polymerases I and III, Virology 3861215
  40. 40. Greco-StewartV. S.ThibaultC. S.PelchatM.2006Binding of the polypyrimidine tract-binding protein-associated splicing factor (PSF) to the hepatitis delta virus RNA, Virology 3563544
  41. 41. GudimaS.DingleK.WuT. T.MoraledaG.TaylorJ.1999Characterization of the 5’ ends for polyadenylated RNAs synthesized during the replication of hepatitis delta virus, J Virol 7365336539
  42. 42. GudimaS.WuS. Y.ChiangC. M.MoraledaG.TaylorJ.2000Origin of hepatitis delta virus mRNA, J. Virol. 7472047210
  43. 43. HanZ.AlvesC.GudimaS.TaylorJ.2009Intracellular localization of hepatitis delta virus proteins in the presence and absence of viral RNA accumulation, J. Virol. 8364576463
  44. 44. HarrisK. S.XiangW.AlexanderL.LaneW. S.PaulA. V.WimmerE.1994Interaction of poliovirus polypeptide 3CDpro with the 5’ and 3’ termini of the poliovirus genome. Identification of viral and cellular cofactors needed for efficient binding, J. Biol. Chem. 2692700427014
  45. 45. HastingsM. L.KrainerA. R.2001Pre-mRNA splicing in the new millennium, Curr Opin Cell Biol 13302309
  46. 46. HausseckerD.CaoD.HuangY.ParameswaranP.FireA. Z.KayM. A.2008Capped small RNAs and MOV10 in human hepatitis delta virus replication, Nat. Struct. Mol. Biol. 15714721
  47. 47. HiscoxJ. A.2002The nucleolus--a gateway to viral infection?, Arch. Virol. 14710771089
  48. 48. HuangW. H.YungB. Y.SyuW. J.LeeY. H.2001The nucleolar phosphoprotein B23 interacts with hepatitis delta antigens and modulates the hepatitis delta virus RNA replication, J. Biol. Chem. 2762516625175
  49. 49. HuangY.SteitzJ. A.2001Splicing factors SRp20 and 9G8 promote the nucleocytoplasmic export of mRNA, Mol Cell 7899905
  50. 50. JoshiR. L.RavelJ. M.HaenniA. L.1986Interaction of turnip yellow mosaic virus Val-RNA with eukaryotic elongation factor EF-1 [alpha]. Search for a function, EMBO J. 511431148
  51. 51. KrogsgaardK.KrygerP.AldershvileJ.AnderssonP.SorensenT. I.NielsenJ. O.1987Delta-infection and suppression of hepatitis B virus replication in chronic HBsAg carriers, Hepatology 74245
  52. 52. KuoM. Y.ChaoM.TaylorJ.1989Initiation of replication of the human hepatitis delta virus genome from cloned DNA: role of delta antigen, J. Virol. 6319451950
  53. 53. KuoM. Y.GoldbergJ.CoatesL.MasonW.GerinJ.TaylorJ.1988Molecular cloning of hepatitis delta virus RNA from an infected woodchuck liver: sequence, structure, and applications, J. Virol. 6218551861
  54. 54. LaiM. M.1998Cellular factors in the transcription and replication of viral RNA genomes: a parallel to DNA-dependent RNA transcription, Virology 244112
  55. 55. LaiM. M.2005RNA replication without RNA-dependent RNA polymerase: surprises from hepatitis delta virus, J. Virol. 7979517958
  56. 56. LeeC. H.ChangS. C.ChenC. J.ChangM. F.1998The nucleolin binding activity of hepatitis delta antigen is associated with nucleolus targeting, J. Biol. Chem. 27376507656
  57. 57. LeeC. Z.ChenP. J.ChenD. S.1995Large hepatitis delta antigen in packaging and replication inhibition: role of the carboxyl-terminal 19 amino acids and amino-terminal sequences, J. Virol. 6953325336
  58. 58. LeeE. J.JoM.ParkJ.ZhangW.LeeJ. H.2006Alternative splicing variants of IRF-1 lacking exons 7, 8, and 9 in cervical cancer, Biochem Biophys Res Commun 347882888
  59. 59. LehmannE.BruecknerF.CramerP.2007Molecular basis of RNA-dependent RNA polymerase II activity, Nature 450445449
  60. 60. LiL.FengT.LianY.ZhangG.GarenA.SongX.2009Role of human noncoding RNAs in the control of tumorigenesis, Proc. Natl. Acad. Sci. USA 1061295612961
  61. 61. LiX.ManleyJ. L.2005Inactivation of the SR protein splicing factor ASF/SF2 results in genomic instability, Cell 122365378
  62. 62. LiX.WangJ.ManleyJ. L.2005Loss of splicing factor ASF/SF2 induces G2 cell cycle arrest and apoptosis, but inhibits internucleosomal DNA fragmentation, Genes Dev 1927052714
  63. 63. LiY. J.MacnaughtonT.GaoL.LaiM. M.2006RNA-templated replication of hepatitis delta virus: genomic and antigenomic RNAs associate with different nuclear bodies, J. Virol. 8064786486
  64. 64. LiY. J.StallcupM. R.LaiM. M.2004Hepatitis delta virus antigen is methylated at arginine residues, and methylation regulates subcellular localization and RNA replication, J. Virol. 781332513334
  65. 65. LinS. S.ChangS. C.WangY. H.SunC. Y.ChangM. F.2000Specific interaction between the hepatitis delta virus RNA and glyceraldehyde 3-phosphate dehydrogenase: an enhancement on ribozyme catalysis, Virology 2714657
  66. 66. LiuX.MertzJ. E.1995HnRNP L binds a cis-acting RNA sequence element that enables intron-dependent gene expression, Genes Dev 917661780
  67. 67. LoK.SheuG. T.LaiM. M.1998Inhibition of Cellular RNA polymerase II transcription by delta antigen of hepatitis delta virus, Virology 247178188
  68. 68. MacNaughton. T. B.GowansE. J.Mc NamaraS. P.BurrellC. J.1991Hepatitis delta antigen is necessary for access of hepatitis delta virus RNA to the cell transcriptional machinery but is not part of the transcriptional complex, Virology 184387390
  69. 69. MacnaughtonT. B.ShiS. T.ModahlL. E.LaiM. M.2002Rolling circle replication of hepatitis delta virus RNA is carried out by two different cellular RNA polymerases, J. Virol. 7639203927
  70. 70. MitsuzawaH.KimuraM.KandaE.IshihamaA.2005Glyceraldehyde-3-phosphate dehydrogenase and actin associate with RNA polymerase II and interact with its Rpb7 subunit, FEBS Lett 5794852
  71. 71. ModahlL. E.MacnaughtonT. B.ZhuN.JohnsonD. L.LaiM. M.2000RNA-Dependent replication and transcription of hepatitis delta virus RNA involve distinct cellular RNA polymerases, Mol. Cell. Biol. 2060306039
  72. 72. MoraledaG.TaylorJ.2001Host RNA polymerase requirements for transcription of the human hepatitis delta virus genome, J Virol 751016110169
  73. 73. MorgeneggG.WinklerG. C.HubscherU.HeizmannC. W.MousJ.KuenzleC. C.1986Glyceraldehyde-3-phosphate dehydrogenase is a nonhistone protein and a possible activator of transcription in neurons, J Neurochem 475462
  74. 74. MotaS.MendesM.FreitasN.PenqueD.CoelhoA. V.CunhaC.2009Proteome analysis of a human liver carcinoma cell line stably expressing hepatitis delta virus ribonucleoproteins, J Proteomics 72616627
  75. 75. MotaS.MendesM.PenqueD.CoelhoA. V.CunhaC.2008Changes in the proteome of Huh7 cells induced by transient expression of hepatitis D virus RNA and antigens, J. Proteomics 717179
  76. 76. MuJ. J.ChenD. S.ChenP. J.2001The conserved serine 177 in the delta antigen of hepatitis delta virus is one putative phosphorylation site and is required for efficient viral RNA replication, J. Virol. 7590879095
  77. 77. MuJ. J.TsayY. G.JuanL. J.FuT. F.HuangW. H.ChenD. S.ChenP. J.2004The small delta antigen of hepatitis delta virus is an acetylated protein and acetylation of lysine 72 may influence its cellular localization and viral RNA synthesis, Virology 3196070
  78. 78. MuJ. J.WuH. L.ChiangB. L.ChangR. P.ChenD. S.ChenP. J.1999Characterization of the phosphorylated forms and the phosphorylated residues of hepatitis delta virus delta antigens, J. Virol. 731054010545
  79. 79. NiranjanakumariS.LasdaE.BrazasR.Garcia-BlancoM. A.2002Reversible cross-linking combined with immunoprecipitation to study RNA-protein interactions in vivo, Methods 26182190
  80. 80. OkuwakiM.2008The structure and functions of NPM1/Nucleophsmin/B23, a multifunctional nucleolar acidic protein, J. Biochem. 143441448
  81. 81. OttoJ. C.CaseyP. J.1996The hepatitis delta virus large antigen is farnesylated both in vitro and in animal cells, J. Biol. Chem. 27145694572
  82. 82. PascarellaS.NegroF.2011Hepatitis D virus: an update, Liver Int 31721
  83. 83. PolsonA. G.LeyH. L.3rd BassB. L.CaseyJ. L.1998Hepatitis delta virus RNA editing is highly specific for the amber/W site and is suppressed by hepatitis delta antigen, Mol. Cell. Biol. 1819191926
  84. 84. PospisilH.HerrmannA.ButherusK.PirsonS.ReichJ. G.KemmnerW.2006Verification of predicted alternatively spliced Wnt genes reveals two new splice variants (CTNNB1 and LRP5) and altered Axin-1 expression during tumour progression, BMC Genomics 7, 148
  85. 85. RizzettoM.CaneseM. G.AricoS.CrivelliO.TrepoC.BoninoF.VermeG.1977Immunofluorescence detection of new antigen-antibody system (delta/anti-delta) associated to hepatitis B virus in liver and in serum of HBsAg carriers, Gut 189971003
  86. 86. RomeoR.Del NinnoE.RumiM.RussoA.SangiovanniA.de FranchisR.RonchiG.ColomboM.2009A 28year Study of the Course of Hepatitis Delta Infection, a Risk Factor for Cirrhosis and Hepatocellular Carcinoma, Gastroenterology
  87. 87. RyuW. S.BayerM.TaylorJ.1992Assembly of hepatitis delta virus particles, J. Virol. 6623102315
  88. 88. RyuW. S.NetterH. J.BayerM.TaylorJ.1993Ribonucleoprotein complexes of hepatitis delta virus, J. Virol. 6732813287
  89. 89. SchultzD. E.HardinC. C.LemonS. M.1996Specific interaction of glyceraldehyde 3-phosphate dehydrogenase with the 5’-nontranslated RNA of hepatitis A virus, J Biol Chem 2711413414142
  90. 90. Shav-TalY.ZiporiD.2002PSF and 54nrb)/NonO--multi-functional nuclear proteins, FEBS Lett. 531, 109-114
  91. 91. SikoraD.Greco-StewartV. S.MironP.PelchatM.2009The hepatitis delta virus RNA genome interacts with eEF1A1, 54nrb), hnRNP-L, GAPDH and ASF/SF2, Virology 390, 71-78
  92. 92. SiroverM. A.2005New nuclear functions of the glycolytic protein, glyceraldehyde-3-phosphate dehydrogenase, in mammalian cells, J. Cell. Biochem. 954552
  93. 93. SongX.SuiA.GarenA.2004Binding of mouse VL30 retrotransposon RNA to PSF protein induces genes repressed by PSF: effects on steroidogenesis and oncogenesis, Proc Natl Acad Sci U S A 101621626
  94. 94. SongX.SunY.GarenA.2005Roles of PSF protein and VL30 RNA in reversible gene regulation, Proc Natl Acad Sci U S A 1021218912193
  95. 95. SuC. W.HuangY. H.HuoT. I.ShihH. H.SheenI. J.ChenS. W.LeeP. C.LeeS. D.WuJ. C.2006Genotypes and viremia of hepatitis B and D viruses are associated with outcomes of chronic hepatitis D patients, Gastroenterology 13016251635
  96. 96. SureauC.MoriartyA. M.ThorntonG. B.LanfordR. E.1992Production of infectious hepatitis delta virus in vitro and neutralization with antibodies directed against hepatitis B virus pre-S antigens, J. Virol. 6612411245
  97. 97. TaylorJ. M.2006Hepatitis delta virus, Virology 3447176
  98. 98. TaylorJ. M.2009Chapter 3. Replication of the hepatitis delta virus RNA genome, Adv. Virus Res. 74103121
  99. 99. TsengC. H.LaiM. C.2009Hepatitis Delta Virus RNA Replication, Viruses 1818831
  100. 100. TutejaR.TutejaN.1998Nucleolin: a multifunctional major nucleolar phosphoprotein, Crit. Rev. Biochem. Mol. Biol. 33407436
  101. 101. WangC. C.ChangT. C.LinC. W.TsuiH. L.ChuP. B.ChenB. S.HuangZ. S.WuH. N.2003Nucleic acid binding properties of the nucleic acid chaperone domain of hepatitis delta antigen, Nucleic Acids Res. 3164816492
  102. 102. WangD.PearlbergJ.LiuY. T.GanemD.2001Deleterious effects of hepatitis delta virus replication on host cell proliferation, J Virol 7536003604
  103. 103. WangG.CuiY.ZhangG.GarenA.SongX.2009Regulation of proto-oncogene transcription, cell proliferation, and tumorigenesis in mice by PSF protein and a VL30 noncoding RNA, Proc. Natl. Acad. Sci. USA 1061679416798
  104. 104. WeiY.GanemD.1998Activation of heterologous gene expression by the large isoform of hepatitis delta antigen, J. Virol. 7220892096
  105. 105. WeinerA. J.ChooQ. L.WangK. S.GovindarajanS.RedekerA. G.GerinJ. L.HoughtonM.1988A single antigenomic open reading frame of the hepatitis delta virus encodes the epitope(s) of both hepatitis delta antigen polypeptides 24delta and p27 delta, J. Virol. 62, 594-599
  106. 106. WenX.TannukitS.PaineM. L.2008TFIP11 interacts with mDEAH9, an RNA helicase involved in spliceosome disassembly, Int J Mol Sci 921052113
  107. 107. WongS. K.LazinskiD. W.2002Replicating hepatitis delta virus RNA is edited in the nucleus by the small form of ADAR1, Proc. Natl. Acad. Sci. USA 991511815123
  108. 108. Wu-BaerF.LaneW. S.GaynorR. B.1996Identification of a group of cellular cofactors that stimulate the binding of RNA polymerase II and TRP-185 to human immunodeficiency virus 1 TAR RNA, J. Biol. Chem. 27142014208
  109. 109. YamaguchiY.FilipovskaJ.YanoK.FuruyaA.InukaiN.NaritaT.WadaT.SugimotoS.KonarskaM. M.HandaH.2001Stimulation of RNA polymerase II elongation by hepatitis delta antigen, Science 293124127
  110. 110. YamaguchiY.MuraT.ChanaratS.OkamotoS.HandaH.2007Hepatitis delta antigen binds to the clamp of RNA polymerase II and affects transcriptional fidelity, Genes Cells 12863875
  111. 111. YudkovskyN.RanishJ. A.HahnS.2000A transcription reinitiation intermediate that is stabilized by activator, Nature 408225229
  112. 112. ZhengL.RoederR. G.LuoY.2003S phase activation of the histone H2B promoter by OCA-S, a coactivator complex that contains GAPDH as a key component, Cell 114255266

Written By

Yasnee Beeharry and Martin Pelchat

Submitted: 12 November 2010 Published: 29 August 2011