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.
2.
Hepatitis C virus (HCV) entry occurs via a pH- and clathrin-dependent endocytic pathway and requires a number of cellular factors, including CD81, the tight-junction proteins claudin 1 (CLDN1) and occludin, and scavenger receptor class B member I (SR-BI). HCV tropism is restricted to the liver, where hepatocytes are tightly packed. Here, we demonstrate that SR-BI and CLDN1 expression is modulated in confluent human hepatoma cells, with both receptors being enriched at cell-cell junctions. Cellular contact increased HCV pseudoparticle (HCVpp) and HCV particle (HCVcc) infection and accelerated the internalization of cell-bound HCVcc, suggesting that the cell contact modulation of receptor levels may facilitate the assembly of receptor complexes required for virus internalization. CLDN1 overexpression in subconfluent cells was unable to recapitulate this effect, whereas increased SR-BI expression enhanced HCVpp entry and HCVcc internalization, demonstrating a rate-limiting role for SR-BI in HCV internalization.Hepatitis C virus (HCV) is an enveloped positive-strand RNA virus, classified in the genus Hepacivirus of the family Flaviviridae. Worldwide, approximately 170 million individuals are persistently infected with HCV, and the majority are at risk of developing chronic liver disease. Hepatocytes in the liver are thought to be the principal reservoir of HCV replication. HCV pseudoparticles (HCVpp) demonstrate a restricted tropism for hepatocyte-derived cells, suggesting that virus-encoded glycoprotein-receptor interactions play an important role in defining HCV tissue specificity.Recent evidence suggests that a number of host cell molecules are important for HCV entry: the tetraspanin CD81; scavenger receptor class B member I (SR-BI) (reviewed in reference 11); members of the tight-junction protein family claudin 1 (CLDN1), CLDN6, and CLDN9 (12, 34, 48, 52); and occludin (OCLN) (2, 33, 40). HCV enters cells via a pH- and clathrin-dependent endocytic pathway; however, the exact role(s) played by each of the host cell molecules in this process is unclear (4, 8, 21, 34, 45).CD81 and SR-BI interact with HCV-encoded E1E2 glycoproteins, suggesting a role in mediating virus attachment to the cell (reviewed in reference 44). In contrast, there is minimal evidence to support direct interaction of CLDN1 or OCLN with HCV particles (12). Evans and colleagues proposed that CLDN1 acts at a late stage in the entry process and facilitates fusion between the virus and host cell membranes (12). We (13, 19) and others (9, 48) have reported that CLDN1 associates with CD81, suggesting a role for CLDN1-CD81 complexes in viral entry. Cukierman et al. recently reported that CLDN1 enrichment at cell-cell contacts may generate specialized membrane domains that promote HCV internalization (9). In this study, we demonstrate that cellular contact modulates SR-BI and CLDN1 expression levels and promotes HCV internalization. CLDN1 overexpression in subconfluent cells was unable to recapitulate this effect, whereas increased SR-BI expression enhanced HCVpp entry and HCVcc internalization rates, demonstrating a critical and rate-limiting role for SR-BI in HCV internalization.  相似文献   

3.
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   

4.
Scavenger receptor class B type I (SR-BI) is an essential receptor for hepatitis C virus (HCV) and a cell surface high-density-lipoprotein (HDL) receptor. The mechanism of SR-BI-mediated HCV entry, however, is not clearly understood, and the specific protein determinants required for the recognition of the virus envelope are not known. HCV infection is strictly linked to lipoprotein metabolism, and HCV virions may initially interact with SR-BI through associated lipoproteins before subsequent direct interactions of the viral glycoproteins with SR-BI occur. The kinetics of inhibition of cell culture-derived HCV (HCVcc) infection with an anti-SR-BI monoclonal antibody imply that the recognition of SR-BI by HCV is an early event of the infection process. Swapping and single-substitution mutants between mouse and human SR-BI sequences showed reduced binding to the recombinant soluble E2 (sE2) envelope glycoprotein, thus suggesting that the SR-BI interaction with the HCV envelope is likely to involve species-specific protein elements. Most importantly, SR-BI mutants defective for sE2 binding, although retaining wild-type activity for receptor oligomerization and binding to the physiological ligand HDL, were impaired in their ability to fully restore HCVcc infectivity when transduced into an SR-BI-knocked-down Huh-7.5 cell line. These findings suggest a specific and direct role for the identified residues in binding HCV and mediating virus entry. Moreover, the observation that different regions of SR-BI are involved in HCV and HDL binding supports the hypothesis that new therapeutic strategies aimed at interfering with virus/SR-BI recognition are feasible.Hepatitis C virus (HCV) is a global blood-borne pathogen, with 3% of the world''s population chronically infected. Most infections are asymptomatic, yet 60 to 80% become persistent and lead to severe fibrosis and cirrhosis, hepatic failure, or hepatocellular carcinoma (3). Currently available therapies are limited to the administration of pegylated alpha interferon in combination with ribavirin, which are expensive and often unsuccessful, with significant side effects (23, 36). Thus, the development of novel therapeutic approaches against HCV remains a high priority (18, 40, 60). Targeting the early steps of HCV infection may represent one such option, and much effort is being devoted to uncovering the mechanism of viral attachment and entry.The current view is that HCV entry into target cells occurs after attachment to specific cellular receptors via its surface glycoproteins E1 and E2 (27). The molecules to which HCV initially binds might constitute a diverse collection of cellular proteins, carbohydrates, and lipids that concentrate viruses on the cell surface and determine to a large extent which cell types, tissues, and organisms HCV can infect.CD81, claudin 1 (CLDN1), occludin (OCLN), and scavenger receptor class B type I (SR-BI) were previously shown to play essential roles in HCV cell entry (15, 22, 26, 35, 42, 43, 50, 63, 64).Recent reports suggest that CD81 engagement triggers intracellular signaling responses, ultimately leading to actin remodeling and the relocalization of CD81 to tight junctions (TJ) (11). Thus, CD81 may function as a bridge between the initial interaction of the virus with receptors on the basolateral surface of the hepatocyte and the TJ where two of the HCV entry molecules, CLDN1 and OCLN, are located. CD81 acts as a postbinding factor, and the TJ proteins CLDN1 and OCLN seem to be involved in late steps of HCV entry, such as HCV glycoprotein-dependent cell fusion (9, 11, 22). The discovery of TJ proteins as entry factors has added complexity to the model of HCV entry, suggesting parallels with other viruses like coxsackievirus B infection, where an initial interaction of the viral particle with the primary receptor decay-accelerating factor induces the lateral movement of the virus from the luminal surface to TJ, where coxsackievirus B binds coxsackievirus-adenovirus receptor and internalization takes place (17).Much less is known about the specific role of SR-BI in virus entry: neither the specific step of the entry pathway that SR-BI is involved in nor the protein determinants that mediate such processes are known. SR-BI is a lipoprotein receptor of 509 amino acids (aa) with cytoplasmic C- and N-terminal domains separated by a large extracellular domain (1, 13, 14). It is expressed primarily in liver and steroidogenic tissues, where it mediates selective cholesteryl ester uptake from high-density lipoprotein (HDL) and may act as an endocytic receptor (45, 46, 51, 52). SR-BI was originally identified as being a putative receptor for HCV because it binds soluble E2 (sE2) through interactions with E2 hypervariable region 1 (HVR1) (8, 50). RNA interference studies as well as the ability to block both HCV pseudoparticles (HCVpp) and cell culture-derived HCV (HCVcc) infections with anti SR-BI antibodies have confirmed its involvement in the HCV entry process (7, 8, 15, 26, 33, 63). Intriguingly, lipoproteins were previously shown to modulate HCV infection through SR-BI (12). It was indeed previously demonstrated that two natural ligands of SR-BI, HDL and oxidized low-density lipoprotein, can improve and inhibit HCV entry, respectively (57, 59). Moreover, small-molecule inhibitors of SR-BI-mediated lipid transfer (block of lipid transfer BLT-3 and BLT-4) abrogate the stimulation of HCV infectivity by human serum or HDL, suggesting that the enhancement of viral infection might be dependent on the lipid exchange activity of SR-BI (20, 58).We previously generated high-affinity monoclonal antibodies (MAbs) specific for human SR-BI and showed that they were capable of inhibiting the binding of SR-BI to sE2 and blocking HCVcc infection of human hepatoma cells (15). The HDL-induced enhancement of infection had no impact on the ability of the anti-SR-BI MAbs to block HCV infection, and the antibodies were effective in counteracting HCV infection even in the absence of lipoproteins. These data demonstrated that SR-BI participates in the HCV infection process as an entry receptor by directly interacting with viral glycoproteins. Here we have used one of the anti-SR-BI MAbs to show that SR-BI participates in an early step of HCV infection. By assays of binding of sE2 to SR-BI molecules from different species and to SR-BI mutants, we identified species-specific SR-BI protein residues that are required for sE2 binding. The functional significance of these observations was confirmed by the finding that SR-BI mutants with reduced binding to sE2 were also impaired in their ability to restore the infectivity of an SR-BI-knocked-down Huh-7.5 cell line. Finally, we demonstrated that SR-BI mutants with impaired sE2 binding can still form oligomeric structures and that they can bind the physiological ligand HDL and mediate cholesterol efflux, suggesting that distinct protein determinants are responsible for the interaction with HDL and the HCV particle.  相似文献   

5.
The precise mechanisms regulating hepatitis C virus (HCV) entry into hepatic cells remain unknown. However, several cell surface proteins have been identified as entry factors for this virus. Of these molecules, claudin-1, a tight junction (TJ) component, is considered a coreceptor required for HCV entry. Recently, we have demonstrated that HCV envelope glycoproteins (HCVgp) promote structural and functional TJ alterations. Additionally, we have shown that the intracellular interaction between viral E2 glycoprotein and occludin, another TJ-associated protein, could be the cause of the mislocalization of TJ proteins. Herein we demonstrated, by using cell culture-derived HCV particles (HCVcc), that interference of occludin expression markedly reduced HCV infection. Furthermore, our results with HCV pseudotyped particles indicated that occludin, but not other TJ-associated proteins, such as junctional adhesion molecule A or zonula occludens protein 1, was required for HCV entry. Using HCVcc, we demonstrated that occludin did not play an essential role in the initial attachment of HCV to target cells. Surface protein labeling experiments showed that both expression levels and cell surface localization of HCV (co)receptors CD81, scavenger receptor class B type I, and claudin-1 were not affected upon occludin knockdown. In addition, immunofluorescence confocal analysis showed that occludin interference did not affect subcellular distribution of the HCV (co)receptors analyzed. However, HCVgp fusion-associated events were altered after occludin silencing. In summary, we propose that occludin plays an essential role in HCV infection and probably affects late entry events. This observation may provide new insights into HCV infection and related pathogenesis.Hepatitis C virus (HCV) is a small enveloped positive-strand RNA virus that belongs to the Flaviviridae family (20). More than 80% of acute infections become chronic, which eventually progress to cirrhosis and hepatocellular carcinoma (28). HCV infects mainly hepatocytes, but the precise mechanisms of infection are largely unknown (11). The HCV particle consists of a nucleocapsid surrounded by a lipid bilayer in which the two envelope glycoproteins (HCVgp), E1 and E2, are anchored as a heterodimer and play a major role in HCV entry (20). The development of an infectious cell culture model based on the production of infective HCV particles (cell culture-derived HCV particles [HCVcc]) (34) and the generation of HCV pseudotyped retroviral particles (HCVpp) (4) have provided powerful tools to study the HCV cycle. HCV entry is a complex multistep process that requires the presence of several factors. There are multiple pieces of evidence for the involvement of host cell proteins in HCV entry, including glycosaminoglycans, the low-density lipoprotein receptor, scavenger receptor class B type I (SR-BI), and the tetraspanin CD81 (11). Recently, claudin-1, a tight junction (TJ) component, has been identified as a coreceptor required for a late step in HCV entry (13).TJs are major components of cell-cell adhesion complexes and are composed of integral membrane proteins, including occludin and claudins, which associate with actin cytoskeleton-interacting proteins, such as zonula occludens protein 1 (ZO-1) (2). These structures maintain cell polarity, separating apical from basolateral membrane domains, and form a paracellular barrier that allows the selective passage of certain solutes (2). In hepatocytes, TJs seal the bile canaliculi and form the intercellular barrier between bile and blood (12). Recently, we have shown that TJ-associated proteins occludin and claudin-1 disappeared from their normal localization in both HCV-infected and genomic HCV replicon-containing Huh7 cells. Furthermore, TJ function was also altered in these cells (5). In this matter, we have reported an intracellular interaction between E2 and occludin (5). Moreover, it has been reported that claudin-1 and several TJ-associated proteins, such as coxsackievirus and adenovirus receptor (35) and junctional adhesion molecule (JAM) (3), act as virus (co)receptors. Since coxsackievirus entry across epithelial TJs requires occludin (10), we have explored the role of occludin in HCV infection.  相似文献   

6.
The variability of the hepatitis C virus (HCV), which likely contributes to immune escape, is most pronounced in hypervariable region 1 (HVR1) of viral envelope protein 2. This domain is the target for neutralizing antibodies, and its deletion attenuates replication in vivo. Here we characterized the relevance of HVR1 for virus replication in vitro using cell culture-derived HCV. We show that HVR1 is dispensable for RNA replication. However, viruses lacking HVR1 (ΔHVR1) are less infectious, and separation by density gradients revealed that the population of ΔHVR1 virions comprises fewer particles with low density. Strikingly, ΔHVR1 particles with intermediate density (1.12 g/ml) are as infectious as wild-type virions, while those with low density (1.02 to 1.08 g/ml) are poorly infectious, despite quantities of RNA and core similar to those in wild-type particles. Moreover, ΔHVR1 particles exhibited impaired fusion, a defect that was partially restored by an E1 mutation (I347L), which also rescues infectivity and which was selected during long-term culture. Finally, ΔHVR1 particles were no longer neutralized by SR-B1-specific immunoglobulins but were more prone to neutralization and precipitation by soluble CD81, E2-specific monoclonal antibodies, and patient sera. These results suggest that HVR1 influences the biophysical properties of released viruses and that this domain is particularly important for infectivity of low-density particles. Moreover, they indicate that HVR1 obstructs the viral CD81 binding site and conserved neutralizing epitopes. These functions likely optimize virus replication, facilitate immune escape, and thus foster establishment and maintenance of a chronic infection.Hepatitis C virus (HCV) is a single-stranded positive-sense RNA virus of the family Flaviviridae that has infected an estimated 130 million people worldwide (1). Acute HCV infection is mostly asymptomatic; however, virus persistence can lead to severe liver disease, and within 20 years ca. 20% of chronically infected adults develop cirrhosis (46). In fact, morbidity associated with chronic HCV infection is the most common indication for orthotopic liver transplantation (7). The mechanisms that permit the virus to establish chronic infection in ca. 55 to 85% of cases (24) despite vigorous immune responses are incompletely understood.A number of studies have highlighted the pivotal role of strong, multispecific, and sustained T-cell responses for control of HCV infection (summarized in reference 53). Although resolution of acute HCV infection can occur in the absence of antibodies (47), mounting evidence indicates that neutralizing antibodies also contribute to protective immunity (summarized in reference 62). Nevertheless, HCV often successfully evades cellular and humoral immune pressure likely at least in part via the constant generation of variants created by an error-prone RNA replication machinery. In line with this notion, a high degree of HCV sequence evolution is associated with chronic disease, while a comparatively static pool of variants correlates with resolution (13, 15, 43).Virus isolates from patients are classified into at least 7 different genetic groups (genotypes [GTs]), which differ from each other by ca. 31 to 33% at the nucleotide level (20, 48). However, genetic variability is not equally distributed across the HCV genome, which encodes a large polyprotein of ca. 3,000 amino acids and contains 5′- and 3′-terminal nontranslated regions (NTR) required for RNA replication. More specifically, the 5′ NTR and the terminal 99 bases of the 3′ NTR are most conserved, while the N-terminal 27 amino acids of the envelope glycoprotein 2 (E2), called HVR1, are most divergent among HCV isolates (48). Notably, HVR1 contains epitopes which are recognized by patients'' antibodies (28, 29, 51, 59) and by antibodies that neutralize infection of chimpanzees (14). Moreover, during an acute infection, sequence changes occur almost exclusively within this region, and these are temporally correlated with antibody seroconversion (13). Therefore, the pronounced variability of this portion of E2 is likely due to strong humoral immune pressure, which drives its rapid evolution. However, variability of HVR1 is not random, as the chemicophysical properties and the conformation of this basic domain are well conserved (39). These findings suggest functional constraints for the evolution of HVR1, and the exposure of this epitope on the surface of HCV particles argues for an important role of this domain during virus entry.In line with this assumption, Forns et al. observed that an HCV mutant lacking HVR1 (ΔHVR1) was infectious for chimpanzees but clearly attenuated (17). Interestingly, an increase in titers of the ΔHVR1 virus coincided with emergence of two mutations in the ectodomain of E2, suggesting that these changes may have compensated for a putative functional impairment of the mutant (17).The development of retroviral particles which carry HCV glycoproteins on their surfaces (HCV pseudoparticles [HCVpp]) and, more recently, cell culture-derived HCV (HCVcc) based on the JFH1 strain provides robust models for dissecting the mechanisms of HCV entry in vitro (3, 25, 35, 58, 64). By means of these systems, the tetraspanin CD81, the lipoprotein receptor SR-BI, and tight junction proteins claudin-1 and occludin were identified as essential host factors for HCV infection (3, 5, 12, 25, 41, 42, 45).Moreover, it was recognized that there is a complex interplay between HCV and lipoproteins. Specifically, high-density lipoprotein (HDL) and oxidized low-density lipoprotein (oxLDL), both ligands of SR-BI, modulate HCVpp infection in an SR-BI-dependent fashion (4, 56, 57). Of note, HVR1 seems to be involved in SR-BI-mediated entry of HCVpp (5), since deletion of this domain ablated stimulation of HCVpp infection by HDL and rendered the virus resistant to inhibition by SR-BI-specific antibodies (4, 5), which prevent infection of HCVpp carrying wild-type HCV glycoproteins. Finally, HCVpp lacking HVR1 are more susceptible to neutralization by patient serum-derived immunoglobulins (4). Thus, altogether these results indicated an important role for HVR1 in viral fitness, likely due to an involvement in HCV entry via SR-BI, and in the interaction of HCV with the humoral immune system. Despite these important observations in the HCVpp system, the role of HVR1 in infection by authentic HCV particles was not defined. In addition, it was unclear if HCVpp produced in 293T cells that are unable to produce lipoproteins reflect natural HCV particles with regard to HVR1 function.Therefore, to better understand the role of HVR1 for virus replication and immune evasion, in this study we analyzed the importance of HVR1 for virus replication and neutralization using authentic, cell culture-derived HCV. We dissected the influence of this domain on HCV receptor interactions and membrane fusion and investigated compensatory mechanisms that permit the virus to regain fitness after deletion of HVR1.  相似文献   

7.
Broadly neutralizing antibodies are commonly present in the sera of patients with chronic hepatitis C virus (HCV) infection. To elucidate possible mechanisms of virus escape from these antibodies, retrovirus particles pseudotyped with HCV glycoproteins (HCVpp) isolated from sequential samples collected over a 26-year period from a chronically infected patient, H, were used to characterize the neutralization potential and binding affinity of a panel of anti-HCV E2 human monoclonal antibodies (HMAbs). Moreover, AP33, a neutralizing murine monoclonal antibody (MAb) to a linear epitope in E2, was also tested against selected variants. The HMAbs used were previously shown to broadly neutralize HCV and to recognize a cluster of highly immunogenic overlapping epitopes, designated domain B, containing residues that are also critical for binding of viral E2 glycoprotein to CD81, a receptor essential for virus entry. Escape variants were observed at different time points with some of the HMAbs. Other HMAbs neutralized all variants except for the isolate 02.E10, obtained in 2002, which was also resistant to MAb AP33. The 02.E10 HCVpp that have reduced binding affinities for all antibodies and for CD81 also showed reduced infectivity. Comparison of the 02.E10 nucleotide sequence with that of the strain H-derived consensus variant, H77c, revealed the former to have two mutations in E2, S501N and V506A, located outside the known CD81 binding sites. Substitution A506V in 02.E10 HCVpp restored binding to CD81, but its antibody neutralization sensitivity was only partially restored. Double substitutions comprising N501S and A506V synergistically restored 02.E10 HCVpp infectivity. Other mutations that are not part of the antibody binding epitope in the context of N501S and A506V were able to completely restore neutralization sensitivity. These findings showed that some nonlinear overlapping epitopes are more essential than others for viral fitness and consequently are more invariant during earlier years of chronic infection. Further, the ability of the 02.E10 consensus variant to escape neutralization by the tested antibodies could be a new mechanism of virus escape from immune containment. Mutations that are outside receptor binding sites resulted in structural changes leading to complete escape from domain B neutralizing antibodies, while simultaneously compromising viral fitness by reducing binding to CD81.Over 170 million people worldwide are infected with hepatitis C virus (HCV). While acute infection is usually silent, the majority of infected individuals develop persistent infections. Approximately 30% of acute infections are spontaneously resolved. Cellular immunity is clearly necessary, as robust and sustained CD4+ and CD8+ T-cell responses are temporally associated with virus clearance leading to disease resolution (7). Persistent infection is associated with an inability to sustain a vigorous CD4+ response. The role of antibodies in disease resolution is increasingly recognized but less understood. Clinical trials with gamma globulin administration prior to the discovery of HCV achieved prophylactic effects on transfusion-associated non-A, non-B hepatitis cases, most of which were subsequently shown to be HCV related (28, 46). Animal studies showed that gamma globulin therapy delayed the onset of acute HCV infection (29). Preincubation of the infectious inoculum with pooled gamma globulin from HCV-positive donors prevented infection in challenged chimpanzees (55). The protection afforded by gamma globulin preparations correlated with antibody titers blocking infection of target cells with retroviral pseudotype particles expressing HCV E1E2 glycoproteins (HCVpp) (4). In addition, chimpanzees vaccinated with recombinant HCV E2 glycoproteins were protected against infection in a manner that correlated with serum antibody titers inhibiting binding of E2 to CD81 (19, 40, 41), a receptor required for entry by both HCVpp and cell culture infectious HCV (HCVcc) (5, 17, 33, 53, 56). Two recent studies observed that patients with strong and progressive neutralizing antibody responses demonstrated decreasing viremia and control of viral replication (31, 39). A third study, however, reported the lack of neutralizing antibodies to heterologous HCVpp isolates in the sera of patients who eventually controlled their viremia during acute HCV infection (21). Furthermore, 104 to 106 virions per milliliter of serum are usually detected during chronic infection in the presence of high titers of serum neutralizing antibodies.A driver of persistent viremia is a high degree of viral variants, or “quasispecies.” Owing to a high viral replication rate (1012 copies per day) and an error-prone viral RNA-dependent polymerase, the estimated mutation rate is 2.0 × 10−3 base substitutions per genome per year (9, 34). This high rate of quasispecies formation contributes to the emergence of escape viral variants from immune surveillance. Mutations within major histocompatibility complex class I-restricted HCV epitopes lead to escape from cytotoxic T-cell responses (7). Mutations leading to escape from humoral immunity, particularly in E2 hypervariable region 1 (HVR1), known to be the target of host neutralizing antibodies, are also documented (10, 22, 30, 45). Protection in chimpanzees is achieved following challenge with an inoculum that had been preincubated with antibodies to autologous HVR1 (10). Yet over time, these isolate-specific antibodies drive the emergence of new viral variants that the concurrent immune response poorly recognizes. A study of sequential HCV isolates obtained from a patient, H, who was meticulously followed for a 26-year period starting 3 weeks after exposure to the virus, showed that the serial HCV variants were poorly neutralized by the concurrent serum antibodies (52). Escape was associated in part with mutations in HVR1 leading to decreased binding and neutralization by monoclonal antibodies (MAbs) to HVR1 that were produced against the first isolate obtained from this patient.Broadly neutralizing antibodies are usually directed against conformational epitopes within E2 (2, 8, 13, 14, 44). We previously described a panel of neutralizing and nonneutralizing human MAbs (HMAbs) to conformational epitopes on HCV E2 that were derived from peripheral B cells of individuals infected with either genotype 1a or 1b HCV. Cross-competition analyses delineated at least three immunogenic clusters of overlapping epitopes with distinct functions and properties (23-25). All nonneutralizing antibodies fell within one cluster, designated domain A (24). Neutralizing HMAbs segregated into two clusters, designated domains B and C, with domain B HMAbs having greater potency than domain C HMAbs in blocking infection with the strain JFH1 genotype 2a HCVcc (23, 25).The epitopes of increasing numbers of anti-HCV E2 neutralizing antibodies include residues that are also critical for binding of E2 to CD81. All of our domain B HMAbs inhibit binding of E2 to CD81. Alanine scanning mutagenesis of E2 regions implicated in binding to CD81 identified two highly conserved residues, G530 and D535, that are needed for all domain B antibodies, with a subset also requiring W529 (25, 26, 36). Other laboratories have isolated similar neutralizing antibodies to epitopes containing these residues (20, 32, 38). A similar panel of E2 mutants was previously used to identify five amino acid residues, W420, Y527, W529, G530, and D535, that are essential for interaction with CD81 (37, 42). These findings show that domain B antibodies exert their potent neutralization of HCV infectivity by directly competing with CD81 for binding to E2. It also explains the breadth of neutralization against different HCV genotypes and subtypes for many of these antibodies, since any changes in their epitopes could affect CD81 binding and virus entry. The conserved nature of this cluster of overlapping epitopes makes them of interest for vaccine and immunotherapeutic development. A critical question involves the likelihood that immune selection could lead to escape from neutralization by domain B HMAbs. The series of sequential HCVpp variants derived from patient H over a span of 26 years (52) provide a unique resource for studying the extent and mechanisms of virus escape from broadly neutralizing antibodies. This report describes evidence of escape from immune containment of some but not other domain B HMAbs. Interestingly, a single H variant with reduced HCVpp infectivity and diminished CD81 binding was resistant to neutralization by all domain B antibodies as well as MAb AP33, recognizing a highly conserved linear epitope spanning residues 413 to 420 (35, 47). Sequence analysis revealed multiple mutations on E2 at a considerable distance from CD81 binding residues that could account for the immune escape, although it is unlikely that they are part of the domain B HMAb or the AP33 epitopes. Site-directed substitutions at these mutations restored neutralization sensitivity to all antibodies and CD81 dependency.  相似文献   

8.
Recently, claudin-1 (CLDN1) was identified as a host protein essential for hepatitis C virus (HCV) infection. To evaluate CLDN1 function during virus entry, we searched for hepatocyte cell lines permissive for HCV RNA replication but with limiting endogenous CLDN1 expression, thus permitting receptor complementation assays. These criteria were met by the human hepatoblastoma cell line HuH6, which (i) displays low endogenous CLDN1 levels, (ii) efficiently replicates HCV RNA, and (iii) produces HCV particles with properties similar to those of particles generated in Huh-7.5 cells. Importantly, naïve cells are resistant to HCV genotype 2a infection unless CLDN1 is expressed. Interestingly, complementation of HCV entry by human, rat, or hamster CLDN1 was highly efficient, while mouse CLDN1 (mCLDN1) supported HCV genotype 2a infection with only moderate efficiency. These differences were observed irrespective of whether cells were infected with HCV pseudoparticles (HCVpp) or cell culture-derived HCV (HCVcc). Comparatively low entry function of mCLDN1 was observed in HuH6 but not 293T cells, suggesting that species-specific usage of CLDN1 is cell type dependent. Moreover, it was linked to three mouse-specific residues in the second extracellular loop (L152, I155) and the fourth transmembrane helix (V180) of the protein. These determinants could modulate the exposure or affinity of a putative viral binding site on CLDN1 or prevent optimal interaction of CLDN1 with other human cofactors, thus precluding highly efficient infection. HuH6 cells represent a valuable model for analysis of the complete HCV replication cycle in vitro and in particular for analysis of CLDN1 function in HCV cell entry.Hepatitis C virus (HCV) is a liver-tropic plus-strand RNA virus of the family Flaviviridae that has chronically infected about 130 million individuals worldwide. During long-term persistent virus replication, many patients develop significant liver disease which can lead to cirrhosis and hepatocellular carcinoma (54). Current treatment of chronic HCV infection consists of a combination of pegylated alpha interferon and ribavirin. However, this regimen is not curative for all treated patients and is associated with severe side effects (37). Therefore, an improved therapy is needed and numerous HCV-specific drugs targeting viral enzymes are currently being developed (47). These efforts have been slowed down by a lack of small-animal models permissive for HCV replication since HCV infects only humans and chimpanzees. Among small animals, only immunodeficient mice suffering from a transgene-induced disease of endogenous liver cells and repopulated with human primary hepatocytes are susceptible to HCV infection (39).The restricted tropism of HCV likely reflects very specific host factor requirements for entry, RNA replication, assembly, and release of virions. Although HCV RNA replication has been observed in nonhepatic human cells and even nonhuman cells, its efficiency is rather low (2, 11, 59, 67). In addition, so far, efficient production of infectious particles has only been reported with Huh-7 human hepatoma cells, Huh-7-derived cell clones, and LH86 cells (33, 61, 65, 66). Although murine cells sustain HCV RNA replication, they do not produce detectable infectious virions (59). Together, these results suggest that multiple steps of the HCV replication cycle may be blocked or impaired in nonhuman or nonhepatic cells.HCV entry into host cells is complex and involves interactions between viral surface-resident glycoproteins E1 and E2 and multiple host factors. Initial adsorption to the cell surface is likely facilitated by interaction with attachment factors like glycosaminoglycans (4, 31) and lectins (13, 35, 36, 51). Beyond these, additional host proteins have been implicated in HCV entry. Since HCV circulates in the blood associated with lipoproteins (3, 43, 57), it has been postulated that HCV enters hepatocytes via the low-density lipoprotein receptor (LDL-R), and evidence in favor of an involvement of LDL-R has been provided (1, 40, 42, 44). Direct interactions between soluble E2 and scavenger receptor class B type I (SR-BI) (53) and CD81 (49) have been reported, and firm experimental proof has accumulated that these host proteins are essential for HCV infection (5, 6, 16, 26, 28, 33, 41, 61). Finally, more recently, claudin-1 (CLDN1) and occludin, two proteins associated with cellular tight junctions, have been identified as essential host factors for infection (20, 34, 50) and an interaction between E2 and these proteins, as revealed by coimmunoprecipitation assays, was reported (7, 34, 63). Although the precise functions of the individual cellular proteins during HCV infection remain poorly defined, based on kinetic studies with antibodies blocking interactions with SR-BI, CD81, or CLDN1, these factors are likely required subsequent to viral attachment (14, 20, 31, 64). Interestingly, viral resistance to antibodies directed against CLDN1 seems to be slightly delayed compared to resistance to antibodies directed against CD81 and SR-BI (20, 64), suggesting that there may be a sequence of events with the virus encountering first SR-BI and CD81 and subsequently CLDN1. Moreover, in Huh-7 cells, engagement of CD81 by soluble E1/E2 induces Rho GTPase-dependent relocalization of these complexes to areas of cell-to-cell contact, where these colocalized with CLDN1 and occludin (9). Together, these findings are consistent with a model where HCV reaches the basolateral, sinusoid-exposed surface of hepatocytes via the circulation. Upon binding to attachment factors SR-BI and CD81, which are highly expressed in this domain (52), the HCV-receptor complex may be ferried to tight-junction-resident CLDN1 and occludin and finally be endocytosed in a clathrin-dependent fashion (8, 38). Once internalized, the viral genome is ultimately delivered into the cytoplasm through a pH-dependent fusion event (24, 26, 31, 58). Recently, Ploss et al. reported that expression of human SR-BI, CD81, CLDN1, and occludin was sufficient to render human and nonhuman cells permissive for HCV infection (50). These results indicate that these four factors are the minimal cell type-specific set of host proteins essential for HCV entry. Interestingly, HCV seems to usurp at least CD81 and occludin in a very species-specific manner since their murine orthologs permit HCV infection with limited efficiency only (22, 50). Recently, it was shown that expression of mouse SR-BI did not fully restore entry function in Huh-7.5 cells with knockdown of endogenous human SR-BI, suggesting that also SR-BI function in HCV entry is, to some extent, species specific (10).In this study, we have developed a receptor complementation system for CLDN1 that permits the assessment of functional properties of this crucial HCV host factor with cell culture-derived HCV (HCVcc) and a human hepatocyte cell line. This novel model is based on HuH6 cells, which were originally isolated from a male Japanese patient suffering from a hepatoblastoma (15). These cells express little endogenous CLDN1, readily replicate HCV RNA, and produce high numbers of infectious HCVcc particles with properties comparable to those of Huh-7 cell-derived HCV. In addition, we identified three mouse-typic residues of CLDN1 that limit receptor function in HuH6 cells. These results suggest that besides CD81 and occludin, and to a minor degree SR-BI, CLDN1 also contributes to the restricted species tropism of HCV.  相似文献   

9.
Nearly all livers transplanted into hepatitis C virus (HCV)-positive patients become infected with HCV, and 10 to 25% of reinfected livers develop cirrhosis within 5 years. Neutralizing monoclonal antibody could be an effective therapy for the prevention of infection in a transplant setting. To pursue this treatment modality, we developed human monoclonal antibodies (HuMAbs) directed against the HCV E2 envelope glycoprotein and assessed the capacity of these HuMAbs to neutralize a broad panel of HCV genotypes. HuMAb antibodies were generated by immunizing transgenic mice containing human antibody genes (HuMAb mice; Medarex Inc.) with soluble E2 envelope glycoprotein derived from a genotype 1a virus (H77). Two HuMAbs, HCV1 and 95-2, were selected for further study based on initial cross-reactivity with soluble E2 glycoproteins derived from genotypes 1a and 1b, as well as neutralization of lentivirus pseudotyped with HCV 1a and 1b envelope glycoproteins. Additionally, HuMAbs HCV1 and 95-2 potently neutralized pseudoviruses from all genotypes tested (1a, 1b, 2b, 3a, and 4a). Epitope mapping with mammalian and bacterially expressed proteins, as well as synthetic peptides, revealed that HuMAbs HCV1 and 95-2 recognize a highly conserved linear epitope spanning amino acids 412 to 423 of the E2 glycoprotein. The capacity to recognize and neutralize a broad range of genotypes, the highly conserved E2 epitope, and the fully human nature of the antibodies make HuMAbs HCV1 and 95-2 excellent candidates for treatment of HCV-positive individuals undergoing liver transplantation.Hepatitis C virus (HCV) is a major cause of liver failure and infects more than 170 million people worldwide. HCV is a member of the Flaviviridae family and contains a 9.6-kb positive-strand RNA genome. The genome is translated into a single polypeptide that is cleaved by viral and cellular proteases into at least nine different proteins. The major HCV surface glycoproteins, E1 and E2, form a noncovalent heterodimer on the virion surface (23) and are believed to mediate viral entry via a complex set of poorly understood interactions with cellular coreceptors, including CD81 (28), claudin-1 (8), occludin (29), scavenger receptor class B type I (30), and others (38). The E2 glycoprotein has been shown to interact directly with receptors (38); currently, no function has been assigned to E1, although it is known to be required for viral infection. These viral glycoproteins provide an obvious target for neutralizing monoclonal antibodies (MAbs).Isolation of potently neutralizing HCV-specific MAbs has been complicated by the lack of an in vitro cell culture system to study the full infection cycle of the virus. Recently, systems have been developed that allow for the generation of infectious viral particles, highlighting the importance of E1 and E2 in viral binding and entry. A novel in vitro infection system employs HCV pseudotyped viral particles (HCVpp) generated from a lentivirus that are devoid of native glycoproteins and engineered to contain HCV glycoproteins E1 and E2 (4, 15). HCVpp specifically infect cell lines derived from human liver cells and can be neutralized by polyclonal and MAbs directed against the HCV envelope glycoproteins.HCVpp have allowed the identification of antibodies that can neutralize HCV infection in cell culture. E1 has proven to be a difficult target for MAb-mediated neutralization, possibly because it appears to have low immunogenicity (32), has no identified binding proteins on the cell surface, and has an undefined role in cell entry. Despite this challenge, two groups have identified HCV neutralizing MAbs directed to E1: these MAbs are H-111, which has moderate neutralizing activity (17), and the recently isolated IGH505 and IGH526, which neutralize numerous HCV genotypes (1a, 1b, 2a, 4a, 5a, and 6a but not 2b and 3a) (22). Although they are predicted to inhibit viral binding or fusion, the mechanism by which these E1-directed MAbs neutralize HCV infection is unclear.A diverse group of mouse anti-E2 antibodies, recognizing both linear and discontinuous epitopes, has been generated. Many of these MAbs showed broad neutralization of multiple HCV genotypes, but not surprisingly, several HCV isolates were refractory to neutralization. In contrast, AP33, a mouse MAb that largely recognizes a highly conserved linear epitope in the N terminus of E2 (amino acids 412 to 423), was identified as a broadly cross-reactive antibody that neutralized strains from all genotypes tested (1a, 1b, 2a, 2b, 3a, 4, 5, and 6), with the exception of one genotype 5 virus (UKN5.14.4; GenBank accession no. AY894682) (24). Recently, several cross-reactive neutralizing MAbs have been identified that are of human origin and have the capacity to neutralize a significant fraction of the genotypes tested (1, 5, 12, 13, 27, 31) or to neutralize all genotypes tested (16, 20, 25). As with the vast majority of previously described human MAbs (HuMAbs), these MAbs recognize conformation-dependent epitopes of E2. One broadly neutralizing human antibody, AR3B, was tested in a mouse model of infection and showed significant protection from viremia (20). Given the known function of the E2 envelope glycoprotein, the high level of immunogenicity, the surface vulnerability, and the abundance of data pertaining to E2 and HCV neutralization, E2 provides a promising target for the development of fully human neutralizing antibodies.Liver deterioration due to HCV infection is the leading reason for liver transplantation in the United States. Unfortunately, it is highly likely that the transplanted liver will also become infected with HCV, and 10 to 25% of these patients develop cirrhosis within 5 years of transplant (9, 40). Here we describe the characterization of HuMAbs directed against the HCV E2 envelope glycoprotein, generated using transgenic mice. Based on epitope conservation and broad neutralization capacity, HuMAbs HCV1 and 95-2 provide excellent candidates for prevention of graft reinfection of HCV-infected individuals undergoing liver transplantation.  相似文献   

10.
Claudin-1, a component of tight junctions between liver hepatocytes, is a hepatitis C virus (HCV) late-stage entry cofactor. To investigate the structural and functional roles of various claudin-1 domains in HCV entry, we applied a mutagenesis strategy. Putative functional intracellular claudin-1 domains were not important. However, we identified seven novel residues in the first extracellular loop that are critical for entry of HCV isolates drawn from six different subtypes. Most of the critical residues belong to the highly conserved claudin motif W30-GLW51-C54-C64. Alanine substitutions of these residues did not impair claudin-1 cell surface expression or lateral protein interactions within the plasma membrane, including claudin-1-claudin-1 and claudin-1-CD81 interactions. However, these mutants no longer localized to cell-cell contacts. Based on our observations, we propose that cell-cell contacts formed by claudin-1 may generate specialized membrane domains that are amenable to HCV entry.Hepatitis C virus (HCV) is a major human pathogen that affects approximately 3% of the global population, leading to cirrhosis and hepatocellular carcinoma in chronically infected individuals (5, 23, 42). Hepatocytes are the major target cells of HCV (11), and entry follows a complex cascade of interactions with several cellular factors (6, 8, 12, 17). Infectious viral particles are associated with lipoproteins and initially attach to target cells via glycosaminoglycans and the low-density lipoprotein receptor (1, 7, 31). These interactions are followed by direct binding of the E2 envelope glycoprotein to the scavenger receptor class B type I (SR-B1) and then to the CD81 tetraspanin (14, 15, 33, 36). Early studies showed that CD81 and SR-B1 were necessary but not sufficient for HCV entry, and claudin-1 was discovered to be a requisite HCV entry cofactor that appears to act at a very late stage of the process (18).Claudin-1 is a member of the claudin protein family that participates in the formation of tight junctions between adjacent cells (25, 30, 37). Tight junctions regulate the paracellular transport of solutes, water, and ions and also generate apical-basal cell polarity (25, 37). In the liver, the apical surfaces of hepatocytes form bile canaliculi, whereas the basolateral surfaces face the underside of the endothelial layer that lines liver sinusoids. Claudin-1 is highly expressed in tight junctions formed by liver hepatocytes as well as on all hepatoma cell lines that are permissive to HCV entry (18, 24, 28). Importantly, nonhepatic cell lines that are engineered to express claudin-1 become permissive to HCV entry (18). Claudin-6 and -9 are two other members of the human claudin family that enable HCV entry into nonpermissive cells (28, 43).The precise role of claudin-1 in HCV entry remains to be determined. A direct interaction between claudins and HCV particles or soluble E2 envelope glycoprotein has not been demonstrated (18; T. Dragic, unpublished data). It is possible that claudin-1 interacts with HCV entry receptors SR-B1 or CD81, thereby modulating their ability to bind to E2. Alternatively, claudin-1 may ferry the receptor-virus complex to fusion-permissive intracellular compartments. Recent studies show that claudin-1 colocalizes with the CD81 tetraspanin at the cell surface of permissive cell lines (22, 34, 41). With respect to nonpermissive cells, one group observed that claudin-1 was predominantly intracellular (41), whereas another reported associations of claudin-1 and CD81 at the cell surface, similar to what is observed in permissive cells (22).Claudins comprise four transmembrane domains along with two extracellular loops and two cytoplasmic domains (19, 20, 25, 30, 37). The first extracellular loop (ECL1) participates in pore formation and influences paracellular charge selectivity (25, 37). It has been shown that the ECL1 of claudin-1 is required for HCV entry (18). All human claudins comprise a highly conserved motif, W30-GLW51-C54-C64, in the crown of ECL1 (25, 37). The exact function of this domain is unknown, and we hypothesized that it is important for HCV entry. The second extracellular loop is required for the holding function and oligomerization of the protein (25). Claudin-1 also comprises various signaling domains and a PDZ binding motif in the intracellular C terminus that binds ZO-1, another major component of tight junctions (30, 32, 37). We further hypothesized that some of these domains may play a role in HCV entry.To understand the role of claudin-1 in HCV infection, we developed a mutagenesis strategy targeting the putative sites for internalization, glycosylation, palmitoylation, and phosphorylation. The functionality of these domains has been described by others (4, 16, 25, 35, 37, 40). We also mutagenized charged and bulky residues in ECL1, including all six residues within the highly conserved motif W30-GLW51-C54-C64. None of the intracellular domains were found to affect HCV entry. However, we identified seven residues in ECL1 that are critical for entry mediated by envelope glycoproteins derived from several HCV subtypes, including all six residues of the conserved motif. These mutants were still expressed at the cell surface and able to form lateral homophilic interactions within the plasma membrane as well as to engage in lateral interactions with CD81. In contrast, they no longer engaged in homophilic trans interactions at cell-cell contacts. We conclude that the highly conserved motif W30-GLW51-C54-C64 of claudin-1 is important for HCV entry into target cells and participates in the formation of cell-cell contacts.  相似文献   

11.
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   

12.
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   

13.
Hepatitis C virus (HCV) is a leading cause of liver disease worldwide. As HCV infects only human and chimpanzee cells, antiviral therapy and vaccine development have been hampered by the lack of a convenient small-animal model. In this study we further investigate how the species tropism of HCV is modulated at the level of cell entry. It has been previously determined that the tight junction protein occludin (OCLN) is essential for HCV host cell entry and that human OCLN is more efficient than the mouse ortholog at mediating HCV cell entry. To further investigate the relationship between OCLN sequence and HCV species tropism, we compared OCLN proteins from a range of species for their ability to mediate infection of naturally OCLN-deficient 786-O cells with lentiviral pseudoparticles bearing the HCV glycoproteins. While primate sequences function equivalently to human OCLN, canine, hamster, and rat OCLN had intermediate activities, and guinea pig OCLN was completely nonfunctional. Through analysis of chimeras between these OCLN proteins and alanine scanning mutagenesis of the extracellular domains of OCLN, we identified the second half of the second extracellular loop (EC2) and specific amino acids within this domain to be critical for modulating the HCV cell entry factor activity of this protein. Furthermore, this critical region of EC2 is flanked by two conserved cysteine residues that are essential for HCV cell entry, suggesting that a subdomain of EC2 may be defined by a disulfide bond.Hepatitis C virus (HCV), a member of the family Flaviviridae, is the causative agent of classically defined non-A, non-B hepatitis and is highly prevalent, with approximately 3% of the worldwide population infected (48). HCV infection often results in a chronic, life-long infection that can have severe health consequences, including hepatitis, cirrhosis, hepatocellular carcinoma, and liver failure. There is no HCV vaccine available, and the currently employed interferon-based treatment is inadequate as it has severe side effects and is effective only in half of the major genotype-infected individuals (22, 32). Specific anti-HCV inhibitors targeting the viral proteases and polymerase are currently being developed and will likely improve therapeutic options substantially. Undoubtedly, however, the emergence of viral resistance to such inhibitors will be a problem facing future HCV treatment options. As such, developing a spectrum of inhibitors targeting diverse steps in the virus life cycle, including HCV cell entry, is a priority for HCV research. Such inhibitors may be particularly useful following liver transplantation. Although HCV is the leading cause of liver transplants worldwide (10), the usefulness of such procedures is limited by subsequent universal graft reinfection and often accelerated disease progression (21). Even transiently inhibiting graft reinfection with HCV cell entry inhibitors could greatly improve the effectiveness of this procedure. Therefore, a greater understanding of HCV cell entry is required for the development of therapies targeting this stage of the viral life cycle.HCV host cell entry is a complex process that culminates in the clathrin-dependent endocytosis of the virion and low-pH-mediated fusion of viral and cellular lipid membranes in an early endosome (9, 12, 26, 27, 36, 51). The entry process requires the two viral envelope glycoproteins, E1 and E2, and many cellular factors, including glycosaminoglycans (GAGs) (3, 27), lipoproteins, the low-density lipoprotein receptor (LDL-R) (1, 38-40), tetraspanin CD81 (43), scavenger receptor class B type I (SR-BI) (47), and two tight junction proteins, claudin-1 (CLDN1) (17) and occludin (OCLN) (31, 44). The polarized nature of hepatocytes and the tight junction roles of OCLN and CLDN1 suggest an entry pathway similar to that of the group B coxsackieviruses, where the virion initially binds readily accessible factors that then provide a mechanism for migration of the virion into the tight junction region, just prior to internalization (14). Indeed, cellular factors are utilized by the incoming HCV virion in a temporal manner. At least GAGs and LDL-R appear to mediate virion binding (1, 3, 27, 38-40). Conflicting evidence has shown that SR-BI acts as either a binding (11) or postbinding entry factor (53), while CD81 (7, 13, 17, 27) and CLDN1 (17, 29) play postbinding roles in the HCV cell entry process. Although the kinetics of OCLN usage have not been clearly defined, this protein does not appear to play a role in virion binding (6). However, recent data showing that CD81 and CLDN1 may form complexes prior to infection (15, 24, 25, 28, 29, 35, 52) and imaging of the cell entry process (12) may contradict such a model.Human hepatocytes are the major target for HCV infection. While multiple blocks at a number of viral life cycle stages likely exist in other cell types, cell entry is one of the events limiting HCV tropism (45). Although species differences in SR-BI and CLDN1 may exert some influence on this selectivity (11, 23), CD81 and OCLN appear to be largely responsible for the restriction of HCV entry to cells from human and chimpanzee origin (7, 8, 20, 44). In fact, overexpression of the human versions of CD81 and OCLN, along with either mouse or human SR-BI and CLDN1, renders a mouse cell able to support HCV cell entry (44).We sought to provide greater insight into the species-specific restrictions of HCV cell entry and to elucidate the mechanism by which OCLN acts to mediate HCV cell entry. We examined the ability of OCLN proteins from a range of species to mediate HCV cell entry and how this function correlated with the degree of similarity to the human protein. A six-amino-acid portion of the second extracellular loop (EC2) of human OCLN was found to be responsible for the species-specific differences in entry factor function. OCLN proteins that were less functional than the human protein could be rendered fully functional by adding the human residues at these positions. Conversely, the ability of the human OCLN protein to mediate HCV cell entry was impaired by swapping this region with the corresponding sequence from species with less functional OCLN proteins. Comprehensive alanine scanning of the extracellular loops of human OCLN confirmed that the second half of EC2 was most important for the HCV cell entry process. Two cysteine residues that flank this region were found to be essential for HCV cell entry, suggesting that these residues may define a disulfide-linked subdomain of EC2. None of these amino acid changes influenced OCLN expression or localization, implying that they may serve to modulate an interaction with either another host protein or the incoming HCV virion.  相似文献   

14.
Redirecting the tropism of viral vectors enables specific transduction of selected cells by direct administration of vectors. We previously developed targeting lentiviral vectors by pseudotyping with modified Sindbis virus envelope proteins. These modified Sindbis virus envelope proteins have mutations in their original receptor-binding regions to eliminate their natural tropisms, and they are conjugated with targeting proteins, including antibodies and peptides, to confer their tropisms on target cells. We investigated whether our targeting vectors interact with DC-SIGN, which traps many types of viruses and gene therapy vectors by binding to the N-glycans of their envelope proteins. We found that these vectors do not interact with DC-SIGN. When these vectors were produced in the presence of deoxymannojirimycin, which alters the structures of N-glycans from complex to high mannose, these vectors used DC-SIGN as their receptor. Genetic analysis demonstrated that the N-glycans at E2 amino acid (aa) 196 and E1 aa 139 mediate binding to DC-SIGN, which supports the results of a previous report of cryoelectron microscopy analysis. In addition, we investigated whether modification of the N-glycan structures could activate serum complement activity, possibly by the lectin pathway of complement activation. DC-SIGN-targeted transduction occurs in the presence of human serum complement, demonstrating that high-mannose structure N-glycans of the envelope proteins do not activate human serum complement. These results indicate that the strategy of redirecting viral vectors according to alterations of their N-glycan structures would enable the vectors to target specific cells types expressing particular types of lectins.The ultimate goal of gene therapy is cell- and tissue-specific targeted delivery of therapeutic genes. A targeted system increases the therapeutic effects of transgenes at the site of action while reducing adverse effects in surrounding cells and tissues that commonly occur through nonspecific modes of gene delivery (5-8). Gene therapy vectors that can home to specific cells and tissues after intravenous administration, also known as targeting vectors, are ideal for targeted delivery (62). In the past, many attempts have been made to develop targeting viral vectors by using adenovirus, adeno-associated virus, oncoretrovirus, lentivirus, measles virus, and alphavirus (70, 89).To create targeting viral vectors, the natural tropisms of the viruses must first be eliminated and new binding specificities conferred (89). The binding of envelope viruses, such as oncoretrovirus, lentivirus, measles virus, and alphavirus, is mediated by envelope proteins. To redirect the tropisms of these viruses, the original receptor-binding regions of their envelope proteins must be eliminated. We have developed targeting oncoretroviral and lentiviral vectors by pseudotyping them with modified Sindbis virus envelope proteins and by mutating the receptor-binding regions of the envelope proteins, thereby reducing the nonspecific transduction of untargeted cells (61, 63-66). The mutated regions of the envelope protein originally interact directly with other receptors, including heparan sulfate, laminin receptor, and/or unknown molecules (10, 46, 67, 90). These mutations reduced the nonspecific transduction of the liver and spleen when the vectors were administered intravenously (66). By conjugating the virus with targeting ligands, including antibodies and peptides, the virus can transduce specific cells and tissues both in vitro and in vivo (53, 61, 63-66, 71, 72). These results demonstrated that we can eliminate the natural tropism of the Sindbis virus envelope protein while maintaining its fusion activity.However, the N-glycans of the envelope proteins are still intact and possibly interact with cell surface lectins. DC-SIGN is the best-known cell surface lectin expressed on dendritic cells, certain macrophages, and activated B cells (27, 29, 30).Structural and biochemical studies show flexible modes of DC-SIGN binding to cognate saccharides. The trimannose core unit of high-mannose N-glycans is the primary binding site for DC-SIGN (23), while nonreducing alpha1-2-linked terminal mannose moieties contribute to the high avidity seen when DC-SIGN binds the Man8 or Man9 structures common to many viral envelope glycoproteins (22). DC-SIGN traps a wide variety of viruses and viral vectors (HIV [29, 30], simian immunodeficiency virus [50], human T-cell leukemia virus type 1 [12], measles virus [17, 18], dengue virus [86], feline corona virus [77], herpes simplex virus type 1 [16], human cytomegalovirus [36], human herpesvirus type 8 [76], Ebola virus [1], West Nile virus [15], influenza virus [91], Marburg virus [57], and severe acute respiratory syndrome virus [93]) by binding to the N-glycans of the viruses and viral vectors. Binding of DC-SIGN with virus and viral vectors results in enhanced infection and/or transduction of DC-SIGN-positive cells (cis infection/transduction) and/or neighboring cells (trans infection/transduction).If any targeting vector can be trapped by DC-SIGN, it is necessary to eliminate its binding to DC-SIGN to increase the targeting specificity of the virus in vivo (28, 49, 73). In addition to enhanced infection/transduction, binding to DC-SIGN causes signaling that can activate DC-SIGN-expressing antigen-presenting cells (32, 38). Activation of antigen-presenting cells can lead to adverse effects, including systemic inflammation and immune reactions to viral vectors and their transgene products (7, 8, 32, 59, 88). Therefore, investigation of the interactions between viral vectors and DC-SIGN, identification of N-glycans that mediate binding to DC-SIGN, and elimination of interactions with DC-SIGN are important aspects of reducing adverse effects of vector administration and prolonging transgene expression.The envelope protein of our targeting lentiviral vectors, the Sindbis virus envelope protein, contains four N-linked glycans (9, 48). Sindbis virus can replicate in insect and mammalian cells, which have different types of enzymes to process N-glycans (3). Therefore, the structures of N-glycans differ between the virus produced in insect cells and that produced in mammalian cells (40, 58). The N-glycans of the virus produced in insect cells have either the high-mannose or the paucimannosidic structure. Paucimannosidic structure N-glycans, as well as high-mannose structure N-glycans, have terminal mannose residues, and all N-glycans produced in insect cells are predicted to be able to bind DC-SIGN (Fig. (Fig.11 a) (39, 47). On the other hand, two N-glycans of the virus produced in mammalian cells have the high-mannose structure, while two others have the complex structure (40, 58). The two complex structure N-glycans have been shown to be exposed on the surface of the envelope protein, while the two high-mannose structure N-glycans are buried within the center of the trimer of the envelope proteins (74, 94). Therefore, the virus produced in insect cells can access DC-SIGN as its receptor while the virus produced in mammalian cells cannot (47). Because our targeting vectors are produced in mammalian cells, they should not bind DC-SIGN efficiently. However, one group demonstrated that lentiviral vectors pseudotyped with a modified Sindbis virus envelope protein bind to DC-SIGN and target DC-SIGN-positive cells (92), in contrast to the results seen with replication-competent Sindbis virus. Both Sindbis virus and the pseudotyped lentiviral vectors were produced in mammalian cells; Sindbis virus was produced in baby hamster kidney (BHK) cells, chicken embryonic fibroblasts, and hamster fibroblast cells; and the pseudotyped vector was produced in human embryonic kidney fibroblast (293T) cells (69). Because it is known that the N-glycans of the HIV envelope protein produced in lymphocytes have structures different from those produced in macrophages, the different producer cells may account for the differences between the N-glycan structures of the virus and Sindbis virus envelope-pseudotyped lentivectors (54, 55). It is also known that the N-glycan structure of dengue virus can be altered by the presence of viral capsid (35). Thus, the capsid of Sindbis virus and HIV could also affect the structures of the N-glycans of envelope proteins differently.Open in a separate windowFIG. 1.(a) N-glycan structures and processing pathway. All N-glycans are first produced as the high-mannose structure in both mammalian cells and insect cells. In mammalian cells, certain N-glycans are further processed to the complex structure. In insect cells, certain N-glycans are further processed to the paucimannosidic structure. DMNJ inhibits mannosidase I, which is necessary for the formation of the complex structure; thus, all N-glycans have the high-mannose structure when generated in the presence of DMNJ. One representative structure of each N-glycan is shown. Man, mannose; GlcNAc, N-acetylglucosamine; SA, sialic acid; Gal, galactose. (b) Schematic representation of chimeric Sindbis virus envelope proteins. The Sindbis virus envelope protein is first synthesized as a polypeptide and subsequently cleaved by cellular proteases to generate the E3, E2, 6K, and E1 proteins. E1 and E2 are incorporated into the viral envelope, and E3 and 6K are leader sequences for E2 and E1, respectively. The N-linked glycosylation sites of the envelope proteins are shown. 2.2 is a modified Sindbis virus envelope protein in which the IgG-binding domain of protein A (ZZ) was inserted into the E2 region at aa 70. 2.2 1L1L has two flexible linkers (Gly-Gly-Gly-Gly-Ser) at aa 70 of the E2 protein. 2.2 ΔE2-196N does not have the N-glycan at E2 aa 196, 2.2 ΔE1-139N does not have the N-glycan at E1 aa 139, and 2.2 ΔE2-196N E1-139N does not have the N-glycans at either E2 aa 196 or E1 aa 139.In this study, we investigated whether our targeting vector binds DC-SIGN. We found that DC-SIGN does not mediate the transduction of our targeting vectors efficiently. The vectors can be redirected to DC-SIGN by modifying the structures of the N-glycans of the envelope proteins by using the mannosidase I inhibitor deoxymannojirimycin (DMNJ) (25, 47, 51).  相似文献   

15.
16.
Binding to the primary receptor CD4 induces conformational changes in the human immunodeficiency virus type 1 (HIV-1) gp120 envelope glycoprotein that allow binding to the coreceptor (CCR5 or CXCR4) and ultimately trigger viral membrane-cell membrane fusion mediated by the gp41 transmembrane envelope glycoprotein. Here we report the derivation of an HIV-1 gp120 variant, H66N, that confers envelope glycoprotein resistance to temperature extremes. The H66N change decreases the spontaneous sampling of the CD4-bound conformation by the HIV-1 envelope glycoproteins, thus diminishing CD4-independent infection. The H66N change also stabilizes the HIV-1 envelope glycoprotein complex once the CD4-bound state is achieved, decreasing the probability of CD4-induced inactivation and revealing the enhancing effects of soluble CD4 binding on HIV-1 infection. In the CD4-bound conformation, the highly conserved histidine 66 is located between the receptor-binding and gp41-interactive surfaces of gp120. Thus, a single amino acid change in this strategically positioned gp120 inner domain residue influences the propensity of the HIV-1 envelope glycoproteins to negotiate conformational transitions to and from the CD4-bound state.Human immunodeficiency virus type 1 (HIV-1), the cause of AIDS (6, 29, 66), infects target cells by direct fusion of the viral and target cell membranes. The viral fusion complex is composed of gp120 and gp41 envelope glycoproteins, which are organized into trimeric spikes on the surface of the virus (10, 51, 89). Membrane fusion is initiated by direct binding of gp120 to the CD4 receptor on target cells (17, 41, 53). CD4 binding creates a second binding site on gp120 for the chemokine receptors CCR5 and CXCR4, which serve as coreceptors (3, 12, 19, 23, 25). Coreceptor binding is thought to lead to further conformational changes in the HIV-1 envelope glycoproteins that facilitate the fusion of viral and cell membranes. The formation of an energetically stable six-helix bundle by the gp41 ectodomain contributes to the membrane fusion event (9, 10, 79, 89, 90).The energy required for viral membrane-cell membrane fusion derives from the sequential transitions that the HIV-1 envelope glycoproteins undergo, from the high-energy unliganded state to the low-energy six-helix bundle. The graded transitions down this energetic slope are initially triggered by CD4 binding (17). The interaction of HIV-1 gp120 with CD4 is accompanied by an unusually large change in entropy, which is thought to indicate the introduction of order into the conformationally flexible unliganded gp120 glycoprotein (61). In the CD4-bound state, gp120 is capable of binding CCR5 with high affinity; moreover, CD4 binding alters the quaternary structure of the envelope glycoprotein complex, resulting in the exposure of gp41 ectodomain segments (27, 45, 77, 92). The stability of the intermediate state induced by CD4 binding depends upon several variables, including the virus (HIV-1 versus HIV-2/simian immunodeficiency virus [SIV]), the temperature, and the nature of the CD4 ligand (CD4 on a target cell membrane versus soluble forms of CD4 [sCD4]) (30, 73). For HIV-1 exposed to sCD4, if CCR5 binding occurs within a given period of time, progression along the entry pathway continues. If CCR5 binding is impeded or delayed, the CD4-bound envelope glycoprotein complex decays into inactive states (30). In extreme cases, the binding of sCD4 to the HIV-1 envelope glycoproteins induces the shedding of gp120 from the envelope glycoprotein trimer (31, 56, 58). Thus, sCD4 generally inhibits HIV-1 infection by triggering inactivation events, in addition to competing with CD4 anchored in the target cell membrane (63).HIV-1 isolates vary in sensitivity to sCD4, due in some cases to a low affinity of the envelope glycoprotein trimer for CD4 and in other cases to differences in propensity to undergo inactivating conformational transitions following CD4 binding (30). HIV-1 isolates that have been passaged extensively in T-cell lines (the tissue culture laboratory-adapted [TCLA] isolates) exhibit lower requirements for CD4 than primary HIV-1 isolates (16, 63, 82). TCLA viruses bind sCD4 efficiently and are generally sensitive to neutralization compared with primary HIV-1 isolates. Differences in sCD4 sensitivity between primary and TCLA HIV-1 strains have been mapped to the major variable loops (V1/V2 and V3) of the gp120 glycoprotein (34, 42, 62, 81). Sensitivity to sCD4 has been shown to be independent of envelope glycoprotein spike density or the intrinsic stability of the envelope glycoprotein complex (30, 35).In general, HIV-1 isolates are more sensitive to sCD4 neutralization than HIV-2 or SIV isolates (4, 14, 73). The relative resistance of SIV to sCD4 neutralization can in some cases be explained by a reduced affinity of the envelope glycoprotein trimer for sCD4 (57); however, at least some SIV isolates exhibit sCD4-induced activation of entry into CD4-negative, CCR5-expressing target cells that lasts for several hours after exposure to sCD4 (73). Thus, for some primate immunodeficiency virus envelope glycoproteins, activated intermediates in the CD4-bound conformation can be quite stable.The HIV-1 envelope glycoprotein elements important for receptor binding, subunit interaction, and membrane fusion are well conserved among different viral strains (71, 91). Thus, these elements represent potential targets for inhibitors of HIV-1 entry. Understanding the structure and longevity of the envelope glycoprotein intermediates along the virus entry pathway is relevant to attempts at inhibition. For example, peptides that target the heptad repeat 1 region of gp41 exhibit major differences in potency against HIV-1 strains related to efficiency of chemokine receptor binding (20, 21), which is thought to promote the conformational transition to the next step in the virus entry cascade. The determinants of the duration of exposure of targetable HIV-1 envelope glycoprotein elements during the entry process are undefined.To study envelope glycoprotein determinants of the movement among the distinct conformational states along the HIV-1 entry pathway, we attempted to generate HIV-1 variants that exhibit improved stability. Historically, labile viral elements have been stabilized by selecting virus to replicate under conditions, such as high temperature, that typically weaken protein-protein interactions (38, 39, 76, 102). Thus, we subjected HIV-1 to repeated incubations at temperatures between 42°C and 56°C, followed by expansion and analysis of the remaining replication-competent virus fraction. In this manner, we identified an envelope glycoprotein variant, H66N, in which histidine 66 in the gp120 N-terminal segment was altered to asparagine. The resistance of HIV-1 bearing the H66N envelope glycoproteins to changes in temperature has been reported elsewhere (37). Here, we examine the effect of the H66N change on the ability of the HIV-1 envelope glycoproteins to negotiate conformational transitions, either spontaneously or in the presence of sCD4. The H66N phenotype was studied in the context of both CD4-dependent and CD4-independent HIV-1 variants.  相似文献   

17.
18.
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   

19.
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   

20.
Recently, complete replication of hepatitis C virus (HCV) in tissue culture was established using the JFH1 isolate. To analyze determinants of HCV genome packaging and virion assembly, we developed a system that supports particle production based on trans-packaging of subgenomic viral RNAs. Using JFH1 helper viruses, we show that subgenomic JFH1 replicons lacking the entire core to NS2 coding region are efficiently encapsidated into infectious virus-like particles. Similarly, chimeric helper viruses with heterologous structural proteins trans-package subgenomic JFH1 replicons. Like authentic cell culture-produced HCV (HCVcc) particles, these trans-complemented HCV particles (HCVTCP) penetrate target cells in a CD81 receptor-dependent fashion. Since HCVTCP production was limited by competition between the helper and subgenomic RNA and to avoid contamination of HCVTCP stocks with helper viruses, we created HCV packaging cells. These cells encapsidate various HCV replicons with high efficiency, reaching infectivity titers up to 106 tissue culture infectious doses 50 per milliliter. The produced particles display a buoyant density comparable to HCVcc particles and can be propagated in the packaging cell line but support only a single-round infection in naïve cells. Together, this work demonstrates that subgenomic HCV replicons are assembly competent, thus excluding cis-acting RNA elements in the core-to-NS2 genomic region essential for RNA packaging. The experimental system described here should be helpful to decipher the mechanisms of HCV assembly and to identify RNA elements and viral proteins involved in particle formation. Similar to other vector systems of plus-strand RNA viruses, HCVTCP may prove valuable for gene delivery or vaccination approaches.Hepatitis C virus (HCV) is an enveloped plus-strand RNA virus of the genus Hepacivirus within the family Flaviviridae (34). The HCV genome is approximately 9.6 kb in length and consists of a single open reading frame encoding a polyprotein of ca. 3,000 amino acids and nontranslated regions (NTRs) located at the 5′ and 3′ termini. These NTRs are highly structured RNA segments encompassing critical cis-active RNA elements essential for genome replication and RNA translation (31). Viral proteins are expressed in a cap-independent manner by means of an internal ribosome entry site (IRES) located in the 5′ NTR. Co- and posttranslational cleavages liberate 10 viral proteins: core; envelope protein 1 (E1) and E2, representing the structural proteins that constitute the virion; a small membrane-associated ion channel protein designated p7 that is essential for virus assembly (16, 22, 43, 57); and six nonstructural (NS) proteins (NSs 2, 3, 4A, 4B, 5A, and 5B). HCV proteins NS3 to NS5B are both necessary and sufficient to establish membrane-bound replication complexes catalyzing RNA replication (13, 36). More recent data indicate that the NS2 protease that catalyzes cleavage at the NS2-NS3 site in addition participates in assembly and release of infectious viruses (22). Finally, ribosomal frame-shifting and internal translation initiation yield a group of additional proteins designated ARFP (alternative reading frame protein) or core+1 proteins. However, their function for the HCV replication cycle is currently not known.One hallmark of HCV is its high propensity to establish a persistent infection, which frequently causes progressive morbidity ranging from hepatic fibrosis to cirrhosis and hepatocellular carcinoma (20). Despite considerable progress in the treatment of HCV infection, the currently available therapy (a combination of pegylated interferon alpha with ribavirin) is not well tolerated and is efficacious in only ca. 50% of patients infected with the most prevalent genotype 1 (38). Therapeutic or prophylactic vaccines are not available. For these reasons and with currently ca. 170 million persistently infected individuals, HCV infection represents a considerable global health problem necessitating pertinent basic and applied research efforts.In recent years three major advances enabled analysis of the HCV replication cycle in tissue culture. First, Lohmann and colleagues developed subgenomic HCV replicons (36). These autonomously replicating RNA molecules carry all the genetic elements necessary for self-replication (the NTRs and NS3 to NS5B), including a selectable marker or a reporter gene in place of the viral structural proteins, and an internal IRES for expression of the HCV replicase genes (reviewed in reference 45). Second, HCV pseudotype particles, i.e., retroviral particles surrounded by an envelope carrying HCV E1-E2 complexes in place of their cognate envelope proteins, were established (3, 21). As these particles carry functional HCV glycoprotein complexes on their surface, HCV pseudotype particles have been instrumental for the analysis of E1-E2 receptor interactions and the early events of HCV infection (reviewed in reference 2). Finally, in 2005 fully permissive cell culture systems which are based on the JFH1 clone were described (33, 66, 72). This isolate replicates with unprecedented efficiency in transfected Huh7 human hepatoma cells and produces particles infectious both in vitro and in vivo, thus providing a model system reproducing the complete HCV replication cycle.Use of these novel models has considerably expanded our knowledge of viral and host cell factors involved in HCV replication (for a recent review, see reference 59). It is now known that similar to virtually all other plus-strand RNA viruses, HCV induces intracellular membrane alterations and replicates its genome in conjunction with vesicular membrane structures, the so-called “membranous web” (10, 13). Presumably as a consequence of this specific, rather secluded architecture of the membrane-associated replication machinery, all viral proteins involved in RNA replication, with the exception of NS5A function in cis, cannot be complemented in trans (1). Restricted trans-complementation of viral replicase proteins has been observed for other plus-strand RNA viruses as well, thus indicating that a rather “closed” replication machinery is a shared feature of these viruses (15, 27, 60). In contrast, for a number of plus-strand RNA viruses from diverse virus families like Picornaviridae (poliovirus), Alphaviridae (Sindbis virus, Semliki Forest virus, and Venezuelan equine encephalitis virus), Coronaviridae (human coronavirus E229), and Flaviviridae (tick-borne encephalitis virus, Kunjin virus, West Nile virus, and yellow fever virus), assembly of progeny viruses can be achieved when structural proteins are expressed in trans and independent from the RNA molecule that encodes the replicase proteins. Similarly, Miyanari recently reported that HCV genomes with lethal mutations in core protein can be rescued by ectopic expression of functional core protein (39). This flexibility has been extensively used to create viral vectors for gene delivery as well as viral vector-based immunization approaches (32, 48, 49, 61, 68) (for a recent review on alphaviral vectors, the most frequently used among plus strand RNA vectors, see reference 37). In these systems the viral genome region encoding the structural proteins is replaced by a transgene. The resulting defective vector genomes are capable of RNA replication but due to the lack of structural proteins are unable to produce progeny virus particles. This defect is rescued by expression of the structural proteins in trans via helper viruses (28, 55) or, in some cases, by DNA constructs stably expressed in packaging cell lines (17). The resulting virus-like particles are infectious but support only single-round infection and are unable to spread, thus improving the safety of the viral transduction system.Given the success of plus-strand RNA vector technology for basic and applied clinical research, in this study we developed a trans-complementation system for HCV that provided new insights into the basic principles of HCV particle assembly.  相似文献   

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