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.
Borna disease virus (BDV), the prototypic member of the family Bornaviridae within the order Mononegavirales, provides an important model for the investigation of viral persistence within the central nervous system (CNS) and of associated brain disorders. BDV is highly neurotropic and enters its target cell via receptor-mediated endocytosis, a process mediated by the virus surface glycoprotein (G), but the cellular factors and pathways determining BDV cell tropism within the CNS remain mostly unknown. Cholesterol has been shown to influence viral infections via its effects on different viral processes, including replication, budding, and cell entry. In this work, we show that cell entry, but not replication and gene expression, of BDV was drastically inhibited by depletion of cellular cholesterol levels. BDV G-mediated attachment to BDV-susceptible cells was cholesterol independent, but G localized to lipid rafts (LR) at the plasma membrane. LR structure and function critically depend on cholesterol, and hence, compromised structural integrity and function of LR caused by cholesterol depletion likely inhibited the initial stages of BDV cell internalization. Furthermore, we also show that viral-envelope cholesterol is required for BDV infectivity.Borna disease virus (BDV) is an enveloped virus with a nonsegmented negative-strand RNA genome whose organization (3′-N-p10/P-M-G-L-5′) is characteristic of mononegaviruses (6, 28, 46, 48). However, based on its unique genetics and biological features, BDV is considered to be the prototypic member of a new virus family, Bornaviridae, within the order Mononegavirales (8, 28, 46, 49).BDV can infect a variety of cell types in cell culture but in vivo exhibits exquisite neurotropism and causes central nervous system (CNS) disease in different vertebrate species, which is frequently manifested in behavioral abnormalities (19, 33, 44, 53). Both host and viral factors contribute to a variable period of incubation and heterogeneity in the symptoms and pathology associated with BDV infection (14, 16, 29, 42, 44). BDV provides an important model for the investigation of both immune-mediated pathological events associated with virus-induced neurological disease and mechanisms whereby noncytolytic viruses induce neurodevelopmental and behavioral disturbances in the absence of inflammation (15, 18, 41). Moreover, serological data and molecular epidemiological studies suggest that BDV, or a BDV-like virus, can infect humans and that it might be associated with certain neuropsychiatric disorders (17, 24), which further underscores the interest in understanding the mechanisms underlying BDV persistence in the CNS and its effect on brain cell functions. The achievement of these goals will require the elucidation of the determinants of BDV cell tropism within the CNS.BDV enters its target cell via receptor-mediated endocytosis, a process in which the BDV G protein plays a central role (1, 5, 13, 14, 39). Cleavage of BDV G by the cellular protease furin generates two functional subunits: GP1 (GPN), involved in virus interaction with a yet-unidentified cell surface receptor (1, 39), and GP2 (GPC), which mediates a pH-dependent fusion event between viral and cellular membranes (13). However, a detailed characterization of cellular factors and pathways involved in BDV cell entry remains to be done.Besides cell surface molecules that serve as viral receptors, many other cell factors, including nonproteinaceous molecules, can influence cell entry by virus (52). In this regard, cholesterol, which plays a critical role in cellular homeostasis (55), has also been identified as a key factor required for productive infection by different viruses. Accordingly, cholesterol participates in a variety of processes in virus-infected cells, including fusion events between viral and cellular membranes (3), viral replication (23), and budding (35, 37), as well as maintenance of lipid rafts (LR) (12) as scaffold structures where the viral receptor and coreceptor associate (11, 26, 32, 36). LR are specialized microdomains within cellular membranes constituted principally of proteins, sphingolipids, and cholesterol. LR facilitate the close proximity and interaction of specific sets of proteins and contribute to different processes associated with virus multiplication (38). Cholesterol can also influence virus infection by contributing to the maintenance of the properties of the viral envelope required for virus particle infectivity (21, 54). Here, we show for the first time that cholesterol plays a critical role in BDV infection. Depletion of cellular cholesterol prior to, but not after, BDV cell entry prevented productive BDV infection, likely due to disruption of plasma membrane LR that appear to be the cell entry point for BDV. In addition, we document that cholesterol also plays an essential role in the properties of the BDV envelope required for virus particle infectivity.  相似文献   

2.
3.
Human immunodeficiency virus type 1 (HIV-1) infects target cells by binding to CD4 and a chemokine receptor, most commonly CCR5. CXCR4 is a frequent alternative coreceptor (CoR) in subtype B and D HIV-1 infection, but the importance of many other alternative CoRs remains elusive. We have analyzed HIV-1 envelope (Env) proteins from 66 individuals infected with the major subtypes of HIV-1 to determine if virus entry into highly permissive NP-2 cell lines expressing most known alternative CoRs differed by HIV-1 subtype. We also performed linear regression analysis to determine if virus entry via the major CoR CCR5 correlated with use of any alternative CoR and if this correlation differed by subtype. Virus pseudotyped with subtype B Env showed robust entry via CCR3 that was highly correlated with CCR5 entry efficiency. By contrast, viruses pseudotyped with subtype A and C Env proteins were able to use the recently described alternative CoR FPRL1 more efficiently than CCR3, and use of FPRL1 was correlated with CCR5 entry. Subtype D Env was unable to use either CCR3 or FPRL1 efficiently, a unique pattern of alternative CoR use. These results suggest that each subtype of circulating HIV-1 may be subject to somewhat different selective pressures for Env-mediated entry into target cells and suggest that CCR3 may be used as a surrogate CoR by subtype B while FPRL1 may be used as a surrogate CoR by subtypes A and C. These data may provide insight into development of resistance to CCR5-targeted entry inhibitors and alternative entry pathways for each HIV-1 subtype.Human immunodeficiency virus type 1 (HIV-1) infects target cells by binding first to CD4 and then to a coreceptor (CoR), of which C-C chemokine receptor 5 (CCR5) is the most common (6, 53). CXCR4 is an additional CoR for up to 50% of subtype B and D HIV-1 isolates at very late stages of disease (4, 7, 28, 35). Many other seven-membrane-spanning G-protein-coupled receptors (GPCRs) have been identified as alternative CoRs when expressed on various target cell lines in vitro, including CCR1 (76, 79), CCR2b (24), CCR3 (3, 5, 17, 32, 60), CCR8 (18, 34, 38), GPR1 (27, 65), GPR15/BOB (22), CXCR5 (39), CXCR6/Bonzo/STRL33/TYMSTR (9, 22, 25, 45, 46), APJ (26), CMKLR1/ChemR23 (49, 62), FPLR1 (67, 68), RDC1 (66), and D6 (55). HIV-2 and simian immunodeficiency virus SIVmac isolates more frequently show expanded use of these alternative CoRs than HIV-1 isolates (12, 30, 51, 74), and evidence that alternative CoRs other than CXCR4 mediate infection of primary target cells by HIV-1 isolates is sparse (18, 30, 53, 81). Genetic deficiency in CCR5 expression is highly protective against HIV-1 transmission (21, 36), establishing CCR5 as the primary CoR. The importance of alternative CoRs other than CXCR4 has remained elusive despite many studies (1, 30, 70, 81). Expansion of CoR use from CCR5 to include CXCR4 is frequently associated with the ability to use additional alternative CoRs for viral entry (8, 16, 20, 63, 79) in most but not all studies (29, 33, 40, 77, 78). This finding suggests that the sequence changes in HIV-1 env required for use of CXCR4 as an additional or alternative CoR (14, 15, 31, 37, 41, 57) are likely to increase the potential to use other alternative CoRs.We have used the highly permissive NP-2/CD4 human glioma cell line developed by Soda et al. (69) to classify virus entry via the alternative CoRs CCR1, CCR3, CCR8, GPR1, CXCR6, APJ, CMKLR1/ChemR23, FPRL1, and CXCR4. Full-length molecular clones of 66 env genes from most prevalent HIV-1 subtypes were used to generate infectious virus pseudotypes expressing a luciferase reporter construct (19, 57). Two types of analysis were performed: the level of virus entry mediated by each alternative CoR and linear regression of entry mediated by CCR5 versus all other alternative CoRs. We thus were able to identify patterns of alternative CoR use that were subtype specific and to determine if use of any alternative CoR was correlated or independent of CCR5-mediated entry. The results obtained have implications for the evolution of env function, and the analyses revealed important differences between subtype B Env function and all other HIV-1 subtypes.  相似文献   

4.
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   

5.
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   

6.
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   

7.
Herpes simplex virus type 1 (HSV-1)-induced cell fusion is mediated by viral glycoproteins and other membrane proteins expressed on infected cell surfaces. Certain mutations in the carboxyl terminus of HSV-1 glycoprotein B (gB) and in the amino terminus of gK cause extensive virus-induced cell fusion. Although gB is known to be a fusogenic glycoprotein, the mechanism by which gK is involved in virus-induced cell fusion remains elusive. To delineate the amino-terminal domains of gK involved in virus-induced cell fusion, the recombinant viruses gKΔ31-47, gKΔ31-68, and gKΔ31-117, expressing gK carrying in-frame deletions spanning the amino terminus of gK immediately after the gK signal sequence (amino acids [aa] 1 to 30), were constructed. Mutant viruses gKΔ31-47 and gKΔ31-117 exhibited a gK-null (ΔgK) phenotype characterized by the formation of very small viral plaques and up to a 2-log reduction in the production of infectious virus in comparison to that for the parental HSV-1(F) wild-type virus. The gKΔ31-68 mutant virus formed substantially larger plaques and produced 1-log-higher titers than the gKΔ31-47 and gKΔ31-117 mutant virions at low multiplicities of infection. Deletion of 28 aa from the carboxyl terminus of gB (gBΔ28syn) caused extensive virus-induced cell fusion. However, the gBΔ28syn mutation was unable to cause virus-induced cell fusion in the presence of the gKΔ31-68 mutation. Transient expression of a peptide composed of the amino-terminal 82 aa of gK (gKa) produced a glycosylated peptide that was efficiently expressed on cell surfaces only after infection with the HSV-1(F), gKΔ31-68, ΔgK, or UL20-null virus. The gKa peptide complemented the gKΔ31-47 and gKΔ31-68 mutant viruses for infectious-virus production and for gKΔ31-68/gBΔ28syn-mediated cell fusion. These data show that the amino terminus of gK modulates gB-mediated virus-induced cell fusion and virion egress.Herpes simplex virus type 1 (HSV-1) specifies at least 11 virally encoded glycoproteins, as well as several nonglycosylated and lipid-anchored membrane-associated proteins, which serve important functions in virion infectivity and virus spread. Although cell-free enveloped virions can efficiently spread viral infection, virions can also spread by causing cell fusion of adjacent cellular membranes. Virus-induced cell fusion, which is caused by viral glycoproteins expressed on infected cell surfaces, enables transmission of virions from one cell to another, avoiding extracellular spaces and exposure of free virions to neutralizing antibodies (reviewed in reference 56). Most mutations that cause extensive virus-induced cell-to-cell fusion (syncytial or syn mutations) have been mapped to at least four regions of the viral genome: the UL20 gene (5, 42, 44); the UL24 gene (37, 58); the UL27 gene, encoding glycoprotein B (gB) (9, 51); and the UL53 gene, coding for gK (7, 15, 35, 53, 54, 57).Increasing evidence suggests that virus-induced cell fusion is mediated by the concerted action of glycoproteins gD, gB, and gH/gL. Recent studies have shown that gD interacts with both gB and gH/gL (1, 2). Binding of gD to its cognate receptors, including Nectin-1, HVEM, and others (12, 29, 48, 59, 60, 62, 63), is thought to trigger conformation changes in gH/gL and gB that cause fusion of the viral envelope with cellular membranes during virus entry and virus-induced cell fusion (32, 34). Transient coexpression of gB, gD, and gH/gL causes cell-to-cell fusion (49, 68). However, this phenomenon does not accurately model viral fusion, because other viral glycoproteins and membrane proteins known to be important for virus-induced cell fusion are not required (6, 14, 31). Specifically, gK and UL20 were shown to be absolutely required for virus-induced cell fusion (21, 46). Moreover, syncytial mutations within gK (7, 15, 35, 53, 54, 57) or UL20 (5, 42, 44) promote extensive virus-induced cell fusion, and viruses lacking gK enter more slowly than wild-type virus into susceptible cells (25). Furthermore, transient coexpression of gK carrying a syncytial mutation with gB, gD, and gH/gL did not enhance cell fusion, while coexpression of the wild-type gK with gB, gD, and gH/gL inhibited cell fusion (3).Glycoproteins gB and gH are highly conserved across all subfamilies of herpesviruses. gB forms a homotrimeric type I integral membrane protein, which is N glycosylated at multiple sites within the polypeptide. An unusual feature of gB is that syncytial mutations that enhance virus-induced cell fusion are located exclusively in the carboxyl terminus of gB, which is predicted to be located intracellularly (51). Single-amino-acid substitutions within two regions of the intracellular cytoplasmic domain of gB were shown to cause syncytium formation and were designated region I (amino acid [aa] positions 816 and 817) and region II (aa positions 853, 854, and 857) (9, 10, 28, 69). Furthermore, deletion of 28 aa from the carboxyl terminus of gB, disrupting the small predicted alpha-helical domain H17b, causes extensive virus-induced cell fusion as well as extensive glycoprotein-mediated cell fusion in the gB, gD, and gH/gL transient-coexpression system (22, 49, 68). The X-ray structure of the ectodomain of gB has been determined and is predicted to assume at least two major conformations, one of which may be necessary for the fusogenic properties of gB. Therefore, perturbation of the carboxyl terminus of gB may alter the conformation of the amino terminus of gB, thus favoring one of the two predicted conformational structures that causes membrane fusion (34).The UL53 (gK) and UL20 genes encode multipass transmembrane proteins of 338 and 222 aa, respectively, which are conserved in all alphaherpesviruses (15, 42, 55). Both proteins have multiple sites where posttranslational modification can occur; however, only gK is posttranslationally modified by N-linked carbohydrate addition (15, 35, 55). The specific membrane topologies of both gK and UL20 protein (UL20p) have been predicted and experimentally confirmed using epitope tags inserted within predicted intracellular and extracellular domains (18, 21, 44). Syncytial mutations in gK map predominantly within extracellular domains of gK and particularly within the amino-terminal portion of gK (domain I) (18), while syncytial mutations of UL20 are located within the amino terminus of UL20p, shown to be located intracellularly (44). A series of recent studies have shown that HSV-1 gK and UL20 functionally and physically interact and that these interactions are necessary for their coordinate intracellular transport and cell surface expression (16, 18, 21, 26, 45). Specifically, direct protein-protein interactions between the amino terminus of HSV-1 UL20 and gK domain III, both of which are localized intracellularly, were recently demonstrated by two-way coimmunoprecipitation experiments (19).According to the most prevalent model for herpesvirus intracellular morphogenesis, capsids initially assemble within the nuclei and acquire a primary envelope by budding into the perinuclear spaces. Subsequently, these virions lose their envelope through fusion with the outer nuclear lamellae. Within the cytoplasm, tegument proteins associate with the viral nucleocapsid and final envelopment occurs by budding of cytoplasmic capsids into specific trans-Golgi network (TGN)-associated membranes (8, 30, 47, 70). Mature virions traffic to cell surfaces, presumably following the cellular secretory pathway (33, 47, 61). In addition to their significant roles in virus-induced cell fusion, gK and UL20 are required for cytoplasmic virion envelopment. Viruses with deletions in either the gK or the UL20 gene are unable to translocate from the cytoplasm to extracellular spaces and accumulated as unenveloped virions in the cytoplasm (5, 15, 20, 21, 26, 35, 36, 38, 44, 55). Current evidence suggests that the functions of gK and UL20 in cytoplasmic virion envelopment and virus-induced cell fusion are carried out by different, genetically separable domains of UL20p. Specifically, UL20 mutations within the amino and carboxyl termini of UL20p allowed cotransport of gK and UL20p to cell surfaces, virus-induced cell fusion, and TGN localization, while effectively inhibiting cytoplasmic virion envelopment (44, 45).In this paper, we demonstrate that the amino terminus of gK expressed as a free peptide of 82 aa (gKa) is transported to infected cell surfaces by viral proteins other than gK or UL20p and facilitates virus-induced cell fusion caused by syncytial mutations in the carboxyl terminus of gB. Thus, functional domains of gK can be genetically separated, as we have shown previously (44, 45), as well as physically separated into different peptide portions that retain functional activities of gK. These results are consistent with the hypothesis that the amino terminus of gK directly or indirectly interacts with and modulates the fusogenic properties of gB.  相似文献   

8.
HIV-1 possesses an exquisite ability to infect cells independently from their cycling status by undergoing an active phase of nuclear import through the nuclear pore. This property has been ascribed to the presence of karyophilic elements present in viral nucleoprotein complexes, such as the matrix protein (MA); Vpr; the integrase (IN); and a cis-acting structure present in the newly synthesized DNA, the DNA flap. However, their role in nuclear import remains controversial at best. In the present study, we carried out a comprehensive analysis of the role of these elements in nuclear import in a comparison between several primary cell types, including stimulated lymphocytes, macrophages, and dendritic cells. We show that despite the fact that none of these elements is absolutely required for nuclear import, disruption of the central polypurine tract-central termination sequence (cPPT-CTS) clearly affects the kinetics of viral DNA entry into the nucleus. This effect is independent of the cell cycle status of the target cells and is observed in cycling as well as in nondividing primary cells, suggesting that nuclear import of viral DNA may occur similarly under both conditions. Nonetheless, this study indicates that other components are utilized along with the cPPT-CTS for an efficient entry of viral DNA into the nucleus.Lentiviruses display an exquisite ability to infect dividing and nondividing cells alike that is unequalled among Retroviridae. This property is thought to be due to the particular behavior or composition of the viral nucleoprotein complexes (NPCs) that are liberated into the cytoplasm of target cells upon virus-to-cell membrane fusion and that allow lentiviruses to traverse an intact nuclear membrane (17, 28, 29, 39, 52, 55, 67, 79). In the case of the human immunodeficiency type I virus (HIV-1), several studies over the years identified viral components of such structures with intrinsic karyophilic properties and thus perfect candidates for mediation of the passage of viral DNA (vDNA) through the nuclear pore: the matrix protein (MA); Vpr; the integrase (IN); and a three-stranded DNA flap, a structure present in neo-synthesized viral DNA, specified by the central polypurine tract-central termination sequence (cPPT-CTS). It is clear that these elements may mediate nuclear import directly or via the recruitment of the host''s proteins, and indeed, several cellular proteins have been found to influence HIV-1 infection during nuclear import, like the karyopherin α2 Rch1 (38); importin 7 (3, 30, 93); the transportin SR-2 (13, 20); or the nucleoporins Nup98 (27), Nup358/RANBP2, and Nup153 (13, 56).More recently, the capsid protein (CA), the main structural component of viral nucleoprotein complexes at least upon their cytoplasmic entry, has also been suggested to be involved in nuclear import or in postnuclear entry steps (14, 25, 74, 90, 92). Whether this is due to a role for CA in the shaping of viral nucleoprotein complexes or to a direct interaction between CA and proteins involved in nuclear import remains at present unknown.Despite a large number of reports, no single viral or cellular element has been described as absolutely necessary or sufficient to mediate lentiviral nuclear import, and important controversies as to the experimental evidences linking these elements to this step exist. For example, MA was among the first viral protein of HIV-1 described to be involved in nuclear import, and 2 transferable nuclear localization signals (NLSs) have been described to occur at its N and C termini (40). However, despite the fact that early studies indicated that the mutation of these NLSs perturbed HIV-1 nuclear import and infection specifically in nondividing cells, such as macrophages (86), these findings failed to be confirmed in more-recent studies (23, 33, 34, 57, 65, 75).Similarly, Vpr has been implicated by several studies of the nuclear import of HIV-1 DNA (1, 10, 21, 43, 45, 47, 64, 69, 72, 73, 85). Vpr does not possess classical NLSs, yet it displays a transferable nucleophilic activity when fused to heterologous proteins (49-51, 53, 77, 81) and has been shown to line onto the nuclear envelope (32, 36, 47, 51, 58), where it can truly facilitate the passage of the viral genome into the nucleus. However, the role of Vpr in this step remains controversial, as in some instances Vpr is not even required for viral replication in nondividing cells (1, 59).Conflicting results concerning the role of IN during HIV-1 nuclear import also exist. Indeed, several transferable NLSs have been described to occur in the catalytic core and the C-terminal DNA binding domains of IN, but for some of these, initial reports of nuclear entry defects (2, 9, 22, 46, 71) were later shown to result from defects at steps other than nuclear import (60, 62, 70, 83). These reports do not exclude a role for the remaining NLSs in IN during nuclear import, and they do not exclude the possibility that IN may mediate this step by associating with components of the cellular nuclear import machinery, such as importin alpha and beta (41), importin 7 (3, 30, 93, 98), and, more recently, transportin-SR2 (20).The central DNA flap, a structure present in lentiviruses and in at least 1 yeast retroelement (44), but not in other orthoretroviruses, has also been involved in the nuclear import of viral DNA (4, 6, 7, 31, 78, 84, 95, 96), and more recently, it has been proposed to provide a signal for viral nucleoprotein complexes uncoating in the proximity of the nuclear pore, with the consequence of providing a signal for import (8). However, various studies showed an absence or weakness of nuclear entry defects in viruses devoid of the DNA flap (24, 26, 44, 61).Overall, the importance of viral factors in HIV-1 nuclear import is still unclear. The discrepancies concerning the role of MA, IN, Vpr, and cPPT-CTS in HIV-1 nuclear import could in part be explained by their possible redundancy. To date, only one comprehensive study analyzed the role of these four viral potentially karyophilic elements together (91). This study showed that an HIV-1 chimera where these elements were either deleted or replaced by their murine leukemia virus (MLV) counterparts was, in spite of an important infectivity defect, still able to infect cycling and cell cycle-arrested cell lines to similar efficiencies. If this result indicated that the examined viral elements of HIV-1 were dispensable for the cell cycle independence of HIV, as infections proceeded equally in cycling and arrested cells, they did not prove that they were not required in nuclear import, because chimeras displayed a severe infectivity defect that precluded their comparison with the wild type (WT).Nuclear import and cell cycle independence may not be as simply linked as previously thought. On the one hand, there has been no formal demonstration that the passage through the nuclear pore, and thus nuclear import, is restricted to nondividing cells, and for what we know, this passage may be an obligatory step in HIV infection in all cells, irrespective of their cycling status. In support of this possibility, certain mutations in viral elements of HIV affect nuclear import in dividing as well as in nondividing cells (4, 6, 7, 31, 84, 95). On the other hand, cell cycle-independent infection may be a complex phenomenon that is made possible not only by the ability of viral DNA to traverse the nuclear membrane but also by its ability to cope with pre- and postnuclear entry events, as suggested by the phenotypes of certain CA mutants (74, 92).Given that the cellular environment plays an important role during the early steps of viral infection, we chose to analyze the role of the four karyophilic viral elements of HIV-1 during infection either alone or combined in a wide comparison between cells highly susceptible to infection and more-restrictive primary cell targets of HIV-1 in vivo, such as primary blood lymphocytes (PBLs), monocyte-derived macrophages (MDM), and dendritic cells (DCs).In this study, we show that an HIV-1-derived virus in which the 2 NLSs of MA are mutated and the IN, Vpr, and cPPT-CTS elements are removed displays no detectable nuclear import defect in HeLa cells independently of their cycling status. However, this mutant virus is partially impaired for nuclear entry in primary cells and more specifically in DCs and PBLs. We found that this partial defect is specified by the cPPT-CTS, while the 3 remaining elements seem to play no role in nuclear import. Thus, our study indicates that the central DNA flap specifies the most important role among the viral elements involved thus far in nuclear import. However, it also clearly indicates that the role played by the central DNA flap is not absolute and that its importance varies depending on the cell type, independently from the dividing status of the cell.  相似文献   

9.
In their vertebrate hosts, arboviruses such as Semliki Forest virus (SFV) (Togaviridae) generally counteract innate defenses and trigger cell death. In contrast, in mosquito cells, following an early phase of efficient virus production, a persistent infection with low levels of virus production is established. Whether arboviruses counteract RNA interference (RNAi), which provides an important antiviral defense system in mosquitoes, is an important question. Here we show that in Aedes albopictus-derived mosquito cells, SFV cannot prevent the establishment of an antiviral RNAi response or prevent the spread of protective antiviral double-stranded RNA/small interfering RNA (siRNA) from cell to cell, which can inhibit the replication of incoming virus. The expression of tombusvirus siRNA-binding protein p19 by SFV strongly enhanced virus spread between cultured cells rather than virus replication in initially infected cells. Our results indicate that the spread of the RNAi signal contributes to limiting virus dissemination.In animals, RNA interference (RNAi) was first described for Caenorhabditis elegans (27). The production or introduction of double-stranded RNA (dsRNA) in cells leads to the degradation of mRNAs containing homologous sequences by sequence-specific cleavage of mRNAs. Central to RNAi is the production of 21- to 26-nucleotide small interfering RNAs (siRNAs) from dsRNA and the assembly of an RNA-induced silencing complex (RISC), followed by the degradation of the target mRNA (23, 84). RNAi is a known antiviral strategy of plants (3, 53) and insects (21, 39, 51). Study of Drosophila melanogaster in particular has given important insights into RNAi responses against pathogenic viruses and viral RNAi inhibitors (31, 54, 83, 86, 91). RNAi is well characterized for Drosophila, and orthologs of antiviral RNAi genes have been found in Aedes and Culex spp. (13, 63).Arboviruses, or arthropod-borne viruses, are RNA viruses mainly of the families Bunyaviridae, Flaviviridae, and Togaviridae. The genus Alphavirus within the family Togaviridae contains several mosquito-borne pathogens: arboviruses such as Chikungunya virus (16) and equine encephalitis viruses (88). Replication of the prototype Sindbis virus and Semliki Forest virus (SFV) is well understood (44, 71, 74, 79). Their genome consists of a positive-stranded RNA with a 5′ cap and a 3′ poly(A) tail. The 5′ two-thirds encodes the nonstructural polyprotein P1234, which is cleaved into four replicase proteins, nsP1 to nsP4 (47, 58, 60). The structural polyprotein is encoded in the 3′ one-third of the genome and cleaved into capsid and glycoproteins after translation from a subgenomic mRNA (79). Cytoplasmic replication complexes are associated with cellular membranes (71). Viruses mature by budding at the plasma membrane (35).In nature, arboviruses are spread by arthropod vectors (predominantly mosquitoes, ticks, flies, and midges) to vertebrate hosts (87). Little is known about how arthropod cells react to arbovirus infection. In mosquito cell cultures, an acute phase with efficient virus production is generally followed by the establishment of a persistent infection with low levels of virus production (9). This is fundamentally different from the cytolytic events following arbovirus interactions with mammalian cells and pathogenic insect viruses with insect cells. Alphaviruses encode host response antagonists for mammalian cells (2, 7, 34, 38).RNAi has been described for mosquitoes (56) and, when induced before infection, antagonizes arboviruses and their replicons (1, 4, 14, 15, 29, 30, 32, 42, 64, 65). RNAi is also functional in various mosquito cell lines (1, 8, 43, 49, 52). In the absence of RNAi, alphavirus and flavivirus replication and/or dissemination is enhanced in both mosquitoes and Drosophila (14, 17, 31, 45, 72). RNAi inhibitors weakly enhance SFV replicon replication in tick and mosquito cells (5, 33), posing the questions of how, when, and where RNAi interferes with alphavirus infection in mosquito cells.Here we use an A. albopictus-derived mosquito cell line to study RNAi responses to SFV. Using reporter-based assays, we demonstrate that SFV cannot avoid or efficiently inhibit the establishment of an RNAi response. We also demonstrate that the RNAi signal can spread between mosquito cells. SFV cannot inhibit cell-to-cell spread of the RNAi signal, and spread of the virus-induced RNAi signal (dsRNA/siRNA) can inhibit the replication of incoming SFV in neighboring cells. Furthermore, we show that SFV expression of a siRNA-binding protein increases levels of virus replication mainly by enhancing virus spread between cells rather than replication in initially infected cells. Taken together, these findings suggest a novel mechanism, cell-to-cell spread of antiviral dsRNA/siRNA, by which RNAi limits SFV dissemination in mosquito cells.  相似文献   

10.
Clade B of the New World arenaviruses contains both pathogenic and nonpathogenic members, whose surface glycoproteins (GPs) are characterized by different abilities to use the human transferrin receptor type 1 (hTfR1) protein as a receptor. Using closely related pairs of pathogenic and nonpathogenic viruses, we investigated the determinants of the GP1 subunit that confer these different characteristics. We identified a central region (residues 85 to 221) in the Guanarito virus GP1 that was sufficient to interact with hTfR1, with residues 159 to 221 being essential. The recently solved structure of part of the Machupo virus GP1 suggests an explanation for these requirements.Arenaviruses are bisegmented, single-stranded RNA viruses that use an ambisense coding strategy to express four proteins: NP (nucleoprotein), Z (matrix protein), L (polymerase), and GP (glycoprotein). The viral GP is sufficient to direct entry into host cells, and retroviral vectors pseudotyped with GP recapitulate the entry pathway of these viruses (5, 13, 24, 31). GP is a class I fusion protein comprising two subunits, GP1 and GP2, cleaved from the precursor protein GPC (4, 14, 16, 18, 21). GP1 contains the receptor binding domain (19, 28), while GP2 contains structural elements characteristic of viral membrane fusion proteins (8, 18, 20, 38). The N-terminal stable signal peptide (SSP) remains associated with the mature glycoprotein after cleavage (2, 39) and plays a role in transport, maturation, and pH-dependent fusion (17, 35, 36, 37).The New World arenaviruses are divided into clades A, B, and C based on phylogenetic relatedness (7, 9, 11). Clade B contains the human pathogenic viruses Junin (JUNV), Machupo (MACV), Guanarito (GTOV), Sabia, and Chapare, which cause severe hemorrhagic fevers in South America (1, 10, 15, 26, 34). Clade B also contains the nonpathogenic viruses Amapari (AMAV), Cupixi, and Tacaribe (TCRV), although mild disease has been reported for a laboratory worker infected with TCRV (29).Studies with both viruses and GP-pseudotyped retroviral vectors have shown that the pathogenic clade B arenaviruses use the human transferrin receptor type 1 (hTfR1) to gain entry into human cells (19, 30). In contrast, GPs from nonpathogenic viruses, although capable of using TfR1 orthologs from other species (1), cannot use hTfR1 (1, 19) and instead enter human cells through as-yet-uncharacterized hTfR1-independent pathways (19). In addition, human T-cell lines serve as useful tools to distinguish these GPs, since JUNV, GTOV, and MACV pseudotyped vectors readily transduce CEM cells, while TCRV and AMAV GP vectors do not (27; also unpublished data). These properties of the GPs do not necessarily reflect a tropism of the pathogenic viruses for human T cells, since viral tropism is influenced by many factors and T cells are not a target for JUNV replication in vivo (3, 22, 25).  相似文献   

11.
Human cytomegalovirus (HCMV) is a widely circulating pathogen that causes severe disease in immunocompromised patients and infected fetuses. By immortalizing memory B cells from HCMV-immune donors, we isolated a panel of human monoclonal antibodies that neutralized at extremely low concentrations (90% inhibitory concentration [IC90] values ranging from 5 to 200 pM) HCMV infection of endothelial, epithelial, and myeloid cells. With the single exception of an antibody that bound to a conserved epitope in the UL128 gene product, all other antibodies bound to conformational epitopes that required expression of two or more proteins of the gH/gL/UL128-131A complex. Antibodies against gB, gH, or gM/gN were also isolated and, albeit less potent, were able to neutralize infection of both endothelial-epithelial cells and fibroblasts. This study describes unusually potent neutralizing antibodies against HCMV that might be used for passive immunotherapy and identifies, through the use of such antibodies, novel antigenic targets in HCMV for the design of immunogens capable of eliciting previously unknown neutralizing antibody responses.Human cytomegalovirus (HCMV) is a member of the herpesvirus family which is widely distributed in the human population and can cause severe disease in immunocompromised patients and upon infection of the fetus. HCMV infection causes clinical disease in 75% of patients in the first year after transplantation (58), while primary maternal infection is a major cause of congenital birth defects including hearing loss and mental retardation (5, 33, 45). Because of the danger posed by this virus, development of an effective vaccine is considered of highest priority (51).HCMV infection requires initial interaction with the cell surface through binding to heparan sulfate proteoglycans (8) and possibly other surface receptors (12, 23, 64, 65). The virus displays a broad host cell range (24, 53), being able to infect several cell types such as endothelial cells, epithelial cells (including retinal cells), smooth muscle cells, fibroblasts, leukocytes, and dendritic cells (21, 37, 44, 54). Endothelial cell tropism has been regarded as a potential virulence factor that might influence the clinical course of infection (16, 55), whereas infection of leukocytes has been considered a mechanism of viral spread (17, 43, 44). Extensive propagation of HCMV laboratory strains in fibroblasts results in deletions or mutations of genes in the UL131A-128 locus (1, 18, 21, 36, 62, 63), which are associated with the loss of the ability to infect endothelial cells, epithelial cells, and leukocytes (15, 43, 55, 61). Consistent with this notion, mouse monoclonal antibodies (MAbs) to UL128 or UL130 block infection of epithelial and endothelial cells but not of fibroblasts (63). Recently, it has been shown that UL128, UL130, and UL131A assemble with gH and gL to form a five-protein complex (thereafter designated gH/gL/UL128-131A) that is an alternative to the previously described gCIII complex made of gH, gL, and gO (22, 28, 48, 63).In immunocompetent individuals T-cell and antibody responses efficiently control HCMV infection and reduce pathological consequences of maternal-fetal transmission (13, 67), although this is usually not sufficient to eradicate the virus. Albeit with controversial results, HCMV immunoglobulins (Igs) have been administered to transplant patients in association with immunosuppressive treatments for prophylaxis of HCMV disease (56, 57), and a recent report suggests that they may be effective in controlling congenital infection and preventing disease in newborns (32). These products are plasma derivatives with relatively low potency in vitro (46) and have to be administered by intravenous infusion at very high doses in order to deliver sufficient amounts of neutralizing antibodies (4, 9, 32, 56, 57, 66).The whole spectrum of antigens targeted by HCMV-neutralizing antibodies remains poorly characterized. Using specific immunoabsorption to recombinant antigens and neutralization assays using fibroblasts as model target cells, it was estimated that 40 to 70% of the serum neutralizing activity is directed against gB (6). Other studies described human neutralizing antibodies specific for gB, gH, or gM/gN viral glycoproteins (6, 14, 26, 29, 34, 41, 52, 60). Remarkably, we have recently shown that human sera exhibit a more-than-100-fold-higher potency in neutralizing infection of endothelial cells than infection of fibroblasts (20). Similarly, CMV hyperimmunoglobulins have on average 48-fold-higher neutralizing activities against epithelial cell entry than against fibroblast entry (10). However, epitopes that are targeted by the antibodies that comprise epithelial or endothelial cell-specific neutralizing activity of human immune sera remain unknown.In this study we report the isolation of a large panel of human monoclonal antibodies with extraordinarily high potency in neutralizing HCMV infection of endothelial and epithelial cells and myeloid cells. With the exception of a single antibody that recognized a conserved epitope of UL128, all other antibodies recognized conformational epitopes that required expression of two or more proteins of the gH/gL/UL128-131A complex.  相似文献   

12.
Herpesviruses can enter host cells using pH-dependent endocytosis pathways in a cell-specific manner. Envelope glycoprotein B (gB) is conserved among all herpesviruses and is a critical component of the complex that mediates membrane fusion and entry. Here we demonstrate that mildly acidic pH triggers specific conformational changes in herpes simplex virus (HSV) gB. The antigenic structure of gB was specifically altered by exposure to low pH both in vitro and during entry into host cells. The oligomeric conformation of gB was altered at a similar pH range. Exposure to acid pH appeared to convert virion gB into a lower-order oligomer. The detected conformational changes were reversible, similar to those in other class III fusion proteins. Exposure of purified, recombinant gB to mildly acidic pH resulted in similar changes in conformation and caused gB to become more hydrophobic, suggesting that low pH directly affects gB. We propose that intracellular low pH induces alterations in gB conformation that, together with additional triggers such as receptor binding, are essential for virion-cell fusion during herpesviral entry by endocytosis.Herpes simplex virus (HSV) is an important human pathogen, causing significant morbidity and mortality worldwide. HSV enters host cells by fusion of the viral envelope with either an endosomal membrane (38) or the plasma membrane (63). The entry pathway taken is thought to be determined by both virus (17, 45) and host cell (4, 17, 35, 39, 45) factors. Based on experiments with lysosomotropic agents, which elevate the normally low pH of endosomes, acidic pH has been implicated in the endocytic entry of HSV into several cell types, including human epithelial cells (37). Low pH has also recently been implicated in cell infection by several other human and veterinary herpesviruses (1, 21, 26, 47). The mechanistic role of endosomal pH in herpesvirus entry into cells is not known.Herpesviruses are a paradigm for membrane fusion mediated by a complex of several glycoproteins. We have proposed that HSV likely encodes machinery to mediate both pH-dependent and pH-independent membrane fusion reactions. Envelope glycoproteins glycoprotein B (gB) and gD and the heterodimer gH-gL are required for both pH-independent and pH-dependent entry pathways (11, 22, 30, 39, 46). Interaction of gD with one of its cognate receptors is an essential trigger for membrane fusion and entry (13, 52), regardless of the cellular pathway. However, engagement of a gD receptor is not sufficient for fusion, and at least one additional unknown trigger involving gB or gH-gL is likely necessary. gB is conserved among all herpesviruses, and in all cases studied to date, it plays roles in viral entry, including receptor binding and membrane fusion. The crystal structure of an ectodomain fragment of HSV type 1 (HSV-1) gB is an elongated, rod-like structure containing hydrophobic internal fusion loops (28). This structure bears striking architectural homology to the low pH, postfusion form of G glycoprotein from vesicular stomatitis virus (VSV-G) (43). Both the gB and G structures have features of class I and class II fusion proteins and are thus designated class III proteins (57).During entry of the majority of virus families, low pH acts directly on glycoproteins to induce membrane fusion (60). In some cases, the low pH trigger is not sufficient, or it plays an indirect role. For example, host cell proteases, such as cathepsins D and L, require intravesicular low pH to cleave Ebola virus and severe acute respiratory syndrome (SARS) glycoproteins to trigger fusion (14, 51).We investigated the role of low pH in the molecular mechanism of herpesviral entry. The results suggest that mildly acidic pH, similar to that found within endosomes, triggers a conformational change in gB. We propose that, together with other cellular cues such as receptor interaction, intracellular low pH can play a direct activating role in HSV membrane fusion and entry.  相似文献   

13.
In animal models of infection, glycoprotein E (gE) is required for efficient herpes simplex virus type 1 (HSV-1) spread from the inoculation site to the cell bodies of innervating neurons (retrograde direction). Retrograde spread in vivo is a multistep process, in that HSV-1 first spreads between epithelial cells at the inoculation site, then infects neurites, and finally travels by retrograde axonal transport to the neuron cell body. To better understand the role of gE in retrograde spread, we used a compartmentalized neuron culture system, in which neurons were infected in the presence or absence of epithelial cells. We found that gE-deleted HSV-1 (NS-gEnull) retained retrograde axonal transport activity when added directly to neurites, in contrast to the retrograde spread defect of this virus in animals. To better mimic the in vivo milieu, we overlaid neurites with epithelial cells prior to infection. In this modified system, virus infects epithelial cells and then spreads to neurites, revealing a 100-fold retrograde spread defect for NS-gEnull. We measured the retrograde spread defect of NS-gEnull from a variety of epithelial cell lines and found that the magnitude of the spread defect from epithelial cells to neurons correlated with epithelial cell plaque size defect, indicating that gE plays a similar role in both types of spread. Therefore, gE-mediated spread between epithelial cells and neurites likely explains the retrograde spread defect of gE-deleted HSV-1 in vivo.Herpes simplex virus type 1 (HSV-1) is an alphaherpesvirus that characteristically infects skin and mucosal surfaces before spreading to sensory neurons, where it establishes a lifelong persistent infection. The virus periodically returns to the periphery via sensory axons and causes recurrent lesions as well as asymptomatic shedding. This life cycle requires viral transport along axons in two directions: toward the neuron cell body (retrograde direction) and away from the neuron cell body (anterograde direction).Many studies of alphaherpesvirus neuronal spread have focused on pseudorabies virus (PRV), a virus whose natural host is the pig. Three PRV proteins, glycoprotein E (gE), gI, and Us9, have been shown to mediate anterograde neuronal spread both in animal models of infection and in cultured neurons. However, these three proteins are dispensable for retrograde spread (3, 8, 11, 12, 31, 46). In contrast, numerous animal models of infection have shown that HSV-1 gE is required for retrograde spread from the inoculation site to the cell bodies of innervating neurons (4, 9, 44, 56). In the murine flank model, wild-type (WT) virus replicates in the skin and then infects sensory neurons and spreads in a retrograde direction to the dorsal root ganglia (DRG). In this model, gE-deleted HSV-1 replicates in the skin but is not detected in the DRG (9, 44). This phenotype differs from gE-deleted PRV, which is able to reach the DRG at WT levels (8). Thus, unlike PRV, gE-deleted HSV-1 viruses have a retrograde spread defect in vivo.HSV-1 gE is a 552-amino-acid type I membrane protein found in the virion membrane as well as in the trans-Golgi and plasma membranes of infected cells (1). gE forms a heterodimer with another viral glycoprotein, gI. The gE/gI complex is important for HSV-1 immune evasion through its Fc receptor activity. gE/gI binds to the Fc domain of antibodies directed against other viral proteins, sequestering these antibodies and blocking antibody effector functions (27, 32, 40). Additionally, gE/gI promotes spread between epithelial cells. Viruses lacking either gE or gI form characteristically small plaques in cell culture and small inoculation site lesions in mice (4, 9, 18, 40, 58). In animal models, gE and gI also mediate viral spread in both anterograde and retrograde directions (4, 19, 44, 56).In order to better understand the role of gE in HSV-1 retrograde neuronal spread, we employed a compartmentalized neuron culture system that has been used to study directional neuronal spread of PRV and West Nile virus (12, 14, 45). In the Campenot chamber system, neurites are contained in a compartment that is separate from their corresponding cell bodies. Therefore, spread in an exclusively retrograde direction can be measured by infecting neurites and detecting spread to neuron cell bodies.HSV-1 replication requires retrograde transport of incoming viral genomes to the nucleus. In neurites, fusion between viral and cellular membranes occurs at the plasma membrane (43, 48). Upon membrane fusion, the capsid and a subset of tegument proteins (the inner tegument) dissociate from glycoproteins and outer tegument proteins, which remain at the plasma membrane (28, 38). Unenveloped capsids and the associated inner tegument proteins are then transported in the retrograde direction to the nucleus (7, 48, 49).For both neurons and epithelial cells, retrograde transport is dependent upon microtubules, ATP, the retrograde microtubule motor dynein, and the dynein cofactor dynactin (22, 34, 49, 52). Several viral proteins interact with components of the dynein motor complex (23, 39, 60). However, none of these proteins suggest a completely satisfactory mechanism by which viral retrograde transport occurs, either because they are not components of the complex that is transported to the nucleus (UL34, UL9, VP11/12) or because capsids lacking that protein retain retrograde transport activity (VP26) (2, 17, 21, 28, 37). This implies that additional viral proteins are involved in retrograde trafficking.We sought to better characterize the role of gE in retrograde spread and found that gE is dispensable for retrograde axonal transport; however, it promotes HSV-1 spread from epithelial cells to neurites. This epithelial cell-to-neuron spread defect provides a plausible explanation for the retrograde spread defect of gE-deleted HSV-1 in animal models of infection.  相似文献   

14.
15.
Upon entry, neuroinvasive herpesviruses traffic from axon terminals to the nuclei of neurons resident in peripheral ganglia, where the viral DNA is deposited. A detailed analysis of herpes simplex virus type 1 (HSV-1) transport dynamics in axons following entry is currently lacking. Here, time lapse fluorescence microscopy was used to compare the postentry viral transport of two neurotropic herpesviruses: HSV-1 and pseudorabies virus (PRV). HSV-1 capsid transport dynamics were indistinguishable from those of PRV and did not differ in neurons of human, mouse, or avian origin. Simultaneous tracking of capsids and tegument proteins demonstrated that the composition of actively transporting HSV-1 is remarkably similar to that of PRV. This quantitative assessment of HSV-1 axon transport following entry demonstrates that HSV-1 and PRV share a conserved mechanism for postentry retrograde transport in axons and provides the foundation for further studies of the retrograde transport process.Herpes simplex virus type 1 (HSV-1) and the veterinary herpesvirus pathogen pseudorabies virus (PRV) establish latent infections within the peripheral nervous systems (PNS) of their hosts. Neurotropic herpesviruses gain access to the PNS at nerve endings present in infected skin or mucosal tissue. Upon entry at the nerve terminal, viral particles are transported in axons toward the neuronal cell body to ultimately deposit the viral genome into the nucleus. This process is referred to as retrograde transport and is critical for the establishment of latency. Following reactivation, progeny viral particles travel anterogradely from the ganglia toward the nerve terminals, resulting in reinfection of the dermis or other innervated tissues. Reactivated infection can manifest in various forms, including asymptomatic virus shedding or mild focal lesions (herpes labialis), or less frequently in more-severe disease (herpes keratitis, encephalitis, and in the case of varicella-zoster virus, shingles).All herpesviruses consist of an icosahedral capsid that contains the viral genome surrounded by a layer of proteins known as the tegument, which is contained within a membrane envelope (33). HSV-1 and PRV capsids disassociate from the viral envelope (2, 13, 14, 22, 23, 25, 28, 30, 40) and several tegument proteins (13, 16, 21, 25) upon fusion-mediated entry into cells. However, following entry into epithelial cell lines, the VP1/2 and UL37 tegument proteins are detected in association with cytosolic capsids of PRV by immunogold electron microscopy (16) and colocalize with HSV-1 capsids at the nuclear membrane by immunofluorescence microscopy (8). In primary sensory neurons, VP1/2 and UL37 are observed to be cotransported with PRV capsids during retrograde transport by time lapse fluorescence microscopy (21), and the kinetics of axon transport have been assessed (39).Although HSV-1 and PRV share similarities in their neurotropism in vivo (reviewed in reference 12), studies of axon transport have indicated possible mechanistic differences relevant to the underlying cell biology of neural transmission (reviewed in reference 10). As a result, a live-cell analysis comparing PRV and HSV-1 is needed to determine if axon transport mechanisms are conserved between the two neuroinvasive herpesvirus genera: Simplexvirus (HSV-1) and Varicellovirus (PRV). In this study, the retrograde transport process that delivers capsids to the nuclei of sensory neurons was compared for HSV-1 (strains KOS and F) and PRV (strain Becker).  相似文献   

16.
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   

17.
Human respiratory syncytial virus (HRSV) fusion (F) protein is an essential component of the virus envelope that mediates fusion of the viral and cell membranes, and, therefore, it is an attractive target for drug and vaccine development. Our aim was to analyze the neutralizing mechanism of anti-F antibodies in comparison with other low-molecular-weight compounds targeted against the F molecule. It was found that neutralization by anti-F antibodies is related to epitope specificity. Thus, neutralizing and nonneutralizing antibodies could bind equally well to virions and remained bound after ultracentrifugation of the virus, but only the former inhibited virus infectivity. Neutralization by antibodies correlated with inhibition of cell-cell fusion in a syncytium formation assay, but not with inhibition of virus binding to cells. In contrast, a peptide (residues 478 to 516 of F protein [F478-516]) derived from the F protein heptad repeat B (HRB) or the organic compound BMS-433771 did not interfere with virus infectivity if incubated with virus before ultracentrifugation or during adsorption of virus to cells at 4°C. These inhibitors must be present during virus entry to effect HRSV neutralization. These results are best interpreted by asserting that neutralizing antibodies bind to the F protein in virions interfering with its activation for fusion. Binding of nonneutralizing antibodies is not enough to block this step. In contrast, the peptide F478-516 or BMS-433771 must bind to F protein intermediates generated during virus-cell membrane fusion, blocking further development of this process.Human respiratory syncytial virus (HRSV), a member of the Pneumovirus genus of the Paramyxoviridae family, is the main cause of severe lower respiratory tract infections in very young children (36), and it is a pathogen of considerable importance in the elderly (24, 26) and in immunocompromised adults (22). Currently, there is no effective vaccine against the virus although it is known that passive administration of neutralizing antibodies to individuals at high risk is an effective immunoprophylaxis (37, 38).The HRSV genome is a single-stranded negative-sense RNA molecule of approximately 15 kb that encodes 11 proteins (16, 53). Two of these proteins are the main surface glycoproteins of the virion. These are (i) the attachment (G) protein, which mediates virus binding to cells (44), and (ii) the fusion (F) protein, which promotes both fusion of the viral and cell membranes at the initial stages of the infectious cycle and fusion of the membrane of infected cells with those of adjacent cells to form characteristic syncytia (72). These two glycoproteins are the only targets of neutralizing antibodies either induced in animal models (19, 63, 65, 70) or present in human sera (62).The G protein is a highly variable type II glycoprotein that shares neither sequence identity nor structural features with the attachment protein of other paramyxoviruses (75). It is synthesized as a precursor of about 300 amino acids (depending on the strain) that is modified posttranslationally by the addition of a large number of N- and O-linked oligosaccharides and is also palmitoylated (17). The G protein is oligomeric (probably a homotetramer) (23) and promotes binding of HRSV to cell surface proteoglycans (35, 40, 49, 67). Whether this is the only interaction of G with cell surface components is presently unknown.The F protein is a type I glycoprotein that is synthesized as an inactive precursor of 574 amino acids (F0) which is cleaved by furin during transport to the cell surface to yield two disulfide-linked polypeptides, F2 from the N terminus and F1 from the C terminus (18). Like other viral type I fusion proteins, the mature F protein is a homotrimer which is in a prefusion, metastable, conformation in the virus particle. After fusion, the F protein adopts a highly stable postfusion conformation. Stability of the postfusion conformation is determined to great extent by two heptad repeat (HR) sequences, HRA and HRB, present in the F1 chain. Mixtures of HRA and HRB peptides form spontaneously heterotrimeric complexes (43, 51) that assemble in six-helix bundles (6HB), consisting of an internal core of three HRA helices surrounded by three antiparallel HRB helices, as determined by X-ray crystallography (79).The three-dimensional (3D) structure of the HRSV F protein has not been solved yet. Nevertheless, the structures of the pre- and postfusion forms of two paramyxovirus F proteins have revealed substantial conformational differences between the pre- and postfusion conformations (77, 78). The present hypothesis about the mechanism of membrane fusion mediated by paramyxovirus F proteins proposes that, following binding of the virus to the cell surface, the prefusion form of the F glycoprotein is activated, and membrane fusion is triggered. The F protein experiences then a series of conformational changes which include the exposure of a hydrophobic region, called the fusion peptide, and its insertion into the target membrane. Subsequent refolding of this intermediate leads to formation of the HRA and HRB six-helix bundle, concomitant with approximation of the viral and cell membranes that finally fuse, placing the fusion peptide and the transmembrane domain in the same membrane (4, 20). The formation of the 6HB and the associated free energy change are tightly linked to the merger of the viral and cellular membranes (60).Antibodies play a major role in protection against HRSV. Animal studies have demonstrated that immunization with either F or G glycoproteins induces neutralizing antibodies and protects against a viral challenge (19, 63, 70). Furthermore, transfer of these antibodies (31, 56) or of anti-F or anti-G monoclonal antibodies (MAbs) protects mice, cotton rats, or calves against either a human or bovine RSV challenge, respectively (65, 68, 73). Likewise, infants at high risk of severe HRSV disease are protected by the prophylactic administration of immunoglobulins with high anti-HRSV neutralizing titers (33). Finally, a positive correlation was found between high titers of serum neutralizing antibodies and protection in adult volunteers challenged with HRSV (34, 74), while an inverse correlation was found between high titers of neutralizing antibodies and risk of infection in children (29) and in the elderly (25).Whereas all the anti-G monoclonal antibodies reported to date are poorly neutralizing (1, 28, 48, 71), some anti-F monoclonal antibodies have strong neutralization activity (1, 3, 5, 28, 46). It is believed that HRSV neutralization by anti-G antibodies requires simultaneous binding of several antibodies to different epitopes, leading to steric hindrance for interaction of the G glycoprotein with the cell surface. Indeed, it has been shown that neutralization is enhanced by mixtures of anti-G monoclonal antibodies (1, 50), mimicking the effect of polyclonal anti-G antibodies. In contrast, highly neutralizing anti-F monoclonal antibodies do not require cooperation by other antibodies to block HRSV infectivity efficiently (1).In addition to neutralizing antibodies, other low-molecular-weight compounds directed against the F protein are potent inhibitors of HRSV infectivity. Synthetic peptides that reproduce sequences of heptad repeat B inhibit both membrane fusion promoted by the F protein and HRSV infectivity (42). Also, other small molecules obtained by chemical synthesis have been shown to interact with F protein and inhibit HRSV infectivity. These HRSV entry inhibitors have been the topic of intense research in recent years (55).This study explores the mechanisms of HRSV neutralization by different inhibitors of membrane fusion, including anti-F monoclonal antibodies, an HRB peptide, and the synthetic compound BMS-433771 (13-15). The results obtained indicate that antibodies and low-molecular-weight compounds block membrane fusion at different stages during virus entry.  相似文献   

18.
Of the four required herpes simplex virus (HSV) entry glycoproteins, the precise role of gH-gL in fusion remains the most elusive. The heterodimer gH-gL has been proposed to mediate hemifusion after the interaction of another required glycoprotein, gD, with a receptor. To identify functional domains of HSV-1 gH, we generated 22 randomized linker-insertion mutants. Analyses of 22 gH mutants revealed that gH is relatively tolerant of insertion mutations, as 15 of 22 mutants permitted normal processing and transport of gH-gL to the cell surface. gH mutants that were not expressed well at the cell surface did not function in fusion or viral entry. The screening of gH mutants for function revealed the following: (i) for wild-type gH and some gH mutants, fusion with nectin-1-expressing target cells occurred more rapidly than with herpesvirus entry mediator (HVEM)-expressing target cells; (ii) some gH mutants reduced the rate of cell fusion without abrogating fusion completely, indicating that gH may play a role in governing the kinetics of fusion and may be responsible for a rate-limiting first stage in HSV-1 fusion; and (iii) only one gH mutant, located within the short cytoplasmic tail, completely abrogated function, indicating that the gH cytoplasmic tail is crucial for cell fusion and viral infectivity.Herpes simplex virus (HSV), an enveloped neurotropic virus, infects target cells via membrane fusion, a process executed by viral fusion proteins capable of inserting into target membranes. Unlike many enveloped viruses that induce fusion through the activity of a single viral fusion protein, HSV requires four glycoproteins, glycoprotein B (gB), glycoprotein D (gD), glycoprotein H (gH), and glycoprotein L (gL), to execute fusion (6, 40, 42). The focus of this study, gH, is expressed as a heterodimer with gL (gH-gL). HSV gH and gL rely on one another for proper folding, posttranslational processing, and transport to the cell and virion surface (5, 23, 35).A sequential model of entry is the prevailing working hypothesis of HSV entry (1-3, 28, 32, 41). Viral attachment is mediated by the binding of glycoprotein C (gC) or gB to cell surface glycosaminoglycans such as heparan sulfate (38). The subsequent fusion between the virion envelope and host cell membrane is thought to result from a series of concerted events. First, gD binds to one of its host cell receptors. These receptors include herpesvirus entry mediator (HVEM), a member of the tumor necrosis factor (TNF) receptor family; nectin-1 and nectin-2, cell adhesion molecules of the Ig superfamily; and heparan sulfate modified by specific 3-O-sulfotransferases (39).It was previously proposed that gD binding a receptor induces a conformational change that allows for interactions between gD, gB, and/or gH-gL (1, 2, 8, 10, 16, 25, 32). It is thought that while gD functions primarily in receptor binding, gB and gH-gL function as the core fusion machinery of HSV.Based on its crystal structure, gB has structural features typical of viral fusion proteins in general and is structurally similar to vesicular stomatitis virus (VSV) glycoprotein G, the fusion protein of VSV (22, 34). In addition to its resemblance to other viral fusogens, gB also binds its own receptor, paired immunoglobulin-like receptor (PILRalpha) (36, 37). Importantly, HSV gB does not successfully execute fusion in the absence of gD or gH-gL (41). Compared to the other required HSV entry glycoproteins, relatively little is known about the specific roles of gH-gL during fusion. The structure of gH-gL is unknown, although in silico analyses and studies of synthetic gH peptides suggested that gH also has fusogenic properties (12, 13, 17-20).gD, a gD receptor, and gH-gL have been shown to be sufficient for inducing hemifusion, the mixing of the proximal leaflets of the viral and host cell bilayers (41). Several lines of research suggest that the subsequent step in fusion is an interaction between gH-gL and gB, with the latter glycoprotein being required for a committed and expanding fusion pore (1-3, 16, 28, 41). However, it is still unclear whether the gB and gH-gL interaction requires that gD first bind a receptor (1, 3), indicating that another viable model of HSV entry may be nonsequential gD-gB-gH-gL complex formation.Several domains important for fusion within HSV gH have been discerned. The only function associated with the N-terminal domain of HSV gH, to date, is gL binding. Residues 377 to 397 within a predicted alpha-helix in the gH ectodomain are required for cell-cell fusion and complementation of a gH-null virus (18). The mutation of a predicted heptad repeat region spanning residues 443 to 471 abrogated cell-cell fusion (17). Insertion mutations within what has been termed the pretransmembrane region of gH have also been shown to abrogate fusion and viral entry (11). The glycine residue at position 812 within the predicted gH transmembrane domain was shown previously to be important for fusion (21). Finally, although the deletion of the final six residues of gH (residues 832 to 838), which are within its short cytoplasmic tail, has no effect on fusion, further deletions were shown to decrease polykaryocyte formation by a syncytial HSV strain (4, 43).We used a transposon-based comprehensive random linker-insertion mutagenesis strategy to generate a library of mutants spanning the entire length of HSV-1 gH, an 838-amino-acid type I membrane protein. A panel of 22 insertion mutants was generated, 15 of which were expressed at near-normal levels on the cell surface. Interestingly, some insertions reduced the rate of cell fusion rather than abrogating cell fusion activity altogether, suggesting that gH may have a role in governing the kinetics of fusion and may be responsible for a rate-limiting first stage in HSV-1 fusion. Additionally, one insertion mutation that completely abrogated cell fusion and viral infectivity is located within the gH cytoplasmic tail, indicating that the short C-terminal tail of gH is critical for cell fusion and entry mediated by HSV-1.  相似文献   

19.
Little is known about the transmission or tropism of the newly discovered human retrovirus, human T-cell lymphotropic virus type 3 (HTLV-3). Here, we examine the entry requirements of HTLV-3 using independently expressed Env proteins. We observed that HTLV-3 surface glycoprotein (SU) binds efficiently to both activated CD4+ and CD8+ T cells. This contrasts with both HTLV-1 SU, which primarily binds to activated CD4+ T cells, and HTLV-2 SU, which primarily binds to activated CD8+ T cells. Binding studies with heparan sulfate proteoglycans (HSPGs) and neuropilin-1 (NRP-1), two molecules important for HTLV-1 entry, revealed that these molecules also enhance HTLV-3 SU binding. However, unlike HTLV-1 SU, HTLV-3 SU can bind efficiently in the absence of both HSPGs and NRP-1. Studies of entry performed with HTLV-3 Env-pseudotyped viruses together with SU binding studies revealed that, for HTLV-1, glucose transporter 1 (GLUT-1) functions at a postbinding step during HTLV-3 Env-mediated entry. Further studies revealed that HTLV-3 SU binds efficiently to naïve CD4+ T cells, which do not bind either HTLV-1 or HTLV-2 SU and do not express detectable levels of HSPGs, NRP-1, and GLUT-1. These results indicate that the complex of receptor molecules used by HTLV-3 to bind to primary T lymphocytes differs from that of both HTLV-1 and HTLV-2.The primate T-cell lymphotropic virus (PTLV) group of deltaretroviruses consists of three types of human T-cell lymphotropic viruses (HTLVs) (HTLV-1, HTLV-2, HTLV-3), their closely related simian T-cell lymphotropic viruses (STLVs) (STLV-1, STLV-2, STLV-3), an HTLV (HTLV-4) for which a simian counterpart has not been yet identified, and an STLV (STLV-5) originally described as a divergent STLV-1 (5-7, 30, 35, 37, 38, 45, 51, 53). HTLV-1 and HTLV-2, which have a 70% nucleotide homology, differ in both their pathobiology and tropism (reviewed in reference 13). While HTLV-1 causes a neurological disorder (tropical spastic paraparesis/HTLV-1-associated myelopathy) and a hematological disease (adult T-cell leukemia/lymphoma) (15, 42, 55), HTLV-2 is only rarely associated with tropical spastic paraparesis/HTLV-1-associated myelopathy-like disease and is not definitively linked to any lymphoproliferative disease (12, 20). In vivo, both HTLV-1 and HTLV-2 infect T cells. Although HTLV-1 is primarily found in CD4+ T cells, other cell types in the peripheral blood of infected individuals have been found to contain HTLV-1, including CD8+ T cells, dendritic cells, and B cells (19, 29, 33, 36, 46).Binding and entry of retroviruses requires specific interactions between the Env glycoproteins on the virus and cell surface receptor complexes on target cells. For HTLV-1, three molecules have been identified as important for entry, as follows: heparan sulfate proteoglycans (HSPGs), neuropilin-1 (NRP-1), and glucose transporter 1 (GLUT-1) (16, 22, 26, 28, 29, 34, 39, 44). Recent studies support a model in which HSPG and NRP-1 function during the initial binding of HTLV-1 to target cells, and GLUT-1 functions at a postattachment stage, most likely to facilitate fusion (29, 34, 49). Efficient HTLV-2 binding and entry requires NRP-1 and GLUT-1 but not HSPGs (16, 26, 39, 49).This difference in the molecules required for binding to target cells reflects differences in the T-cell tropisms of these two viruses. Activated CD4+ T cells express much higher levels of HSPGs than CD8+ T cells (26). In infected individuals, HTLV-1 is primarily found in CD4+ T cells, while HTLV-2 is primarily found in CD8+ T cells (21, 43, 46). In vitro, HTLV-1 preferentially transforms CD4+ T cells while HTLV-2 preferentially transforms CD8+ T cells, and this difference has been mapped to the Env proteins (54).We and others have reported the discovery of HTLV-3 in two Cameroonese inhabitants (6, 7, 53). We recently uncovered the presence of a third HTLV-3 strain in a different population living several hundred kilometers away from the previously identified groups (5), suggesting that this virus may be common in central Africa. Since the HTLV-3 sequences were obtained by PCR amplification of DNA isolated from peripheral blood mononuclear cells (PBMCs) of infected individuals, little is known about its tropism and pathobiology in vivo. Based on the correlation between HSPG expression levels and viral tropisms of HTLV-1 and HTLV-2, we reasoned that knowledge about the HTLV-3 receptors might provide insight into the tropism of this virus. We therefore generated vectors expressing HTLV-3 Env proteins and used them to begin to characterize the receptor complex used by HTLV-3 to bind and enter cells.  相似文献   

20.
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   

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