首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 46 毫秒
1.
Replication of hepatitis C virus (HCV) RNA occurs on intracellular membranes, and the replication complex (RC) contains viral RNA, nonstructural proteins, and cellular cofactors. We previously demonstrated that cyclophilin A (CyPA) is an essential cofactor for HCV infection and the intracellular target of cyclosporine''s anti-HCV effect. Here we investigate the mechanism by which CyPA facilitates HCV replication. Cyclosporine treatment specifically blocked the incorporation of NS5B into the RC without affecting either the total protein level or the membrane association of the protein. Other nonstructural proteins or viral RNAs in the RC were not affected. NS5B from the cyclosporine-resistant replicon was resistant to this disruption of RC incorporation. We also isolated membrane fractions from both naïve and HCV-positive cells and found that CyPA is recruited into membrane fractions in HCV-replicating cells via an interaction with RC-associated NS5B, which is sensitive to cyclosporine treatment. Finally, we introduced point mutations in the prolyl-peptidyl isomerase (PPIase) motif of CyPA and demonstrated a critical role of this motif in HCV replication in cDNA rescue experiments. We propose a model in which the incorporation of the HCV polymerase into the RC depends on its interaction with a cellular chaperone protein and in which cyclosporine inhibits HCV replication by blocking this critical interaction and the PPIase activity of CyPA. Our results provide a mechanism of action for the cyclosporine-mediated inhibition of HCV and identify a critical role of CyPA''s PPIase activity in the proper assembly and function of the HCV RC.Hepatitis C virus (HCV), of the family Flaviviridae, is an enveloped, positive-stranded RNA virus. Spread mostly by blood-borne transmission, HCV infects more than 170 million people worldwide. The viral genome is composed of a single open reading frame (ORF) plus 5′- and 3′-nontranslated regions. The ORF encodes a large polyprotein that is cleaved by cellular and viral proteases into 10 viral proteins. The structural proteins, including the capsid protein (core), two glycoproteins (E1 and E2), and a small ion channel protein (p7), reside in the N-terminal half of the polyprotein. The rest of the ORF encodes six nonstructural (NS) proteins: NS2, NS3, NS4A, NS4B, NS5A, and NS5B. NS3 through NS5B assemble into a replication complex (RC) and are necessary and sufficient for HCV RNA replication in cell culture (8, 42). NS3 is a multifunctional protein with both a serine protease and an RNA helicase activity. The protease activity is responsible for cleavage at the NS3-NS4A, NS4A-NS4B, NS4B-NS5A, and NS5A-NS5B junctions (5), and the helicase activity is probably required to unwind the double-stranded RNA intermediates formed during replication (38). NS4A serves as an essential cofactor for the NS3 protease and anchors the NS3 protein to intracellular membranes (25, 36, 39). NS4B induces the formation of a “membranous web” that is probably the site of HCV replication (16). It also contains a GTP-binding motif that is required for replication (17). The web is derived from the endoplasmic reticulum (ER) compartment, although proteins of early-endosome origin have also been found to locate to the web (62). NS5A is a phosphoprotein and an integral component of the viral RC. The precise function of NS5A in replication is still unknown but appears to be regulated by phosphorylation and its interaction with several cellular proteins (19, 22, 24, 51, 52, 59, 63, 67). In addition, it may be involved in the transition from replication and particle formation (4, 45, 64). NS5B is the RNA-dependent RNA polymerase that is responsible for copying the RNA genome of the virus during replication. Several cellular cofactors interact with NS5B and modulate its activity in the context of the viral RC (22, 24, 35, 69, 71).Positive-stranded RNA viruses alter the intracellular membranes of host cells to form an RC in which RNA replication occurs. Modifications include the proliferation and reorganization of certain cellular membranes (1). HCV forms an RC associated with altered cellular membranes (16, 23), and crude RCs (CRCs) that maintain the replicase activity in vitro can be isolated by membrane sedimentation or flotation techniques (2, 3, 18, 27, 37).Cyclosporine is a widely used immunosuppressive and anti-inflammatory drug for organ transplant patients. It functions by forming an inhibitory complex with cyclophilins (CyPs) that inhibits the phosphatase activity of calcineurin, which is important for T-cell activation. In recent years, cyclosporine and its derivatives have been shown to be highly effective in suppressing HCV replication in vitro (44, 49, 53, 68) and in vivo (30). The mechanism of this inhibition is independent of its immunosuppressive function and distinct from that of interferon (IFN) (44, 53, 56, 68).We recently showed that HCV infection in vitro is inhibited when CyPA, a major intracellular target of cyclosporine, is downregulated by RNA interference, and mutations in NS5B that confer cyclosporine-resistant binding to CyPA contribute to the cyclosporine resistance of the replicons harboring these mutations (56, 71). Here we report that CyPA is recruited into the HCV RC together with NS5B in HCV replicon or in HCV-infected cells. Cyclosporine disrupts the association between RC-incorporated NS5B and CyPA and results in an exclusion of the polymerase from the viral RC. We also show that the prolyl-peptidyl isomerase (PPIase) motif of CyPA is essential for HCV replication.  相似文献   

2.
Studies of the hepatitis C virus (HCV) life cycle have been aided by development of in vitro systems that enable replication of viral RNA and production of infectious virus. However, the functions of the individual proteins, especially those engaged in RNA replication, remain poorly understood. It is considered that NS4B, one of the replicase components, creates sites for genome synthesis, which appear as punctate foci at the endoplasmic reticulum (ER) membrane. In this study, a panel of mutations in NS4B was generated to gain deeper insight into its functions. Our analysis identified five mutants that were incapable of supporting RNA replication, three of which had defects in production of foci at the ER membrane. These mutants also influenced posttranslational modification and intracellular mobility of another replicase protein, NS5A, suggesting that such characteristics are linked to focus formation by NS4B. From previous studies, NS4B could not be trans-complemented in replication assays. Using the mutants that blocked RNA synthesis, defective NS4B expressed from two mutants could be rescued in trans-complementation replication assays by wild-type protein produced by a functional HCV replicon. Moreover, active replication could be reconstituted by combining replicons that were defective in NS4B and NS5A. The ability to restore replication from inactive replicons has implications for our understanding of the mechanisms that direct viral RNA synthesis. Finally, one of the NS4B mutations increased the yield of infectious virus by five- to sixfold. Hence, NS4B not only functions in RNA replication but also contributes to the processes engaged in virus assembly and release.Recent estimates predict that the prevalence of hepatitis C virus (HCV) infection is approximately 2.2% worldwide, equivalent to about 130 million persons (22). The virus typically establishes a chronic infection that frequently leads to serious liver disease (1), and current models indicate that both morbidity and mortality as a consequence of HCV infection will continue to rise for about the next 20 years (10, 11, 29).HCV is the only assigned species of the Hepacivirus genus within the family Flaviviridae. The virus can be classified into six genetic groups or clades (numbered 1 to 6) and then further separated into subtypes (e.g., 1a, 1b, 2a, 2b, etc.) (53, 55). HCV has a single-stranded, positive-sense RNA genome that is approximately 9.6 kb in length (reviewed in reference 46). Genomic RNA carries a single open reading frame flanked by 5′ and 3′ nontranslated regions, which are important for both replication and translation (19, 20, 34, 47, 56). Viral RNA is translated by the host ribosomal machinery, and the resultant polyprotein is co- and posttranslationally cleaved to generate the mature viral proteins. The structural proteins (core, E1, and E2) and a small hydrophobic polypeptide called p7 are produced by the cellular proteases signal peptidase and signal peptide peptidase (28, 45, 54). Two virus-encoded proteases, the NS2-3 autoprotease and the NS3 serine protease (5, 13, 26), are responsible for maturation of the nonstructural (NS) proteins (NS2, NS3, NS4A, NS4B, NS5A, and NS5B). With the exception of NS2, the NS proteins are necessary for genome replication (8, 40) and form replication complexes (RCs), which are located at the endoplasmic reticulum (ER) membrane (14, 24, 52, 57, 59). The functions of all viral constituents of RCs have not been characterized in detail. It is known that NS5B is the RNA-dependent RNA polymerase (6), while NS3 possesses helicase and nucleoside triphosphatase activities in addition to acting as a protease (32, 58). However, the precise roles of the other proteins remain to be firmly established.Expression of NS4B, one of the replicase proteins, generates rearrangements at the ER membrane that have been termed the “membranous web” (14, 24) and “membrane-associated foci” (MAFs) (25). Detection of viral RNA at such foci suggests that NS4B is involved in creating the sites where genome synthesis occurs (18, 24, 59). It is predicted that NS4B has an amphipathic α-helix within its N-terminal region, which is followed by four transmembrane domains (TMDs) in the central portion of the protein (17, 42). As a result, the majority of NS4B is likely to be tightly anchored to membranes, and experimental evidence indicates that it has characteristics consistent with an integral membrane protein (27). It is thought that after membrane association, NS4B rearranges membranes into a network, thereby generating foci which act as a “scaffold” to facilitate RNA replication. The mechanisms engaged in formation of foci are not known but include the notion that the NS4B N terminus can translocate into the ER lumen, resulting in rearrangement of cellular membranes (41, 42). Alternatively, palmitoylation, a lipid modification, might facilitate polymerization of NS4B, in turn promoting formation of RCs on the ER membrane (68).Apart from inducing membranous changes required for replication, NS4B may perform other tasks in HCV RNA synthesis. For example, studies of cell culture adaptive mutations in subgenomic replicons (SGRs) have identified amino acid changes that can stimulate RNA production (39), suggesting that NS4B may exert a regulatory role in determining replication efficiency. In support of a regulatory function, replacement of NS4B sequences in an SGR from strain H77 (a genotype 1a strain) with those from strain Con-1 (a genotype 1b strain) gave higher levels of replication than for a wild-type (wt) strain H77 SGR (7). The corresponding replacement of strain Con-1 NS4B sequences with those from strain H77 reduced the replication efficiency of a Con-1 SGR (7). Moreover, interactions of NS4B with the RC can affect the behavior of other replicase proteins. For example, NS4B is needed for hyperphosphorylation of NS5A (35, 48) and restricts its intracellular movement (30).To try to gain greater insight into the functional organization of the components that constitute RCs, trans-complementation assays using defective and helper SGRs have been established (2, 64). Such studies reveal that the only protein capable of trans-complementation is NS5A, while active replication cannot be restored for replicons harboring deleterious mutations in NS3, NS4B, and NS5B. These data led to the conclusion that functional NS5A may be able to exchange between RCs (2), whereas, by inference, such exchange would not be possible for other HCV replicase proteins. In transient-replication assays, complementation by NS5A also relied on its expression as part of a polyprotein (minimally NS3-NS5A), and production of the protein alone failed to restore replication for an inactive SGR (2). However, in a separate study, stable expression of wt NS5A was capable of complementing a defective replicon (64). Thus, different assay systems can give dissimilar results for complementation by NS5A.In this study, we have created a series of mutations in the NS4B gene of HCV strain JFH1 (31) to explore the function of the protein in the HCV life cycle. We focused our attention on the C-terminal portion of NS4B, downstream from the predicted TMD regions, since it is relatively well conserved and is predicted to lie on the cytosolic side of the ER membrane (15, 42). Our analysis examines the impact of mutations on replication efficiency and the intracellular characteristics of the mutants compared to the behavior of the wt protein. In addition, we have utilized this series of mutants to reassess trans-complementation of NS4B in replication assays. Finally, we also analyze the impact of mutations which do not affect replication on the production of infectious virus to determine whether NS4B plays a role in virus assembly and release.  相似文献   

3.
Persistent infection with hepatitis C virus (HCV) is a major cause of chronic liver diseases. The aim of this study was to identify host cell factor(s) participating in the HCV replication complex (RC) and to clarify the regulatory mechanisms of viral genome replication dependent on the host-derived factor(s) identified. By comparative proteome analysis of RC-rich membrane fractions and subsequent gene silencing mediated by RNA interference, we identified several candidates for RC components involved in HCV replication. We found that one of these candidates, creatine kinase B (CKB), a key ATP-generating enzyme that regulates ATP in subcellular compartments of nonmuscle cells, is important for efficient replication of the HCV genome and propagation of infectious virus. CKB interacts with HCV NS4A protein and forms a complex with NS3-4A, which possesses multiple enzyme activities. CKB upregulates both NS3-4A-mediated unwinding of RNA and DNA in vitro and replicase activity in permeabilized HCV replicating cells. Our results support a model in which recruitment of CKB to the HCV RC compartment, which has high and fluctuating energy demands, through its interaction with NS4A is important for efficient replication of the viral genome. The CKB-NS4A association is a potential target for the development of a new type of antiviral therapeutic strategy.Hepatitis C virus (HCV) infection represents a significant global healthcare burden, and current estimates suggest that a minimum of 3% of the world''s population is chronically infected (4, 19). The virus is responsible for many cases of severe chronic liver diseases, including cirrhosis and hepatocellular carcinoma (4, 16, 19). HCV is a positive-stranded RNA virus belonging to the family Flaviviridae. Its ∼9.6-kb genome is translated into a single polypeptide of about 3,000 amino acids (aa), in which the nonstructural (NS) proteins NS2, NS3, NS4A, NS4B, NS5A, and NS5B reside in the C-terminal half region (6, 34, 44). NS4A, a small 7-kDa protein, functions as a cofactor for NS3 to enhance NS3 enzyme activities such as serine protease and helicase activities. The hydrophobic N-terminal region of NS4A, which is predicted to form a transmembrane α-helix, is responsible for membrane anchorage of the NS3-4A complex (8, 44, 50), and the central region of NS4A is important for the interaction with NS3 (10, 44). A recent study demonstrated the involvement of the C terminus of NS4A in the regulation of NS5A hyperphosphorylation and viral replication (28).The development of HCV replicon technology several years ago accelerated research on viral RNA replication (7, 44). Furthermore, a robust cell culture system for propagation of infectious HCV particles was developed using a viral genome of HCV genotype 2a, JFH-1 strain, enabling us to study every process in the viral life cycle (27, 47, 54). RNA derived from genotype 1a, HCV H77, containing cell-culture adaptive mutations, also produces infectious viruses (52). Using these systems, it has been reported that the HCV genome replicates in a distinct, subcellular replication complex (RC) compartment, which includes NS3-5B and the viral RNA (2, 14, 33). The RC forms in a distinct compartment with high concentrations of viral and cellular components located on detergent-resistant membrane (DRM) structures, possibly a lipid-raft structure (2, 41), which may protect the RC from external proteases and nucleases. Almost all processes in viral replication are dependent on the host cell''s machinery and involve intimate interaction between viral and host proteins. However, the functional roles of host factors interacting with the HCV RC in viral genome replication remain ambiguous.To gain a better understanding of cellular factors that are components of the HCV RC and that function as regulators of viral replication, a comparative proteomic analysis of DRM fractions from HCV replicon and parental cells and subsequent RNA interference (RNAi) silencing of selected genes were performed. We identified creatine kinase B (CKB) as a key factor for the HCV genome replication. CKB catalyzes the reversible transfer of the phosphate group of phosphocreatine (pCr) to ADP to yield ATP and creatine and is known to play important roles in local delivery and cellular compartmentalization of ATP (48, 51). The findings obtained here suggest that recruitment of CKB to the HCV RC, through CKB interaction with NS4A, is essential for maintenance or enhancement of viral replicase activity.  相似文献   

4.
5.
Hepatitis C virus (HCV) replication involves many viral and host factors. Here, we employed a lentivirus-based RNA interference (RNAi) screening approach to search for possible cellular factors. By using a kinase-phosphatase RNAi library and an HCV replicon reporter system, we identified a serine-threonine kinase, Polo-like kinase 1 (Plk1), as a potential host factor regulating HCV replication. Knockdown of Plk1 reduced both HCV RNA replication and nonstructural (NS) protein production in both HCV replicon cells and HCV-infected cells while it did not significantly affect host cellular growth or cell cycle. Overexpression of Plk1 in the knockdown cells rescued HCV replication. Interestingly, the ratio between the hyperphosphorylated form (p58) and the basal phosphorylated form (p56) of NS5A was lower in the Plk1 knockdown cells and Plk1 kinase inhibitor-treated cells than in the control groups. Further studies showed that Plk1 could be immunoprecipitated together with NS5A. Both proteins partially colocalized in the perinuclear region. Furthermore, Plk1 could phosphorylate NS5A to both the p58 and p56 forms in an in vitro assay system; the phosphorylation efficiency was comparable to that of the reported casein kinase. Taken together, this study shows that Plk1 is an NS5A phosphokinase and thereby indirectly regulates HCV RNA replication. Because of the differential effects of Plk1 on HCV replication and host cell growth, Plk1 could potentially serve as a target for anti-HCV therapy.Hepatitis C virus (HCV) is the major causative agent of non-A/non-B hepatitis (26). More than 170 million people, or 3% of the population in the world, are infected with HCV (29). It establishes chronic infection in at least 85% of infected individuals and is associated with liver cirrhosis and hepatocellular carcinoma. Current treatment, which combines polyethylene glycol-interferon (PEG-IFN) and ribavirin, is ineffective in 22% of patients with non-genotype 1 and in 45% of patients with genotype 1 HCV (1, 16, 23, 55). Therefore, identification of new targets for HCV therapy is an important issue, and cellular genes involved in the HCV life cycle may serve as good candidates.HCV is a positive-strand RNA virus and the only known member of Hepacivirus genus in the family Flaviviridae. Its genome has a length of about 9,600 nucleotides coding for a single polyprotein. The long polyprotein is further processed into at least 10 different products, including four structural proteins (core, E1, E2, and p7) and six nonstructural (NS) proteins (NS2, NS3, NS4A, NS4B, NS5A, and NS5B). Nonstructural proteins NS3-NS5B are components of the membrane-associated HCV replication complex (8, 13, 36, 45). NS3 is a bifunctional protein containing an N-terminal protease domain and a C-terminal helicase/NTPase domain, and NS4A serves as a cofactor for NS3 protease. NS4B protein is known to induce intracellular membrane changes that probably serve as the site for viral RNA replication (8). NS5A is required for RNA replication, but little is known about its function. NS5B is the RNA-dependent RNA polymerase (reviewed in reference 47).NS5A is phosphorylated on multiple serine and threonine residues and exists in basal phosphorylated (p56) and hyperphosphorylated (p58) forms (49). Increasing evidence suggests that the regulation of NS5A phosphorylation is important for HCV RNA replication. Adaptive mutations or kinase inhibitors, which reduce NS5A hyperphosphorylation, increased the replication of an HCV replicon in cell culture (HCVcc) systems (2, 4, 38). However, when an adaptive replicon with reduced p58 was further treated with the same kinase inhibitor or introduced with a second adaptive mutation, RNA replication was completely blocked (32, 38). Furthermore, the mutations that reduce NS5A hyperphosphorylation and promote RNA replication in cell culture, paradoxically, prevented productive replication in the chimpanzee model (6). These results imply that the tight control of the p58/p56 ratio is important for HCV replication. The detailed mechanism is still not clear, but a clue was provided by the finding of differential association of NS5A phospho-forms with the host vesicle-associated membrane protein-associated protein A (VAP-A) protein, which is an essential molecule for HCV replicase (9, 12). On the other hand, NS5A phosphorylation was recently found to regulate the production of infectious virus (34, 50). Alanine substitutions in the C-terminal domain III of NS5A impaired NS5A phosphorylation, leading to a decrease in NS5A-core protein interaction, disturbance of subcellular localization of NS5A, and disruption of virion production (3, 34, 50). In summary, phosphorylation on NS5A is not only important for HCV RNA replication but also critical for infectious virus production.Since the phosphorylation state of NS5A is correlated with HCV RNA replication and virion production, cellular kinases responsible for NS5A phosphorylation may serve as good candidates for drug targets. Several kinases have been shown to target NS5A in vitro, including casein kinase I (CKI), CKII, MEK1, MKK6, MKK7, AKT, and p70S6K (7, 24). Among these proteins, CKI and CKII are better characterized for NS5A phosphorylation. CKIα has been identified as the target of kinase inhibitors which decrease the hyperphosphorylation of NS5A and was further confirmed as a direct kinase of NS5A (41, 42). CKI requires prephosphorylation of residues near the predicted phosphorylation site in NS5A for effective modification, suggesting that other kinases are also involved in this process (42). CKII has been shown to bind to the C-terminal domain of NS5A and phosphorylate NS5A in vitro (24). Inhibition of CKII with chemical compounds or small interfering RNA (siRNA) did not significantly affect HCV RNA replication but severely disrupted virus production (50).In this study, using lentivirus-based RNA interference (RNAi) screening, we identified a serine/threonine kinase, Polo-like kinase 1 (Plk1), which is involved in HCV replication. Expression of short hairpin RNAs (shRNAs) targeting Plk1 decreased HCV replication and virus production. Moreover, silencing of Plk1 decreased the hyperphosphorylated form of NS5A. In cells treated with a Plk1-specific kinase inhibitor, HCV replication and NS5A hyperphosphorylation were significantly reduced, indicating that Plk1 kinase activity is required for this process. Further studies showed that Plk1 was coimmunoprecipitated and partially colocalized with NS5A, suggesting NS5A as a possible substrate for Plk1. Finally, NS5A is hyperphosphorylated by Plk1 in vitro, supporting the proposition that Plk1 regulates HCV replication through hyperphosphorylation of NS5A.  相似文献   

6.
Cell culture-adaptive mutations within the hepatitis C virus (HCV) E2 glycoprotein have been widely reported. We identify here a single mutation (N415D) in E2 that arose during long-term passaging of HCV strain JFH1-infected cells. This mutation was located within E2 residues 412 to 423, a highly conserved region that is recognized by several broadly neutralizing antibodies, including the mouse monoclonal antibody (MAb) AP33. Introduction of N415D into the wild-type (WT) JFH1 genome increased the affinity of E2 to the CD81 receptor and made the virus less sensitive to neutralization by an antiserum to another essential entry factor, SR-BI. Unlike JFH1WT, the JFH1N415D was not neutralized by AP33. In contrast, it was highly sensitive to neutralization by patient-derived antibodies, suggesting an increased availability of other neutralizing epitopes on the virus particle. We included in this analysis viruses carrying four other single mutations located within this conserved E2 region: T416A, N417S, and I422L were cell culture-adaptive mutations reported previously, while G418D was generated here by growing JFH1WT under MAb AP33 selective pressure. MAb AP33 neutralized JFH1T416A and JFH1I422L more efficiently than the WT virus, while neutralization of JFH1N417S and JFH1G418D was abrogated. The properties of all of these viruses in terms of receptor reactivity and neutralization by human antibodies were similar to JFH1N415D, highlighting the importance of the E2 412-423 region in virus entry.Hepatitis C virus (HCV), which belongs to the Flaviviridae family, has a positive-sense single-stranded RNA genome encoding a polyprotein that is cleaved by cellular and viral proteases to yield mature structural and nonstructural proteins. The structural proteins consist of core, E1 and E2, while the nonstructural proteins are p7, NS2, NS3, NS4A, NS4B, NS5A, and NS5B (42). The hepatitis C virion comprises the RNA genome surrounded by the structural proteins core (nucleocapsid) and E1 and E2 (envelope glycoproteins). The HCV glycoproteins lie within a lipid envelope surrounding the nucleocapsid and play a major role in HCV entry into host cells (21). The development of retrovirus-based HCV pseudoparticles (HCVpp) (3) and the cell culture infectious clone JFH1 (HCVcc) (61) has provided powerful tools to study HCV entry.HCV entry is initiated by the binding of virus particles to attachment factors which are believed to be glycosaminoglycans (2), low-density lipoprotein receptor (41), and C-type lectins such as DC-SIGN and L-SIGN (12, 37, 38). Upon attachment at least four entry factors are important for particle internalization. These include CD81 (50), SR-BI (53) and the tight junction proteins claudin-1 (15) and occludin (6, 36, 51).CD81, a member of the tetraspanin family, is a cell surface protein with various functions including tissue differentiation, cell-cell adhesion and immune cell maturation (34). It consists of a small and a large extracellular loop (LEL) with four transmembrane domains. Viral entry is dependent on HCV E2 binding to the LEL of CD81 (3, 50). The importance of HCV glycoprotein interaction with CD81 is underlined by the fact that many neutralizing antibodies compete with CD81 and act in a CD81-blocking manner (1, 5, 20, 45).SR-BI is a multiligand receptor expressed on liver cells and on steroidogenic tissue. It binds to high-density lipoproteins (HDL), low-density lipoproteins (LDL), and very low-density lipoproteins (VLDL) (31). The SR-BI binding site is mapped to the hypervariable region 1 (HVR-1) of HCV E2 (53). SR-BI ligands, such as HDL and oxidized LDL have been found to affect HCV infectivity (4, 14, 58-60). Indeed, HDL has been shown to enhance HCV infection in an SR-BI-dependent manner (4, 14, 58, 59). Antibodies against SR-BI and knockdown of SR-BI in cells result in a significant inhibition of viral infection in both the HCVpp and the HCVcc systems (5, 25, 32).Although clearly involved in entry and immune recognition, the more downstream function(s) of HCV glycoproteins are poorly understood, as their structure has not yet been solved. Nonetheless, mutational analysis and mapping of neutralizing antibody epitopes have delineated several discontinuous regions of E2 that are essential for HCV particle binding and entry (24, 33, 45, 47). One of these is a highly conserved sequence spanning E2 residues 412 to 423 (QLINTNGSWHIN). Several broadly neutralizing monoclonal antibodies (MAbs) bind to this epitope. These include mouse monoclonal antibody (MAb) AP33, rat MAb 3/11, and the human MAbs e137, HCV1, and 95-2 (8, 16, 44, 45, 49). Of these, MAbs AP33, 3/11, and e137 are known to block the binding of E2 to CD81.Cell culture-adaptive mutations within the HCV glycoproteins are valuable for investigating the virus interaction(s) with cellular receptors (18). In the present study, we characterize an asparagine-to-aspartic acid mutation at residue 415 (N415D) in HCV strain JFH1 E2 that arose during the long-term passaging of infected human hepatoma Huh-7 cells. Alongside N415D, we also characterize three adjacent cell culture adaptive mutations reported previously and a novel substitution generated in the present study by propagating virus under MAb AP33 selective pressure to gain further insight into the function of this region of E2 in viral infection.  相似文献   

7.
Using a cell-based replicon screen, we identified a class of compounds with a thiazolidinone core structure as inhibitors of hepatitis C virus (HCV) replication. The concentration of one such compound, BMS-824, that resulted in a 50% inhibition of HCV replicon replication was ∼5 nM, with a therapeutic index of >10,000. The compound showed good specificity for HCV, as it was not active against several other RNA and DNA viruses. Replicon cells resistant to BMS-824 were isolated, and mutations were identified. A combination of amino acid substitutions of leucine to valine at residue 31 (L31V) and glutamine to leucine at residue 54 (Q54L) in NS5A conferred resistance to this chemotype, as did a single substitution of tyrosine to histidine at amino acid 93 (Y93H) in NS5A. To further explore the region(s) of NS5A involved in inhibitor sensitivity, genotype-specific NS5A inhibitors were used to evaluate a series of genotype 1a/1b hybrid replicons. Our results showed that, consistent with resistance mapping, the inhibitor sensitivity domain also mapped to the N terminus of NS5A, but it could be distinguished from the key resistance sites. In addition, we demonstrated that NS5A inhibitors, as well as an active-site inhibitor that specifically binds NS3 protease, could block the hyperphosphorylation of NS5A, which is believed to play an essential role in the viral life cycle. Clinical proof of concept has recently been achieved with derivatives of these NS5A inhibitors, indicating that small molecules targeting a nontraditional viral protein like NS5A, without any known enzymatic activity, can also have profound antiviral effects on HCV-infected subjects.Hepatitis C virus (HCV) is the major causative agent for non-A, non-B hepatitis worldwide, which affects more than 3% of the world population. HCV establishes chronic infections in a large percentage of infected individuals, increasing the risk for developing liver cirrhosis and, in some cases, hepatocellular carcinoma. Although the current standard of care for HCV infection involves the use of PEGylated interferon and ribavirin, a large proportion of patients fail to respond to this therapy, and treatment is associated with frequent and sometimes serious side effects (9). Given the limited efficacy of the current therapy, the development of safer and more effective therapies is of tremendous importance.HCV is a positive-strand RNA virus belonging to the family Flaviviridae. The HCV genome consists of a ∼9.6-kb RNA with a large open reading frame encoding a polyprotein of ∼3,010 amino acids. The polyprotein is cleaved co- and posttranslationally by both cellular and viral proteases into at least 10 different products (10, 11). The viral proteins required for RNA replication include NS3, NS4A, NS4B, NS5A, and NS5B (4, 19). NS3 consists of an amino-terminal protease domain required for the cleavage of the remaining nonstructural proteins and a carboxyl-terminal helicase/NTPase domain (8, 11, 30). NS4A serves as a cofactor for NS3 protease and helicase activities (8). NS4B is an integral membrane protein involved in the formation of the membranous web, where HCV replication complexes are thought to assemble (7). NS5A is a membrane-associated phosphoprotein present in basally phosphorylated (p56) and hyperphosphorylated (p58) forms (15, 31). It was previously reported that only p58-defective mutants could be complemented in trans (1), and NS5A is involved in HCV virion production (22, 34), suggesting that different forms of NS5A exert multiple functions at various stages of the viral life cycle. The N terminus of NS5A (domain I) has been crystallized in alternative dimer forms and contains zinc- and RNA-binding domains (20, 33). The ability of NS5A to bind to zinc (32) and RNA (14) has been demonstrated in vitro. NS5A has been shown to interact with a number of host proteins, is implicated in interferon resistance in vivo, and has been the subject of several reviews (13, 21). NS5B functions as the viral RNA-dependent RNA polymerase (2). Previous studies have shown that the NS3-NS5B proteins are all essential for HCV replication and are believed to form the HCV replicase complex (4, 18, 19).The development of the cell-based HCV replicon system provides a means for the large-scale screening of HCV inhibitors against multiple viral targets. The use of a cell-based replication assay likely includes essential functions that previously could not be evaluated with in vitro enzyme assays. The disadvantages for the advancement of HCV inhibitors targeting nonenzymatic proteins are (i) the potential for structure-activity relationships (SAR) to be difficult to interpret based on the complexity of cell-based systems, (ii) the lack of a system for validation, and (iii) difficulty in predicting if in vitro potency can translate into in vivo effect. Therefore, during the process of developing HCV NS5A inhibitors, we established a series of assays and checkpoints prior to entering the clinic. This is the first report in a series of articles detailing the development of HCV NS5A inhibitors that has culminated in the demonstration of clinical efficacy for this novel mechanistic class of HCV inhibitor (25).In this report, we have used a previously described cell-based approach (26) to identify a novel compound that specifically inhibits HCV RNA replication. Through the use of resistance selection, we have demonstrated that the inhibitor targets the HCV NS5A protein, thereby establishing that the function of NS5A in replication can be inhibited by small molecules. In addition, using genotype-specific inhibitors, we have further shown that the N terminus of NS5A plays an essential role in compound activity by both 50% effective concentration (EC50) determinations as well as a functional assay to evaluate NS5A hyperphosphorylation.  相似文献   

8.
Hepatitis C virus (HCV) RNA genome replicates within the ribonucleoprotein (RNP) complex in the modified membranous structures extended from endoplasmic reticulum. A proteomic analysis of HCV RNP complexes revealed the association of oxysterol binding protein (OSBP) as one of the components of these complexes. OSBP interacted with the N-terminal domain I of the HCV NS5A protein and colocalized to the Golgi compartment with NS5A. An OSBP-specific short hairpin RNA that partially downregulated OSBP expression resulted in a decrease of the HCV particle release in culture supernatant with little effect on viral RNA replication. The pleckstrin homology (PH) domain located in the N-terminal region of OSBP targeted this protein to the Golgi apparatus. OSBP deletion mutation in the PH (ΔPH) domain failed to localize to the Golgi apparatus and inhibited the HCV particle release. These studies suggest a possible functional role of OSBP in the HCV maturation process.Hepatitis C virus (HCV) infection is one of the leading causes of chronic hepatitis. HCV infection is associated with cirrhosis, steatosis, and hepatocellular carcinoma (33). The HCV RNA genome of ∼9.6 kb is translated via an internal ribosome entry site element on the rough endoplasmic reticulum (ER) as a polyprotein precursor of about 3,010 amino acids that is co- and posttranslationally processed by cellular and viral proteases into mature structural and nonstructural (NS) proteins (33). HCV replicates within ribonucleoprotein (RNP) complexes associated with modified ER membranous structures (15). Recent work implicated lipid droplets that emanate from the ER as sites of RNA replication (28, 44). Almost all of the HCV NS proteins along with a variety of cellular factors are associated with the RNP complexes engaged in viral RNA replication (37). It is likely that these NS proteins not only participate in replication process but also are involved in the various steps of virion morphogenesis and assembly. Membrane-associated RNP complexes are generally composed of viral proteins, replicating RNA, host proteins, and altered cellular membranes (1). In this respect, a growing body of evidence implicates the functional role of NS5A in early steps of virion assembly and morphogenesis (3, 27, 45). NS5A is a phosphoprotein that migrates in sodium dodecyl sulfate gels as 56-kDa (basally phosphorylated) and 58-kDa (hyperphosphorylated) forms of proteins. The C-terminal domain III region of NS5A and the phosphorylated residue (Ser457) are important for virion maturation (3, 27, 45). NS5A domain III contains the binding site for viral core protein, indicating the possible involvement of NS5A protein in virus assembly (27). NS5A anchors to the ER membrane by an N-terminal hydrophobic α-helix, and this attachment is needed for its key role(s) in viral replication (10). Studies suggest that phosphorylation of NS5A plays a functional role in viral replication (12). The hyperphosphorylated NS5A reduces its interaction with the human vesicle-associated membrane protein-associated protein A (VAP-A) (12). VAP-A binds both NS5A and NS5B (13, 17). These associations are important for RNA replication (13, 17).HCV alters lipid homeostasis to benefit its infectious processes. Host lipids and their synthesis affect viral infectious process (21, 40, 51, 57). HCV RNA replication can be induced by saturated and monounsaturated fatty acids and inhibited by polyunsaturated fatty acids (18, 21). HCV gene expression induces lipogenesis by stimulating the activation of the sterol regulatory element binding proteins, the master regulators of lipid/fatty acid biosynthetic pathways (51). Reagents that interfere with host lipid biosynthetic pathways abrogate viral replication (21, 57). It has been suggested that HCV utilizes the very-low-density lipoprotein (VLDL) secretion pathway for its viral particle release (14, 19). These studies collectively suggest that host lipid metabolism plays a key role in the viral life cycle including replication, virion assembly, and secretion (56).In the present study, we focus on the functional role of oxysterol binding protein (OSBP) that was identified by proteomic analysis as one of the host factors associated with the HCV RNP complexes. OSBP belongs to a family of the OSBP-related proteins. Originally discovered as a major cytosolic receptor for oxidized cholesterols, it undergoes translocation from the cytosolic/vesicular compartment to the Golgi apparatus upon ligand (hydroxycholesterol) binding (38). OSBP also binds to VAP-A via its FFAT motif (53). Golgi apparatus translocation of OSBP is regulated by the pleckstrin homology (PH) domain. This domain also harbors binding sites for phosphatidylinositol 4-phosphate (PI4P) and phosphatidylinositol 4,5-biphosphate (PI4,5P2) (25). OSBP and OSBP-related proteins are implicated in cholesterol homeostasis, phospholipid metabolism, vesicular transport, and cell signaling (55). OSBP functions as sterol sensor that regulates the transport of ceramide from the ER to the Golgi apparatus for de novo synthesis of sphingomyelin by coordinated action with ceramide transport protein (CERT) (36). OSBP also functions as a scaffolding protein for two phosphatases (phosphatase 2A/HePTP) (49). This complex regulates the activity of extracellular signal-regulate kinase. This cytosolic 440-kDa complex disassembles by the addition of 25-hydroxycholesterol (25-HC) or depletion of cholesterol, both of which cause OSBP translocation to the Golgi compartment (49). Thus, in addition to its role in intracellular trafficking, OSBP appears to regulate cell signaling. We investigated the functional significance of OSBP association with HCV RNP complexes. RNA interference studies support a functional role of OSBP in virion morphogenesis and release process. The OSBP PH domain deletion mutant (ΔPH) failed to localize to the Golgi apparatus and caused an inhibition of the HCV particle release. Our work described herein also demonstrates that the association of OSBP with NS5A may also contribute to the overall HCV maturation process.  相似文献   

9.
Hepatitis C virus (HCV) is a liver-tropic pathogen with severe health consequences for infected individuals. Chronic HCV infection can progress to cirrhosis and hepatocellular carcinoma and is a leading indicator for liver transplantation. The HCV core protein is an essential component of the infectious virus particle, but many aspects of its role remain undefined. The C-terminal region of the core protein acts as a signal sequence for the E1 glycoprotein and undergoes dual processing events during infectious virus assembly. The exact C terminus of the mature, virion-associated core protein is not known. Here, we performed genetic analyses to map the essential determinants of the HCV core C-terminal region, as well as to define the minimal length of the protein that can function for infectious virus production in trans.Hepatitis C virus (HCV) is a major contributor to the development of human liver diseases, infecting approximately 2% of the population, or 130 million people, worldwide (2). Up to 80% of HCV infections progress to chronic hepatitis and can lead to cirrhosis and hepatocellular carcinoma (38). No vaccine exists to prevent HCV infection, and current treatments are frequently inadequate.HCV is an enveloped virus of the genus Hepacivirus in the family Flaviviridae (30). The single-stranded, positive-sense RNA genome encodes a polyprotein of about 3,000 amino acids, which is processed by viral and host proteases into three structural proteins (the core protein, E1, and E2) and seven presumed nonstructural proteins (p7, NS2, NS3, NS4A, NS4B, NS5A, and NS5B). The core protein is thought to encapsidate the RNA genome within the virion, forming a complex that is surrounded by a host cell-derived lipid bilayer displaying the envelope glycoproteins, E1 and E2. Although not thought to be components of the virion, p7 and NS2 have recently been implicated in the production of infectious virus (5, 12, 29, 33, 39). The remaining nonstructural proteins, NS3 to NS5B, are essential for genome replication and have additional emerging roles in virus assembly. NS3 possesses RNA helicase/NTPase activities and together with its cofactor, NS4A, forms the major viral protease. NS5B is the RNA-dependent RNA polymerase (reviewed in references 16 and 27).The core protein is the first protein produced during translation of the incoming viral genome. A signal sequence in its C-terminal region targets the nascent E1 glycoprotein to the endoplasmic reticulum (ER) membrane and is the substrate for processing by two host proteases. Cleavage by signal peptidase (SP) following core amino acid 191 (31) is thought to precede processing by signal peptide peptidase (SPP) (20, 26), an integral membrane aspartyl protease that cleaves within transmembrane segments (37). The C terminus of the mature, infectious-virion-associated core protein has not been determined, but it is speculated to lie between amino acids 173 and 182 (24, 31). SPP processing has been shown to mobilize the core from the ER membrane and enable it to traffic to lipid droplets (20). These triglyceride-rich storage organelles have recently been shown to be the sites of HCV particle assembly (21). Consistent with this finding, impaired SPP activity leads to decreased HCV infectious titers (34). Dual processing of the core proteins is a common feature of the Flaviviridae family. GB virus B, a hepacivirus, and classical swine fever virus, a related pestivirus, encode core proteins that undergo SP and SPP processing during maturation (8, 35). In the genus Flavivirus, the capsid protein undergoes regulated cleavage by the viral NS2B-3 protease and SP; this stepwise processing has been shown to be essential for proper encapsidation of genomes into infectious particles (3).The development of an infectious cell culture system for HCV has been a major breakthrough in the field (7). Many details of virus morphogenesis and infectivity, however, are still unknown. In this study, we examined the role of the C-terminal portion of the HCV core protein and identified individual amino acids that are essential for infectious virus assembly and core protein stability. Findings from alanine-scanning and transcomplementation studies suggest that at least 177 residues of the core protein are needed to produce infectious particles.  相似文献   

10.
We analyzed the biochemical and ultrastructural properties of hepatitis C virus (HCV) particles produced in cell culture. Negative-stain electron microscopy revealed that the particles were spherical (∼40- to 75-nm diameter) and pleomorphic and that some of them contain HCV E2 protein and apolipoprotein E on their surfaces. Electron cryomicroscopy revealed two major particle populations of ∼60 and ∼45 nm in diameter. The ∼60-nm particles were characterized by a membrane bilayer (presumably an envelope) that is spatially separated from an internal structure (presumably a capsid), and they were enriched in fractions that displayed a high infectivity-to-HCV RNA ratio. The ∼45-nm particles lacked a membrane bilayer and displayed a higher buoyant density and a lower infectivity-to-HCV RNA ratio. We also observed a minor population of very-low-density, >100-nm-diameter vesicular particles that resemble exosomes. This study provides low-resolution ultrastructural information of particle populations displaying differential biophysical properties and specific infectivity. Correlative analysis of the abundance of the different particle populations with infectivity, HCV RNA, and viral antigens suggests that infectious particles are likely to be present in the large ∼60-nm HCV particle populations displaying a visible bilayer. Our study constitutes an initial approach toward understanding the structural characteristics of infectious HCV particles.Hepatitis C virus (HCV) is a major cause of chronic hepatitis worldwide, with approximately 170 million humans chronically infected. Persistent HCV infection often leads to fibrosis, cirrhosis, and hepatocellular carcinoma (27). There is no vaccine against HCV, and the most widely used therapy involves the administration of type I interferon (IFN-α2Α) combined with ribavirin. However, this treatment is often associated with severe adverse effects and is often ineffective (53).HCV is a member of the Flaviviridae family and is the sole member of the genus Hepacivirus (43). HCV is an enveloped virus with a single-strand positive RNA genome that encodes a unique polyprotein of ∼3,000 amino acids (14, 15). A single open reading frame is flanked by untranslated regions (UTRs), the 5′ UTR and 3′ UTR, that contain RNA sequences essential for RNA translation and replication, respectively (17, 18, 26). Translation of the single open reading frame is driven by an internal ribosomal entry site (IRES) sequence residing within the 5′ UTR (26). The resulting polyprotein is processed by cellular and viral proteases into its individual components (reviewed in reference 55). The E1, E2, and core structural proteins are required for particle formation (5, 6) but not for viral RNA replication or translation (7, 40). These processes are mediated by the nonstructural (NS) proteins NS3, NS4A, NS4B, NS5A, and NS5B, which constitute the minimal viral components necessary for efficient viral RNA replication (7, 40).Expression of the viral polyprotein leads to the formation of virus-like particles (VLPs) in HeLa (48) and Huh-7 cells (23). Furthermore, overexpression of core, E1, and E2 is sufficient for the formation of VLPs in insect cells (3, 4). In the context of a viral infection, the viral structural proteins (65), p7 (31, 49, 61), and all of the nonstructural proteins (2, 29, 32, 41, 44, 63, 67) are required for the production of infectious particles, independent of their role in HCV RNA replication. It is not known whether the nonstructural proteins are incorporated into infectious virions.The current model for HCV morphogenesis proposes that the core protein encapsidates the viral genome in areas where endoplasmic reticulum (ER) cisternae are in contact with lipid droplets (47), forming HCV RNA-containing particles that acquire the viral envelope by budding through the ER membrane (59). We along with others showed recently that infectious particle assembly requires microsomal transfer protein (MTP) activity and apolipoprotein B (apoB) (19, 28, 50), suggesting that these two components of the very-low-density lipoprotein (VLDL) biosynthetic machinery are essential for the formation of infectious HCV particles. This idea is supported by the reduced production of infectious HCV particles in cells that express short hairpin RNAs (shRNAs) targeting apolipoprotein E (apoE) (12, 30).HCV RNA displays various density profiles, depending on the stage of the infection at which the sample is obtained (11, 58). The differences in densities and infectivities have been attributed to the presence of host lipoproteins and antibodies bound to the circulating viral particles (24, 58). In patients, HCV immune complexes that have been purified by protein A affinity chromatography contain HCV RNA, core protein, triglycerides, apoB (1), and apoE (51), suggesting that these host factors are components of circulating HCV particles in vivo.Recent studies using infectious molecular clones showed that both host and viral factors can influence the density profile of infectious HCV particles. For example, the mean particle density is reduced by passage of cell culture-grown virus through chimpanzees and chimeric mice whose livers contain human hepatocytes (39). It has also been shown that a point mutation in the viral envelope protein E2 (G451R) increases the mean density and specific infectivity of JFH-1 mutants (70).HCV particles exist as a mixture of infectious and noninfectious particles in ratios ranging from 1:100 to 1:1,000, both in vivo (10) and in cell culture (38, 69). Extracellular infectious HCV particles have a lower average density than their noninfectious counterparts (20, 24, 38). Equilibrium sedimentation analysis indicates that particles with a buoyant density of ∼1.10 to 1.14 g/ml display the highest ratio of infectivity per genome equivalent (GE) both in cell culture (20, 21, 38) and in vivo (8). These results indicate that these samples contain relatively more infectious particles than any other particle population. Interestingly, mutant viruses bearing the G451R E2 mutation display an increased infectivity-HCV RNA ratio only in fractions with a density of ∼1.1 g/ml (21), reinforcing the notion that this population is selectively enriched in infectious particles.The size of infectious HCV particles has been estimated in vivo by filtration (50 to 80 nm) (9, 22) and by rate-zonal centrifugation (54 nm) (51) and in cell culture by calculation of the Stokes radius inferred from the sedimentation velocity of infectious JFH-1 particles (65 to 70 nm) (20). Previous ultrastructural studies using patient-derived material report particles with heterogeneous diameters ranging from 35 to 100 nm (33, 37, 42, 57, 64). Cell culture-derived particles appear to display a diameter within that range (∼55 nm) (65, 68).In this study we exploited the increased growth capacity of a cell culture-adapted virus bearing the G451R mutation in E2 (70) and the enhanced particle production of the hyperpermissive Huh-7 cell subclone Huh-7.5.1 clone 2 (Huh-7.5.1c2) (54) to produce quantities of infectious HCV particles that were sufficient for electron cryomicroscopy (cryoEM) analyses. These studies revealed two major particle populations with diameters of ∼60 and ∼45 nm. The larger-diameter particles were distinguished by the presence of a membrane bilayer, characterized by electron density attributed to the lipid headgroups in its leaflets. Isopycnic ultracentrifugation showed that the ∼60-nm particles are enriched in fractions with a density of ∼1.1 g/ml, where optimal infectivity-HCV RNA ratios are observed. These results indicate that the predominant morphology of the infectious HCV particle is spherical and pleomorphic and surrounded by a membrane envelope.  相似文献   

11.
12.
The hepatitis C virus NS2 protein has been recently implicated in virus particle assembly. To further understand the role of NS2 in this process, we conducted a reverse genetic analysis of NS2 in the context of a chimeric genotype 2a infectious cell culture system. Of 32 mutants tested, all were capable of RNA replication and 25 had moderate-to-severe defects in virus assembly. Through forward genetic selection for variants capable of virus spread, we identified second-site mutations in E1, E2, NS2, NS3, and NS4A that suppressed NS2 defects in assembly. Two suppressor mutations, E1 A78T and NS3 Q221L, were further characterized by additional genetic and biochemical experiments. Both mutations were shown to suppress other NS2 defects, often with mutual exclusivity. Thus, several NS2 mutants were enhanced by NS3 Q221L and inhibited by E1 A78T, while others were enhanced by E1 A78T and inhibited by NS3 Q221L. Furthermore, we show that the NS3 Q221L mutation lowers the affinity of native, full-length NS3-NS4A for functional RNA binding. These data reveal a complex network of interactions involving NS2 and other viral structural and nonstructural proteins during virus assembly.Hepatitis C virus (HCV) is a major cause of acute and chronic liver disease and contributes to the development of hepatocellular carcinoma. HCV is an enveloped, positive-strand RNA virus, the type member of the Hepacivirus genus in the family Flaviviridae (43). HCV exhibits high levels of sequence diversity that cluster into seven major genotypes and numerous subtypes (21).HCV genomes are 9.6 kb and encode a single long open reading frame of ∼3,011 codons (43). Translation of this genome produces a large polyprotein that is co- and posttranslationally processed by viral and host proteases into 10 distinct products. The N-terminal one-third of the polyprotein encodes the structural proteins, which are thought to compose the virus particle. These include an RNA-binding nucleocapsid protein, core (C), and two viral envelope glycoproteins, E1 and E2. E1 and E2 are type I membrane proteins that coordinately fold into a heterodimer complex (36). The remainder of the genome encodes the nonstructural (NS) proteins NS2, NS3, NS4A, NS4B, NS5A, and NS5B, which mediate the intracellular aspects of the viral life cycle. In addition, a small viroporin-like protein, p7, resides between the structural and NS genes.HCV encodes two proteases, the NS2-NS3 cysteine autoprotease and the NS3-NS4A serine protease. The only known substrate of the NS2-NS3 autoprotease is the NS2/3 junction. This enzyme is encoded by the C-terminal 121 amino acids (aa) of NS2, which forms a homodimer with twin composite active sites composed of two residues from one chain and one residue from the other (45). In addition, the serine protease domain of NS3 plays a noncatalytic role in stimulating NS2/3 cleavage (69). Upstream of the cysteine protease domain, the N-terminal hydrophobic region of NS2 mediates interaction with cellular membranes. While the membrane topology of NS2 is not yet fully known (67, 80), N-terminal cleavage by endoplasmic reticulum-resident signal peptidase and C-terminal cleavage by the cytosolic NS2-NS3 cysteine protease indicate that NS2 likely contains one or three transmembrane (TM) domains.The NS3-NS4A serine protease is encoded by the N-terminal domain of NS3 and is responsible for downstream cleavages at the NS3/4A, NS4A/B, NS4B/5A, and NS5A/B junctions. NS4A, a small (54-aa), membrane-anchored protein, acts as a cofactor for the serine protease activity by helping to complete the chymotrypsin-like fold of NS3 (14, 46). In addition to polyprotein processing, NS3-NS4A serine protease helps to dampen the innate antiviral response by cleaving cellular proteins involved in signal transduction (65).The C-terminal region of NS3 encodes an RNA helicase/NTPase activity that is essential for viral replication, although it is not yet clear which specific step(s) of the replication cycle requires this activity (29, 33). Interestingly, the NS3 serine protease and RNA helicase domains enhance each other''s activities, suggesting that proteolysis and RNA replication may be functionally coordinated (5, 6). In addition, NS4A helps to promote RNA-stimulated ATP hydrolysis by the NS3 helicase domain (4).In addition to their role in polyprotein processing, emerging evidence indicates that NS2 and NS3-NS4A participate in virus particle assembly (52). Prior work showed that NS2 is not essential for RNA replication of subgenomic replicons engineered to express NS3 through NS5B (44). The first evidence for an additional function of NS2 came from the construction of improved chimeric genotype 2a cDNA clones that replicated to high titers in cell culture (HCVcc). Pietschmann and colleagues showed that the Jc1 chimera containing a J6-JFH1 junction between the first and second putative TM domains of NS2 yielded higher-titer viruses than the original infectious J6/JFH chimera (41, 58). Furthermore, a number of adaptive mutations that improve virus production have been mapped to NS2 and NS3 (22, 23, 27, 53, 64, 68, 82). By using bicistronic constructs to express NS2 and NS3 independently of NS2/3 cleavage, two groups showed that full-length NS2, but not uncleaved NS2-NS3 or the NS2 cysteine protease active sites, was required for virus production (24, 25). Moreover, a limited number of mutations in NS2 were shown to inhibit virus assembly (24, 79, 83).Despite these observations, the role of NS2 in virus assembly remains unclear. We have therefore undertaken a genetic analysis to target conserved residues in NS2 for site-directed mutagenesis and identified a number of key residues that are important for virus assembly. Further analysis revealed that a network of genetic interactions among NS2, E1-E2, and NS3-NS4A helps to direct virus assembly. Finally, a suppressor mutation in NS3 was shown to influence functional RNA binding by the RNA helicase/ATPase.  相似文献   

13.
Hepatitis C virus (HCV) is an important human pathogen affecting 170 million chronically infected individuals. In search for cellular proteins involved in HCV replication, we have developed a purification strategy for viral replication complexes and identified annexin A2 (ANXA2) as an associated host factor. ANXA2 colocalized with viral nonstructural proteins in cells harboring genotype 1 or 2 replicons as well as in infected cells. In contrast, we found no obvious colocalization of ANXA2 with replication sites of other positive-strand RNA viruses. The silencing of ANXA2 expression showed no effect on viral RNA replication but resulted in a significant reduction of extra- and intracellular virus titers. Therefore, it seems likely that ANXA2 plays a role in HCV assembly rather than in genome replication or virion release. Colocalization studies with individually expressed HCV nonstructural proteins indicated that NS5A specifically recruits ANXA2, probably by an indirect mechanism. By the deletion of individual NS5A subdomains, we identified domain III (DIII) as being responsible for ANXA2 recruitment. These data identify ANXA2 as a novel host factor contributing, with NS5A, to the formation of infectious HCV particles.Hepatitis C virus (HCV) infections are characterized by a mostly unapparent acute phase leading to persistence in ca. 70% of all infected individuals. Currently, 170 million people suffer from chronic hepatitis C, and they have a high risk to develop severe liver disease. It has been estimated that HCV accounts for 27% of cirrhosis and 25% of hepatocellular carcinoma cases worldwide (2).HCV is an enveloped positive-strand RNA virus belonging to the genus Hepacivirus in the family Flaviviridae. The genome of HCV encompasses a single ∼9,600-nucleotide (nt)-long RNA molecule containing one large open reading frame (ORF) that is flanked by nontranslated regions (NTRs), which are important for viral translation and replication. HCV proteins generated from the polyprotein precursor are cleaved by cellular and viral proteases into at least 10 different products (for a review of polyprotein cleavage and the function of the individual proteins, see reference 4). The structural proteins Core, E1, and E2 are located in the amino-terminal portion of the polyprotein, followed by p7, a hydrophobic peptide that is supposed to be a viroporin, and the nonstructural proteins (NS) NS2, NS3, NS4A, NS4B, NS5A, and NS5B. Only the nonstructural proteins NS3 to NS5B are involved in viral RNA replication. NS3 is a multifunctional protein, consisting of an amino-terminal protease domain required for the processing of the NS3 to NS5B region and a carboxyterminal helicase/nucleoside triphosphatase domain. NS4A is a cofactor that activates the NS3 protease function by forming a heterodimer. The hydrophobic protein NS4B induces vesicular membrane alterations involved in RNA replication. NS5A is a phosphoprotein that seems to play an important role in viral replication and assembly (3, 35, 58). NS5B is the RNA-dependent RNA polymerase of HCV.Positive-strand RNA viruses replicate their RNA in vesicular structures originating from different cellular organelles (36). In the case of HCV, particular membrane alterations have been identified by electron microscopy, designated the membranous web, consisting of accumulations of vesicles primarily derived from the endoplasmic reticulum (17). Important insights into the organization of HCV replication complexes were obtained by the in vitro analysis of viral RNA synthesis in membrane preparations of cells harboring subgenomic HCV replicons, so-called crude replication complexes (CRCs) (1, 20). A current model based on a stoichiometric analysis of CRCs suggests that each vesicular structure contains multiple copies of viral nonstructural proteins and has a connection to the cytoplasm, allowing the constant supply of nucleotides for RNA synthesis (45), presumably analogously to the replication complex of the closely related dengue virus (DV) (64). Viral RNA synthesis in CRCs is highly resistant to proteinases and nucleases (39), and the membranes are detergent resistant at 4°C, resembling features of lipid rafts (54).Several purification techniques have been established to identify relevant HCV host factors by proteomics, based on either the extraction of detergent-resistant membranes (19, 34) or the immunoprecipitation of vesicles (24), revealing different sets of cellular proteins potentially involved in viral replication. In most of these studies, cell lines harboring persistent subgenomic replicons were utilized (33); however, with the availability of a fully permissive cell culture system supporting the complete HCV replication cycle (31, 63, 66), it became evident that viral RNA replication and assembly are closely linked. Recent work revealed an intimate connection of viral replication complexes and assembly sites in close proximity to cytoplasmic lipid droplets (38), with Core and especially NS5A functioning as central regulators by a poorly defined mechanism. NS5A is phosphorylated at multiple serine and threonine residues, binds RNA, and is composed of three domains, which are separated by trypsin-sensitive low-complexity regions (LCS I and II) (59). An N-terminal amphipathic alpha helix tightly associates NS5A with intracellular membranes. Domain I and LCS1 most likely are involved in viral RNA replication, since replication-enhancing mutations primarily mapped to this region (8, 32). The role of domain II is unknown, while domain III recently has been shown to be dispensable for RNA replication but essential for viral particle assembly (3, 35, 58). One of the proposed mechanisms points to a critical interaction with the Core protein, for which phosphorylation in the C-terminal part of domain III of NS5A appears to be required (35). The interaction of Core and NS5A has been proposed to be important for the recruitment of the replication complexes to lipid droplets (3), thereby allowing a coordinated packaging of the newly synthesized RNA.In this study, we identified annexin A2 (also called annexin II, calpactin 1, and ANXA2) as an HCV host factor by a proteomic analysis. ANXA2 belongs to a family of proteins characterized by their Ca2+-dependent binding to negatively charged phospholipids. The annexin proteins consist of two principle domains, a variable N-terminal and a conserved C-terminal domain, which harbors the Ca2+ and membrane binding sites (for a review, see references 14 and 15). All annexins show cytosolic and membrane localizations. Membrane recruitment probably is regulated by intracellular Ca2+ fluctuations, and target membrane selection differs for different annexins.In addition to showing a cytosolic distribution, ANXA2 can associate with the plasma membrane and the membrane of early endosomes. Plasma membrane-associated ANXA2 typically is found in a tight heterotetrameric complex with the S100 protein S100A10 (p11). ANXA2 specifically interacts with phosphatidylinositol(4,5)bisphosphate (PIP2) (22, 48) and binds to membranes enriched in cholesterol, supporting a role in the organization of lipid raft-like membrane microdomains. Due to the direct binding of ANXA2 to F-actin, the protein has been proposed to provide a direct link between cytoskeletal elements and PIP2/cholesterol-rich membrane domains (47).ANXA2 has been implicated in several cellular transport processes, including the internalization and transport of cholesteryl esters, the biogenesis of multivesicular bodies, the recycling of plasma membrane receptors, and the Ca2+-induced exocytosis of certain secretory granules (14). Here, we show that ANXA2 is present at HCV replication sites within the membranous web. The recruitment of ANXA2 is mediated by domain III of NS5A and probably is required for efficient virus assembly.  相似文献   

14.
The hepatitis C virus (HCV) nonstructural protein 2 (NS2) is a dimeric multifunctional hydrophobic protein with an essential but poorly understood role in infectious virus production. We investigated the determinants of NS2 function in the HCV life cycle. On the basis of the crystal structure of the postcleavage form of the NS2 protease domain, we mutated conserved features and analyzed the effects of these changes on polyprotein processing, replication, and infectious virus production. We found that mutations around the protease active site inhibit viral RNA replication, likely by preventing NS2-3 cleavage. In contrast, alterations at the dimer interface or in the C-terminal region did not affect replication, NS2 stability, or NS2 protease activity but decreased infectious virus production. A comprehensive deletion and mutagenesis analysis of the C-terminal end of NS2 revealed the importance of its C-terminal leucine residue in infectious particle production. The crystal structure of the NS2 protease domain shows that this C-terminal leucine is locked in the active site, and mutation or deletion of this residue could therefore alter the conformation of NS2 and disrupt potential protein-protein interactions important for infectious particle production. These studies begin to dissect the residues of NS2 involved in its multiple essential roles in the HCV life cycle and suggest NS2 as a viable target for HCV-specific inhibitors.An estimated 130 million people are infected with hepatitis C virus (HCV), the etiologic agent of non-A, non-B viral hepatitis. Transmission of the virus occurs primarily through blood or blood products. Acute infections are frequently asymptomatic, and 70 to 80% of the infected individuals are unable to eliminate the virus. Of the patients with HCV-induced chronic hepatitis, 15 to 30% progress to cirrhosis within years to decades after infection, and 3 to 4% of patients develop hepatocellular carcinoma (17). HCV infection is a leading cause of cirrhosis, end-stage liver disease, and liver transplantation in Europe and the United States (7), and reinfection after liver transplantation occurs almost universally. There is no vaccine available, and current HCV therapy of pegylated alpha interferon in combination with ribavirin leads to a sustained response in only about 50% of genotype 1-infected patients.The positive-stranded RNA genome of HCV is about 9.6 kb in length and encodes a single open reading frame flanked by 5′ and 3′ nontranslated regions (5′ and 3′ NTRs). The translation product of the viral genome is a large polyprotein containing the structural proteins (core, envelope proteins E1 and E2) in the N-terminal region and the nonstructural proteins (p7, nonstructural protein 2 [NS2], NS3, NS4A, NS4B, NS5A, and NS5B) in the C-terminal region. The individual proteins are processed from the polyprotein by various proteases. The host cellular signal peptidase cleaves between core/E1, E1/E2, E2/p7, and p7/NS2, and signal peptide peptidase releases core from the E1 signal peptide. Two viral proteases, the NS2-3 protease and the NS3-4A protease, cleave the remainder of the viral polyprotein in the nonstructural region (22, 27). The structural proteins package the genome into infectious particles and mediate virus entry into a naïve host cell; the nonstructural proteins NS3 through NS5B form the RNA replication complex. p7 and NS2 are not thought to be incorporated into the virion but are essential for the assembly of infectious particles (14, 36); however, their mechanisms of action are not understood.NS2 (molecular mass of 23 kDa) is a hydrophobic protein containing several transmembrane segments in the N-terminal region (5, 9, 32, 39). The C-terminal half of NS2 and the N-terminal third of NS3 form the NS2-3 protease (10, 11, 26, 37). NS2 is not required for the replication of subgenomic replicons, which span NS3 to NS5B (20). However, cleavage at the NS2/3 junction is necessary for replication in chimpanzees (16), the full-length replicon (38), and in the infectious tissue culture system (HCVcc) (14). Although cleavage can occur in vitro in the absence of microsomal membranes, synthesis of the polyprotein precursor in the presence of membranes greatly increases processing at the NS2/3 site (32). In vitro studies indicate that purified NS2-3 protease is active in the absence of cellular cofactors (11, 37). In addition to its role as a protease, NS2 has been shown to be required for assembly of infectious intracellular virus (14). The N-terminal helix of NS2 was first implicated in infectivity by the observation that an intergenotypic breakpoint following this transmembrane segment resulted in higher titers of infectious virus (28). Structural and functional characterization of the NS2 transmembrane region has shown that this domain is essential for infectious virus production (13). In particular, a central glycine residue in the first NS2 helix plays a critical role in HCV infectious virus assembly (13). The NS2 protease domain, but not its catalytic activity, is also essential for infectious virus assembly, whereas the unprocessed NS2-3 precursor is not required (13, 14).The crystal structure of the postcleavage NS2 protease domain (NS2pro, residues 94 to 217), revealed a dimeric cysteine protease containing two composite active sites (Fig. 2C; [21]). Two antiparallel α-helices make up the N-terminal subdomain, followed by an extended crossover region, which positions the β-sheet-rich C-terminal subdomain near the N-terminal region of the partner monomer. Two of the conserved residues of the catalytic triad (His 143, Glu 163) are located in the loop region after the second N-terminal helix of one monomer, while the third catalytic residue, Cys 184, is located in the C-terminal subdomain of the other monomer. Creation of this unusual pair of composite active sites through NS2 dimerization has been shown to be essential for autoproteolytic cleavage (21). The structure of NS2pro further demonstrated that the C-terminal residue of NS2 remains bound in the active site after cleavage, suggesting a possible mechanism for restriction of this enzyme to a single proteolytic event (21). Here we have used the crystal structure of NS2pro, along with sequence alignments, to target conserved residues in each of the NS2pro structural regions. Our mutational analysis revealed that the residues in the dimer crossover region and the C-terminal subdomain are important for infectious virus production. In contrast, the majority of amino acids in the active site pocket were not required for infectivity. Interestingly, we observed that the extreme C-terminal leucine of NS2 is absolutely essential for generation of infectious virus, as mutations, deletions, and extensions into NS3 are very poorly tolerated. This analysis begins to dissect the determinants of the multiple functions of this important protease in the HCV life cycle.  相似文献   

15.
For Bovine viral diarrhea virus (BVDV), the type species of the genus Pestivirus in the family Flaviviridae, cytopathogenic (cp) and noncytopathogenic (ncp) viruses are distinguished according to their effect on cultured cells. It has been established that cytopathogenicity of BVDV correlates with efficient production of viral nonstructural protein NS3 and with enhanced viral RNA synthesis. Here, we describe generation and characterization of a temperature-sensitive (ts) mutant of cp BVDV strain CP7, termed TS2.7. Infection of bovine cells with TS2.7 and the parent CP7 at 33°C resulted in efficient viral replication and a cytopathic effect. In contrast, the ability of TS2.7 to cause cytopathogenicity at 39.5°C was drastically reduced despite production of high titers of infectious virus. Further experiments, including nucleotide sequencing of the TS2.7 genome and reverse genetics, showed that a Y1338H substitution at residue 193 of NS2 resulted in the temperature-dependent attenuation of cytopathogenicity despite high levels of infectious virus production. Interestingly, TS2.7 and the reconstructed mutant CP7-Y1338H produced NS3 in addition to NS2-3 throughout infection. Compared to the parent CP7, NS2-3 processing was slightly decreased at both temperatures. Quantification of viral RNAs that were accumulated at 10 h postinfection demonstrated that attenuation of the cytopathogenicity of the ts mutants at 39.5°C correlated with reduced amounts of viral RNA, while the efficiency of viral RNA synthesis at 33°C was not affected. Taken together, the results of this study show that a mutation in BVDV NS2 attenuates viral RNA replication and suppresses viral cytopathogenicity at high temperature without altering NS3 expression and infectious virus production in a temperature-dependent manner.The pestiviruses Bovine viral diarrhea virus-1 (BVDV-1), BVDV-2, Classical swine fever virus (CSFV), and Border disease virus (BDV) are causative agents of economically important livestock diseases. Together with the genera Flavivirus, including several important human pathogens like Dengue fever virus, West Nile virus, Yellow fever virus, and Tick-borne encephalitis virus, and Hepacivirus (human Hepatitis C virus [HCV]), the genus Pestivirus constitutes the family Flaviviridae (8, 20). All members of this family are enveloped viruses with a single-stranded positive-sense RNA genome encompassing one large open reading frame (ORF) flanked by 5′ and 3′ nontranslated regions (NTR) (see references 8 and 28 for reviews). The ORF encodes a polyprotein which is co- and posttranslationally processed into the mature viral proteins by viral and cellular proteases. For BVDV, the RNA genome is about 12.3 kb in length and encodes a polyprotein of about 3,900 amino acids. The first third of the ORF encodes a nonstructural (NS) autoprotease and four structural proteins, while the remaining part of the genome encodes NS proteins which share many common characteristics and functions with the corresponding NS proteins encoded by the HCV genome (8, 28). NS2 of BVDV represents a cysteine autoprotease which is distantly related to the HCV NS2-3 protease (26). NS3, NS4A, NS4B, NS5A, and NS5B are essential components of the pestivirus replicase (7, 10, 49). NS3 possesses multiple enzymatic activities, namely serine protease (48, 52, 53), NTPase (46), and helicase activity (51). NS4A acts as an essential cofactor for the NS3 proteinase. NS5B represents the RNA-dependent RNA polymerase (RdRp) (22, 56). The functions of NS4B and NS5A remain to be determined. NS5A has been shown to be a phosphorylated protein that is associated with cellular serine/threonine kinases (44).According to their effects in tissue culture, two biotypes of pestiviruses are distinguished: cytopathogenic (cp) and noncytopathogenic (ncp) viruses (17, 27). The occurrence of cp BVDV in cattle persistently infected with ncp BVDV is directly linked to the induction of lethal mucosal disease in cattle (12, 13). Previous studies have shown that cp BVDV strains evolved from ncp BVDV strains by different kinds of mutations. These include RNA recombination with various cellular mRNAs, resulting in insertions of cellular protein-coding sequences into the viral genome, as well as insertions, duplications, and deletions of viral sequences, and point mutations (1, 2, 9, 24, 33, 36, 37, 42). A common consequence of all these genetic changes in cp BVDV genomes is the efficient production of NS3 at early and late phases of infection. In contrast, NS3 cannot be detected in cells at late time points after infection with ncp BVDV. An additional major difference is that the cp viruses produce amounts of viral RNA significantly larger than those of their ncp counterparts (7, 32, 50). While there is clear evidence that cell death induced by cp BVDV is mediated by apoptosis, the molecular mechanisms involved in pestiviral cytopathogenicity are poorly understood. In particular, the role of NS3 in triggering apoptosis remains unclear. It has been hypothesized that the NS3 serine proteinase might be involved in activation of the apoptotic proteolytic cascade (21, 55). Furthermore, it has been suggested that the NS3-mediated, enhanced viral RNA synthesis of cp BVDV and subsequently larger amounts of viral double-stranded RNAs may play a crucial role in triggering apoptosis (31, 54).In this study, we describe generation and characterization of a temperature-sensitive (ts) cp BVDV mutant whose ability to cause viral cytopathogenicity at high temperature is strongly attenuated. Our results demonstrate that a single amino acid substitution in NS2 attenuates BVDV cytopathogenicity at high temperature without affecting production of infectious viruses and expression of NS3 in a temperature-dependent manner.  相似文献   

16.
Hepatitis C virus (HCV) is a causative agent of chronic hepatitis, liver cirrhosis, and hepatocellular carcinoma. HCV in circulating blood associates with lipoproteins such as very low density lipoprotein (VLDL) and low-density lipoprotein (LDL). Although these associations suggest that lipoproteins are important for HCV infectivity, the roles of lipoproteins in HCV production and infectivity are not fully understood. To clarify the roles of lipoprotein in the HCV life cycle, we analyzed the effect of apolipoprotein E (ApoE), a component of lipoprotein, on virus production and infectivity. The production of infectious HCV was significantly reduced by the knockdown of ApoE. When an ApoE mutant that fails to be secreted into the culture medium was used, the amount of infectious HCV in the culture medium was dramatically reduced; the infectious HCV accumulated inside these cells, suggesting that infectious HCV must associate with ApoE prior to virus release. We performed rescue experiments in which ApoE isoforms were ectopically expressed in cells depleted of endogenous ApoE. The ectopic expression of the ApoE2 isoform, which has low affinity for the LDL receptor (LDLR), resulted in poor recovery of infectious HCV, whereas the expression of other isoforms, ApoE3 and ApoE4, rescued the production of infectious virus, raising it to an almost normal level. Furthermore, we found that the infectivity of HCV required both the LDLR and scavenger receptor class B, member I (SR-BI), ligands for ApoE. These findings indicate that ApoE is an essential apolipoprotein for HCV infectivity.Hepatitis C virus (HCV) infection is a major global health problem. More than 170 million people worldwide are infected with HCV. HCV causes chronic hepatitis, liver cirrhosis, and hepatocellular carcinoma (18). A member of the family Flaviviridae, HCV has a positive-sense, single-stranded RNA genome that is packaged into an enveloped viral particle. The genome encodes a large precursor polyprotein, which is cleaved by host and viral proteases to generate at least 10 functional viral proteins: core, envelope protein 1 (E1), E2, p7, nonstructural protein 2 (NS2), NS3, NS4A, NS4B, NS5A, and NS5B (12, 13). Core associates with the lipid droplet (LD). The role of this association remained elusive until robust HCV replication systems became available (32). We previously showed that the LD is an important organelle for HCV production (23). In hepatocytes, the LD is physiologically important as a lipid source for the production of lipoproteins such as very low density lipoprotein (VLDL) (11). VLDL is synthesized in the liver as a triglyceride/cholesterol ester-rich particle (diameter, 30 to 100 nm) surrounded by apolipoproteins such as apolipoprotein B100 (abbreviated as ApoB throughout), ApoC''s, and ApoE. VLDL is released into blood vessels to be delivered as a lipid source to peripheral cells, and it is also readsorbed by liver cells after processing (5).HCV particles circulating in the blood of HCV carriers associate with lipoproteins, such as low-density lipoprotein (LDL), VLDL, and chylomicrons; thus, these are termed lipo-viro particles (LVPs) (1, 26). Purified LVPs from circulating blood contain triglyceride, ApoB, ApoB48, ApoCII, ApoCIII, ApoE, and virus components such as HCV RNA and core (8), indicating that the LVP has dual viral and lipoprotein characteristics. The HCVcc strain, which contains a chimeric HCV-2a genome with a structural region from HCV-J6 and nonstructural/noncoding regions from an infectious JFH1 virus, can establish long-term infection in chimpanzees. Viruses recovered from the chimpanzee contain infectious virus particles with a slightly low density, suggesting that an in vivo association with low-density factors influences infectivity (19). However, the role of a lipoprotein-like component of LVPs in virus replication is not clear. Moreover, the mechanism by which LVPs are generated during HCV production is unknown.When HCV-producing cells are treated with an inhibitor of microsomal triglyceride transfer protein (MTP) or with ApoB-specific small interfering RNA (siRNA), the production of HCV particles is suppressed (10, 14, 25). Therefore, lipoprotein biosynthesis appears to play an important role in the production of infectious HCV and its egress from infected cells. ApoB, ApoC1, and ApoE associate with infectious virus particles in the HCVcc infection/replication system (4, 6, 15, 22, 27). Furthermore, ApoE depletion suppresses the production of infectious HCV (4, 6, 15, 27). These reports strongly suggest the importance of lipoprotein function to the HCV life cycle. However, the precise roles of lipoproteins and apolipoproteins in virus production and infectivity are not fully understood.We analyzed the production of HCV from cells in which apolipoprotein production was knocked down with siRNA. We found that ApoE is required for the infectivity of HCV, a finding consistent with other reports (4, 6, 15). ApoE is a polymorphic protein with three major isoforms: ApoE2, ApoE3, and ApoE4. The three isoforms differ by amino acid substitutions at one or two sites (residues 130 and 176) on the 317-amino-acid chain of the ApoE molecule. The polymorphism of ApoE influences its multiple functions due to isoform-dependent differences in receptor-binding activity and lipoprotein association preference. For example, ApoE2 has drastically lower LDL receptor (LDLR) binding activity than ApoE3 and ApoE4 (7). In the present study, we investigated the role of ApoE isoforms in virus production and infectivity.(Part of this study was presented at the 16th International Symposium on Hepatitis C Virus and Related Viruses, Nice, France, 3 to 7 October 2009.)  相似文献   

17.
We have recently demonstrated that human apolipoprotein E (apoE) is required for the infectivity and assembly of hepatitis C virus (HCV) (K. S. Chang, J. Jiang, Z. Cai, and G. Luo, J. Virol. 81:13783-13793, 2007; J. Jiang and G. Luo, J. Virol. 83:12680-12691, 2009). In the present study, we have determined the molecular basis underlying the importance of apoE in HCV assembly. Results derived from mammalian two-hybrid studies demonstrate a specific interaction between apoE and HCV nonstructural protein 5A (NS5A). The C-terminal third of apoE per se is sufficient for interaction with NS5A. Progressive deletion mutagenesis analysis identified that the C-terminal α-helix domain of apoE is important for NS5A binding. The N-terminal receptor-binding domain and the C-terminal 20 amino acids of apoE are dispensable for the apoE-NS5A interaction. The NS5A-binding domain of apoE was mapped to the middle of the C-terminal α-helix domain between amino acids 205 and 280. Likewise, deletion mutations disrupting the apoE-NS5A interaction resulted in blockade of HCV production. These findings demonstrate that the specific apoE-NS5A interaction is required for assembly of infectious HCV. Additionally, we have determined that using different major isoforms of apoE (E2, E3, and E4) made no significant difference in the apoE-NS5A interaction. Likewise, these three major isoforms of apoE are equally compatible with infectivity and assembly of infectious HCV, suggesting that apoE isoforms do not differentially modulate the infectivity and/or assembly of HCV in cell culture.Hepatitis C virus (HCV) remains a major global health problem, chronically infecting approximately 170 million people worldwide, with severe consequences such as hepatitis, fibrosis/cirrhosis, and hepatocellular carcinoma (HCC) (2, 57). The current standard therapy for hepatitis C is pegylated alpha interferon in combination with ribavirin. However, this anti-HCV regimen has limited efficacy (<50% sustained antiviral response for the dominant genotype 1 HCV) and causes severe side effects (17, 39). Recent clinical studies on the HCV protease- and polymerase-specific inhibitors showed promising results but also found that drug-resistant HCV mutants emerged rapidly (3, 27), undermining the efficacy of specific antiviral therapy for hepatitis C. Therefore, future antiviral therapies for hepatitis C likely require a combination of several safer and more efficacious antiviral drugs that target different steps of the HCV life cycle. The lack of knowledge about the molecular details of the HCV life cycle has significantly impeded the discovery of antiviral drugs and development of HCV vaccines.HCV is a small enveloped RNA virus classified as a member of the Hepacivirus genus in the family Flaviviridae (46, 47). It contains a single positive-sense RNA genome that encodes a large viral polypeptide, which is proteolytically processed by cellular peptidases and viral proteases into different structural and nonstructural proteins in the order of C, E1, E2, p7, NS2, NS3, NS4A, NS4B, NS5A, and NS5B (30, 31). Other novel viral proteins derived from the C-coding region have also been discovered (11, 13, 55, 59). The nucleotides at both the 5′ and 3′ untranslated regions (UTR) are highly conserved and contain cis-acting RNA elements important for internal ribosome entry site (IRES)-mediated initiation of protein translation and viral RNA replication (15, 16, 33, 56, 60).The success in the development of HCV replicon replication systems has made enormous contributions to the determination of the roles of the conserved RNA sequences/structures and viral NS proteins in HCV RNA replication (4, 5, 7, 32). However, the molecular mechanisms of HCV assembly, morphogenesis, and egression have not been well understood. A breakthrough advance has been the development of robust cell culture systems for HCV infection and propagation, which allow us to determine the roles of viral and cellular proteins in the HCV infectious cycle (9, 29, 54, 63). We have recently demonstrated that infectious HCV particles are enriched in apolipoprotein E (apoE) and that apoE is required for HCV infection and assembly (10, 23). apoE-specific monoclonal antibodies efficiently neutralized HCV infectivity. The knockdown of endogenous apoE expression by a specific small interfering RNA (siRNA) and the blockade of apoE secretion by microsomal triglyceride transfer protein (MTP) inhibitors remarkably suppressed HCV assembly (10, 23). More importantly, apoE was found to interact with the HCV NS5A in the cell and purified HCV particles, as determined by yeast two-hybrid and coimmunoprecipitation (co-IP) studies (6, 23). These findings suggest that apoE has dual functions in HCV infection and assembly via distinct interactions with cell surface receptors and HCV NS5A. To further understand the molecular mechanism of apoE in HCV assembly, we carried out a mutagenesis analysis of apoE and determined the importance of the apoE-NS5A interaction in HCV assembly. Progressive deletion mutagenesis analysis has mapped the NS5A-binding domain of apoE to the C-terminal α-helix region between amino acid residues 205 and 280. Mutations disrupting the apoE-NS5A interaction also blocked HCV production. Additionally, we have determined the effects of three major isoforms of apoE on HCV infection and assembly. Our results demonstrate that apoE isoforms do not determine the infectivity and assembly of infectious HCV in cell culture.  相似文献   

18.
19.
Naturally occurring hepatitis C virus (HCV) subgenomic RNAs have been found in several HCV patients. These subgenomic deletion mutants, mostly lacking the genes encoding envelope glycoproteins, were found in both liver and serum, where their relatively high abundance suggests that they are capable of autonomous replication and can be packaged and secreted in viral particles, presumably harboring the envelope proteins from wild type virus coinfecting the same cell. We recapitulated some of these natural subgenomic deletions in the context of the isolate JFH-1 and confirmed these hypotheses in vitro. In Huh-7.5 cells, these deletion-containing genomes show robust replication and can be efficiently trans-packaged and infect naïve Huh-7.5 cells when cotransfected with the full-length wild-type J6/JFH genome. The genome structure of these natural subgenomic deletion mutants was dissected, and the maintenance of both core and NS2 regions was proven to be significant for efficient replication and trans-packaging. To further explore the requirements needed to achieve trans-complementation, we provided different combinations of structural proteins in trans. Optimal trans-complementation was obtained when fragments of the polyprotein encompassing core to p7 or E1 to NS2 were expressed. Finally, we generated a stable helper cell line, constitutively expressing the structural proteins from core to p7, which efficiently supports trans-complementation of a subgenomic deletion encompassing amino acids 284 to 732. This cell line can produce and be infected by defective particles, thus representing a powerful tool to investigate the life cycle and relevance of natural HCV subgenomic deletion mutants in vivo.Hepatitis C virus (HCV) is an enveloped virus belonging to the family Flaviviridae. The virus genome is a positive-stranded RNA of about 9,600 nucleotides, which contains a single open reading frame (ORF) encoding both structural (core, E1, E2, and p7) and nonstructural (NS2, NS3, NS4A, NS4B, NS5A, and NS5B) proteins (47). Two highly conserved untranslated regions (UTRs) are found at the 5′ and 3′ ends, which play critical roles in both viral translation and replication (13, 14, 23).HCV is estimated to infect 170 million people worldwide (1, 43) and in a high percentage of individuals causes a chronic liver infection that frequently evolves into an array of diseases, including cirrhosis (12, 15, 38) and hepatocellular carcinoma (4, 7, 17, 30, 38, 49).The HCV RNA-dependent RNA polymerase (NS5B) has a high frequency of incorrect nucleotide insertions, in the range of 10−4 to 10−5 base substitutions per site, which can result in the rapid generation of HCV quasispecies (37, 45). Because of this huge genetic diversity, HCV is currently classified into six major genotypes and more than 80 subtypes (44). Recombination may be another mechanism exploited by HCV to increase genetic diversity: naturally occurring intergenotypic recombinant viruses that often have their recombination points in the trans-membrane domains of NS2 were recently identified (20, 21, 26, 27, 33, 35).Recent publications have reported the presence of natural HCV subgenomic RNAs in serum and liver of infected patients, mostly containing large in-frame deletions from E1 up to NS2, always found together with the full-length wild-type (wt) RNAs (5, 16, 36, 54). These mutant viral genomes persist for a long time (at least 2 years), and sequence analysis suggests that subgenomic (the predominant species during this period) and full-length HCV evolve independently (54). The relative abundance and persistence of such subgenomic RNAs in vivo suggests that (i) they are capable of autonomous replication and (ii) they can be packaged and secreted in infectious viral particles, presumably harboring the envelope proteins from wt virus coinfecting the same cell.Analysis of the genetic structure of 18 independent subgenomic deletion-containing RNAs (5, 16, 36, 54) strongly suggests that the possibility of recombination and/or deletion is restricted to specific regions.As expected, the 5′ UTR, the 3′ UTR, and the region coding from NS3 to NS5B are always conserved, in line with the notion that these regions are the minimal requirements for RNA replication. In addition to the regions required for RNA replication, however, naturally occurring subgenomic HCV RNA invariably contains an intact core region and the protease domain of NS2. In fact, the 5′ deletion boundary falls between the codons for amino acid 189 of core and amino acid 4 of E1 in 33% of cases and between the codons for amino acids 21 and 29 of E1 in 39% of cases, such that in 72% of cases overall, the 5′ boundary is between the codons for amino acid 189 of core and amino acid 29 of E1 (see Fig. S1 in the supplemental material). Likewise, the 3′ boundary shows a distinct localization, occurring between the codons for amino acids 51 and 79 of NS2 in 56% of cases.In the present study, we modified the infectious isolate JFH-1 in order to recapitulate in vitro the genetic structure of two of the most representative in-frame natural subgenomic deletion-containing RNAs found circulating in patients (28, 51, 57). Using this system, we analyzed natural subgenomic variants for their ability to replicate autonomously and demonstrated that the natural subgenomic deletion mutants are replication competent and are trans-packaged into infectious virions when coexpressed together with wt virus. Furthermore, our data suggest that the presence of the NS2 protease domain is required in order to generate the correct NS3 N terminus, required for RNA replication. Unexpectedly, the presence of NS2 generates, in turn, a strict cis requirement for the core region in order to allow efficient trans-packaging of the subgenomic RNA, revealing a complex interplay between the NS2 and the core viral genes. Finally, we performed trans-complementation studies of the subgenomic deletions with different HCV structural proteins to gain insight into the minimal requirements for the assembly and release of HCV virions.  相似文献   

20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号