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.
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.
Hantaviruses infect endothelial cells and cause 2 vascular permeability-based diseases. Pathogenic hantaviruses enhance the permeability of endothelial cells in response to vascular endothelial growth factor (VEGF). However, the mechanism by which hantaviruses hyperpermeabilize endothelial cells has not been defined. The paracellular permeability of endothelial cells is uniquely determined by the homophilic assembly of vascular endothelial cadherin (VE-cadherin) within adherens junctions, which is regulated by VEGF receptor-2 (VEGFR2) responses. Here, we investigated VEGFR2 phosphorylation and the internalization of VE-cadherin within endothelial cells infected by pathogenic Andes virus (ANDV) and Hantaan virus (HTNV) and nonpathogenic Tula virus (TULV) hantaviruses. We found that VEGF addition to ANDV- and HTNV-infected endothelial cells results in the hyperphosphorylation of VEGFR2, while TULV infection failed to increase VEGFR2 phosphorylation. Concomitant with the VEGFR2 hyperphosphorylation, VE-cadherin was internalized to intracellular vesicles within ANDV- or HTNV-, but not TULV-, infected endothelial cells. Addition of angiopoietin-1 (Ang-1) or sphingosine-1-phosphate (S1P) to ANDV- or HTNV-infected cells blocked VE-cadherin internalization in response to VEGF. These findings are consistent with the ability of Ang-1 and S1P to inhibit hantavirus-induced endothelial cell permeability. Our results suggest that pathogenic hantaviruses disrupt fluid barrier properties of endothelial cell adherens junctions by enhancing VEGFR2-VE-cadherin pathway responses which increase paracellular permeability. These results provide a pathway-specific mechanism for the enhanced permeability of hantavirus-infected endothelial cells and suggest that stabilizing VE-cadherin within adherens junctions is a primary target for regulating endothelial cell permeability during pathogenic hantavirus infection.Hantaviruses cause 2 human diseases: hemorrhagic fever with renal syndrome (HFRS) and hantavirus pulmonary syndrome (HPS) (50). HPS and HFRS are multifactorial in nature and cause thrombocytopenia, immune and endothelial cell responses, and hypoxia, which contribute to disease (7, 11, 31, 42, 62). Although these syndromes sound quite different, they share common components which involve the ability of hantaviruses to infect endothelial cells and induce capillary permeability. Edema, which results from capillary leakage of fluid into tissues and organs, is a common finding in both HPS and HFRS patients (4, 7, 11, 31, 42, 62). In fact, both diseases can present with renal or pulmonary sequelae, and the renal or pulmonary focus of hantavirus diseases is likely to result from hantavirus infection of endothelial cells within vast glomerular and pulmonary capillary beds (4, 7, 11, 31, 42, 62). All hantaviruses predominantly infect endothelial cells which line capillaries (31, 42, 44, 61, 62), and endothelial cells have a primary role in maintaining fluid barrier functions of the vasculature (1, 12, 55). Although hantaviruses do not lyse endothelial cells (44, 61), this primary cellular target underlies hantavirus-induced changes in capillary integrity. As a result, understanding altered endothelial cell responses following hantavirus infection is fundamental to defining the mechanism of permeability induced by pathogenic hantaviruses (1, 12, 55).Pathogenic, but not nonpathogenic, hantaviruses use β3 integrins on the surface of endothelial cells and platelets for attachment (19, 21, 23, 39, 46), and β3 integrins play prominent roles in regulating vascular integrity (3, 6, 8, 24, 48). Pathogenic hantaviruses bind to basal, inactive conformations of β3 integrins (35, 46, 53) and days after infection inhibit β3 integrin-directed endothelial cell migration (20, 46). This may be the result of cell-associated virus (19, 20, 22) which keeps β3 in an inactive state but could also occur through additional regulatory processes that have yet to be defined. Interestingly, the nonpathogenic hantaviruses Prospect Hill virus (PHV) and Tula virus (TULV) fail to alter β3 integrin functions, and their entry is consistent with the use of discrete α5β1 integrins (21, 23, 36).On endothelial cells, αvβ3 integrins normally regulate permeabilizing effects of vascular endothelial growth factor receptor-2 (VEGFR2) (3, 24, 48, 51). VEGF was initially identified as an edema-causing vascular permeability factor (VPF) that is 50,000 times more potent than histamine in directing fluid across capillaries (12, 14). VEGF is responsible for disassembling adherens junctions between endothelial cells to permit cellular movement, wound repair, and angiogenesis (8, 10, 12, 13, 17, 26, 57). Extracellular domains of β3 integrins and VEGFR2 reportedly form a coprecipitable complex (3), and knocking out β3 causes capillary permeability that is augmented by VEGF addition (24, 47, 48). Pathogenic hantaviruses inhibit β3 integrin functions days after infection and similarly enhance the permeability of endothelial cells in response to VEGF (22).Adherens junctions form the primary fluid barrier of endothelial cells, and VEGFR2 responses control adherens junction disassembly (10, 17, 34, 57, 63). Vascular endothelial cadherin (VE-cadherin) is an endothelial cell-specific adherens junction protein and the primary determinant of paracellular permeability within the vascular endothelium (30, 33, 34). Activation of VEGFR2, another endothelial cell-specific protein, triggers signaling responses resulting in VE-cadherin disassembly and endocytosis, which increases the permeability of endothelial cell junctions (10, 12, 17, 34). VEGF is induced by hypoxic conditions and released by endothelial cells, platelets, and immune cells (2, 15, 38, 52). VEGF acts locally on endothelial cells through the autocrine or paracrine activation of VEGFR2, and the disassembly of endothelial cell adherens junctions increases the availability of nutrients to tissues and facilitates leukocyte trafficking and diapedesis (10, 12, 17, 55). The importance of endothelial cell barrier integrity is often in conflict with requirements for endothelial cells to move in order to permit angiogenesis and repair or cell and fluid egress, and as a result, VEGF-induced VE-cadherin responses are tightly controlled (10, 17, 18, 32, 33, 59). This limits capillary permeability while dynamically responding to a variety of endothelial cell-specific factors and conditions. However, if unregulated, this process can result in localized capillary permeability and edema (2, 9, 10, 12, 14, 17, 29, 60).Interestingly, tissue edema and hypoxia are common findings in both HPS and HFRS patients (11, 31, 62), and the ability of pathogenic hantaviruses to infect human endothelial cells provides a means for hantaviruses to directly alter normal VEGF-VE-cadherin regulation. In fact, the permeability of endothelial cells infected by pathogenic Andes virus (ANDV) or Hantaan virus (HTNV) is dramatically enhanced in response to VEGF addition (22). This response is absent from endothelial cells comparably infected with the nonpathogenic TULV and suggests that enhanced VEGF-induced endothelial cell permeability is a common underlying response of both HPS- and HFRS-causing hantaviruses (22). In these studies, we comparatively investigate responses of human endothelial cells infected with pathogenic ANDV and HTNV, as well as nonpathogenic TULV.  相似文献   

4.
Hantavirus cardiopulmonary syndrome (HCPS) is a highly pathogenic emerging disease (40% case fatality rate) caused by New World hantaviruses. Hantavirus infections are transmitted to humans mainly by inhalation of virus-contaminated aerosol particles of rodent excreta and secretions. At present, there are no antiviral drugs or immunotherapeutic agents available for the treatment of hantaviral infection, and the survival rates for infected patients hinge largely on early virus recognition and hospital admission and aggressive pulmonary and hemodynamic support. In this study, we show that Andes virus (ANDV) interacts with human apolipoprotein H (ApoH) and that ApoH-coated magnetic beads or ApoH-coated enzyme-linked immunosorbent assay plates can be used to capture and concentrate the virus from complex biological mixtures, such as serum and urine, allowing it to be detected by both immunological and molecular approaches. In addition, we report that ANDV-antigens and infectious virus are shed in urine of HCPS patients.Hantaviruses are segmented RNA viruses belonging to the genus Hantavirus in the family Bunyaviridae (43). Hantaviral genomes are tripartite, consisting of three different single-stranded RNA segments, designated large (L), medium (M), and small (S), that are packed into helical nucleocapsids (39, 42). These segments encode the RNA polymerase, a glycoprotein precursor that is cotranslationally processed to yield two envelope glycoproteins (Gc and Gn) and a nucleocapsid (N) protein. Hantaviruses are maintained in various rodent reservoirs, in which the hosts are persistently infected but lack disease symptoms (28, 32). Virus transmission to humans does not require direct human-to-rodent contact. Instead, human hantaviral infections are acquired by the respiratory route, most commonly through inhalation of virus-contaminated aerosol particles of rodent excreta (feces, saliva, or urine). Hantaviruses are known to cause two serious and often fatal human diseases, hemorrhagic fever with renal syndrome and hantavirus cardiopulmonary syndrome (HCPS) (19, 31). Of the two diseases, HCPS is more severe, with a mortality rate of approximately 40% (19). Hemorrhagic fever with renal syndrome is a mild-to-severe disease characterized by the development of an acute influenza-like febrile illness that may lead to hemorrhagic manifestations, and renal failure is caused by pathogenic Old World hantaviruses, which include Seoul, Hantaan, Dobrava, Tula, and Puumala viruses (28, 32, 42). The New World hantaviruses are responsible for HCPS, which is characterized by a febrile phase (prodrome) and pulmonary infection, cardiac depression, and hematologic manifestations (18, 31). HCPS pathogenesis generally includes capillary leak syndrome, which selectively involves the pulmonary bed, noncardiogenic pulmonary edema, thrombocytopenia, hypotension, and/or cardiogenic shock (19, 31). The pathogenesis of HCPS, like that of many other viral hemorrhagic fevers, is poorly understood. However, the long incubation period for illness, the generally well-advanced adaptive immune response at the time of onset of the disease, and the apparent absence of direct lytic damage to vascular endothelium, all characteristics shared with other hemorrhagic fevers, are among the findings that strongly suggest that HCPS pathogenesis is largely immune mediated (22, 27). The lack of an FDA-approved vaccine for HCPS, the absence of specific antiviral drugs or immunotherapeutic agents, and the high overall mortality rate for hantavirus infection highlight the medical significance of New World hantavirus (5, 8, 19, 28, 32).Survival rates for patients with hantaviral infection hinge largely on early virus recognition and hospital admission and aggressive pulmonary and hemodynamic support (19, 31). The diagnosis, clinical course, and supportive care of patients with New World hantaviral infections have recently been reviewed (19, 31). Unfortunately, early diagnosis of New World hantaviral infections is complex, as the prodrome leading to acute cardiopulmonary deterioration in HCPS can be confused with febrile phases produced by, for example, mycoplasmas and chlamydophilial infections (52).Human apolipoprotein H (ApoH), also known as beta 2-glycoprotein I, is a constituent of human plasma (0.2 mg/ml) notorious for binding to negatively charged surfaces (3, 7, 14, 17, 44, 45). Several reports show that ApoH also interacts with viral proteins, such as the hepatitis B virus (HBV) antigen and proteins p18, p26, and gp160 of the human immunodeficiency virus (12, 30, 46, 47). Interestingly, studies involving binding to the HBV antigen suggest that ApoH specifically binds DNA-containing HBV particles, thus discriminating, through an undefined mechanism, between active replicating virus and empty noninfectious particles (47). These findings prompted us to assess a possible interaction between ApoH and Andes virus (ANDV), which is the major etiological agent of HCPS in South America and is unique among hantaviruses in its reported ability to be transmitted from person to person (19, 29, 34). The mechanism of person-to-person dissemination of ANDV remains to be elucidated, yet it is likely that person-to-person transmission of ANDV could be explained by mechanisms similar to those described for rodent-to-rodent and rodent-to-human transmission. If so, a compulsory condition for ANDV dissemination among humans is that the infected host must shed the pathogen in, for example, urine.In this study, we show that when fixed to a solid matrix, ApoH can be used to capture and concentrate ANDV from complex biological samples, including serum and urine, allowing virus detection by both immunological and molecular approaches. Furthermore, we took advantage of the ApoH-ANDV interaction to develop a high-throughput screening assay and show for the first time ANDV-antigen shedding in the urine of patients with acute HCPS. We also report the presence of infectious viral particles in the urine of two patients with HCPS.  相似文献   

5.
Measles virus (MV) entry requires at least 2 viral proteins, the hemagglutinin (H) and fusion (F) proteins. We describe the rescue and characterization of a measles virus with a specific mutation in the stalk region of H (I98A) that is able to bind normally to cells but infects at a lower rate than the wild type due to a reduction in fusion triggering. The mutant H protein binds to F more avidly than the parent H protein does, and the corresponding virus is more sensitive to inhibition by fusion-inhibitory peptide. We show that after binding of MV to its receptor, H-F dissociation is required for productive infection.Measles virus (MV) infection requires binding of the hemagglutinin (H) protein to its cognate receptors (9, 20, 21, 29, 41) while the fusion (F) protein triggers membrane lipid mixing and fusion. The H protein is a type II transmembrane homodimeric, disulfide-linked glycoprotein (33). The F protein is a type I membrane glycoprotein that exists as a homotrimeric complex. The protein is cleaved by furin in the trans-Golgi network into a metastable heterodimer with a membrane-spanning F1 domain and a membrane-distal F2 domain (16). Expressed alone, neither H nor F leads to membrane fusion, and therefore, both proteins are required and have to interact for productive infection of a target cell (46). There is evidence that these interactions start within the endoplasmic reticulum (34).The H proteins of Paramyxoviridae family members have a globular head with a six-blade β-propellor structure that is responsible for receptor binding (4, 7, 13), a stalk region composed of alpha-helical coiled coils (18, 48) that anchors the complex to the plasma membrane, and a short cytoplasmic domain that can interact with the matrix (M) protein and modulate fusion (2). Given that the F protein does not interact with a receptor on the target cell but undergoes conformational changes to enable membrane fusion, it seems likely that the F protein must interact with the H protein that enables fusion (14, 19, 23, 24, 35, 47). The molecular interactions between the F and H proteins are being increasingly understood (6, 8, 24, 25, 30, 35, 42). Hummel and Bellini have described a mutation in the H glycoprotein where threonine replaced isoleucine 98, which led to loss of fusion in chronically infected cells, but the virus was not rescued (15). Corey and Iorio performed alanine-scanning mutagenesis to determine the role of specific, membrane-proximal residues in the stalk region of the H protein responsible for H-F interactions (6). Substitution of alanine for specific residues in this region altered cell-to-cell fusion and the strength of the H-F interaction in transient-transfection experiments (6). Replacement of isoleucine with alanine at position 98 reduced fusion but did not significantly alter hemadsorption, implying that binding of the mutant H protein to CD46 was not affected (6). More recently, Paal et al. showed that the H protein can tolerate significant additions to its alpha-helical coiled coils without loss of binding or fusion in transient-transfection assays (30). Although these studies confirm the importance of the interactions between the H protein stalk and the metastable F protein for enabling fusion after receptor binding, the exact steps leading to fusion are still unclear. Moreover, studies evaluating H-F interactions were performed with transient protein expression and not in the presence of the actual virus. This is potentially an important shortcoming since the M protein can modulate infection and fusion (1).  相似文献   

6.
Poxviruses produce complement regulatory proteins to subvert the host''s immune response. Similar to the human pathogen variola virus, ectromelia virus has a limited host range and provides a mouse model where the virus and the host''s immune response have coevolved. We previously demonstrated that multiple components (C3, C4, and factor B) of the classical and alternative pathways are required to survive ectromelia virus infection. Complement''s role in the innate and adaptive immune responses likely drove the evolution of a virus-encoded virulence factor that regulates complement activation. In this study, we characterized the ectromelia virus inhibitor of complement enzymes (EMICE). Recombinant EMICE regulated complement activation on the surface of CHO cells, and it protected complement-sensitive intracellular mature virions (IMV) from neutralization in vitro. It accomplished this by serving as a cofactor for the inactivation of C3b and C4b and by dissociating the catalytic domain of the classical pathway C3 convertase. Infected murine cells initiated synthesis of EMICE within 4 to 6 h postinoculation. The levels were sufficient in the supernatant to protect the IMV, upon release, from complement-mediated neutralization. EMICE on the surface of infected murine cells also reduced complement activation by the alternative pathway. In contrast, classical pathway activation by high-titer antibody overwhelmed EMICE''s regulatory capacity. These results suggest that EMICE''s role is early during infection when it counteracts the innate immune response. In summary, ectromelia virus produced EMICE within a few hours of an infection, and EMICE in turn decreased complement activation on IMV and infected cells.Poxviruses encode in their large double-stranded DNA genomes many factors that modify the immune system (30, 56). The analysis of these molecules has revealed a delicate balance between viral pathogenesis and the host''s immune response (2, 21, 31, 61). Variola, vaccinia, monkeypox, cowpox, and ectromelia (ECTV) viruses each produce an orthologous complement regulatory protein (poxviral inhibitor of complement enzymes [PICE]) that has structural and functional homology to host proteins (14, 29, 34, 38, 41, 45, 54). The loss of the regulatory protein resulted in smaller local lesions with vaccinia virus lacking the vaccinia virus complement control protein (VCP) (29) and in a greater local inflammatory response in the case of cowpox lacking the inflammation-modulatory protein (IMP; the cowpox virus PICE) (35, 45, 46). Additionally, the complete loss of the monkeypox virus inhibitor of complement enzymes (MOPICE) may account for part of the reduced mortality observed in the West African compared to Congo basin strains of monkeypox virus (12).The complement system consists of proteins on the cell surface and in blood that recognize and destroy invading pathogens and infected host cells (36, 52). Viruses protect themselves from the antiviral effects of complement activation in a variety of ways, including hijacking the host''s complement regulatory proteins or producing their own inhibitors (7, 8, 15, 20, 23). Another effective strategy is to incorporate the host''s complement regulators in the outermost viral membrane, which then protects the virus from complement attack (62). The extracellular enveloped virus (EEV) produced by poxviruses acquires a unique outer membrane derived from the Golgi complex or early endosomes that contain the protective host complement regulators (58, 62). Poxviruses have multiple infectious forms, and the most abundant, intracellular mature virions (IMV), are released when infected cells lyse (58). The IMV lacks the outermost membrane found on EEV and is sensitive to complement-mediated neutralization. The multiple strategies viruses have evolved to evade the complement system underscore its importance to innate and adaptive immunity (15, 36).The most well-characterized PICE is VCP (24-29, 34, 49, 50, 53, 55, 59, 60). Originally described as a secreted complement inhibitor (34), VCP also attaches to the surface of infected cells through an interaction with the viral membrane protein A56 that requires an unpaired N-terminal cysteine (26). This extra cysteine also adds to the potency of the inhibitor by forming function-enhancing dimers (41). VCP and the smallpox virus inhibitor of complement enzymes (SPICE) bind heparin in vitro, and this may facilitate cell surface interactions (24, 38, 50, 59). The coevolution of variola virus with its only natural host, humans, likely explains the enhanced activity against human complement observed with SPICE compared to the other PICEs (54, 64).Our recent work with ECTV, the causative agent of mousepox infection, demonstrated that the classical and alternative pathways of the complement system are required for host survival (48). The mouse-specific pathogen ECTV causes severe disease in most strains and has coevolved with its natural host, analogous to variola virus in humans (9). This close host-virus relationship is particularly important for evaluating the role of the complement system, given the species specificity of many complement proteins, receptors, and regulators (10, 47, 62). Additionally, the availability of complement-deficient mice permits dissection of the complement activation pathways involved. Naïve C57BL/6 mouse serum neutralizes the IMV of ECTV in vitro, predominately through opsonization (48). Maximal neutralization requires natural antibody, classical-pathway activation, and amplification by the alternative pathway. C3 deficiency in the normally resistant C57BL/6 strain results in acute mortality, similar to immunodeficiencies in important elements of the antiviral immune response, including CD8+ T cells (19, 32), natural killer cells (18, 51), and gamma interferon (33). During ECTV infection, the complement system acts in the first few hours and days to delay the spread of infection, resulting in lower levels of viremia and viral burden in tissues (48).This study characterized the PICE produced by ECTV, ectromelia virus inhibitor of complement enzymes (EMICE), and assessed its complement regulatory activity. Recombinant EMICE (rEMICE) decreased activation of both human and mouse complement. Murine cells produced EMICE at 4 to 6 h postinfection prior to the release of the majority of the complement-sensitive IMV from infected cells. rEMICE protected ECTV IMV from complement-mediated neutralization. Further, EMICE produced during natural infection inhibited complement deposition on infected cells by the alternative pathway. ECTV likely produces this abundance of EMICE to protect both the IMV and infected cells.  相似文献   

7.
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   

8.
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   

9.
Adeno-associated virus (AAV) type 2 and 5 proteins Rep52 and Rep40 were polyubiquitinated during AAV-adenovirus type 5 (Ad5) coinfection and during transient transfection in either the presence or absence of Ad5 E4orf6 and E1b-55k. Polyubiquitination of small Rep proteins via lysine 48 (K48) linkages, normally associated with targeting of proteins for proteasomal degradation, was detected only in the presence of E4orf6. The small Rep proteins were ubiquitinated via lysine 63 (K63) following transfection in either the presence or absence of E4orf6 or following coinfection with Ad5. E4orf6/E1b-55k-dependent K48-specific polyubiquitination of small Rep proteins could be inhibited using small interfering RNA (siRNA) to cullin 5.Together, adenovirus type 5 (Ad5) early gene products E1a, E1b-55k, E2a, E4orf6, and virus-associated (VA) RNA can support efficient replication of adeno-associated virus (AAV) (4, 31). E4orf6 and E1b-55k are known to interact with cellular cullin 5 (cul5), elongins B and C, and the ring box protein Rbx1 to form an E3 ubiquitin ligase complex that specifically targets a small population of cellular proteins for degradation by the proteasome (1, 7, 21, 22, 24, 27). This property has been implicated in a number of functions presumed to be required for both Ad and AAV replication (3, 8-10, 17, 23, 24, 34, 35).Previously, only p53, Mre11, DNA ligase IV, and integrin α3 had been shown to be substrates of the Ad5 E3 ubiquitin ligase complex (1, 7, 21, 22, 24, 27); however, we have recently shown (16, 17) that the small Rep proteins and capsid proteins of AAV5 are also degraded in the presence of Ad E4orf6 and E1b-55k in a proteasome-dependent manner. These proteins were restored to levels required during infection by the action of VA RNA (17). The targeting for degradation of AAV5 protein by the E4orf6/E1b-55k E3 ubiquitin ligase complex required functional BC-box motifs in E4orf6 and could be inhibited by depletion of the scaffolding protein cullin 5 using directed small interfering RNA (siRNA) (16). In addition, the degradation of AAV5 protein was partially prevented by overexpression of pUBR7, a plasmid that generates a dominant-negative ubiquitin (16). The role this targeted degradation plays in the life cycle of AAV has not yet been clarified; however, E4orf6 mutants that cannot function in this regard do not support AAV replication as well as wild-type E4orf6 (R. Nayak and D. J. Pintel, unpublished data). Degradation of Mre11 by the Ad5 E3 ligase has also been implicated in allowing efficient Ad5 and AAV replication (24). Ubiquitination of AAV Rep proteins during viral infection, however, has not previously been reported.  相似文献   

10.
To facilitate the release of infectious progeny virions, human immunodeficiency virus type 1 (HIV-1) exploits the Endosomal Sorting Complex Required for Transport (ESCRT) pathway by engaging Tsg101 and ALIX through late assembly (L) domains in the C-terminal p6 domain of Gag. However, the L domains in p6 are known to be dispensable for efficient particle production by certain HIV-1 Gag constructs that have the nucleocapsid (NC) domain replaced by a foreign dimerization domain to substitute for the assembly function of NC. We now show that one such L domain-independent HIV-1 Gag construct (termed ZWT) that has NC-p1-p6 replaced by a leucine zipper domain is resistant to dominant-negative inhibitors of the ESCRT pathway that block HIV-1 particle production. However, ZWT became dependent on the presence of an L domain when NC-p1-p6 was restored to its C terminus. Furthermore, when the NC domain was replaced by a leucine zipper, the p1-p6 region, but not p6 alone, conferred sensitivity to inhibition of the ESCRT pathway. In an authentic HIV-1 Gag context, the effect of an inhibitor of the ESCRT pathway on particle production could be alleviated by deleting a portion of the NC domain together with p1. Together, these results indicate that the ESCRT pathway dependence of HIV-1 budding is determined, at least in part, by the NC-p1 region of Gag.Human immunodeficiency virus type 1 (HIV-1) and other retroviruses hijack the cellular Endosomal Sorting Complex Required for Transport (ESCRT) pathway to promote the detachment of virions from the cell surface and from each other (3, 21, 42, 44, 47). The ESCRT pathway was initially identified based on its requirement for the sorting of ubiquitinated cargo into multivesicular bodies (MVB) (50, 51). During MVB biogenesis, the ESCRT pathway drives the membrane deformation and fission events required for the inward vesiculation of the limiting membrane of this organelle (26, 29, 50, 51). More recently, it emerged that the ESCRT pathway is also essential for the normal abscission of daughter cells during the final stage of cell division (10, 43). Most of the components of the ESCRT pathway are involved in the formation of four heteromeric protein complexes termed ESCRT-0, ESCRT-I, ESCRT-II, and ESCRT-III. Additional components include ALIX, which interacts both with ESCRT-I and ESCRT-III, and the AAA ATPase Vps4, which mediates the disassembly of ESCRT-III (29, 42).The deformation and scission of endocytic membranes is thought to be mediated by ESCRT-III, which, together with Vps4, constitutes the most conserved element of the pathway (23, 26, 42). Indeed, it was recently shown that purified yeast ESCRT-III induces membrane deformation (52), and in another study three subunits of yeast ESCRT-III were sufficient to promote the formation of intralumenal vesicles in an in vitro assay (61). In mammals, ESCRT-III is formed by the charged MVB proteins (CHMPs), which are structurally related and tightly regulated through autoinhibition (2, 33, 46, 53, 62). The removal of an inhibitory C-terminal domain induces polymerization and association with endosomal membranes and converts CHMPs into potent inhibitors of retroviral budding (34, 46, 53, 60, 62). Alternatively, CHMPs can be converted into strong inhibitors of the ESCRT pathway and of HIV-1 budding through the addition of a bulky tag such as green fluorescent protein (GFP) or red fluorescent protein (RFP) (27, 36, 39, 54). Retroviral budding in general is also strongly inhibited by catalytically inactive Vps4 (22, 41, 55), or upon Vsp4B depletion (31), confirming the crucial role of ESCRT-III.Retroviruses engage the ESCRT pathway through the activity of so-called late assembly (L) domains in Gag. In the case of HIV-1, the primary L domain maps to a conserved PTAP motif in the C-terminal p6 domain of Gag (24, 28) and interacts with the ESCRT-I component Tsg101 (15, 22, 40, 58). HIV-1 p6 also harbors an auxiliary L domain of the LYPxnL type, which interacts with the V domain of ALIX (20, 35, 39, 54, 59, 63). Interestingly, Tsg101 binding site mutants of HIV-1 can be fully rescued through the overexpression of ALIX, and this rescue depends on the ALIX binding site in p6 (20, 56). In contrast, the overexpression of a specific splice variant of the ubiquitin ligase Nedd4-2 has been shown to rescue the release and infectivity of HIV-1 mutants lacking all known L domains in p6 (12, 57). Nedd4 family ubiquitin ligases had previously been implicated in the function of PPxY-type L domains, which also depend on an intact ESCRT pathway for function (4, 32, 38). However, HIV-1 Gag lacks PPxY motifs, and the WW domains of Nedd4-2, which mediate its interaction with PPxY motifs, are dispensable for the rescue of HIV-1 L domain mutants (57).ALIX also interacts with the nucleocapsid (NC) region of HIV-1 Gag (18, 49), which is located upstream of p6 and the p1 spacer peptide. ALIX binds HIV-1 NC via its Bro1 domain, and the capacity to interact with NC and to stimulate the release of a minimal HIV-1 Gag construct is shared among widely divergent Bro1 domain proteins (48). Based on these findings and the observation that certain mutations in NC cause a phenotype that resembles that of L domain mutants, it has been proposed that NC cooperates with p6 to recruit the machinery required for normal HIV-1 budding (18, 49).NC also plays a role in Gag polyprotein multimerization, and this function of NC depends on its RNA-binding activity (5-8). It has been proposed that the role of the NC-nucleic acid interaction during assembly is to promote the formation of Gag dimers (37), and HIV-1 assembly in the absence of NC can indeed be efficiently rescued by leucine zipper dimerization domains (65). Surprisingly, in this setting the L domains in p6 also became dispensable, since particle production remained efficient even when the entire NC-p1-p6 region of HIV-1 Gag was replaced by a leucine zipper (1, 65). These findings raised the possibility that the reliance of wild-type (WT) HIV-1 Gag on a functional ESCRT pathway is, at least in part, specified by NC-p1-p6. However, it also remained possible that the chimeric Gag constructs engaged the ESCRT pathway in an alternative manner.In the present report, we provide evidence supporting the first of those two possibilities. Particle production became independent of ESCRT when the entire NC-p1-p6 region was replaced by a leucine zipper, and reversion to ESCRT dependence was shown to occur as a result of restoration of p1-p6 but not of p6 alone. Furthermore, although the deletion of p1 alone had little effect in an authentic HIV-1 Gag context, the additional removal of a portion of NC improved particle production in the presence of an inhibitor of the ESCRT pathway. Together, these data imply that the NC-p1 region plays an important role in the ESCRT-dependence of HIV-1 particle production.  相似文献   

11.
12.
The filovirus VP40 protein is capable of budding from mammalian cells in the form of virus-like particles (VLPs) that are morphologically indistinguishable from infectious virions. Ebola virus VP40 (eVP40) contains well-characterized overlapping L domains, which play a key role in mediating efficient virus egress. L domains represent only one component required for efficient budding and, therefore, there is a need to identify and characterize additional domains important for VP40 function. We demonstrate here that the 96LPLGVA101 sequence of eVP40 and the corresponding 84LPLGIM89 sequence of Marburg virus VP40 (mVP40) are critical for efficient release of VP40 VLPs. Indeed, deletion of these motifs essentially abolished the ability of eVP40 and mVP40 to bud as VLPs. To address the mechanism by which the 96LPLGVA101 motif of eVP40 contributes to egress, a series of point mutations were introduced into this motif. These mutants were then compared to the eVP40 wild type in a VLP budding assay to assess budding competency. Confocal microscopy and gel filtration analyses were performed to assess their pattern of intracellular localization and ability to oligomerize, respectively. Our results show that mutations disrupting the 96LPLGVA101 motif resulted in both altered patterns of intracellular localization and self-assembly compared to wild-type controls. Interestingly, coexpression of either Ebola virus GP-WT or mVP40-WT with eVP40-ΔLPLGVA failed to rescue the budding defective eVP40-ΔLPLGVA mutant into VLPs; however, coexpression of eVP40-WT with mVP40-ΔLPLGIM successfully rescued budding of mVP40-ΔLPLGIM into VLPs at mVP40-WT levels. In sum, our findings implicate the LPLGVA and LPLGIM motifs of eVP40 and mVP40, respectively, as being important for VP40 structure/stability and budding.Ebola and Marburg viruses are members of the family Filoviridae. Filoviruses are filamentous, negative-sense, single-stranded RNA viruses that cause lethal hemorrhagic fevers in both humans and nonhuman primates (5). Filoviruses encode seven viral proteins including: NP (major nucleoprotein), VP35 (phosphoprotein), VP40 (matrix protein), GP (glycoprotein), VP30 (minor nucleoprotein), VP24 (secondary matrix protein), and L (RNA-dependent RNA polymerase) (2, 5, 10, 12, 45). Numerous studies have shown that expression of Ebola virus VP40 (eVP40) alone in mammalian cells leads to the production of virus-like particles (VLPs) with filamentous morphology which is indistinguishable from infectious Ebola virus particles (12, 17, 18, 25, 26, 27, 30, 31, 34, 49). Like many enveloped viruses such as rhabdovirus (11) and arenaviruses (44), Ebola virus encodes late-assembly or L domains, which are sequences required for the membrane fission event that separates viral and cellular membranes to release nascent virion particles (1, 5, 7, 10, 12, 18, 25, 27, 34). Thus far, four classes of L domains have been identified which were defined by their conserved amino acid core sequences: the Pro-Thr/Ser-Ala-Pro (PT/SAP) motif (25, 27), the Pro-Pro-x-Tyr (PPxY) motif (11, 12, 18, 19, 41, 53), the Tyr-x-x-Leu (YxxL) motif (3, 15, 27, 37), and the Phe-Pro-Ile-Val (FPIV) motif (39). Both PTAP and the PPxY motifs are essential for efficient particle release for eVP40 (25, 27, 48, 49), whereas mVP40 contains only a PPxY motif. L domains are believed to act as docking sites for the recruitment of cellular proteins involved in endocytic trafficking and multivesicular body biogenesis to facilitate virus-cell separation (8, 13, 14, 16, 28, 29, 33, 36, 43, 50, 51).In addition to L domains, oligomerization, and plasma-membrane localization of VP40 are two functions of the protein that are critical for efficient budding of VLPs and virions. Specific sequences involved in self-assembly and membrane localization have yet to be defined precisely. However, recent reports have attempted to identify regions of VP40 that are important for its overall function in assembly and budding. For example, the amino acid region 212KLR214 located at the C-terminal region was found to be important for efficient release of eVP40 VLPs, with Leu213 being the most critical (30). Mutation of the 212KLR214 region resulted in altered patterns of cellular localization and oligomerization of eVP40 compared to those of the wild-type genotype (30). In addition, the proline at position 53 was also implicated as being essential for eVP40 VLP release and plasma-membrane localization (54).In a more recent study, a YPLGVG motif within the M protein of Nipah virus (NiV) was shown to be important for stability, membrane binding, and budding of NiV VLPs (35). Whether this NiV M motif represents a new class of L domain remains to be determined. However, it is clear that this YPLGVG motif of NiV M is important for budding, perhaps involving a novel mechanism (35). Our rationale for investigating the corresponding, conserved motifs present within the Ebola and Marburg virus VP40 proteins was based primarily on these findings with NiV. In addition, Ebola virus VP40 motif maps close to the hinge region separating the N- and C-terminal domains of VP40 (4). Thus, the 96LPLGVA101 motif of eVP40 is predicted to be important for the overall stability and function of VP40 during egress. Findings presented here indicate that disruption of these filovirus VP40 motifs results in a severe defect in VLP budding, due in part to impairment in overall VP40 structure, stability and/or intracellular localization.  相似文献   

13.
14.
15.
The process by which nonenveloped viruses cross cell membranes during host cell entry remains poorly defined; however, common themes are emerging. Here, we use correlated in vivo and in vitro studies to understand the mechanism of Flock House virus (FHV) entry and membrane penetration. We demonstrate that low endocytic pH is required for FHV infection, that exposure to acidic pH promotes FHV-mediated disruption of model membranes (liposomes), and particles exposed to low pH in vitro exhibit increased hydrophobicity. In addition, FHV particles perturbed by heating displayed a marked increase in liposome disruption, indicating that membrane-active regions of the capsid are exposed or released under these conditions. We also provide evidence that autoproteolytic cleavage, to generate the lipophilic γ peptide (4.4 kDa), is required for membrane penetration. Mutant, cleavage-defective particles failed to mediate liposome lysis, regardless of pH or heat treatment, suggesting that these particles are not able to expose or release the requisite membrane-active regions of the capsid, namely, the γ peptides. Based on these results, we propose an updated model for FHV entry in which (i) the virus enters the host cell by endocytosis, (ii) low pH within the endocytic pathway triggers the irreversible exposure or release of γ peptides from the virus particle, and (iii) the exposed/released γ peptides disrupt the endosomal membrane, facilitating translocation of viral RNA into the cytoplasm.Flock House virus (FHV), a nonenveloped, positive-sense RNA virus, has been employed as a model system in several important studies to address a wide range of biological questions (reviewed in reference 55). FHV has been instrumental in understanding virus structure and assembly (17, 19, 45), RNA replication (2, 3, 37), and specific packaging of the genome (33, 44, 53, 54). Studies of FHV infection in Drosophila melanogaster flies have provided valuable information about the antiviral innate immune response in invertebrate hosts (29, 57). FHV is also used in nanotechnology applications as an epitope-presenting platform to develop novel vaccines and medical therapies (31, 48). In this report, we use FHV as a model system to further elucidate the means by which nonenveloped viruses enter host cells and traverse cellular membranes.During cell entry enveloped and nonenveloped viral capsid proteins undergo structural rearrangements that enable the virus to breach the membrane bilayer, ultimately releasing the viral genome or nucleocapsid into the cytoplasm. These entry-related conformational changes have been well characterized for enveloped viruses, which use membrane fusion to cross membrane bilayers (reviewed in reference 59). However, the mechanisms nonenveloped viruses employ to breach cellular membranes are poorly defined. Recently, significant parallels in the mechanisms of cell entry have emerged for a diverse group of nonenveloped viruses. Specifically, programmed capsid disassembly and release of small membrane-interacting peptides appear to be a common theme (reviewed in references 4 and 50).The site of membrane penetration depends upon the route of virus entry into the cell. Viruses can enter host cells via several distinct pathways, including clathrin-mediated endocytosis, caveolae-mediated endocytosis, lipid raft-mediated endocytosis, and macropinocytosis (reviewed in reference 40). The two primary routes of virus entry are clathrin-mediated endocytosis, where viruses encounter an acidic environment, and caveolae-mediated endocytosis, which is pH neutral. Many nonenveloped viruses, including adenovirus (24, 52), parvovirus (6), and reovirus (34, 49), require acidic pH during entry. However, numerous nonenveloped viruses have acid-independent entry mechanisms, including rotavirus (28), polyomavirus (43), simian virus 40 (41, 51), and several members of the picornavirus family (7, 14, 32, 42).Upon reaching the appropriate site of membrane penetration, nonenveloped virus capsid proteins are triggered by cellular factors, such as receptor binding and/or exposure to low pH within endosomes, to undergo conformational changes necessary for membrane interactions. These tightly regulated structural rearrangements may include capsid disassembly, exposure of hydrophobic regions, and/or release of membrane-lytic factors. For example, low pH within endosomes triggers adenovirus capsid disassembly, leading to the release of the membrane lytic protein VI (24, 60). In contrast, poliovirus is activated for membrane penetration by a pH-independent mechanism. Receptor binding triggers the poliovirus capsid to undergo a conformational change, resulting in the exposure of the N terminus of VP1 and the release of VP4 (18, 23), both of which facilitate membrane interactions (20). Notably, even though some viruses, such as reovirus, enter cells via an acidic endocytic pathway, membrane penetration is not acid activated (16), indicating that exposure to low pH and membrane penetration are not always mutual events.The overall simplicity of the FHV capsid, composed of a single gene product, along with the wealth of available high-resolution structural information (reviewed in reference 45) make FHV an ideal candidate for understanding nonenveloped virus entry and infection. FHV, a member of the family Nodaviridae, is a nonenveloped insect virus with a bipartite RNA genome surrounded by an icosahedral protein capsid. The quasi-equivalent T=3 virion (∼300-Å diameter) is initially composed of 180 copies of a single coat precursor protein α (44 kDa). Following capsid assembly the α protein undergoes autocatalytic cleavage to generate two particle-associated cleavage products, a large N-terminal fragment, β (39 kDa), and a small C-terminal fragment, γ (4.4 kDa) (22), creating the infectious virion (46). Mutant FHV particles that do not undergo autocatalytic cleavage, and therefore cannot release the γ peptide, are not infectious (46). It has been hypothesized that these particles are noninfectious because they cannot mediate membrane penetration, but this has never been shown directly.The FHV X-ray structure revealed that the γ peptides were located inside the capsid shell with residues 364 to 385 forming amphipathic helices (19). Subsequent studies showed that the FHV capsid is dynamic, with γ transiently exposed to the exterior of the capsid (11). These findings led to a structure-based model of FHV membrane disruption in which the dynamic γ peptides are reversibly exposed to the surface of the capsid (11), “sampling” the environment until they encounter the appropriate cellular trigger. The virus is then activated to undergo an irreversible conformational change in which the γ helical bundles located at each fivefold axis are externalized and released from the virus particle (17, 19). Upon release, the γ pentameric helical bundles are predicted to insert into and create a local disruption of the membrane bilayer to allow the RNA to enter the cytoplasm (10).While biochemical and structural studies have provided the foundation for a model of FHV cell entry, more rigorous in vivo and in vitro studies are necessary to confirm the ideas put forth in this model. Here, we clarify the route of FHV entry and characterize the tightly regulated events required for FHV membrane penetration. We demonstrate for the first time that low endocytic pH is required for FHV infection, that acidic pH promotes FHV membrane penetration, and that particles exposed to low pH exhibit increased hydrophobicity. In addition, we provide evidence that mutant, cleavage-defective particles are blocked specifically at the membrane penetration step during cell entry. Taken together, these findings offer an experimentally supported model of FHV entry into host cells. In addition, these results add to the accumulating evidence that nonenveloped viruses employ common mechanisms to traverse cellular membranes.  相似文献   

16.
The human immunodeficiency virus type 1 structural polyprotein Pr55Gag is necessary and sufficient for the assembly of virus-like particles on cellular membranes. Previous studies demonstrated the importance of the capsid C-terminal domain (CA-CTD), nucleocapsid (NC), and membrane association in Gag-Gag interactions, but the relationships between these factors remain unclear. In this study, we systematically altered the CA-CTD, NC, and the ability to bind membrane to determine the relative contributions of, and interplay between, these factors. To directly measure Gag-Gag interactions, we utilized chimeric Gag-fluorescent protein fusion constructs and a fluorescence resonance energy transfer (FRET) stoichiometry method. We found that the CA-CTD is essential for Gag-Gag interactions at the plasma membrane, as the disruption of the CA-CTD has severe impacts on FRET. Data from experiments in which wild-type (WT) and CA-CTD mutant Gag molecules are coexpressed support the idea that the CA-CTD dimerization interface consists of two reciprocal interactions. Mutations in NC have less-severe impacts on FRET between normally myristoylated Gag proteins than do CA-CTD mutations. Notably, when nonmyristoylated Gag interacts with WT Gag, NC is essential for FRET despite the presence of the CA-CTD. In contrast, constitutively enhanced membrane binding eliminates the need for NC to produce a WT level of FRET. These results from cell-based experiments suggest a model in which both membrane binding and NC-RNA interactions serve similar scaffolding functions so that one can functionally compensate for a defect in the other.The human immunodeficiency virus type 1 (HIV-1) structural precursor polyprotein Pr55Gag is necessary and sufficient for the assembly of virus-like particles (VLPs). Gag is composed of four major structural domains, matrix (MA), capsid (CA), nucleocapsid (NC), and p6, as well as two spacer peptides, SP1 and SP2 (3, 30, 94). Following particle assembly and release, cleavage by HIV-1 protease separates these domains. However, these domains must work together in the context of the full-length Gag polyprotein to drive particle assembly.Previous studies have mapped two major functional domains involved in the early steps of assembly: first, Gag associates with cellular membranes via basic residues and N-terminal myristoylation of the MA domain (10, 17, 20, 35, 39, 87, 91, 106); second, the Gag-Gag interaction domains that span the CA C-terminal domain (CA-CTD) and NC domain promote Gag multimerization (3, 11, 14, 16, 18, 23, 27, 29, 30, 33, 36, 46, 64, 88, 94, 102, 103). Structural and genetic studies have identified two residues (W184 and M185) within a dimerization interface in the CA-CTD that are critical to CA-CA interactions (33, 51, 74, 96). Analytical ultracentrifugation of heterodimers formed between wild-type (WT) Gag and Gag mutants with changes at these residues suggests that the dimerization interface consists of two reciprocal interactions, one of which can be disrupted to form a “half-interface” (22).In addition to the CA-CTD, NC contributes to assembly via 15 basic residues (8, 9, 11, 14, 18, 23, 25, 28, 34, 40, 43, 54, 57, 58, 74, 79, 88, 97, 104, 105), although some researchers have suggested that NC instead contributes to the stability of mature virions after assembly (75, 98, 99). It is thought that the contribution of NC to assembly is due to its ability to bind RNA, since the addition of RNA promotes the formation of particles in vitro (14-16, 37, 46), and RNase treatment disrupts Gag-Gag interactions (11) and immature viral cores (67). However, RNA is not necessary per se, since dimerization motifs can substitute for NC (1, 4, 19, 49, 105). This suggests a model in which RNA serves a structural role, such as a scaffold, to promote Gag-Gag interactions through NC. Based on in vitro studies, it has been suggested that this RNA scaffolding interaction facilitates the low-order Gag multimerization mediated by CA-CTD dimerization (4, 37, 49, 62, 63, 85). Despite a wealth of biochemical data, the relative contributions of the CA-CTD and NC to Gag multimerization leading to assembly are yet to be determined in cells.Mutations in Gag interaction domains alter membrane binding in addition to affecting Gag multimerization. In particular, mutations or truncations of CA reduce membrane binding (21, 74, 82), and others previously reported that mutations or truncations of NC affect membrane binding (13, 78, 89, 107). These findings are consistent with a myristoyl switch model of membrane binding in which Gag can switch between high- and low-membrane-affinity states (38, 71, 76, 83, 86, 87, 92, 95, 107). Many have proposed, and some have provided direct evidence (95), that Gag multimerization mediated by CA or NC interactions promotes the exposure of the myristoyl moiety to facilitate membrane associations.Gag membrane binding and multimerization appear to be interrelated steps of virus assembly, since membrane binding also facilitates Gag multimerization. Unlike betaretroviruses that fully assemble prior to membrane targeting and envelopment (type B/D), lentiviruses, such as HIV, assemble only on cellular membranes at normal Gag expression levels (type C), although non-membrane-bound Gag complexes exist (45, 58, 60, 61, 65). Consistent with this finding, mutations that reduce Gag membrane associations cause a defect in Gag multimerization (59, 74). Therefore, in addition to their primary effects on Gag-Gag interactions, mutations in Gag interaction domains cause a defect in membrane binding, which, in turn, causes a secondary multimerization defect. To determine the relative contributions of the CA-CTD and the NC domain to Gag-Gag interactions at the plasma membrane, it is essential to eliminate secondary effects due to a modulation of membrane binding.Except for studies using a His-tag-mediated membrane binding system (5, 46), biochemical studies of C-type Gag multimerization typically lack membranes. Therefore, these studies do not fully represent particle assembly, which occurs on biological membranes in cells. Furthermore, many biochemical and structural approaches are limited to isolated domains or truncated Gag constructs. Thus, some of these studies are perhaps more relevant to the behavior of protease-cleaved Gag in mature virions. With few exceptions (47, 74), cell-based studies of Gag multimerization have typically been limited to measuring how well mutant Gag is incorporated into VLPs when coexpressed or not with WT Gag. Since VLP production is a complex multistep process, effects of mutations on other steps in the process can confound this indirect measure. For example, NC contributes to VLP production by both promoting multimerization and interacting with the host factor ALIX to promote VLP release (26, 80). To directly assay Gag multimerization in cells, several groups (24, 45, 52, 56) developed microscopy assays based on fluorescence resonance energy transfer (FRET). These assays measure the transfer of energy between donor and acceptor fluorescent molecules that are brought within ∼5 nm by the association of the proteins to which they are attached (41, 48, 90). However, these microscopy-based Gag FRET assays have not been used to fully elucidate several fundamental aspects of HIV-1 Gag multimerization at the plasma membrane of cells, such as the relative contributions of the CA-CTD and NC and the effect of membrane binding on Gag-Gag interactions. In this study, we used a FRET stoichiometry method based on calibrated spectral analysis of fluorescence microscopy images (41). This algorithm determines the fractions of both donor and acceptor fluorescent protein-tagged Gag molecules participating in FRET. For cells expressing Gag molecules tagged with donor (cyan fluorescent protein [CFP]) and acceptor (yellow fluorescent protein [YFP]) molecules, this method measures the apparent FRET efficiency, which is proportional to the mole fraction of Gag constructs in complex. By measuring apparent FRET efficiencies, quantitative estimates of the mole fractions of interacting proteins can be obtained.Using this FRET-based assay, we aim to answer two questions: (i) what are the relative contributions of CA-CTD and NC domains to Gag multimerization when secondary effects via membrane binding are held constant, and (ii) what is the effect of modulating membrane binding on the ability of Gag mutants to interact with WT Gag?Our data demonstrate that the CA-CTD dimerization interface is essential for Gag multimerization at the plasma membrane, as fully disrupting the CA-CTD interaction abolishes FRET, whereas a modest level of FRET is still detected in the absence of NC. We also present evidence that the CA-CTD dimerization interface consists of two reciprocal interactions, allowing the formation of a half-interface that can still contribute to Gag multimerization. Notably, when Gag derivatives with an intact CA-CTD were coexpressed with WT Gag, either membrane binding ability or NC was required for the Gag mutants to interact with WT Gag, suggesting functional compensation between these factors.  相似文献   

17.
Newcastle disease virus (NDV), an avian paramyxovirus, is tumor selective and intrinsically oncolytic because of its potent ability to induce apoptosis. Several studies have demonstrated that NDV is selectively cytotoxic to tumor cells but not normal cells due to defects in the interferon (IFN) antiviral responses of tumor cells. Many naturally occurring strains of NDV have an intact IFN-antagonistic function and can still replicate in normal human cells. To avoid potential toxicity issues with NDV, especially in cancer patients with immunosuppression, safe NDV-oncolytic vectors are needed. We compared the cell killing abilities of (i) a recombinant NDV (rNDV) strain, Beaudette C, containing an IFN-antagonistic, wild-type V protein (rBC), (ii) an isogenic recombinant virus with a mutant V protein (rBC-Edit virus) that induces increased IFN in infected cells and whose replication is restricted in normal human cells, and (iii) a recombinant LaSota virus with a virulent F protein cleavage site that is as interferon sensitive as rBC-Edit virus (LaSota V.F. virus). Our results indicated that the tumor-selective replication of rNDV is determined by the differential regulation of IFN-α and downstream antiviral genes induced by IFN-α, especially through the IRF-7 pathway. In a nude mouse model of human fibrosarcoma, we show that the IFN-sensitive NDV variants are as effective as IFN-resistant rBC virus in clearing the tumor burden. In addition, mice treated with rNDV exhibited no signs of toxicity to the viruses. These findings indicate that augmentation of innate immune responses by NDV results in selective oncolysis and offer a novel and safe virotherapy platform.Several naturally occurring or engineered oncolytic viruses are emerging as novel tools for selective growth in and killing of a variety of tumor cells (1, 21, 34, 41). It has been consistently reported that during tumor evolution, diminished interferon (IFN) responsiveness coevolves as a frequent genetic defect (4, 31, 32, 41). Any defects in responsiveness to interferon will afford permissiveness of tumors for replication of oncolytic viruses by blunting the antiviral innate immune system. Thus, it was suggested that oncolytic viruses could be engineered to induce strong IFN response and/or to be defective in antagonizing the IFN signaling. This would result in virus replication in tumor cells with IFN defects but in reduced or crippled virus replication in normal cells, with the absence of toxicity (42). A variety of oncolytic viruses have been engineered to exploit tumor-specific genetic defects (3, 12, 24, 42, 46) and shown to be potent oncolytic agents.Newcastle disease virus (NDV), an avian paramyxovirus, is a promising broad-spectrum oncolytic agent (27, 29, 30, 37). Nonengineered, naturally occurring strains of NDV such as 73-T (6), MTH68 (7), PV701 (28, 35), and NDV-HUJ (11) have been successfully employed in several clinical studies for tumor regression. NDV is inherently oncolytic and tumor selective, sparing normal cells (9, 15, 37). The tumor selectivity of NDV is considered to be due to a defective IFN response in tumor cells (10, 23, 37). NDV is a strong inducer of type I IFN in many types of cells (18). In normal cells, a robust IFN-mediated antiviral response limits the replication of NDV (9, 23). This known sensitivity of NDV to cellular antiviral mechanisms affords a wide safety margin for its use in humans.Recent studies have indicated that improved therapeutic vectors of NDV could be engineered through reverse genetics for enhanced oncolytic efficacy from an increased anti-tumor response and interleukin 2 (IL-2) receptor-mediated targeting (5, 9, 44, 46). Therefore, we reasoned that recombinant NDVs (rNDVs) that are susceptible to cellular innate immune responses would be safer and more effective oncolytic agents. Even though NDV is an avian virus and induces a strong IFN response in normal human cells, it still expresses IFN-antagonizing activity. Ablation of the expression of V protein, which is responsible for this anti-IFN activity, may further reduce the ability of NDV to infect and kill normal human cells without affecting tumor cell infection and lysis. Here, we describe the relative oncolytic efficacies of three rNDV strains differing in IFN antagonism. The rNDV variants with an IFN-sensitive phenotype had parallel therapeutic efficacies in xenotransplanted human fibrosarcoma cells in a nude mouse model and offer great potential as recombinant vectors in therapy of human malignancies.  相似文献   

18.
19.
20.
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   

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