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.
Murine noroviruses (MNV) are closely related to the human noroviruses (HuNoV), which cause the majority of nonbacterial gastroenteritis. Unlike HuNoV, MNV grow in culture and in a small-animal model that represents a tractable model to study norovirus biology. To begin a detailed investigation of molecular events that occur during norovirus binding to cells, the crystallographic structure of the murine norovirus 1 (MNV-1) capsid protein protruding (P) domain has been determined. Crystallization of the bacterially expressed protein yielded two different crystal forms (Protein Data Bank identifiers [PDB ID], 3LQ6 and 3LQE). Comparison of the structures indicated a large degree of structural mobility in loops on the surface of the P2 subdomain. Specifically, the A′-B′ and E′-F′ loops were found in open and closed conformations. These regions of high mobility include the known escape mutation site for the neutralizing antibody A6.2 and an attenuation mutation site, which arose after serial passaging in culture and led to a loss in lethality in STAT1−/− mice, respectively. Modeling of a Fab fragment and crystal structures of the P dimer into the cryoelectron microscopy three-dimensional (3D) image reconstruction of the A6.2/MNV-1 complex indicated that the closed conformation is most likely bound to the Fab fragment and that the antibody contact is localized to the A′-B′ and E′-F′ loops. Therefore, we hypothesize that these loop regions and the flexibility of the P domains play important roles during MNV-1 binding to the cell surface.Murine noroviruses (MNV) are members of the family Caliciviridae, which contains small icosahedral viruses with positive-sense, single-stranded RNA genomes (18). MNV is related to human noroviruses (HuNoV), which cause most of the sporadic cases and outbreaks of infectious nonbacterial gastroenteritis worldwide in people of all ages (4, 15, 28, 36, 38, 64). However, noroviruses are an understudied group of viruses due to the previous lack of a tissue culture system and small-animal model. Since its discovery in 2003 (23), MNV has become an increasingly important model to study norovirus biology (66). The availability of a small-animal model, cell culture, and reverse-genetics system, combined with many shared characteristics of human and murine noroviruses, allows detailed studies of norovirus biology (7, 23, 63, 65, 66).The norovirus genome is organized into 3 major open reading frames (ORFs), which encode the nonstructural polyprotein (∼200 kDa) and the major (VP1; ∼58-kDa) and minor (VP2; ∼20-kDa) capsid proteins (18). Recently, a putative ORF-4 was identified in MNV, but the existence of that product and its function remain unknown (60). Norovirus capsids are formed from 180 copies of VP1 arranged with T=3 icosahedral symmetry (9, 25, 46-48). Each capsid protein is divided into an N-terminal arm (N), a shell (S), and a C-terminal protruding (P) domain, with the last two domains connected by a short hinge. VP1 self-assembles into virus-like particles (VLPs) in baculovirus, mammalian, and plant expression systems (21, 22, 50, 57, 67). The S domain forms a smooth shell around the viral genome but is unable to bind to receptors (3, 55). The P domain dimerizes, forming arch-like structures on the capsid surface, and is subdivided into P1 (the stem of the arch) and P2 (the top of the arch) subdomains. The sequence of the P2 subdomain is the least conserved, followed by the P1 and S domains with the highest degree of conservation. While the S domain of Norwalk virus (NV) is required in order to form VLPs in a baculovirus expression system, the P domains contribute to stability by intermolecular interactions (3, 24). The homodimeric interactions of the HuNoV P domain, observed by crystallographic studies of VLPs, is retained when the protein region is expressed in a bacterial expression system (55). In addition, the norovirus P domain, specifically the P2 subdomain, contains the sites for antigenicity, immune-driven evolution, and cell binding (13a, 20, 25, 32, 41, 51, 56). For MNV-1, the Fab fragment of the neutralizing antibody A6.2 binds to the outermost tip of the P2 subdomain and is thought to prevent infection by blocking capsid-receptor interaction (25).Early steps in the norovirus life cycle are determinants of norovirus tropism (19) and thereby determine the outcome of a viral infection. While the tropism of HuNoV remains unknown, MNV-1 has a tropism for murine macrophages and dendritic cells in vitro and in vivo (62, 65). Recent studies from our laboratory demonstrated that MNV-1 binds to sialic acid on murine macrophages, in particular on the ganglioside GD1a (58). It subsequently enters murine macrophages and dendritic cells in a pH-independent manner (43). To better understand MNV-cell surface binding, we expressed, purified, and determined the high-resolution structure of the MNV-1 P domain at 2.0-Å resolution. Here, we show that, similar to HuNoV P domains (10, 55), recombinant MNV-1 P domains can be expressed and fold in a biologically correct manner. This was shown by the ability of the recombinant MNV-1 P domain to bind murine macrophages, to competitively inhibit MNV-1 infection, and to be recognized by the neutralizing antibody A6.2, which interferes with macrophage binding. Expressed P domain yielded different crystal forms with significant structural differences in the outermost loops of the P2 subdomains. Overall, the MNV-1 P-domain crystal structures show tertiary structures similar to those of HuNoV P domains, with the greatest structural variation in the polypeptide loops on the outer surface of the P domain corresponding to the mobile regions among the various crystal forms. In particular, one of these loops, E′-F′, was observed in “open” and “closed” conformations. Modeling of a Fab fragment and the crystal structures of the P domain into the cryoelectron microscopy three-dimensional (3D) reconstruction of the Fab/MNV-1 complex indicated that the “closed” conformation is the form likely being bound by the neutralizing antibody A6.2. Two sequences located in the A′-B′ and E′-F′ loops were identified as epitopes for A6.2. Biological support for the in silico modeling data comes from a recombinant MNV-1 in which amino acids of the Norwalk virus E′-F′ loop replaced those of MNV-1 and that was no longer neutralized by A6.2. We hypothesize that flexibility in the E′-F′ loop is important for virus-cell interaction and that A6.2 might sterically block viral binding to the cell surface and/or prevent structural changes in the viral capsid required during receptor interaction. In addition, a channel at the interphase of the P dimer was identified that is stabilized by an “ionic lock” (i.e., a bridge formed by two sets of opposing arginine and glutamic acid residues). We hypothesize that the ionic lock may act as a trigger for structural changes important during infection, possibly at the level of host cell entry. Together, these data identify several potential movements within the MNV-1 P domain, which points to the flexibility of the MNV-1 capsid.  相似文献   

2.
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   

3.
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   

4.
Human noroviruses in the Caliciviridae family are the major cause of nonbacterial epidemic gastroenteritis worldwide. Primary human norovirus infection does not elicit lasting protective immunity, a fact that could greatly affect the efficacy of vaccination strategies. Little is known regarding the pathogenesis of human noroviruses or the immune responses that control them because there has previously been no small-animal model or cell culture system of infection. Using the only available small-animal model of norovirus infection, we found that primary high-dose murine norovirus 1 (MNV-1) infection fails to afford protection against a rechallenge with a homologous virus. Thus, MNV-1 represents a valuable model with which to dissect the pathophysiological basis for the lack of lasting protection against human norovirus infection. Interestingly, the magnitude of protection afforded by a primary MNV-1 infection inversely correlates with the inoculum dose. Future studies will elucidate the mechanisms by which noroviruses avoid the induction of protective immunity and the role played by the inoculum dose in this process, ultimately translating this knowledge into successful vaccination approaches.Human noroviruses (NVs) are estimated to be responsible for >95% of the nonbacterial epidemic gastroenteritis that occurs worldwide. The course of the disease is rapid, with symptoms including vomiting, diarrhea, and nausea arising approximately 24 h following infection and typically resolving 24 to 48 h later. NV outbreaks occur most commonly in semiclosed communities such as nursing homes, schools, hospitals, cruise ships, and military settings (11, 24, 31). Persons of all ages are susceptible to NV infection. Human NVs are thus associated with considerable morbidity and have a significant economic impact. Numerous human volunteer challenge studies have demonstrated that long-term immunity is not induced following primary NV infection of some volunteers (13, 20, 26). The pathophysiological basis for this lack of protection is unclear, since virus-specific adaptive immune responses are generated (1, 7, 9, 10). A similar lack of immunity has been observed in some individuals for a number of other viral pathogens that infect at mucosal surfaces, such as rhinoviruses (32) and respiratory syncytial virus (RSV) (16). Importantly, typical vaccination strategies have been unsuccessful at eliciting protective anti-RSV immunity and studies with animal models to understand the lack of immunity to either natural or vaccinating virus have been uninformative because protection is induced in animals (27). Extrapolating from RSV studies, it may be difficult to vaccinate against NVs and it will be important to understand the underlying cause in order to design more efficacious treatment regimens. Studies with a small-animal model recapitulating this atypical immune outcome would be extremely valuable.  相似文献   

5.
6.
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   

7.
For recognition of infected cells by CD8 T cells, antigenic peptides are presented at the cell surface, bound to major histocompatibility complex class I (MHC-I) molecules. Downmodulation of cell surface MHC-I molecules is regarded as a hallmark function of cytomegalovirus-encoded immunoevasins. The molecular mechanisms by which immunoevasins interfere with the MHC-I pathway suggest, however, that this downmodulation may be secondary to an interruption of turnover replenishment and that hindrance of the vesicular transport of recently generated peptide-MHC (pMHC) complexes to the cell surface is the actual function of immunoevasins. Here we have used the model of murine cytomegalovirus (mCMV) infection to provide experimental evidence for this hypothesis. To quantitate pMHC complexes at the cell surface after infection in the presence and absence of immunoevasins, we generated the recombinant viruses mCMV-SIINFEKL and mCMV-Δm06m152-SIINFEKL, respectively, expressing the Kb-presented peptide SIINFEKL with early-phase kinetics in place of an immunodominant peptide of the viral carrier protein gp36.5/m164. The data revealed ∼10,000 Kb molecules presenting SIINFEKL in the absence of immunoevasins, which is an occupancy of ∼10% of all cell surface Kb molecules, whereas immunoevasins reduced this number to almost the detection limit. To selectively evaluate their effect on preexisting pMHC complexes, cells were exogenously loaded with SIINFEKL peptide shortly after infection with mCMV-SIINFEKA, in which endogenous presentation is prevented by an L174A mutation of the C-terminal MHC-I anchor residue. The data suggest that pMHC complexes present at the cell surface in advance of immunoevasin gene expression are downmodulated due to constitutive turnover in the absence of resupply.CD8 T cells recognize infected cells by interaction of their T-cell receptor (TCR) with a cell surface presentation complex composed of a cognate antigenic peptide bound to a presenting allelic form of a major histocompatibility complex class I (MHC-I) glycoprotein (77, 85, 97, 98). The number of such “peptide receptors” per cell has been estimated to be on the order of 105 to 106 for each MHC-I allomorph (for a review, see reference 82). Viral antigenic peptides are generated within infected cells by proteolytic processing of viral proteins, usually in the proteasome, and associate with nascent MHC-I proteins in the endoplasmic reticulum (ER) before the peptide-MHC (pMHC) complexes travel to the cell surface with the cellular vesicular flow (for reviews, see references 13, 87, 92, and 93). CD8 T cells have long been known to protect against cytomegalovirus (CMV) infection and disease in animal models (60, 72; reviewed in references 33 and 36) and in humans (9, 61, 67, 75, 76). As shown only recently in the murine CMV (mCMV) model of infection of immunocompromised mice by adoptive transfer of epitope-specific CD8 T cells, antiviral protection against CMV is indeed TCR mediated and epitope dependent. Specifically, memory cells purified by TCR-based epitope-specific cell sorting, as well as cells of a peptide-selected cytolytic T-lymphocyte line, protected against mCMV expressing the cognate antigenic peptide, the IE1 peptide 168-YPHFMPTNL-176 in this example, but failed to control infection with a recombinant mCMV expressing a peptide analogue in which the C-terminal MHC-I anchor residue leucine was replaced with alanine (3).Interference with the MHC-I pathway of antigen presentation has evolved as a viral immune evasion mechanism of CMVs and other viruses, mediated by virally encoded proteins that inhibit MHC-I trafficking to the cell surface (for reviews, see references 1, 24, 27, 29, 63, 70, 71, 84, and 95). These molecules are known as immunoevasins (50, 70, 89), as “viral proteins interfering with antigen presentation” (VIPRs) (95), or as negative “viral regulators of antigen presentation” (vRAPs) (34). Although the detailed molecular mechanisms differ between different CMV species in their respective hosts, the common biological outcome is the inhibition of antigen presentation. Accordingly, downmodulation of MHC-I cell surface expression is a hallmark of molecular immune evasion and actually led to the discovery of this class of molecules. Since CD8 T cells apparently protect against infection with wild-type CMV strains despite the expression of immunoevasins, the in vivo relevance of these molecules is an issue of current interest and investigation (for a review, see reference 14). As shown recently with the murine model, antigen presentation in infected host cells is not completely blocked for all epitopes, because pMHC complexes that are constitutively formed in sufficiently large amounts can exhaust the inhibitory capacity of the immunoevasins (40). Likewise, enhancing antigen processing conditionally with gamma interferon (IFN-γ) aids in peptide presentation in the presence of immunoevasins (18, 28). Thus, by raising the threshold of the amount of peptide required for presentation, immunoevasins determine whether a particular viral peptide can function as a protective epitope—an issue of relevance for rational vaccine design as well (94). Whereas deletion of immunoevasin genes gives only incremental improvement to the control of infection in immunocompetent mice (22, 51), expression of immunoevasins reduces the protective effect of adoptively transferred CD8 T cells in immunocompromised recipients (37, 40, 47, 48). In a bone marrow transplantation model, immunoevasins were recently found to contribute to enhanced and prolonged virus replication during hematopoietic reconstitution and, consequently, also to higher latent viral genome loads in the lungs and a higher incidence of virus recurrence (4). Notably, however, immunoevasins do not inhibit but, rather, enhance CD8 T-cell priming (5, 21, 22, 56), due to higher viral replication levels in draining lymph nodes associated with sustained antigen supply for the cross-priming of CD8 T cells by uninfected antigen-presenting cells (5).For mCMV, three molecules are proposed to function as vRAPs, only two of which are confirmed negative regulators that downmodulate cell surface MHC-I (34, 62, 89) and inhibit the presentation of antigenic peptides to CD8 T cells (34, 62). Immunoevasin gp40/m152 transiently interacts with MHC-I molecules and mediates their retention in a cis-Golgi compartment (96), whereas gp48/m06 stably binds to MHC-I molecules in the ER and mediates sorting of the complexes for lysosomal degradation by a mechanism that involves the cellular cargo sorting adaptor proteins AP1-A and AP3-A (73, 74). The third proposed immunoevasin of mCMV, gp34/m04 (46), also binds stably to MHC-I molecules. A function as a CD8 T-cell immunoevasin was predicted from some alleviation of immune evasion for certain epitopes and MHC-I molecules in cells infected with the deletion mutant mCMV-Δm04 (34, 42, 89), but gp34/m04 does not reduce the steady-state level of cell surface class I molecules and does not inhibit peptide presentation when expressed selectively after infection with mCMV-Δm06m152 (34, 62). The m04-MHC-I complexes are expressed on the cell surface (46) and appear to be involved in the modulation of natural killer cell activity (45).Here we give the first report on quantitating the efficacy of immunoevasins in terms of absolute numbers of pMHC complexes displayed at the cell surface. By comparing the fate of pMHC complexes already present at the cell surface in advance of immunoevasin gene expression with that of newly formed pMHC complexes, our data provide direct evidence to conclude that downmodulation of cell surface MHC-I molecules is secondary to an interruption of the flow of newly formed pMHC complexes to the cell surface.(Part of this work was presented at the 12th International CMV/Betaherpesvirus Workshop, 10 to 14 May 2009, Boston, MA.)  相似文献   

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

9.
Moloney murine leukemia virus (MoMLV) Gag utilizes its late (L) domain motif PPPY to bind members of the Nedd4-like ubiquitin ligase family. These interactions recruit components of the cell''s budding machinery that are critical for virus release. MoMLV Gag contains two additional L domains, PSAP and LYPAL, that are believed to drive residual MoMLV release via interactions with cellular proteins Tsg101 and Alix, respectively. We found that overexpression of Tsg101 or Alix failed to rescue the release of PPPY-deficient MoMLV via these other L domains. However, low-level expression of the ubiquitin ligase Itch potently rescued the release and infectivity of MoMLV lacking PPPY function. In contrast, other ubiquitin ligases such as WWP1, Nedd4.1, Nedd4.2, and Nedd4.2s did not rescue this release-deficient virus. Efficient rescue required the ubiquitin ligase activity of Itch and an intact C2 domain but not presence of the endophilin-binding site. Additionally, we found Itch to immunoprecipitate with MoMLV Gag lacking the PPPY motif and to be incorporated into rescued MoMLV particles. The PSAP and LYPAL motifs were dispensable for Itch-mediated virus rescue, and their absence did not affect the incorporation of Itch into the rescued particles. Itch-mediated rescue of release-defective MoMLV was sensitive to inhibition by dominant-negative versions of ESCRT-III components and the VPS4 AAA ATPase, indicating that Itch-mediated correction of MoMLV release defects requires the integrity of the host vacuolar sorting protein pathway. RNA interference knockdown of Itch suppressed the residual release of the MoMLV lacking the PPPY motif. Interestingly, Itch stimulation of the PPPY-deficient MoMLV release was accompanied by the enhancement of Gag ubiquitination and the appearance of new ubiquitinated Gag proteins in virions. Together, these results suggest that Itch can facilitate MoMLV release in an L domain-independent manner via a mechanism that requires the host budding machinery and involves Gag ubiquitination.Retroviruses require access to the host budding machinery to exit the cell (5, 13, 40). To this end, retroviral Gag polyproteins use short sequences called late (L) domains to promote virus release by recruiting members of the host vacuolar protein sorting (vps) machinery. In the cell, vps proteins are involved in membrane dynamics that facilitate the separation of daughter cells at the completion of cytokinesis (9, 39) and the budding of vesicles into endosomal compartments or multivesicular bodies (MVB) (2, 23), a process topologically similar to virus budding (57). Class E vps proteins are organized into three heteromeric endosomal complexes (called endosomal sorting complexes) required for transport, namely, ESCRT-I, -II, and -III (2). In the current model for budding, sequential recruitment of ESCRT components on the cytoplasmic face of the membrane facilitates vesicle invagination into MVB compartments and viral egress from the cell (2). The disassembly of ESCRT-III components is catalyzed by the activity of VPS4 AAA-type ATPase, which in turn is presumed to trigger membrane fission events (3, 50). Any disruption in this sequence, such as mutations in L domain motifs or dominant-negative interference with the function of ESCRT-III members or the VPS4 ATPase, adversely affects virus release. This indicates that Gag interactions with the ESCRT machinery are necessary for virus budding and separation from the cell (19, 21, 34, 49, 57).Currently, three types of L domain motifs have been identified: PT/SAP, LYPXnL, and PPPY. All retroviral Gag molecules contain at least one of these motifs, as multiple L domains are believed to synergistically function to ensure efficient viral release. Moloney murine leukemia virus (MoMLV) Gag carries all three L domain motifs, PSAP, LYPAL, and PPPY, which bind the vps protein Tsg101, the ESCRT-associated protein Alix (46), and members of the Nedd4-ubiquitin ligase family (33), respectively. In HIV-1, the PTAP motif in the p6 region of Gag binds Tsg101 (16, 56), which functions in viral budding (16, 35) as a member of ESCRT-I (16, 36, 57). The LYPXnL motif is also located in p6 and is the binding site for Alix (49, 57), a protein that also interacts with the nucleocapsid domain of HIV-1 Gag (14, 43) and links Gag to components of ESCRT-III (14). Similarly, the human T-cell leukemia virus (HTLV-I) Gag carries PPPY and PTAP L domains, which both contribute to efficient HTLV-1 release (6, 7, 21). The PPPY L domain motif, which is found in numerous retroviral Gag polyproteins (6, 7, 19, 21, 27, 28, 61, 62), plays a critical role in MoMLV release, as mutations disrupting its sequence lead to significant decreases in virus budding and release (33, 62). PSAP and LYPAL, the additional L domain motifs, are believed to serve little to no role in the release of MoMLV Gag virus-like particles (45, 46).The role of Nedd4-like ubiquitin ligases in budding events was initially established by data obtained with the yeast Nedd4-like ligase Rsp5, an enzyme that ubiquitinates surface proteins, thus signaling their incorporation into the MVB pathway (26). From retroviral budding studies, multiple findings support the notion that Nedd4-like ubiquitin ligases link PPPY-containing Gag proteins to the host ESCRT machinery. For example, mutations in the PPPY motif or expression of dominant-negative versions of Nedd4-like ligases resulted in budding defects similar to those seen upon interference with the function of ESCRT-III members (7, 21, 27, 28, 33, 62). Overexpression of Nedd4-like ligases WWP1 and Itch corrected the budding defects of a MoMLV PPPY mutant that retained residual binding to both ligases (33). Also, when transplanted to a heterologous retroviral Gag, the PPPY L domain creates a requirement for Nedd4-like ubiqutin ligase activity to facilitate viral release that is dependent on the presence of a functional ESCRT pathway (63). Collectively, these observations support the notion that Nedd4-like ubiquitin ligases link retroviral Gag polyproteins to components of the ESCRT pathway necessary for budding.Both endosomal and viral budding require the ubiquitin conjugation properties of Nedd4-like ligases, indicating that ubiquitin transfer to a key protein(s) is necessary to promote budding. A role for Gag ubiquitination in viral budding has been suggested (8, 20, 22, 48). In fact, ubiquitin attachment to equine infectious anemia virus (EIAV) Gag can substitute for the lack of L domains and rescue viral budding (25), suggesting that ubiquitin molecules conjugated to Gag can signal the recruitment of the host ESCRT machinery. For feline immunodeficiency virus, efficient budding seems to require L domain-dependent ubiquitination of Gag proteins (8) that is independent of the L domain ability to directly recruit Nedd4-like ubiquitin ligases (i.e., by means of the PT/SAP L domain motif) (8). Similarly, ubiquitination of HTLV-1 Gag was also shown to play a significant role in viral release (22). Conversely, data arguing in favor of a role for the ubiquitination of transacting factors, but not Gag, in the facilitation of viral budding have also been reported (10, 63). Thus Gag polyproteins recruit, in a PPPY-dependent or -independent manner, enzymatically active Nedd4-like ubiquitin ligases that conjugate ubiquitin molecules to Gag or to Gag-binding host factors. Such interactions, whether direct or indirect, are believed to link the viral protein to the host ESCRT pathway and facilitate release.In addition to the well-characterized cellular proteins that bind primary L domain motifs, retroviral Gag can recruit other host factors, either via secondary L domains or independently of L domains (10, 24, 29, 55, 59). These cellular factors are believed to promote virus production by facilitating Gag protein trafficking to the plasma membrane and/or providing additional L domain-independent links to the host vps pathway. Examples of these parallel pathways are illustrated in the rescue of a budding-defective HIV-1 lacking the PTAP domain by overexpression of Alix (15, 54) and in the remarkably potent rescue of HIV-1 lacking all known L domains by the overexpression of Nedd4.2s, a Nedd4.2 isoform that belongs to the Nedd4-like ubiquitin ligase family (10, 55). In this study, we sought to identify host cell factors that rescue budding defects of the MoMLV mutant lacking the PPPY motif (MoMLV AAAY mutant). Our studies provide evidence that Itch overexpression rescued budding and infectivity defects of the MoMLV AAAY mutant virus, indicating that Gag can recruit the ubiquitin ligase Itch in an L domain-independent manner to facilitate MoMLV release via a mechanism that involves Gag ubiquitination.  相似文献   

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

11.
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   

12.
13.
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   

14.
15.
Murine cytomegalovirus (MCMV) functions interfere with protein trafficking in the secretory pathway. In this report we used Δm138-MCMV, a recombinant virus with a deleted viral Fc receptor, to demonstrate that MCMV also perturbs endosomal trafficking in the early phase of infection. This perturbation had a striking impact on cell surface-resident major histocompatibility complex class I (MHC-I) molecules due to the complementary effect of MCMV immunoevasins, which block their egress from the secretory pathway. In infected cells, constitutively endocytosed cell surface-resident MHC-I molecules were arrested and retained in early endosomal antigen 1 (EEA1)-positive and lysobisphosphatidic acid (LBPA)-negative perinuclear endosomes together with clathrin-dependent cargo (transferrin receptor, Lamp1, and epidermal growth factor receptor). Their progression from these endosomes into recycling and degradative routes was inhibited. This arrest was associated with a reduction of the intracellular content of Rab7 and Rab11, small GTPases that are essential for the maturation of recycling and endolysosomal domains of early endosomes. The reduced recycling of MHC-I in Δm138-MCMV-infected cells was accompanied by their accelerated loss from the cell surface. The MCMV function that affects cell surface-resident MHC-I was activated in later stages of the early phase of viral replication, after the expression of known immunoevasins. MCMV without the three immunoevasins (the m04, m06, and m152 proteins) encoded a function that affects endosomal trafficking. This function, however, was not sufficient to reduce the cell surface expression of MHC-I in the absence of the transport block in the secretory pathway.Herpesviruses are well known to interfere with major histocompatibility complex class I (MHC-I) molecules in order to ensure evasion from immune recognition. A majority of evidence so far indicates that they target MHC-I maturation events and MHC-I trafficking in the secretory pathway (33), although evidence exists suggesting that herpesviruses could also interfere with MHC-I functions in endosomal pathways (8). Murine cytomegalovirus (MCMV), a member of the herpesvirus family, dedicates a substantial part of its genome to encoding nonessential genes for the modulation of cellular functions (40), including MHC-I trafficking in the secretory pathway (24, 27, 45, 48, 49, 52). All known immune evasion functions encoded by MCMV are based on a direct interaction of viral gene products with MHC-I complexes in the secretory pathway. The egress of nascent MHC-I complexes to the cell surface of MCMV-infected cells is abolished as a consequence of their retention in the endoplasmic reticulum (ER)-cis-Golgi intermediate compartment (ERGIC) by the m152 gene product (10, 19, 24, 52, 56) as well as redirection of those that escape into the Golgi compartment toward late endosomes (LEs) for degradation by the m06 MCMV gene product (45). These effects are opposed by gp34, a product of the MCMV m04 gene, which associates with MHC-I complexes and reaches the cell surface (24, 27).The loss of MHC-I from the cell surface is an expected consequence of the activity of m152 and m06, which act in the secretory pathway. The level of cell surface MHC-I is substantially reduced at later times of infection (10, 19, 24, 48, 52), and cells stably transfected with either the m152 or m06 gene do not display MHC-I at the cell surface (20, 24). If the loss of MHC-I from the cell surface is a consequence of the prevented egress from the secretory pathway, then the cell surface loss should follow the kinetics of the constitutive internalization of MHC-I complexes in the endosomal pathway. Given that the constitutive internalization is the net result of cell surface supply from the secretory pathway, endocytic uptake, and endocytic recycling, it is a slow process that occurs in normal fibroblasts at a rate of ∼6 to 8% per hour (36). Therefore, the effect of MCMV immunoevasins on cell surface MHC-I should be expected at later times of infection. However, several reports demonstrated that the level of MHC-I surface expression was already reduced in the early phase of infection (10, 45, 48, 52). Thus, it would be reasonable to expect that MCMV contributes with a function that causes the accelerated retrieval of cell surface-resident MHC-I complexes.In this report we demonstrate that MCMV perturbs endosomal trafficking very early in infection by acting on distal parts of early endosome (EE) route and affecting the trafficking of both clathrin-dependent and clathrin-independent cargoes. Clathrin-dependent cargo does not share primary endocytic carriers with MHC-I proteins (12, 14), which enter the cell via the nonclathrin Arf6-associated endocytic carriers (12, 14, 41, 42, 53), but they meet in the proximal part of the common early endocytic route and redirect to distal endocytic carriers around the cell center (12, 14). The perturbation of the distal part of the EE route has dramatic consequences on MHC-I, since it supplements the viral mechanisms that act in the secretory pathway. The net result of this perturbation is a complete loss of MHC-I molecules from the cell surface.  相似文献   

16.
We observed that the nonfusogenic mouse hepatitis virus (MHV) strain MHV-2 reached a titer of ∼2 log10 higher than that of the fusogenic strain A59 in astrocytoma DBT cells. To determine whether the spike protein is responsible for the difference, a recombinant virus, Penn-98-1, that contains the A59 genome with a spike from MHV-2 was used to infect DBT cells. Results showed that Penn-98-1 behaved like MHV-2, thus establishing a role for the spike protein in viral growth. The inverse correlation between viral fusogenicity and growth was further established in four different cell types and with a fusogenic mutant, the S757R mutant, derived from isogenic Penn-98-1. While both A59 and Penn-98-1 entered cells at similar levels, viral RNA and protein syntheses were significantly delayed for A59. Interestingly, when the genomic RNAs were delivered directly into the cells via transfection, the levels of gene expression for these viruses were similar. Furthermore, cell fractionation experiments revealed that significantly more genomic RNAs for the nonfusogenic MHVs were detected in the endoplasmic reticulum (ER) within the first 2 h after infection than for the fusogenic MHVs. Pretreatment of Penn-98-1 with trypsin reversed its properties in syncytium formation, virus production, and genome transport to the ER. These findings identified a novel role for the spike protein in regulating the uncoating and delivery of the viral genome to the ER after internalization.Murine coronavirus mouse hepatitis virus (MHV) is a member of the family Coronaviridae. It is an enveloped, positive-strand-RNA virus. The viral envelope contains three or four structural proteins, depending on the virus strain (21). The spike (S) protein is a glycoprotein with a molecular mass of approximately 180 kDa. For some MHV strains, such as JHM and A59, the S protein is cleaved by a furin-like proteinase into two subunits, the amino-terminal S1 and the carboxyl-terminal S2. The S1 subunit is thought to form the globular head of the spike and is responsible for the initial attachment of the virus to the receptor on the cell surface. The S2 subunit, which forms the stalk portion of the spike and which anchors the S protein to the viral envelope, facilitates the fusion between the viral envelope and the cell membrane and cell-cell fusion (4, 7, 20, 25, 39). In contrast, the S protein of some other MHV strains, such as MHV-2, does not undergo cleavage and usually does not cause cell-cell fusion (15, 34). It appears that the cleavability of the MHV S protein is associated usually, though not always, with its fusogenicity (10, 36). It has been suggested that the fusogenicity of the S protein may determine the route of virus entry, i.e., via direct fusion with plasma membranes or following endocytosis (11, 34), although the mechanism for virus-induced cell-cell fusion may differ from that for virus-cell fusion during entry (8). The S protein also elicits the induction of neutralizing antibodies and cell-mediated immunity in infected hosts (3). It is therefore an important determinant for viral infectivity, pathogenicity, and virulence (2, 5, 31, 38). The hemagglutinin-esterase (HE) protein is present only in certain MHV strains (22, 42) and may play a role in viral pathogenesis (44, 45). The small envelope (E) protein and the membrane (M) protein play a key role in virus assembly (40). The nucleocapsid (N) protein is a phosphoprotein of approximately 50 kDa and is associated with the RNA genome to form the nucleocapsid inside the envelope (21, 37).Infection of host cells by MHV is mediated through the interaction between the S protein and the cellular receptors that are members of the carcinoembryonic antigen (CEA) family of the immunoglobulin superfamily (9). This interaction then triggers fusion between the viral envelope and the plasma membrane or the endosomal membrane, the latter of which follows receptor-mediated endocytosis, thus allowing the nucleocapsid to deliver into the cytoplasm. Direct entry from the plasma membrane appears to be the predominant route for most MHV strains (19, 28), although entry by some mutant MHVs, such as OBLV60 and MHV-2, is low pH dependent, i.e., via endocytosis (11, 34). However, nothing is known about how the genomic RNA is transported to the rough endoplasmic reticulum (ER) for translation. Once on the ER, the viral genomic RNA is translated into a polymerase polyprotein from the 5′-end two open reading frames (two-thirds of the genome) via ribosomal frameshifting. The polymerase polyproteins in turn synthesize genomic and multiple species of subgenomic mRNAs. These mRNAs are then translated into nonstructural and structural proteins, the latter of which are essential for generation of progeny viruses.MHV can infect rodents, causing hepatitis, enteritis, nephritis, and central nervous system diseases. In the mouse central nervous system, some MHV strains, such as JHM and A59, are neurovirulent, causing acute encephalitis and chronic demyelination (1, 13), while others, such as MHV-2, exhibit extremely low neurovirulence, causing only meningitis without apparent encephalitis and demyelination (6, 16, 41). Extensive mutagenesis studies in combination with targeted RNA recombination have identified that the S protein is the major determinant of MHV pathogenicity in animals, although other viral genes also appear to modulate viral pathogenicity (17, 32). For example, the recombinant MHV Penn-98-1, which contains the S protein of MHV-2 in an A59 genome background, causes acute meningoencephalitis similar to that caused by A59 but does not cause demyelination similar to that observed for MHV-2 (6). It has also been shown that the amounts of antigen staining and necrosis in the liver correlate with the viral titer, which is determined largely by the S protein (29). However, how the S protein affects viral titer in cell culture and in animals is not known.In the present study, we initially observed that the levels of production of infectious viruses in an astrocytoma DBT cell line were markedly different among three MHV strains. Using the recombinant MHV Penn-98-1 and its isogenic S757R mutant, we further established that the S protein is responsible for the observed difference. The difference in virus production between A59 and Penn-98-1 was detected as early as 4 to 6 h postinfection (p.i.) and likely occurred during the early stages of the virus life cycle but after virus internalization. Interestingly, when the genomic RNAs were delivered directly into the cells via transfection, the levels of gene expression for these viruses were similar. Furthermore, cell fractionation experiments revealed that significantly more genomic RNAs for nonfusogenic MHVs were delivered to the ER within the first 2 h after infection than for fusogenic MHVs. These results demonstrate that the spike protein of MHV can regulate the intracellular transport of the viral genome to the ER following internalization. To our knowledge, this is the first study identifying a role for a coronavirus S protein in genome delivery in addition to its well-established role in receptor binding and virus-cell and cell-cell fusions during infection.  相似文献   

17.
18.
19.
Soil substrate membrane systems allow for microcultivation of fastidious soil bacteria as mixed microbial communities. We isolated established microcolonies from these membranes by using fluorescence viability staining and micromanipulation. This approach facilitated the recovery of diverse, novel isolates, including the recalcitrant bacterium Leifsonia xyli, a plant pathogen that has never been isolated outside the host.The majority of bacterial species have never been recovered in the laboratory (1, 14, 19, 24). In the last decade, novel cultivation approaches have successfully been used to recover “unculturables” from a diverse range of divisions (23, 25, 29). Most strategies have targeted marine environments (4, 23, 25, 32), but soil offers the potential for the investigation of vast numbers of undescribed species (20, 29). Rapid advances have been made toward culturing soil bacteria by reformulating and diluting traditional media, extending incubation times, and using alternative gelling agents (8, 21, 29).The soil substrate membrane system (SSMS) is a diffusion chamber approach that uses extracts from the soil of interest as the growth substrate, thereby mimicking the environment under investigation (12). The SSMS enriches for slow-growing oligophiles, a proportion of which are subsequently capable of growing on complex media (23, 25, 27, 30, 32). However, the SSMS results in mixed microbial communities, with the consequent difficulty in isolation of individual microcolonies for further characterization (10).Micromanipulation has been widely used for the isolation of specific cell morphotypes for downstream applications in molecular diagnostics or proteomics (5, 15). This simple technology offers the opportunity to select established microcolonies of a specific morphotype from the SSMS when combined with fluorescence visualization (3, 11). Here, we have combined the SSMS, fluorescence viability staining, and advanced micromanipulation for targeted isolation of viable, microcolony-forming soil bacteria.  相似文献   

20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

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