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 31 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.
Borna disease virus (BDV), the prototypic member of the Bornaviridae family within the order Mononegavirales, exhibits high neurotropism and provides an important and unique experimental model system for studying virus-cell interactions within the central nervous system. BDV surface glycoprotein (G) plays a critical role in virus cell entry via receptor-mediated endocytosis, and therefore, G is a critical determinant of virus tissue and cell tropism. However, the specific cell pathways involved in BDV cell entry have not been determined. Here, we provide evidence that BDV uses a clathrin-mediated, caveola-independent cell entry pathway. We also show that BDV G-mediated fusion takes place at an optimal pH of 6.0 to 6.2, corresponding to an early-endosome compartment. Consistent with this finding, BDV cell entry was Rab5 dependent but Rab7 independent and exhibited rapid fusion kinetics. Our results also uncovered a key role for microtubules in BDV cell entry, whereas the integrity and dynamics of actin cytoskeleton were not required for efficient cell entry of BDV.Borna disease virus (BDV) causes central nervous system disease in a variety of vertebrate species that is frequently manifested by behavioral abnormalities (27, 59). BDV is the causative agent of Borna disease, an often fatal immune-mediated neurological disease naturally occurring mainly in horses and sheep (21, 26, 47). However, current evidence indicates that the natural host range, prevalence, and geographic distribution of BDV are wider than originally thought (25, 31). Experimentally, BDV has a wide host range, and both host and viral factors contribute to a variable period of incubation and heterogeneity in the symptoms and pathology associated with BDV infection (20, 23, 34, 50). Notably, cases of proventricular dilatation disease affecting different species of psittacine birds have recently been linked to infection with avian bornaviruses (24, 29), a finding that expands the natural host range of bornavirus infections associated with clinical manifestations.BDV is an enveloped virus with a nonsegmented negative-strand RNA genome (11, 33, 53, 55) whose gene organization [3′-N-P-p10 (X)-M-G-L-5′] is characteristic of mononegaviruses. However, on the basis of its unique genetic and biological features, BDV is considered to be the prototypic member of a new virus family, Bornaviridae, within the order Mononegavirales.The BDV surface glycoprotein G plays a key role in receptor recognition and cell entry (20, 46). The G gene directs the synthesis of a precursor, GPC, with a predicted Mr of ca. 56 kDa, but due to its extensive glycosylation, GPC migrates with an Mr of 84 to 94 kDa. GPC is posttranslationally cleaved by the cellular protease furin into GP-1 and GP-2, corresponding to the N-and C-terminal regions, respectively, of G (2, 8, 19, 49). GP-1 has been shown to be sufficient for virus cell entry via receptor-mediated endocytosis (46), whereas GP-2 likely mediates the pH-dependent fusion event between BDV and cell membranes required for a BDV productive infection (19). In vivo, neurons are the initial target of BDV, suggesting a restricted expression pattern of a yet-unidentified virus receptor. Late in infection, BDV is detected in many tissues and organs as a consequence of its centrifugal spread via the axoplasm of peripheral nerve tissues. Receptor-independent mechanisms also contribute to cell-to-cell propagation of BDV (8).The paucity of cell-free virus associated with BDV infection has hindered studies aimed at the elucidation of the mechanisms involved in BDV cell entry. To overcome this problem, we generated a replication-competent recombinant vesicular stomatitis virus expressing BDV G (rVSVΔG*/BDVG) (45). Cells infected with rVSVΔG*/BDVG produced high titers (107 PFU/ml) of cell-free virus progeny. Notably, rVSVΔG*/BDVG recreated the cell tropism and entry pathway of bona fide BDV, thus providing a unique tool for the investigation of BDV G-mediated cell entry.Viruses that enter cells via receptor-mediated endocytosis mainly use trafficking pathways mediated by either clathrin or caveola, although alternative entry pathways have been also reported (36). Nevertheless, clathrin-mediated endocytosis (CME) is the route most commonly used by enveloped viruses for cell internalization (35). The initial virus-cell surface receptor interaction results in the activation of different signaling pathways leading to the accumulation of clathrin coated-pits and subsequent formation of endocytotic vesicles (43). Another major endocytotic pathway used by several viruses, including Ebola virus (16) and SV40 (44), uses caveolae for viral internalization into the cell. This endocytotic pathway is strictly dependent on recruitment of lipid rafts to the cell surface, an event mediated by cholesterol. In this regard, we have recently documented the requirements of cholesterol and structural integrity of cell surface lipid rafts for efficient cell entry of BDV (9).In this work, we provide evidence for the first time that BDV cell entry follows a CME-dependent, caveola-independent pathway. Moreover, we show that BDV entry is Rab5 dependent but Rab7 independent and that BDV G-mediated fusion has a rapid kinetics and an optimal pH between 6.0 and 6.2. These findings indicate that BDV G-mediated fusion occurs within the early-endosome compartment. We also provide evidence that microtubules, but not actin dynamics, play a role in BDV cell entry likely by mediating trafficking of BDV-containing endosomes to the subcellular location where viral and endosomal membranes fuse.  相似文献   

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.
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   

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.
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   

7.
Mature glycoprotein spikes are inserted in the Lassa virus envelope and consist of the distal subunit GP-1, the transmembrane-spanning subunit GP-2, and the signal peptide, which originate from the precursor glycoprotein pre-GP-C by proteolytic processing. In this study, we analyzed the oligomeric structure of the viral surface glycoprotein. Chemical cross-linking studies of mature glycoprotein spikes from purified virus revealed the formation of trimers. Interestingly, sucrose density gradient analysis of cellularly expressed glycoprotein showed that in contrast to trimeric mature glycoprotein complexes, the noncleaved glycoprotein forms monomers and oligomers spanning a wide size range, indicating that maturation cleavage of GP by the cellular subtilase SKI-1/S1P is critical for formation of the correct oligomeric state. To shed light on a potential relation between cholesterol and GP trimer stability, we performed cholesterol depletion experiments. Although depletion of cholesterol had no effect on trimerization of the glycoprotein spike complex, our studies revealed that the cholesterol content of the viral envelope is important for the infectivity of Lassa virus. Analyses of the distribution of viral proteins in cholesterol-rich detergent-resistant membrane areas showed that Lassa virus buds from membrane areas other than those responsible for impaired infectivity due to cholesterol depletion of lipid rafts. Thus, derivation of the viral envelope from cholesterol-rich membrane areas is not a prerequisite for the impact of cholesterol on virus infectivity.Lassa virus (LASV) is a member of the family Arenaviridae, of which Lymphocytic choriomeningitis virus (LCMV) is the prototype. Arenaviruses comprise more than 20 species, divided into the Old World and New World virus complexes (19). The Old World arenaviruses include the human pathogenic LASV strains, Lujo virus, which was first identified in late 2008 and is associated with an unprecedented high case fatality rate in humans, the nonhuman pathogenic Ippy, Mobala, and Mopeia viruses, and the recently described Kodoko virus (10, 30, 49). The New World virus complex contains, among others, the South American hemorrhagic fever-causing viruses Junín virus, Machupo virus, Guanarito virus, Sabiá virus, and the recently discovered Chapare virus (22).Arenaviruses contain a bisegmented single-stranded RNA genome encoding the polymerase L, matrix protein Z, nucleoprotein NP, and glycoprotein GP. The bipartite ribonucleoprotein of LASV is surrounded by a lipid envelope derived from the plasma membrane of the host cell. The matrix protein Z has been identified as a major budding factor, which lines the interior of the viral lipid membrane, in which GP spikes are inserted (61, 75). The glycoprotein is synthesized as precursor protein pre-GP-C and is cotranslationally cleaved by signal peptidase into GP-C and the signal peptide, which exhibits unusual length, stability, and topology (3, 27, 28, 33, 70, 87). Moreover, the arenaviral signal peptide functions as trans-acting maturation factor (2, 26, 33). After processing by signal peptidase, GP-C of both New World and Old World arenaviruses is cleaved by the cellular subtilase subtilisin kexin isozyme-1/site-1 protease (SKI-1/S1P) into the distal subunit GP-1 and the membrane-anchored subunit GP-2 within the secretory pathway (5, 52, 63). For LCMV, it has been shown that GP-1 subunits are linked to each other by disulfide bonds and are noncovalently connected to GP-2 subunits (14, 24, 31). GP-1 is responsible for binding to the host cell receptor, while GP-2 mediates fusion between the virus envelope and the endosomal membrane at low pH due to a bipartite fusion peptide near the amino terminus (24, 36, 44). Sequence analysis of the LCMV GP-2 ectodomain revealed two heptad repeats that most likely form amphipathic helices important for this process (34, 86).In general, viral class I fusion proteins have triplets of α-helical structures in common, which contain heptad repeats (47, 73). In contrast, class II fusion proteins are characterized by β-sheets that form dimers in the prefusion status and trimers in the postfusion status (43). The class III fusion proteins are trimers that, unlike class I fusion proteins, were not proteolytically processed N-terminally of the fusion peptide, resulting in a fusion-active membrane-anchored subunit (39, 62). Previous studies with LCMV described a tetrameric organization of the glycoprotein spikes (14), while more recent data using a bacterially expressed truncated ectodomain of the LCMV GP-2 subunit pointed toward a trimeric spike structure (31). Due to these conflicting data regarding the oligomerization status of LCMV GP, it remains unclear to which class of fusion proteins the arenaviral glycoproteins belong.The state of oligomerization and the correct conformation of viral glycoproteins are crucial for membrane fusion during virus entry. The early steps of infection have been shown for several viruses to be dependent on the cholesterol content of the participating membranes (i.e., either the virus envelope or the host cell membrane) (4, 9, 15, 20, 21, 23, 40, 42, 53, 56, 76, 78, 79). In fact, it has been shown previously that entry of both LASV and LCMV is susceptible to cholesterol depletion of the target host cell membrane using methyl-β-cyclodextrin (MβCD) treatment (64, 71). Moreover, cholesterol not only plays an important role in the early steps during entry in the viral life cycle but also is critical in the virus assembly and release process. Several viruses of various families, including influenza virus, human immunodeficiency virus type 1 (HIV-1), measles virus, and Ebola virus, use the ordered environment of lipid raft microdomains. Due to their high levels of glycosphingolipids and cholesterol, these domains are characterized by insolubility in nonionic detergents under cold conditions (60, 72). Recent observations have suggested that budding of the New World arenavirus Junin virus occurs from detergent-soluble membrane areas (1). Assembly and release from distinct membrane microdomains that are detergent soluble have also been described for vesicular stomatitis virus (VSV) (12, 38, 68). At present, however, it is not known whether LASV requires cholesterol in its viral envelope for successful virus entry or whether specific membrane microdomains are important for LASV assembly and release.In this study, we first investigated the oligomeric state of the premature and mature LASV glycoprotein complexes. Since it has been shown for several membrane proteins that the oligomerization and conformation are dependent on cholesterol (58, 59, 76, 78), we further analyzed the dependence of the cholesterol content of the virus envelope on glycoprotein oligomerization and virus infectivity. Finally, we characterized the lipid membrane areas from which LASV is released.  相似文献   

8.
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   

9.
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   

10.
Hepatitis C virus (HCV) entry occurs via a pH- and clathrin-dependent endocytic pathway and requires a number of cellular factors, including CD81, the tight-junction proteins claudin 1 (CLDN1) and occludin, and scavenger receptor class B member I (SR-BI). HCV tropism is restricted to the liver, where hepatocytes are tightly packed. Here, we demonstrate that SR-BI and CLDN1 expression is modulated in confluent human hepatoma cells, with both receptors being enriched at cell-cell junctions. Cellular contact increased HCV pseudoparticle (HCVpp) and HCV particle (HCVcc) infection and accelerated the internalization of cell-bound HCVcc, suggesting that the cell contact modulation of receptor levels may facilitate the assembly of receptor complexes required for virus internalization. CLDN1 overexpression in subconfluent cells was unable to recapitulate this effect, whereas increased SR-BI expression enhanced HCVpp entry and HCVcc internalization, demonstrating a rate-limiting role for SR-BI in HCV internalization.Hepatitis C virus (HCV) is an enveloped positive-strand RNA virus, classified in the genus Hepacivirus of the family Flaviviridae. Worldwide, approximately 170 million individuals are persistently infected with HCV, and the majority are at risk of developing chronic liver disease. Hepatocytes in the liver are thought to be the principal reservoir of HCV replication. HCV pseudoparticles (HCVpp) demonstrate a restricted tropism for hepatocyte-derived cells, suggesting that virus-encoded glycoprotein-receptor interactions play an important role in defining HCV tissue specificity.Recent evidence suggests that a number of host cell molecules are important for HCV entry: the tetraspanin CD81; scavenger receptor class B member I (SR-BI) (reviewed in reference 11); members of the tight-junction protein family claudin 1 (CLDN1), CLDN6, and CLDN9 (12, 34, 48, 52); and occludin (OCLN) (2, 33, 40). HCV enters cells via a pH- and clathrin-dependent endocytic pathway; however, the exact role(s) played by each of the host cell molecules in this process is unclear (4, 8, 21, 34, 45).CD81 and SR-BI interact with HCV-encoded E1E2 glycoproteins, suggesting a role in mediating virus attachment to the cell (reviewed in reference 44). In contrast, there is minimal evidence to support direct interaction of CLDN1 or OCLN with HCV particles (12). Evans and colleagues proposed that CLDN1 acts at a late stage in the entry process and facilitates fusion between the virus and host cell membranes (12). We (13, 19) and others (9, 48) have reported that CLDN1 associates with CD81, suggesting a role for CLDN1-CD81 complexes in viral entry. Cukierman et al. recently reported that CLDN1 enrichment at cell-cell contacts may generate specialized membrane domains that promote HCV internalization (9). In this study, we demonstrate that cellular contact modulates SR-BI and CLDN1 expression levels and promotes HCV internalization. CLDN1 overexpression in subconfluent cells was unable to recapitulate this effect, whereas increased SR-BI expression enhanced HCVpp entry and HCVcc internalization rates, demonstrating a critical and rate-limiting role for SR-BI in HCV internalization.  相似文献   

11.
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   

12.
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   

13.
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   

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

15.
16.
The human scavenger receptor gp340 has been identified as a binding protein for the human immunodeficiency virus type 1 (HIV-1) envelope that is expressed on the cell surface of female genital tract epithelial cells. This interaction allows such epithelial cells to efficiently transmit infective virus to susceptible targets and maintain viral infectivity for several days. Within the context of vaginal transmission, HIV must first traverse a normally protective mucosa containing a cell barrier to reach the underlying T cells and dendritic cells, which propagate and spread the infection. The mechanism by which HIV-1 can bypass an otherwise healthy cellular barrier remains an important area of study. Here, we demonstrate that genital tract-derived cell lines and primary human endocervical tissue can support direct transcytosis of cell-free virus from the apical to basolateral surfaces. Further, this transport of virus can be blocked through the addition of antibodies or peptides that directly block the interaction of gp340 with the HIV-1 envelope, if added prior to viral pulsing on the apical side of the cell or tissue barrier. Our data support a role for the previously described heparan sulfate moieties in mediating this transcytosis but add gp340 as an important facilitator of HIV-1 transcytosis across genital tract tissue. This study demonstrates that HIV-1 actively traverses the protective barriers of the human genital tract and presents a second mechanism whereby gp340 can promote heterosexual transmission.Through correlative studies with macaques challenged with simian immunodeficiency virus (SIV), the initial targets of infection in nontraumatic vaginal exposure to human immunodeficiency virus type 1 (HIV-1) have been identified as subepithelial T cells and dendritic cells (DCs) (18, 23, 31, 36-38). While human transmission may differ from macaque transmission, the existing models of human transmission remain controversial. For the virus to successfully reach its CD4+ targets, HIV must first traverse the columnar mucosal epithelial cell barrier of the endocervix or uterus or the stratified squamous barrier of the vagina or ectocervix, whose normal functions include protection of underlying tissue from pathogens. This portion of the human innate immune defense system represents a significant impediment to transmission. Studies have placed the natural transmission rate of HIV per sexual act between 0.005 and 0.3% (17, 45). Breaks in the epithelial barrier caused by secondary infection with other sexual transmitted diseases or the normal physical trauma often associated with vaginal intercourse represent one potential means for viral exposure to submucosal cells and have been shown to significantly increase transmission (reviewed in reference 11). However, studies of nontraumatic exposure to SIV in macaques demonstrate that these disruptions are not necessary for successful transmission to healthy females. This disparity indicates that multiple mechanisms by which HIV-1 can pass through mucosal epithelium might exist in vivo. Identifying these mechanisms represents an important obstacle to understanding and ultimately preventing HIV transmission.Several host cellular receptors, including DC-specific intercellular adhesion molecule-grabbing integrin, galactosyl ceramide, mannose receptor, langerin, heparan sulfate proteoglycans (HSPGs), and chondroitin sulfate proteoglycans, have been identified that facilitate disease progression through binding of HIV virions without being required for fusion and infection (2, 3, 12, 14, 16, 25, 29, 30, 43, 46, 50). These host accessory proteins act predominately through glycosylation-based interactions between HIV envelope (Env) and the host cellular receptors. These different host accessory factors can lead to increased infectivity in cis and trans or can serve to concentrate and expose virus at sites relevant to furthering its spread within the body. The direct transcytosis of cell-free virus through primary genital epithelial cells and the human endometrial carcinoma cell line HEC1A has been described (7, 9); this is, in part, mediated by HSPGs (7). Within the HSPG family, the syndecans have been previously shown to facilitate trans infection of HIV in vitro through binding of a specific region of Env that is moderately conserved (7, 8). This report also demonstrates that while HSPGs mediate a portion of the viral transcytosis that occurs in these two cell types, a significant portion of the observed transport occurs through an HSPG-independent mechanism. Other host cell factors likely provide alternatives to HSPGs for HIV-1 to use in subverting the mucosal epithelial barrier.gp340 is a member of the scavenger receptor cysteine-rich (SRCR) family of innate immune receptors. Its numerous splice variants can be found as a secreted component of human saliva (34, 41, 42) and as a membrane-associated receptor in a large number of epithelial cell lineages (22, 32, 40). Its normal cellular function includes immune surveillance of bacteria (4-6, 44), interaction with influenza A virus (19, 20, 32, 51) and surfactant proteins in the lung (20, 22, 33), and facilitating epithelial cell regeneration at sites of cellular inflammation and damage (27, 32). The secreted form of gp340, salivary agglutinin (SAG), was identified as a component of saliva that inhibits HIV-1 transmission in the oral pharynx through a specific interaction with the viral envelope protein that serves to agglutinate the virus and target it for degradation (34, 35, 41). Interestingly, SAG was demonstrated to form a direct protein-protein interaction with HIV Env (53, 54). Later, a cell surface-associated variant of SAG called gp340 was characterized as a binding partner for HIV-1 in the female genital tract that could facilitate virus transmission to susceptible targets of infection (47) and as a macrophage-expressed enhancer of infection (10).  相似文献   

17.
18.
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   

19.
The vaccinia virus (VACV) complement control protein (VCP) is an immunomodulatory protein that is both secreted from and expressed on the surface of infected cells. Surface expression of VCP occurs though an interaction with the viral transmembrane protein A56 and is dependent on a free N-terminal cysteine of VCP. Although A56 and VCP have been shown to interact in infected cells, the mechanism remains unclear. To investigate if A56 is sufficient for surface expression, we transiently expressed VCP and A56 in eukaryotic cell lines and found that they interact on the cell surface in the absence of other viral proteins. Since A56 contains three extracellular cysteines, we hypothesized that one of the cysteines may be unpaired and could therefore form a disulfide bridge with VCP. To test this, we generated a series of A56 mutants in which each cysteine was mutated to a serine, and we found that mutation of cysteine 162 abrogated VCP cell surface expression. We also tested the ability of other poxvirus complement control proteins to bind to VACV A56. While the smallpox homolog of VCP is able to bind VACV A56, the ectromelia virus (ECTV) VCP homolog is only able to bind the ECTV homolog of A56, indicating that these proteins may have coevolved. Surface expression of poxvirus complement control proteins may have important implications in viral pathogenesis, as a virus that does not express cell surface VCP is attenuated in vivo. This suggests that surface expression of VCP may contribute to poxvirus pathogenesis.Poxviruses, including vaccinia virus (VACV), encode large numbers of immunomodulatory proteins that help them establish an infection and combat the host''s immune response (10, 32). One of these is the vaccinia virus complement control protein (VCP), which is both secreted from and expressed on the surface of infected cells (9, 14, 16, 17). VCP acts against the complement system, a series of soluble proteins that is an important early component of the innate immune system and also shapes adaptive immune responses (15, 42, 43). In response to viral infection, complement can opsonize or inactivate virions and can lyse enveloped virus or infected cells (1, 3, 7, 12). Because of these pressures, a number of viruses, including herpes simplex virus, flaviviruses, and poxviruses, encode novel or host-derived regulators of complement, while others, including HIV and poxviruses, incorporate host complement regulatory proteins into virus particles (7, 11, 31, 39). Many orthopoxviruses encode a complement regulator (8, 20, 23, 29), and the most studied of these is VCP. Structurally, VCP is made up of four short consensus repeats (SCR) that are the basic units of mammalian complement regulators (17, 25), and VCP has been shown to interfere with the complement cascade at multiple steps (2, 16, 20-22, 25, 28-30, 33). Additionally, a VCP knockout virus generates smaller lesions in animal models (14, 16). While some host complement control proteins (CCPs) are secreted, many contain transmembrane domains (or a glycophosphatidylinositol anchor) and are thus expressed on the cell surface (42, 43). Thus, when we found that VCP is also expressed on the infected cell surface and protects infected cells from complement-mediated lysis in vitro (9), we believed this to be an important interaction that required further investigation. We previously found that the N-terminal cysteine on VCP was needed for surface expression and that the VACV transmembrane protein A56 was also required (9). The vaccinia virus A56 protein is a type 1 transmembrane glycoprotein that is found on the surface of infected cells and on extracellular virus particles (4, 18, 26, 27, 36). It interacts with another viral protein, K2 (19, 37, 45), which lacks a transmembrane domain and binds to A56 noncovalently (36). The A56/K2 complex prevents syncytium formation between infected cells and superinfection by interacting with the vaccinia virus entry/fusion complex on virions (24, 38, 40, 41). Here we provide evidence that the N-terminal cysteine on VCP forms an intermolecular disulfide bond with cysteine 162 on the ectodomain of A56. We also demonstrate that similar interactions can occur with other poxvirus CCPs, as the smallpox virus and ectromelia virus homologs of VCP also exhibit A56-dependent surface expression.  相似文献   

20.
Although noroviruses cause the vast majority of nonbacterial gastroenteritis in humans, little is known about their life cycle, including viral entry. Murine norovirus (MNV) is the only norovirus to date that efficiently infects cells in culture. To elucidate the productive route of infection for MNV-1 into murine macrophages, we used a neutral red (NR) infectious center assay and pharmacological inhibitors in combination with dominant-negative (DN) and small interfering RNA (siRNA) constructs to show that clathrin- and caveolin-mediated endocytosis did not play a role in entry. In addition, we showed that phagocytosis or macropinocytosis, flotillin-1, and GRAF1 are not required for the major route of MNV-1 uptake. However, MNV-1 genome release occurred within 1 h, and endocytosis was significantly inhibited by the cholesterol-sequestering drugs nystatin and methyl-β-cyclodextrin, the dynamin-specific inhibitor dynasore, and the dominant-negative dynamin II mutant K44A. Therefore, we conclude that the productive route of MNV-1 entry into murine macrophages is rapid and requires host cholesterol and dynamin II.Murine noroviruses (MNV) are closely related to human noroviruses (HuNoV), the causative agent of most outbreaks of infectious nonbacterial gastroenteritis worldwide in people of all ages (4, 8, 19, 31, 43, 46, 83). Although a major public health concern, noroviruses have been an understudied group of viruses due to the lack of a tissue culture system and small animal model. Since the discovery of MNV-1 in 2003 (27), reverse genetics systems (10, 81), a cell culture model (84), and a small animal model (27) have provided the tools necessary for detailed study of noroviruses.One largely unexplored aspect of norovirus biology is the early events during viral infection that are essential during viral pathogenesis. One of these early events is the attachment of the virus particle to the host. Attachment is mediated by the protruding domain of the MNV-1 capsid (29, 30, 73). For at least three strains (MNV-1, WU-11, and S99), the attachment receptor on the cell surface of murine macrophages is terminal sialic acids, including those found on the ganglioside GD1a (72). The use of carbohydrate receptors for cell attachment is shared with HuNoV, which utilize mostly histo-blood group antigens (HBGA) (18, 34, 70, 71). These carbohydrates are present in body fluids (saliva, breast milk, and intestinal contents) and on the surface of red blood cells and intestinal epithelial cells (33). Some HuNoV strains also bind to sialic acid or heparan sulfate (60, 69). However, despite evidence that for HuNoV HBGA are a genetic susceptibility marker (35), the presence of attachment receptors is not sufficient for a productive infection for either HuNoV (24) or MNV-1 (72). Although the cellular tropism of HuNoV is unknown, MNV infects murine macrophages and dendritic cells in vitro and in vivo (80, 84). Following attachment, MNV-1 infection of murine macrophages and dendritic cells can proceed in the presence of the endosome acidification inhibitor chloroquine or bafilomycin A1, suggesting that MNV-1 entry occurs independently of endosomal pH (54). However, the cellular pathway(s) utilized by MNV-1 during entry remains unclear.Viruses are obligate intracellular pathogens that hijack cellular processes to deliver their genome into cells. The most commonly used endocytic pathway during virus entry is clathrin-mediated endocytosis (41). Clathrin-coated vesicles form at the plasma membrane, pinch off by the action of the small GTPase dynamin II, and deliver their contents to early endosomes (12). For example, vesicular stomatitis virus (VSV) enters cells in this manner (66). However, viruses can also use several clathrin-independent pathways to enter cells, some of which require cholesterol-rich microdomains (i.e., lipid rafts) in the plasma membrane (56). The best studied of these is mediated by caveolin and was initially elucidated through studies of simian virus 40 (SV40) entry (1). SV40 uptake occurs via caveolin-containing vesicles that are released from the plasma membrane in a dynamin II-dependent manner and later fuse with pH-neutral caveosomes (28, 48, 53). Although caveolin-mediated endocytosis is a well-characterized form of cholesterol-dependent endocytosis, other entry mechanisms exist that are clathrin and caveolin independent (5, 14, 55, 57-59, 64, 78). In addition, macropinocytosis and/or phagocytosis can also play a role in viral entry (11, 13, 21, 36, 40, 42, 44, 45). However, the requirement for dynamin II in these processes is not fully understood.Viral entry has been addressed primarily by pharmacologic inhibitor studies, immunofluorescence and electron microscopy, transfections of dominant-negative (DN) constructs, and more recently by small interfering RNA (siRNA) knockdown. Each of these approaches has some limitations; thus, a combination of approaches is needed to elucidate the mechanism of viral entry into host cells. For example, using electron and fluorescence microscopy, which require a high particle number, does not allow the differentiation of infectious and noninfectious particles. Alternatively, the use of pharmacological inhibitors can result in off-target effects, including cytotoxicity. A recent approach used the photoreactive dye neutral red (NR) in an infectious focus assay to determine the mechanism of poliovirus entry (6). Cells were infected in the dark in the presence of neutral red, and virus particles passively incorporated the dye. Upon exposure to light, the neutral red dye cross-linked the viral genome to the viral capsid, thus inactivating the virus. Infectious foci were counted several days later. This assay was performed in the presence of various pharmacologic inhibitors of endocytosis. When an inhibitor blocked a productive route of infection, the number of infectious foci was significantly less than that for an untreated control. Major advantages of this technique over traditional assays are the ability to treat cells with pharmacologic inhibitors only during the viral entry process, the reduction of cytotoxicity, and the ability to infect with a low multiplicity of infection (MOI). Furthermore, infectious virus that is prohibited from uncoating is inactivated by illumination. Therefore, only virus particles leading to a productive infection in the presence or absence of the various inhibitors are measured. We successfully adapted this assay for use with MNV-1. Together with the use of pharmacological inhibitors, DN constructs, and siRNA knockdown, we demonstrate that the major MNV-1 entry pathway into murine macrophages resulting in a productive infection occurred by endocytosis and not phagocytosis or macropinocytosis in a manner that was clathrin and caveolin 1, flotillin 1, and GRAF1 independent but required dynamin II and cholesterol.  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