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 812 毫秒
1.
The first morphological evidence of African swine fever virus (ASFV) assembly is the appearance of precursor viral membranes, thought to derive from the endoplasmic reticulum, within the assembly sites. We have shown previously that protein p54, a viral structural integral membrane protein, is essential for the generation of the viral precursor membranes. In this report, we study the role of protein p17, an abundant transmembrane protein localized at the viral internal envelope, in these processes. Using an inducible virus for this protein, we show that p17 is essential for virus viability and that its repression blocks the proteolytic processing of polyproteins pp220 and pp62. Electron microscopy analyses demonstrate that when the infection occurs under restrictive conditions, viral morphogenesis is blocked at an early stage, immediately posterior to the formation of the viral precursor membranes, indicating that protein p17 is required to allow their progression toward icosahedral particles. Thus, the absence of this protein leads to an accumulation of these precursors and to the delocalization of the major components of the capsid and core shell domains. The study of ultrathin serial sections from cells infected with BA71V or the inducible virus under permissive conditions revealed the presence of large helicoidal structures from which immature particles are produced, suggesting that these helicoidal structures represent a previously undetected viral intermediate.African swine fever virus (ASFV) (61, 72) is the only known DNA-containing arbovirus and the sole member of the Asfarviridae family (24). Infection by this virus of its natural hosts, the wild swine warthogs and bushpigs and the argasid ticks of the genus Ornithodoros, results in a mild disease, often asymptomatic, with low viremia titers, that in many cases develops into a persistent infection (3, 43, 71). In contrast, infection of domestic pigs leads to a lethal hemorrhagic fever for which the only available methods of disease control are the quarantine of the affected area and the elimination of the infected animals (51).The ASFV genome is a lineal molecule of double-stranded DNA of 170 to 190 kbp in length with convalently closed ends and terminal inverted repeats. The genome encodes more than 150 open reading frames, half of which lack any known or predictable function (16, 75).The virus particle, with an overall icosahedral shape and an average diameter of 200 nm (11), is organized in several concentric layers (6, 11, 15) containing more than 50 structural proteins (29). Intracellular particles are formed by an inner viral core, which contains the central nucleoid surrounded by a thick protein coat, referred to as core shell. This core is enwrapped by an inner lipid envelope (7, 34) on top of which the icosahedral capsid is assembled (26, 27, 31). Extracellular virions possess an additional membrane acquired during the budding from the plasma membrane (11). Both forms of the virus, intracellular and extracellular, are infective (8).The assembly of ASFV particles occurs in the cytoplasm of the infected cell, in viral factories located close to the cell nucleus (6, 13, 49). ASFV factories possess several characteristics similar to those of the cellular aggresomes (35), which are accumulations of aggregates of cellular proteins that form perinuclear inclusions (44).Current models propose that ASFV assembly begins with the modification of endoplasmic reticulum (ER) membranes, which are subsequently recruited to the viral factories and transformed into viral precursor membranes. These ER-derived viral membranes represent the precursors of the inner viral envelope and are the first morphological evidence of viral assembly (7, 60). ASFV viral membrane precursors evolve into icosahedral intermediates and icosahedral particles by the progressive assembly of the outer capsid layer at the convex face of the precursor membranes (5, 26, 27, 31) through an ATP- and calcium-dependent process (19). At the same time, the core shell is formed underneath the concave face of the viral envelope, and the viral DNA and nucleoproteins are packaged and condensed to form the innermost electron-dense nucleoid (6, 9, 12, 69). However, the assembly of the capsid and the internal envelope appears to be largely independent of the components of the core of the particle, since the absence of the viral polyprotein pp220 during assembly produces empty virus-like particles that do not contain the core (9).Comparative genome analysis suggests that ASFV shares a common origin with the members of the proposed nucleocytoplasmic large DNA viruses (NCLDVs) (40, 41). The reconstructed phylogeny of NCLDVs as well as the similitude in the structures and organizations of the genomes indicates that ASFV is more closely related to poxviruses than to other members of the NCLDVs. A consensus about the origin and nature of the envelope of the immature form of vaccinia virus (VV), the prototypical poxvirus, seems to be emerging (10, 17, 20, 54). VV assembly starts with the appearance of crescent-shaped structures within specialized regions of the cytoplasm also known as viral factories (21, 23). The crescent membranes originate from preexisting membranes derived from some specialized compartment of the ER (32, 37, 52, 53, 67), and an operative pathway from the ER to the crescent membrane has recently been described (38, 39). VV crescents apparently grow in length while maintaining the same curvature until they become closed circles, spheres in three dimensions, called immature virions (IV) (22). The uniform curvature is produced by a honeycomb lattice of protein D13L (36, 70), which attaches rapidly to the membranes so that nascent viral membranes always appear to be coated over their entirety. The D13L protein is evolutionarily related to the capsid proteins of the other members of the NCLDV group, including ASFV, but lacks the C-terminal jelly roll motif (40). This structural difference is probably related to the fact that poxviruses are the only member of this group without an icosahedral capsid; instead, the spherical D13L coat acts as a scaffold during the IV stage but is discarded in subsequent steps of morphogenesis (10, 28, 46, 66). Thus, although crescents in VV and precursors of the inner envelope in ASFV are the first morphogenetic stages discernible in the viral factories of these viruses, they seem to be different in nature. Crescents are covered by the D13L protein and are more akin to the icosahedral intermediates of ASFV assembly, whereas ASFV viral membrane precursors are more similar to the naked membranes seen when VV morphogenesis is arrested by rifampin treatment (33, 47, 48, 50) or when the expression of the D13L and A17L proteins are repressed during infection with lethal conditional VV viruses (45, 55, 56, 68, 74, 76).Although available evidence strongly supports the reticular origin of the ASFV inner envelope (7, 60), the mechanism of acquisition remains unknown, and the number of membranes present in the inner envelope is controversial. The traditional view of the inner envelope as formed by two tightly opposed membranes derived from ER collapsed cisternae (7, 59, 60) has recently been challenged by the careful examination of the width of the internal membrane of viral particles and the single outer mitochondrial membrane, carried out using chemical fixation, cryosectioning, and high-pressure freezing (34). The results suggest that the inner envelope of ASFV is a single lipid bilayer, which raises the question of how such a structure can be generated and stabilized in the precursors of the ASFV internal envelope. In the case of VV, the coat of the D13L protein has been suggested to play a key role in the stabilization of the single membrane structure of the crescent (10, 17, 36), but the ASFV capsid protein p72 is not a component of the viral membrane precursors. The identification and functional characterization of the proteins involved in the generation of these structures are essential for the understanding of the mechanisms involved in these early stages of viral assembly. For this reason, we are focusing our interest on the study of abundant structural membrane proteins that reside at the inner envelope of the viral particle. We have shown previously that one of these proteins, p54, is essential for the recruitment of ER membranes to the viral factory (59). Repression of protein p54 expression has a profound impact on virus production and leads to an early arrest in virion morphogenesis, resulting in the virtual absence of membranes in the viral factory.Protein p17, encoded by the late gene D117L in the BA71V strain, is an abundant structural protein (60, 65). Its sequence, which is highly conserved among ASFV isolates (16), does not show any significant similarity with the sequences present in the databases. Protein p17 is an integral membrane protein (18) that is predicted to insert in membranes with a Singer type I topology and has been localized in the envelope precursors as well as in both intracellular and extracellular mature particles (60), suggesting that it resides at the internal envelope, the only membranous structure of the intracellular particles.In this work, we analyze the role of protein p17 in viral assembly by means of an IPTG (isopropyl-β-d-thiogalactopyranoside)-dependent lethal conditional virus. The data presented indicate that protein p17 is essential for viral morphogenesis. The repression of this protein appears to block assembly at the level of viral precursor membranes, resulting in their accumulation at the viral factory.From the electron microscopy analysis of serial sections of viral factories at very early times during morphogenesis, we present experimental evidence that suggests that, during assembly, viral precursor membranes and core material organize into large helicoidal intermediates from which icosahedral particles emerge. The possible role of these structures during ASFV morphogenesis is discussed.  相似文献   

2.
3.
4.
5.
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   

6.
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   

7.
The African swine fever virus (ASFV) protein pE248R, encoded by the gene E248R, is a late structural component of the virus particle. The protein contains intramolecular disulfide bonds and has been previously identified as a substrate of the ASFV-encoded redox system. Its amino acid sequence contains a putative myristoylation site and a hydrophobic transmembrane region near its carboxy terminus. We show here that the protein pE248R is myristoylated during infection and associates with the membrane fraction in infected cells, behaving as an integral membrane protein. Furthermore, the protein localizes at the inner envelope of the virus particles in the cytoplasmic factories. The function of the protein pE248R in ASFV replication was investigated by using a recombinant virus that inducibly expresses the gene E248R. Under repressive conditions, the ASFV polyproteins pp220 and pp62 are normally processed and virus particles with morphology indistinguishable from that of those produced in a wild-type infection or under permissive conditions are generated. Moreover, the mutant virus particles can exit the cell as does the parental virus. However, the infectivity of the pE248R-deficient virions was reduced at least 100-fold. An investigation of the defect of the mutant virus indicated that neither virus binding nor internalization was affected by the absence of the protein pE248R, but a cytopathic effect was not induced and early and late gene expression was impaired, indicating that the protein is required for some early postentry event.African swine fever virus (ASFV) is a large enveloped deoxyvirus that causes a severe hemorrhagic disease in domestic pigs (38). The ASFV genome is a double-stranded DNA molecule of 170 to 190 kbp that encodes more than 150 polypeptides (47). The icosahedral virus particle contains more than 50 polypeptides and is composed of several concentric domains, including an internal DNA-containing nucleoid surrounded by a protein layer designated the core shell, an inner envelope, and an outer icosahedral capsid (8, 10, 20). An additional membrane acquired by budding through the plasma membrane envelops the extracellular virion (14).The complex process of virus assembly occurs at specialized cytoplasmic sites, designated viral factories, and is initiated by the recruitment and modification of endoplasmic reticulum (ER) cisternae, which collapse to form the virus inner envelope, where the viral membrane proteins p54 and p17 are localized (8, 16, 21, 32, 37). This model, however, has been recently questioned, and based on data obtained using samples prepared by high-pressure freezing, it has been suggested that the inner envelope of ASFV consists of a single lipid bilayer (28). The icosahedral capsid layer, formed by protein p72, is then progressively assembled on one side of this envelope, while on the other side, the core shell domain, mainly constituted by the processing products of the polyproteins pp220 and pp62, is simultaneously constructed (6, 7, 20, 26). Finally, the viral DNA and nucleoproteins are packaged and condensed to form the nucleoid (15).The functions of several virus proteins in the formation of the different domains of the virus particle have been investigated in recent years. Thus, the structural proteins p72 and pB438L and the nonstructural pB602L protein, described as a chaperone of p72 (22), have been shown to be required for the construction of the icosahedral capsid (24, 25, 26), while the polyprotein pp220 is essential for the formation of the inner core, constituted by the nucleoid and core shell domains (7). It has also been demonstrated that the processing of the polyproteins pp220 and pp62 by the virus-encoded protease is necessary for the assembly of a proper core (5). In addition, it is known that the transmembrane protein p54 is critical for the recruitment of envelope precursors to assembly sites (35), although the mechanisms underlying the conversion of ER cisternae into functional viral envelopes are mostly unknown. Studies of other transmembrane proteins detected as structural components of the virus particle could shed light on this matter. Some of the virion membrane proteins could also play a role in virus entry, as has been described for the proteins p12, identified as a viral attachment protein (11, 19), and p54, also involved in binding of virus to target cells (27).The ASFV protein pE248R is a late structural component of the virus particle (33) that belongs to a class of myristoylated membrane proteins related to vaccinia virus L1 (30), one of the substrates of the pathway for the formation of disulfide bonds encoded by this virus (41). The protein pE248R also contains intramolecular disulfide bridges and has been recently identified as a possible final substrate of the ASFV-encoded redox system (33). In the present study, we investigated the membrane association, the localization in the virion, and the role of the protein pE248R in ASFV replication. Our results indicate that pE248R is a myristoylated integral membrane protein localized at the inner envelope of the virus particle. By using a conditional lethal ASFV mutant, vE248Ri, with an inducible copy of the gene E248R, we showed that the protein pE248R is required for virus infectivity and an early postentry event but not virus assembly.  相似文献   

8.
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   

9.
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   

10.
Human papillomavirus type 16 (HPV16) has been identified as being the most common etiological agent leading to cervical cancer. Despite having a clear understanding of the role of HPV16 in oncogenesis, details of how HPV16 traffics during infection are poorly understood. HPV16 has been determined to enter via clathrin-mediated endocytosis, but the subsequent steps of HPV16 infection remain unclear. There is emerging evidence that several viruses take advantage of cross talk between routes of endocytosis. Specifically, JCV and bovine papillomavirus type 1 have been shown to enter cells by clathrin-dependent endocytosis and then require caveolin-1-mediated trafficking for infection. In this paper, we show that HPV16 is dependent on caveolin-1 after clathrin-mediated endocytosis. We provide evidence for the first time that HPV16 infection is dependent on trafficking to the endoplasmic reticulum (ER). This novel trafficking may explain the requirement for the caveolar pathway in HPV16 infection because clathrin-mediated endocytosis typically does not lead to the ER. Our data indicate that the infectious route for HPV16 following clathrin-mediated entry is caveolin-1 and COPI dependent. An understanding of the steps involved in HPV16 sorting and trafficking opens up the possibility of developing novel approaches to interfere with HPV16 infection and reduce the burden of papillomavirus diseases including cervical cancer.Human papillomavirus (PV) type 16 (HPV16) is a member of the family Papillomaviridae, a group of double-stranded DNA (dsDNA) viruses with a tropism for squamous epithelia (70). Most PV infections result in benign lesions, although a subset of high-risk HPVs are capable of malignant transformation, resulting in various cancers including cervical carcinoma (21, 38). Infection with HPV16 is responsible for causing approximately half of the cases of invasive cervical cancer (7). In spite of the link between HPV16 and cervical cancer, the intracellular movement of HPV16 through target keratinocyte cells during infection has not been defined in detail.Viruses can enter into target cells by taking advantage of the cell''s natural endocytosis machinery (60). One of the best-characterized modes of internalization is by receptor-mediated, clathrin-dependent endocytosis. In this mode of entry, clathrin-coated pits internalize cargo into clathrin-coated vesicles, which are pinched from the plasma membrane by dynamin-2 in order to internalize (68). The process of clathrin-mediated endocytosis occurs rapidly, resulting in the delivery of cargo to early/sorting endosomes within seconds to minutes (23, 31). From the sorting endosome, most clathrin-dependent ligands are trafficked back to the plasma membrane in recycling endosomes or to lysosomes for degradation (35, 56). Another well-studied model of ligand entry is caveolin-1-mediated endocytosis. The caveolar pathway typically involves entry via cholesterol-rich caveolae at the plasma membrane, which deliver their contents to pH-neutral organelles known as caveosomes (44, 65). The delivery of cargo from caveosomes to the Golgi apparatus and the endoplasmic reticulum (ER) was demonstrated previously (44, 46, 50). The traffickings of cargo internalized via clathrin- and caveolin-1-mediated endocytosis were once thought to be separate; however, it is becoming evident that viruses including bovine PV type 1 (BPV1), JCV, HPV31, and BKV rely on both pathways depending on the stage of infection (29, 32, 50, 63).PV internalization is preceded by virion attachment to the extracellular matrix, followed by binding to heparan sulfate (14, 15, 25). The involvement of a secondary receptor has been suggested, putatively an alpha-6 integrin (24, 37). Postbinding, a conformational change in the PV capsid results in a furin cleavage event at the N terminus of the minor capsid protein L2, which has been suggested to play a role in the endosomal escape of the viral genome (19, 30, 52). An increasing body of evidence supports the entry of HPV16 by clathrin-mediated endocytosis (9, 27, 62). Electron microscopy of HPV16 infection in COS-7 cells demonstrated HPV16 pseudovirions in clathrin-coated vesicles 20 min after entry and within structures resembling endosomes by 1 h postentry (9). HPV16 infection of HaCaT keratinocyte, COS-7, and 293TT cells has been blocked by chlorpromazine, an inhibitor of the formation of clathrin-coated pits (9, 27, 62, 67). Importantly, those studies showed that two inhibitors of caveolin-1-mediated internalization, filipin and nystatin, did not interfere with HPV16 infection (9, 27, 62). Our laboratory demonstrated the importance of dynamin in HPV16 infection, presumably in the scission of clathrin-coated vesicles from the plasma membrane (1). Recently, a clathrin-, caveolin-, and dynamin-independent endocytosis of HPV16 was suggested, although the use of the HPV18-positive, heteroploid HeLa cell line calls into question the relevance of this finding to natural infection (64).In a previous study, we described the postentry trafficking of BPV1 from endosomes to caveolin-1-positive vesicles, similarly to a related nonenveloped dsDNA virus, JCV (32, 50). Our data demonstrated that the infectious route of BPV1 involved entry by clathrin-mediated endocytosis followed by transport to the caveolar pathway in order to traffic to the ER (32). We found that BPV1 infection was neutralized by an antibody that prevented viral particle transport to the ER (33). The movement of BPV1 from the endosome to the caveosome provides a possible explanation for why BPV1 trafficking is so slow compared to those of other ligands of clathrin-mediated endocytosis (20, 26). The kinetics of BPV1 and HPV16 entry were previously reported to be identical, and the coincident internalization of HPV16 and BPV1 virus-like particles (VLPs) showed colocalization between the VLPs during infection (20, 62). These data suggest that HPV16 and BPV1 infection may be occurring by a similar mechanism.Our goal in the present study was to determine the intracellular trafficking events leading to HPV16 infection. The use of reporter virion technology has allowed the production of high-titer HPV16 virions by a method previously shown to yield virions that are infectious in vivo (16). In this study, we used HPV16 reporter virions to study HPV16 infection in the spontaneously immortalized human HaCaT keratinocyte cell line. Our data show that the infectious route of HPV16 is from early endosomes to caveolin-1-positive vesicles and then to the ER. Using immunofluorescence and short hairpin RNA (shRNA) against caveolin-1, we demonstrate the importance of the caveolar pathway after HPV16 has been internalized. We show that HPV16 infection was blocked by inhibiting the formation of COPI transport vesicles, which function in trafficking between the ER and the Golgi apparatus and from caveosomes to the ER (5, 39). We provide evidence that after reaching the caveosome, HPV16 requires passage to the ER for successful infection, a trafficking event made possible by COPI vesicle-mediated movement from the caveosome to the ER.  相似文献   

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

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

14.
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   

15.
16.
17.
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   

18.
19.
Rhesus rhadinovirus (RRV) is a gammaherpesvirus closely related to Kaposi''s sarcoma-associated herpesvirus (KSHV), an oncogenic virus linked to the development of Kaposi''s sarcoma and several other lymphoproliferative diseases, including primary effusion lymphoma and multicentric Castleman''s disease. RRV naturally infects rhesus macaques and induces lymphoproliferative diseases under experimental conditions, making it an excellent model for the study of KSHV. Unlike KSHV, which grows poorly in cell culture, RRV replicates efficiently in rhesus fibroblasts (RFs). In this study, we have characterized the entry pathway of RRV in RFs. Using a luciferase-expressing recombinant RRV (RRV-luciferase), we show that the infectivity of RRV is reduced by inhibitors of endosomal acidification. RRV infectivity is also reduced by inhibitors of clathrin-mediated but not caveola-mediated endocytosis, indicating that RRV enters into RFs via clathrin-mediated endocytosis. Using a red fluorescent protein (RFP)-expressing recombinant RRV (RRV-RFP), we show that RRV particles are colocalized with markers of endocytosis (early endosome antigen 1) and clathrin-mediated endocytosis (clathrin heavy chain) during entry into RFs. RRV particles are also colocalized with transferrin, which enters cells by clathrin-mediated endocytosis, but not with cholera toxin B, which enters cells by caveola-mediated endocytosis. Inhibition of clathrin-mediated endocytosis with a dominant-negative construct of EPS15, an essential component of clathrin-coated pits, blocked the entry of RRV into RFs. Together, these results indicate that RRV entry into RFs is mediated by clathrin-mediated endocytosis.Kaposi''s sarcoma-associated herpesvirus (KSHV), also known as human herpesvirus 8 (HHV8), is a gammaherpesvirus associated with the development of Kaposi''s sarcoma, a malignancy commonly found in AIDS patients (13). KSHV is also associated with the development of multicentric Castleman''s disease (MCD) and primary effusion lymphoma (PEL), two rare lymphoproliferative diseases. KSHV has a restricted host range, making it difficult to study KSHV and its related malignances directly in an animal model (25). Rhesus rhadinovirus (RRV) is closely related to KSHV. RRV infects its natural host and induces lymphoproliferative diseases resembling MCD and PEL; thus, it has been proposed as an animal model for the study of KSHV (19, 26, 39). Two isolates of RRV (26-95 and 17577) have been independently isolated and sequenced so far (3, 7, 32).To establish a successful infection, a virus needs to enter the target cells and release its genome (20). Thus, defining the entry and trafficking pathway of RRV can help us understand its mechanism of infection and replication in vitro and in vivo. Herpesviruses bind to the cell surface through complex interactions between viral glycoproteins and receptor molecules, leading to either plasma membrane fusion or endocytosis (35). Plasma membrane fusion is a pH-independent event between the viral envelope and the host cell plasma membrane (23). Enveloped viruses also take advantage of cellular endocytosis pathways for their internalization (34). Endocytosis leads to fusion between the membrane of the internalized vesicle and the viral envelope at low pHs and to the release of the viral particle into the cytoplasm. Following membrane fusion, the nucleocapsid traffics to the perinuclear space and delivers the viral genome to the nucleus. Thus, endocytosis offers a convenient and fast transit system enabling the virus to enter and traffic across the plasma membrane and cytoplasm of the infected cell.In mammalian cells, there are several endocytic pathways, including clathrin-mediated endocytosis, caveola-mediated endocytosis, clathrin- and caveola-independent endocytosis, and macropinocytosis (34). These endocytic pathways differ in the nature and size of the cargo. The clathrin-mediated pathway is the most commonly observed uptake pathway for viruses (30). A viral particle is internalized into a clathrin-coated vesicle, which then loses the clathrin-coated subunits before fusing with the early endosome. An activation step occurs in the endosome, leading to the fusion of the viral envelope with the endosomal membrane and the delivery of the viral capsid to the cytosol. The acidic pH in the endosome is thought to play an essential role in triggering the fusion event. Therefore, pH sensitivity is often considered an indication that a virus enters the cell by endocytosis (30).KSHV has been shown to use clathrin-mediated endocytosis to enter human foreskin fibroblasts, activated primary human B cells, and primary human umbilical vein endothelial cells (1, 12, 29); however, the macropinocytic pathway and plasma membrane fusion pathway have also been implicated (17, 28). The mechanism of RRV entry into cells has not been defined. In this study, using two recombinant RRVs expressing luciferase (RRV-luciferase) and red fluorescent protein (RRV-RFP), respectively, we have characterized the entry pathway of RRV in rhesus fibroblasts (RFs), a cell type that RRV can infect efficiently and in which it can replicate. The results show that RRV entry into RFs occurs primarily via clathrin-mediated endocytosis.  相似文献   

20.
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   

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