首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 46 毫秒
1.
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   

2.
Only a few archaeal viruses have been subjected to detailed structural analyses. Major obstacles have been the extreme conditions such as high salinity or temperature needed for the propagation of these viruses. In addition, unusual morphotypes of many archaeal viruses have made it difficult to obtain further information on virion architectures. We used controlled virion dissociation to reveal the structural organization of Halorubrum pleomorphic virus 1 (HRPV-1) infecting an extremely halophilic archaeal host. The single-stranded DNA genome is enclosed in a pleomorphic membrane vesicle without detected nucleoproteins. VP4, the larger major structural protein of HRPV-1, forms glycosylated spikes on the virion surface and VP3, the smaller major structural protein, resides on the inner surface of the membrane vesicle. Together, these proteins organize the structure of the membrane vesicle. Quantitative lipid comparison of HRPV-1 and its host Halorubrum sp. revealed that HRPV-1 acquires lipids nonselectively from the host cell membrane, which is typical of pleomorphic enveloped viruses.In recent years there has been growing interest in viruses infecting hosts in the domain Archaea (43). Archaeal viruses were discovered 35 years ago (52), and today about 50 such viruses are known (43). They represent highly diverse virion morphotypes in contrast to the vast majority (96%) of head-tail virions among the over 5,000 described bacterial viruses (1). Although archaea are widespread in both moderate and extreme environments (13), viruses have been isolated only for halophiles and anaerobic methanogenes of the kingdom Euryarchaeota and hyperthermophiles of the kingdom Crenarchaeota (43).In addition to soil and marine environments, high viral abundance has also been detected in hypersaline habitats such as salterns (i.e., a multipond system where seawater is evaporated for the production of salt) (19, 37, 50). Archaea are dominant organisms at extreme salinities (36), and about 20 haloarchaeal viruses have been isolated to date (43). The majority of these are head-tail viruses, whereas electron microscopic (EM) studies of highly saline environments indicate that the two other described morphotypes, spindle-shaped and round particles, are the most abundant ones (19, 37, 43). Thus far, the morphological diversity of the isolated haloarchaeal viruses is restricted compared to viruses infecting hyperthermophilic archaea, which are classified into seven viral families (43).All of the previously described archaeal viruses have a double-stranded DNA (dsDNA) genome (44). However, a newly characterized haloarchaeal virus, Halorubrum pleomorphic virus 1 (HRPV-1), has a single-stranded DNA (ssDNA) genome (39). HRPV-1 and its host Halorubrum sp. were isolated from an Italian (Trapani, Sicily) solar saltern. Most of the studied haloarchaeal viruses lyse their host cells, but persistent infections are also typical (40, 44). HRPV-1 is a nonlytic virus that persists in the host cells. In liquid propagation, nonsynchronous infection cycles of HRPV-1 lead to continuous virus production until the growth of the host ceases, resulting in high virus titers in the growth medium (39).The pleomorphic virion of HRPV-1 represents a novel archaeal virus morphotype constituted of lipids and two major structural proteins VP3 (11 kDa) and VP4 (65 kDa). The genome of HRPV-1 is a circular ssDNA molecule (7,048 nucleotides [nt]) containing nine putative open reading frames (ORFs). Three of them are confirmed to encode structural proteins VP3, VP4, and VP8, which is a putative ATPase (39). The ORFs of the HRPV-1 genome show significant similarity, at the amino acid level, to the minimal replicon of plasmid pHK2 of Haloferax sp. (20, 39). Furthermore, an ∼4-kb region, encoding VP4- and VP8-like proteins, is found in the genomes of two haloarchaea, Haloarcula marismortui and Natronomonas pharaonis, and in the linear dsDNA genome (16 kb) of spindle-shaped haloarchaeal virus His2 (39). The possible relationship between ssDNA virus HRPV-1 and dsDNA virus His2 challenges the classification of viruses, which is based on the genome type among other criteria (15, 39).HRPV-1 is proposed to represent a new lineage of pleomorphic enveloped viruses (39). A putative representative of this lineage among bacterial viruses might be L172 of Acholeplasma laidlawii (14). The enveloped virion of L172 is pleomorphic, and the virus has a circular ssDNA genome (14 kb). In addition, the structural protein pattern of L172 with two major structural proteins, of 15 and 53 kDa, resembles that of HRPV-1.The structural approach has made it possible to reveal relationships between viruses where no sequence similarity can be detected. It has been realized that several icosahedral viruses infecting hosts in different domains of life share common virion architectures and folds of their major capsid proteins. These findings have consequences for the concept of the origin of viruses. A viral lineage hypothesis predicts that viruses within the same lineage may have a common ancestor that existed before the separation of the cellular domains of life (3, 5, 8, 26). Currently, limited information is available on the detailed structures of viruses infecting archaea. For example, the virion structures of nontailed icosahedral Sulfolobus turreted icosahedral virus (STIV) and SH1 have been determined (21, 23, 46). However, most archaeal viruses represent unusual, sometimes nonregular, morphotypes (43), which makes it difficult to apply structural methods that are based on averaging techniques.A biochemical approach, i.e., controlled virion dissociation, gives information on the localization and interaction of virion components. In the present study, controlled dissociation was used to address the virion architecture of HRPV-1. A comparative lipid analysis of HRPV-1 and its host was also carried out. Our results show that the unique virion type is composed of a flexible membrane decorated with the glycosylated spikes of VP4 and internal membrane protein VP3. The circular ssDNA genome resides inside the viral membrane vesicle without detected association to any nucleoproteins.  相似文献   

3.
Ebola virus (EBOV) protein VP35 is a double-stranded RNA (dsRNA) binding inhibitor of host interferon (IFN)-α/β responses that also functions as a viral polymerase cofactor. Recent structural studies identified key features, including a central basic patch, required for VP35 dsRNA binding activity. To address the functional significance of these VP35 structural features for EBOV replication and pathogenesis, two point mutations, K319A/R322A, that abrogate VP35 dsRNA binding activity and severely impair its suppression of IFN-α/β production were identified. Solution nuclear magnetic resonance (NMR) spectroscopy and X-ray crystallography reveal minimal structural perturbations in the K319A/R322A VP35 double mutant and suggest that loss of basic charge leads to altered function. Recombinant EBOVs encoding the mutant VP35 exhibit, relative to wild-type VP35 viruses, minimal growth attenuation in IFN-defective Vero cells but severe impairment in IFN-competent cells. In guinea pigs, the VP35 mutant virus revealed a complete loss of virulence. Strikingly, the VP35 mutant virus effectively immunized animals against subsequent wild-type EBOV challenge. These in vivo studies, using recombinant EBOV viruses, combined with the accompanying biochemical and structural analyses directly correlate VP35 dsRNA binding and IFN inhibition functions with viral pathogenesis. Moreover, these studies provide a framework for the development of antivirals targeting this critical EBOV virulence factor.Ebola viruses (EBOVs) are zoonotic, enveloped negative-strand RNA viruses belonging to the family Filoviridae which cause lethal viral hemorrhagic fever in humans and nonhuman primates (47). Currently, information regarding EBOV-encoded virulence determinants remains limited. This, coupled with our lack of understanding of biochemical and structural properties of virulence factors, limits efforts to develop novel prophylactic or therapeutic approaches toward these infections.It has been proposed that EBOV-encoded mechanisms to counter innate immune responses, particularly interferon (IFN) responses, are critical to EBOV pathogenesis (7). However, a role for viral immune evasion functions in the pathogenesis of lethal EBOV infection has yet to be demonstrated. Of the eight major EBOV gene products, two viral proteins have been demonstrated to counter host IFN responses. The VP35 protein is a viral polymerase cofactor and structural protein that also inhibits IFN-α/β production by preventing the activation of interferon regulatory factor (IRF)-3 and -7 (3, 4, 8, 24, 27, 34, 41). VP35 also inhibits the activation of PKR, an IFN-induced, double-stranded RNA (dsRNA)-activated kinase with antiviral activity, and inhibits RNA silencing (17, 20, 48). The VP24 protein is a minor structural protein implicated in virus assembly and regulation of viral RNA synthesis, and changes in VP24 coding sequences are also associated with adaptation of EBOVs to mice and guinea pigs (2, 13, 14, 27, 32, 37, 50, 52). Further, VP24 inhibits cellular responses to both IFN-α/β and IFN-γ by preventing the nuclear accumulation of tyrosine-phosphorylated STAT1 (44, 45). The functions of VP35 and VP24 proteins are manifested in EBOV-infected cells by the absence of IRF-3 activation, impaired production of IFN-α/β, and severely reduced expression of IFN-induced genes, even after treatment of infected cells with IFN-α (3, 19, 21, 22, 24, 25, 28).Previous studies proposed that VP35 basic residues 305, 309, and 312 are required for VP35 dsRNA binding activity (26). VP35 residues K309 and R312 were subsequently identified as critical for binding to dsRNA, and mutation of these residues impaired VP35 suppression of IFN-α/β production (8). In vivo, an EBOV engineered to carry a VP35 R312A point mutation exhibited reduced replication in mice (23). However, because the parental recombinant EBOV into which the mutation was built did not cause disease in these animals, the impact of the mutation on viral pathogenesis could not be fully evaluated. Further, the lack of available structural and biochemical data to explain how the R312A mutation affects VP35 function limited avenues for the therapeutic targeting of critical VP35 functions. Recent structural analyses of the VP35 carboxy-terminal interferon inhibitory domain (IID) suggested that additional residues from the central basic patch may contribute to VP35 dsRNA binding activity and IFN-antagonist function (30). However, a direct correlation between dsRNA and IFN inhibitory functions of VP35 with viral pathogenesis is currently lacking.In order to further define the molecular basis for VP35 dsRNA binding and IFN-antagonist function and to define the contribution of these functions to EBOV pathogenesis, an integrated molecular, structural, and virological approach was taken. The data presented below identify two VP35 carboxy-terminal basic amino acids, K319 and R322, as required for its dsRNA binding and IFN-antagonist functions. Interestingly, these residues are outside the region originally identified as being important for dsRNA binding and IFN inhibition (26). However, they lie within the central basic patch identified by prior structural studies (26, 30). Introduction of these mutations (VP35 with these mutations is designated KRA) into recombinant EBOV renders this otherwise fully lethal virus avirulent in guinea pigs. KRA-infected animals also develop EBOV-specific antibodies and become fully resistant to subsequent challenge with wild-type (WT) virus. Our data further reveal that the KRA EBOV is immunogenic and likely replicates to low levels early after infection in vivo. However, the mutant virus is subsequently cleared by host immune responses. These data demonstrate that the VP35 central basic patch is important not only for IFN-antagonist function but also for EBOV immune evasion and pathogenesis in vivo. High-resolution structural analysis, coupled with our in vitro and in vivo analyses of the recombinant Ebola viruses, provides the molecular basis for loss of function by the VP35 mutant and highlights the therapeutic potential of targeting the central basic patch with small-molecule inhibitors and for future vaccine development efforts.  相似文献   

4.
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   

5.
6.
7.
Hepatitis A virus (HAV), an atypical member of the Picornaviridae, grows poorly in cell culture. To define determinants of HAV growth, we introduced a blasticidin (Bsd) resistance gene into the virus genome and selected variants that grew at high concentrations of Bsd. The mutants grew fast and had increased rates of RNA replication and translation but did not produce significantly higher virus yields. Nucleotide sequence analysis and reverse genetic studies revealed that a T6069G change resulting in a F42L amino acid substitution in the viral polymerase (3Dpol) was required for growth at high Bsd concentrations whereas a silent C7027T mutation enhanced the growth rate. Here, we identified a novel determinant(s) in 3Dpol that controls the kinetics of HAV growth.Hepatitis A virus (HAV) is an atypical member of the Picornaviridae that replicates poorly in cell culture and generally does not cause cytopathic effect (CPE). The HAV positive-strand RNA genome of about 7.5 kb is encapsidated in a 27- to 32-nm icosahedral shell (12). The HAV genome contains a long open reading frame (ORF) that codes for a polyprotein of approximately 250 kDa, which undergoes co- and posttranslational processing by the virus-encoded protease 3Cpro into structural (VP0, VP3, and VP1-2A) and nonstructural proteins (11, 13, 14, 18). VP0 undergoes structural cleavage into VP2 and VP4, and an unknown cellular protease cleaves the VP1-2A precursor (9, 23).HAV replicates inefficiently in cell culture and in general establishes persistent infections (3, 4, 7, 8) without causing CPE. However, some strains of HAV that replicate quickly can induce cell death (5, 19, 27). Due to the growth limitations, experimentation with HAV is difficult and the biology of this virus is poorly understood. To facilitate genetic studies, we recently introduced a blasticidin (Bsd) resistance gene at the 2A-2B junction of wild-type (wt) HAV (16). Bsd, an antibiotic that blocks translation in prokaryotes and eukaryotes and thus affects HAV translation, is inactivated by the Bsd-deaminase encoded in the Bsd resistance gene (15). Cells infected with the wt HAV construct carrying the Bsd resistance gene (HAV-Bsd) grew in the presence of Bsd. We have recently used the wt HAV-Bsd construct to select human hepatoma cell lines that support the stable growth of wt HAV (16) and to establish simple and rapid neutralization and virus titration assays (17). In this study, we developed a genetic approach to study determinants of HAV replication based on the selection of HAV-Bsd variants grown under increased concentrations of Bsd. We hypothesized that by increasing the concentration of Bsd, we would select HAV variants that grew better and allowed the survival of persistently infected cells at higher concentrations of the antibiotic. We also reasoned that we would need a robust HAV-Bsd replication system to provide enough Bsd-deaminase for cell survival. Therefore, we used attenuated HAV grown in rhesus monkey fetal kidney FRhK4 cells as an experimental system because (i) the virus grows 100-fold better in this system than wt HAV in human hepatoma cells (16), and (ii) it already contains cell culture-adapting mutations (3, 4, 7, 8) that are likely to accumulate during passage of wt HAV at high concentrations of Bsd and confound our analysis.  相似文献   

8.
Mature glycoprotein spikes are inserted in the Lassa virus envelope and consist of the distal subunit GP-1, the transmembrane-spanning subunit GP-2, and the signal peptide, which originate from the precursor glycoprotein pre-GP-C by proteolytic processing. In this study, we analyzed the oligomeric structure of the viral surface glycoprotein. Chemical cross-linking studies of mature glycoprotein spikes from purified virus revealed the formation of trimers. Interestingly, sucrose density gradient analysis of cellularly expressed glycoprotein showed that in contrast to trimeric mature glycoprotein complexes, the noncleaved glycoprotein forms monomers and oligomers spanning a wide size range, indicating that maturation cleavage of GP by the cellular subtilase SKI-1/S1P is critical for formation of the correct oligomeric state. To shed light on a potential relation between cholesterol and GP trimer stability, we performed cholesterol depletion experiments. Although depletion of cholesterol had no effect on trimerization of the glycoprotein spike complex, our studies revealed that the cholesterol content of the viral envelope is important for the infectivity of Lassa virus. Analyses of the distribution of viral proteins in cholesterol-rich detergent-resistant membrane areas showed that Lassa virus buds from membrane areas other than those responsible for impaired infectivity due to cholesterol depletion of lipid rafts. Thus, derivation of the viral envelope from cholesterol-rich membrane areas is not a prerequisite for the impact of cholesterol on virus infectivity.Lassa virus (LASV) is a member of the family Arenaviridae, of which Lymphocytic choriomeningitis virus (LCMV) is the prototype. Arenaviruses comprise more than 20 species, divided into the Old World and New World virus complexes (19). The Old World arenaviruses include the human pathogenic LASV strains, Lujo virus, which was first identified in late 2008 and is associated with an unprecedented high case fatality rate in humans, the nonhuman pathogenic Ippy, Mobala, and Mopeia viruses, and the recently described Kodoko virus (10, 30, 49). The New World virus complex contains, among others, the South American hemorrhagic fever-causing viruses Junín virus, Machupo virus, Guanarito virus, Sabiá virus, and the recently discovered Chapare virus (22).Arenaviruses contain a bisegmented single-stranded RNA genome encoding the polymerase L, matrix protein Z, nucleoprotein NP, and glycoprotein GP. The bipartite ribonucleoprotein of LASV is surrounded by a lipid envelope derived from the plasma membrane of the host cell. The matrix protein Z has been identified as a major budding factor, which lines the interior of the viral lipid membrane, in which GP spikes are inserted (61, 75). The glycoprotein is synthesized as precursor protein pre-GP-C and is cotranslationally cleaved by signal peptidase into GP-C and the signal peptide, which exhibits unusual length, stability, and topology (3, 27, 28, 33, 70, 87). Moreover, the arenaviral signal peptide functions as trans-acting maturation factor (2, 26, 33). After processing by signal peptidase, GP-C of both New World and Old World arenaviruses is cleaved by the cellular subtilase subtilisin kexin isozyme-1/site-1 protease (SKI-1/S1P) into the distal subunit GP-1 and the membrane-anchored subunit GP-2 within the secretory pathway (5, 52, 63). For LCMV, it has been shown that GP-1 subunits are linked to each other by disulfide bonds and are noncovalently connected to GP-2 subunits (14, 24, 31). GP-1 is responsible for binding to the host cell receptor, while GP-2 mediates fusion between the virus envelope and the endosomal membrane at low pH due to a bipartite fusion peptide near the amino terminus (24, 36, 44). Sequence analysis of the LCMV GP-2 ectodomain revealed two heptad repeats that most likely form amphipathic helices important for this process (34, 86).In general, viral class I fusion proteins have triplets of α-helical structures in common, which contain heptad repeats (47, 73). In contrast, class II fusion proteins are characterized by β-sheets that form dimers in the prefusion status and trimers in the postfusion status (43). The class III fusion proteins are trimers that, unlike class I fusion proteins, were not proteolytically processed N-terminally of the fusion peptide, resulting in a fusion-active membrane-anchored subunit (39, 62). Previous studies with LCMV described a tetrameric organization of the glycoprotein spikes (14), while more recent data using a bacterially expressed truncated ectodomain of the LCMV GP-2 subunit pointed toward a trimeric spike structure (31). Due to these conflicting data regarding the oligomerization status of LCMV GP, it remains unclear to which class of fusion proteins the arenaviral glycoproteins belong.The state of oligomerization and the correct conformation of viral glycoproteins are crucial for membrane fusion during virus entry. The early steps of infection have been shown for several viruses to be dependent on the cholesterol content of the participating membranes (i.e., either the virus envelope or the host cell membrane) (4, 9, 15, 20, 21, 23, 40, 42, 53, 56, 76, 78, 79). In fact, it has been shown previously that entry of both LASV and LCMV is susceptible to cholesterol depletion of the target host cell membrane using methyl-β-cyclodextrin (MβCD) treatment (64, 71). Moreover, cholesterol not only plays an important role in the early steps during entry in the viral life cycle but also is critical in the virus assembly and release process. Several viruses of various families, including influenza virus, human immunodeficiency virus type 1 (HIV-1), measles virus, and Ebola virus, use the ordered environment of lipid raft microdomains. Due to their high levels of glycosphingolipids and cholesterol, these domains are characterized by insolubility in nonionic detergents under cold conditions (60, 72). Recent observations have suggested that budding of the New World arenavirus Junin virus occurs from detergent-soluble membrane areas (1). Assembly and release from distinct membrane microdomains that are detergent soluble have also been described for vesicular stomatitis virus (VSV) (12, 38, 68). At present, however, it is not known whether LASV requires cholesterol in its viral envelope for successful virus entry or whether specific membrane microdomains are important for LASV assembly and release.In this study, we first investigated the oligomeric state of the premature and mature LASV glycoprotein complexes. Since it has been shown for several membrane proteins that the oligomerization and conformation are dependent on cholesterol (58, 59, 76, 78), we further analyzed the dependence of the cholesterol content of the virus envelope on glycoprotein oligomerization and virus infectivity. Finally, we characterized the lipid membrane areas from which LASV is released.  相似文献   

9.
10.
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   

11.
12.
The mechanism by which herpesviruses acquire their tegument is not yet clear. One model is that outer tegument proteins are recruited by the cytoplasmic tails of viral glycoproteins. In the case of herpes simplex virus tegument protein VP22, interactions with the glycoproteins gE and gD have been shown. We have previously shown that the C-terminal half of VP22 contains the necessary signal for assembly into the virus. Here, we show that during infection VP22 interacts with gE and gM, as well as its tegument partner VP16. However, by using a range of techniques we were unable to demonstrate VP22 binding to gD. By using pulldown assays, we show that while the cytoplasmic tails of both gE and gM interact with VP22, only gE interacts efficiently with the C-terminal packaging domain of VP22. Furthermore, gE but not gM can recruit VP22 to the Golgi/trans-Golgi network region of the cell in the absence of other virus proteins. To examine the role of the gE-VP22 interaction in infection, we constructed a recombinant virus expressing a mutant VP22 protein with a 14-residue deletion that is unable to bind gE (ΔgEbind). Coimmunoprecipitation assays confirmed that this variant of VP22 was unable to complex with gE. Moreover, VP22 was no longer recruited to its characteristic cytoplasmic trafficking complexes but exhibited a diffuse localization. Importantly, packaging of this variant into virions was abrogated. The mutant virus exhibited poor growth in epithelial cells, similar to the defect we have observed for a VP22 knockout virus. These results suggest that deletion of just 14 residues from the VP22 protein is sufficient to inhibit binding to gE and hence recruitment to the viral envelope and assembly into the virus, resulting in a growth phenotype equivalent to that produced by deleting the entire reading frame.The herpesvirus tegument is the virion compartment located between the DNA-containing capsid and the virus envelope (6). Although it is well defined that the viral capsid assembles in the nucleus (37, 38) and the viral envelope is acquired from cellular membranes (3, 24), the mechanism of tegument protein acquisition is still to be established. At least 20 virus-encoded components are recruited into the herpes simplex virus type 1 (HSV-1) tegument (32), and there is increasing evidence to suggest that subsets of these proteins may be added as assembly progresses along the maturation pathway (28). To ensure efficient incorporation, it is likely that individual tegument proteins are specifically targeted to their cellular site of recruitment. Such targeting could involve interaction with a viral partner, a cellular partner, or both. A clearer understanding of how individual tegument proteins are acquired by newly assembling virions will help to define the herpesvirus assembly pathway.A number of protein-protein interactions between individual tegument proteins (13, 40, 42), and between tegument proteins and glycoproteins (19, 20, 22, 32), have been described that may provide useful insight into the assembly process. In particular, the interaction of tegument proteins with the cytoplasmic tails of virus glycoproteins provides an attractive mechanism for the virion recruitment of at least the outer components of the tegument. In the case of VP22, the homologues from pseudorabies virus (PRV) and HSV-1 have been shown to interact with the cytoplasmic tail of gE (19, 20, 32). However, the role of this interaction in virus infection has not yet been clearly defined and the fact that additional glycoprotein interactions have been described, with gM in the case of PRV and gD in the case of HSV-1, may point to potential redundancy in the mechanism of VP22 packaging (4, 19, 20). In addition, we and others have previously shown that HSV-1 VP22 interacts directly with a second tegument protein, namely, VP16 (13, 33), an interaction that could provide an alternative route for VP22 to enter the virion. In a previous study, we concluded that the region of VP22 containing its VP16 interaction domain was required but not sufficient for optimal VP22 packaging into the assembling virion, with an additional C-terminal determinant also involved (23). We also demonstrated that the same region of VP22 that was required for virion packaging was essential to target the protein to its characteristic cytoplasmic trafficking complexes, suggesting that these specific sites may be the location in the cell for VP22 assembly into the virion (23). Since that study, O''Regan and coworkers have reported that the C-terminal half of HSV-1 VP22 also contains the binding site for gE (32), providing a possible candidate for an additional VP22 binding partner. Furthermore, as HSV-1 VP22 has been shown to bind to gD and PRV VP22 interacts with gM, it is possible that the C terminus of VP22 contains a gD and/or a gM binding site (4, 20).In the present study, we aimed to clarify the molecular mechanism by which VP22 is recruited into the virus particle. We show that HSV-1 VP22 binds efficiently to VP16, gE, and gM in the infected cell, but we cannot detect an interaction with gD. We show that the packaging domain of VP22 binds to the cytoplasmic tail of gE but not gM and that the same region of VP22 is recruited to the secretory pathway by gE in the absence of other virus proteins. Finally, we show that a mutant VP22 protein lacking a 14-residue peptide from its packaging domain is unable to interact with gE during infection, exhibits a different subcellular localization, and fails to assemble into the virus particle. This is the first characterization of a single protein-protein interaction essential for the packaging of an HSV-1 tegument protein.  相似文献   

13.
VP40, the major matrix protein of Marburg virus, is the main driving force for viral budding. Additionally, cellular factors are likely to play an important role in the release of progeny virus. In the present study, we characterized the influence of the vacuolar protein sorting (VPS) pathway on the release of virus-like particles (VLPs), which are induced by Marburg virus VP40. In the supernatants of HEK 293 cells expressing VP40, different populations of VLPs with either a vesicular or a filamentous morphology were detected. While the filaments were almost completely composed of VP40, the vesicular particles additionally contained considerable amounts of cellular proteins. In contrast to that in the vesicles, the VP40 in the filaments was regularly organized, probably inducing the elimination of cellular proteins from the released VLPs. Vesicular particles were observed in the supernatants of cells even in the absence of VP40. Mutation of the late-domain motif in VP40 resulted in reduced release of filamentous particles, and likewise, inhibition of the VPS pathway by expression of a dominant-negative (DN) form of VPS4 inhibited the release of filamentous particles. In contrast, the release of vesicular particles did not respond significantly to the expression of DN VPS4. Like the budding of VLPs, the budding of Marburg virus particles was partially inhibited by the expression of DN VPS4. While the release of VLPs from VP40-expressing cells is a valuable tool with which to investigate the budding of Marburg virus particles, it is important to separate filamentous VLPs from vesicular particles, which contain many cellular proteins and use a different budding mechanism.In recent years, virus-like particles (VLPs), which are formed upon recombinant expression of the viral matrix and/or surface glycoproteins, have been recognized as representing powerful tools for developing novel vaccines and investigating certain aspects of the viral replication cycle (24, 44, 59, 63). Matrix proteins from many enveloped RNA viruses, including retroviruses, rhabdoviruses, filoviruses, paramyxoviruses, orthomyxoviruses, and arenaviruses, are able to induce VLPs (10, 14, 18, 28-30, 48, 49, 52). Increasing evidence also indicates that budding activity, and thus the release of VLPs, is often influenced by a complex interplay with components of the endosomal sorting complexes required for transport (ESCRTs), which mainly constitute the vacuolar protein sorting (VPS) pathway (16, 38, 42, 54). ESCRTs trigger the formation and budding of vesicles into the lumina of multivesicular bodies (MVBs), and the constituents of the ESCRTs are recycled by the activity of VPS4, an AAA-type ATPase. Expression of dominant-negative (DN) VPS4 mutants, which lack the ability to bind or hydrolyze ATP, blocks recycling of the ESCRTs and induces the formation of enlarged endosomes lacking internal vesicle accumulation (2, 3, 7). The inward budding of vesicles into the MVBs is topologically similar to the budding of viruses, since the vesicles bud away from the cytosol and into the lumen (reviewed in references 1, 20, and 26). Therefore, it is not entirely surprising that viruses use the cellular ESCRT machinery to organize the budding of viral progeny. Interactions between viral matrix proteins and ESCRTs occur through tetrapeptide motifs, known as late domains, which were first identified in retroviruses. Known late domains consist of the amino acid sequence P(T/S)AP, PPxY, or YxxL, where “x” represents any amino acid (19, 25, 62). The P(T/S)AP motif, for example, mediates interaction with tumor susceptibility gene 101 (Tsg101) (16, 36, 57); the PPxY motif mediates binding to WW domains of Nedd4-like ubiquitin ligases (9, 22); and the YxxL motif mediates interaction with AIP1/Alix (35, 47, 58). Recently, a novel late-domain motif, FPIV, has been identified in paramyxoviruses (46), and it is thought that additional late-domain motifs remain to be discovered (for a review, see reference 5).Inhibition of the VPS pathway has been shown to inhibit the budding of various viruses that are released with the help of ESCRTs. However, the budding of viruses and VLPs depends on the activity of ESCRTs to different degrees. Downregulation of Tsg101, a member of the ESCRT-I complex, inhibited the release of VLPs mediated by lymphocytic choriomeningitis virus Z protein and Marburg virus (MARV) VP40 (42, 54) but did not substantially inhibit the release of Gag-induced VLPs of Moloney murine leukemia virus and Rous sarcoma virus or that of matrix protein-induced VLPs of rabies virus (16, 27, 38). Expression of DN VPS4 inhibited the release of VLPs induced by the Gag proteins of Rous sarcoma virus and Moloney murine leukemia virus (16, 38) as well as that of VLPs induced by Lassa virus Z protein (55) but had no effect on the budding of rabies virus and cytomegalovirus (13, 27). These data indicate that in spite of the presence of late-domain motifs, a block in the VPS pathway may not always be critical for the budding of VLPs. In addition, the lack of known late domains in many enveloped viruses raises the question of whether they use other entry points into the VPS pathway or whether they exploit entirely different mechanisms of budding (60). To date, knowledge of how viral matrix proteins engage cellular machineries, such as the VPS pathway, to induce viral budding at the plasma membrane is very limited (8).The matrix protein VP40 of MARV contains only one known late-domain motif, PPPY, and a recent study showed that mutation of this late domain inhibited the release of VP40-induced VLPs. In addition, depletion of Tsg101 reduced the release of VP40-induced VLPs, suggesting that ESCRT-I is involved in this process (54). Whether a functional VPS pathway is important for the release of MARV VP40-induced VLPs or MARV particles remains unknown.VLPs induced by many viral matrix proteins have a morphology similar to that of cellular vesicles, which makes it difficult to separate the spherical VLPs from released cellular vesicles (4, 17, 53). In contrast, VLPs induced by the filovirus matrix protein VP40 are elongated and similar in morphology to viral particles (30, 49). Nevertheless, we observed that the supernatants of cells expressing VP40 contained various populations of particles with different morphologies. This raised the questions of whether the different particles are released by the same mechanism, whether they are all induced by VP40, and whether they are dependent on the same cellular pathways.The aim of the present study was to analyze the populations of particles released from cells expressing the MARV matrix protein VP40 and to gain further insights into the interaction between MARV and the cellular machinery involved in the budding of VLPs and MARV particles.We found that cells expressing VP40 released vesicular and filamentous particles, which could be separated by gradient centrifugation. Fractions with mainly vesicular particles represented a mixture of vesicles containing exclusively cellular proteins and vesicles also containing VP40 and few short filamentous particles. Longer filamentous particles, whose morphology resembled that of MARV particles but which displayed a much higher variability in length (400 nm to 5 μm), were found in denser gradient fractions. Filamentous VP40-induced VLPs were able to sort out cellular proteins efficiently. Release of VP40-induced filamentous VLPs was supported by the late-domain motif present in VP40, and inhibition of the cellular ESCRT machinery reduced the amount of these VLPs in the supernatant. Interestingly, the release of VLPs induced by a mutant of VP40 lacking the late domain was also reduced by inhibition of the cellular ESCRT machinery. Expression of a DN mutant of VPS4 diminished the budding of infectious MARV particles by 50%, a finding consistent with the idea that the activity of the ESCRT machinery supports viral budding but is not essential.  相似文献   

14.
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   

15.
16.
17.
18.
Bluetongue (BT), caused by Bluetongue virus (BTV), is an economically important disease affecting sheep, deer, cattle, and goats. Since 1998, a series of BT outbreaks have spread across much of southern and central Europe. To study why the epidemiology of the virus happens to change, it is important to fully know the mechanisms resulting in its genetic diversity. Gene mutation and segment reassortment have been considered as the key forces driving the evolution of BTV. However, it is still unknown whether intragenic recombination can occur and contribute to the process in the virus. We present here several BTV groups containing mosaic genes to reveal that intragenic recombination can take place between the virus strains and play a potential role in bringing novel BTV lineages.Bluetongue (BT) is an economically significant disease that seriously threatens sheep, some species of deer, and to a lesser extent cattle and goats. As a vector-borne viral disease of ruminants, BT is endemic in tropical and subtropical countries (46). However, a series of BT outbreaks have spread across much of southern and central Europe since 1998 (29). Thus, it is of great importance to fully understand the molecular basis driving the change of its epidemiology so as to prevent or limit future BT pandemics.Bluetongue virus (BTV), the pathogen of BT, belongs to the Orbivirus genus of the Reoviridae family (46). The virus has a segmented double-stranded RNA (dsRNA) genome that is packaged in a nonenveloped, icosahedral particle (46). Its 10 dsRNA segments encode 11 proteins, VP1 to VP7 (encoded by segments 1, 2, 3, 4, 6, 9, and 7, respectively), NS1 to SN3 (encoded by segments 5, 8, and 10, respectively), and NS3A (encoded by segment 10) (46). Two structural proteins, VP2 and VP5, form the outer layer of the virion particle and are responsible for cell attachment and virus entry (18, 31, 32), neutralizing epitope (14, 21), and virus virulence (36). Both of them are highly variable and generate 24 serotypes of the virus (44). The inner layers contain VP1, VP3, VP4, VP6, and VP7, and form the “core” of the BTV capsid. VP1 and VP6 are involved in RNA replication as the RNA-dependent RNA polymerase (54) and helicase/NTPase, respectively (49). VP7 forms the surface of the core and functions during the entry of the core into insect cells (44) and also can react with “core neutralizing” antibodies as a major serogroup-specific antigen (32, 44). These core proteins and two nonstructural proteins, NS1 and NS2, are thought to be relatively conservative, so that antigenic cross-reaction can take place between different BTV strains and serotypes, whereas NS3/N3a is more variable than the other nonstructural or core proteins (46).The genetic diversity and variation in sequences of different BTV genome segments were initially identified by RNA oligonucleotide fingerprint analysis of BTV field samples (47). Until now, reassortment and dynamic gene mutation, regarded as the key factors responsible for the genetic diversity of BTV, have been studied in details (46). The two mechanisms can result in both genetic drift and genetic shift and contribute to BTV evolution (47). It has been revealed that high-frequency genome segment reassortment occurs readily between different BTV serotypes (16). Thus, segment reassortment is an important factor in generation of genetic diversity in orbivirus populations in nature (45). In addition, it has been shown that homologous recombination can also play a role in the genetic diversity and evolution of some RNA viruses (24, 33) and bring on virulent variants of these viruses at last (8, 56). Although homologous recombination has been observed in rotavirus, a member of the Reoviridae (39, 40), it is still unknown whether the intragenic recombination can occur and play a role in the generation of genetic diversity in orbivirus populations.To determine whether homologous recombination shaped the evolution of BTV and to provide some insights into the recombination itself in the virus, we analyzed roughly 690 complete segments of BTV deposited in GenBank to see whether some of them underwent intragenic recombination event. Several BTV groups isolated at different time points and in different countries were found containing the same (or similar) mosaic segments, demonstrating that intragenic recombination had occurred in the field and that these viruses with mosaic segments had become prevailing strains. That is, intragenic recombination can play a potential role in generating genetic diversity of BTV and exert its influence on the change of BTV epidemiology.  相似文献   

19.
The acid-dependent disassembly of foot-and-mouth disease virus (FMDV) is required for viral RNA release from endosomes to initiate replication. Although the FMDV capsid disassembles at acid pH, mutants escaping inhibition by NH4Cl of endosomal acidification were found to constitute about 10% of the viruses recovered from BHK-21 cells infected with FMDV C-S8c1. For three of these mutants, the degree of NH4Cl resistance correlated with the sensitivity of the virion to acid-induced inactivation of its infectivity. Capsid sequencing revealed the presence in each of these mutants of a different amino acid substitution (VP3 A123T, VP3 A118V, and VP2 D106G) that affected a highly conserved residue among FMDVs located close to the capsid interpentameric interfaces. These residues may be involved in the modulation of the acid-induced dissociation of the FMDV capsid. The substitution VP3 A118V present in mutant c2 was sufficient to confer full resistance to NH4Cl and concanamycin A (a V-ATPase inhibitor that blocks endosomal acidification) as well as to increase the acid sensitivity of the virion to an extent similar to that exhibited by mutant c2 relative to the sensitivity of the parental virus C-S8c1. In addition, the increased propensity to dissociation into pentameric subunits of virions bearing substitution VP3 A118V indicates that this replacement also facilitates the dissociation of the FMDV capsid.Foot-and-mouth disease virus (FMDV) is a member of the Aphthovirus genus in the family Picornaviridae. FMDV displays epithelial tropism and is responsible for a highly contagious disease of cloven-hoofed animals (23, 60). FMDV populations are quasispecies and exhibit a high potential for variation and adaptation, one consequence of which is the extensive antigenic diversity of this virus, reflected in the existence of seven serotypes and multiple antigenic variants (reviewed in references 17 and 60). Different cellular receptors, including αvβ integrins and heparan sulfate (HS) glycosaminoglycans, have been described for natural isolates and tissue culture-adapted FMDVs (3, 4, 6, 28-31, 56). However, viruses that are infectious in vivo use integrins as receptors (28). The interaction between FMDV and the integrin molecule is mediated by an Arg-Gly-Asp (RGD) triplet located at the G-H loop of capsid protein VP1 (9, 47). FMDV isolates interacting with integrins gain entry into the cell following clathrin-mediated endocytosis (8, 39, 52). On the other hand, it has been described that a genetically engineered HS-binding mutant uses caveolae to enter into cultured cells (51). After internalization, FMDV must release its genomic RNA molecule of positive polarity into the host cell cytoplasm to establish a productive infection. Early work showed that a variety of lysosomotropic agents, such as weak bases and ionophores that block acidification of endosomes, inhibit FMDV infection (5, 11-13), indicating that genome release is dependent on endosomal acidification. In addition, internalized FMDV particles colocalize with markers from early and recycling endosomes (8, 51, 52) and FMDV infection is reduced by expression of a dominant negative mutant of Rab5 (33), suggesting that FMDV may release its genome from these compartments.The FMDV capsid comprises 60 copies of each of the four structural proteins (VP1 to VP4) arranged in an icosahedral lattice of 12 pentameric subunits. FMDV particles are highly acid labile and disassemble at pH values slightly below neutrality (13). Acid lability is not a feature of the capsids of other picornaviruses, such as Enterovirus. Pentameric subunits are intermediates of FMDV assembly and disassembly (64). A high density of His residues is found close to the interpentameric interface. Protonation of these residues at the acidic pH in the endosomes has been proposed to trigger acid-induced capsid disassembly by electrostatic repulsion between the protonated His side chains (1). His 142 (H142) in VP3 of type A FMDV is involved in a His-α-helix dipole interaction, which is likely to influence the acid lability of FMDV (13). In silico predictions suggested that H142 and H145 in VP3 may have the greatest effect on this process (63). Experimental evidence of the involvement of H142 of VP3 in acid-induced disassembly of FMDV has also been reported (20). Concomitantly with capsid disassembly into pentameric intermediates, internal protein VP4 and viral RNA are released. VP4 is a highly hydrophobic and myristoylated protein (7) whose release has been suggested to mediate membrane permeabilization and ion channel formation, thus facilitating the endosomal exit of viral RNA (15, 16, 34).Besides providing information about the endosomal pH requirements for the release of virus genomes, drugs modifying endosomal acidification can reveal the molecular changes associated with viral resistance to their action. These analyses may also address whether the balance between acid lability and capsid stability required for completion of virus replication allows FMDV, which disassembles at a pH close to neutrality, to escape inhibition by drugs raising the endosomal pH. In this work, we have isolated and characterized FMDV mutants that are able to escape from the inhibition of endosomal acidification exerted by NH4Cl, a lysosomotropic weak base that raises endolysosomal pH and impairs uncoating and infection of viruses that require transit through acidic endosomal compartments for penetration (5, 26, 53). These mutants showed an increased acid lability, which is likely to allow them to uncoat at more-alkaline pH values. A single amino acid substitution close to the interpentameric interfaces in the capsid of one of these mutants was responsible for a total resistance to the elevation in endosomal pH caused by NH4Cl treatment and for the acid-labile phenotype.  相似文献   

20.
Studies of the hepatitis C virus (HCV) life cycle have been aided by development of in vitro systems that enable replication of viral RNA and production of infectious virus. However, the functions of the individual proteins, especially those engaged in RNA replication, remain poorly understood. It is considered that NS4B, one of the replicase components, creates sites for genome synthesis, which appear as punctate foci at the endoplasmic reticulum (ER) membrane. In this study, a panel of mutations in NS4B was generated to gain deeper insight into its functions. Our analysis identified five mutants that were incapable of supporting RNA replication, three of which had defects in production of foci at the ER membrane. These mutants also influenced posttranslational modification and intracellular mobility of another replicase protein, NS5A, suggesting that such characteristics are linked to focus formation by NS4B. From previous studies, NS4B could not be trans-complemented in replication assays. Using the mutants that blocked RNA synthesis, defective NS4B expressed from two mutants could be rescued in trans-complementation replication assays by wild-type protein produced by a functional HCV replicon. Moreover, active replication could be reconstituted by combining replicons that were defective in NS4B and NS5A. The ability to restore replication from inactive replicons has implications for our understanding of the mechanisms that direct viral RNA synthesis. Finally, one of the NS4B mutations increased the yield of infectious virus by five- to sixfold. Hence, NS4B not only functions in RNA replication but also contributes to the processes engaged in virus assembly and release.Recent estimates predict that the prevalence of hepatitis C virus (HCV) infection is approximately 2.2% worldwide, equivalent to about 130 million persons (22). The virus typically establishes a chronic infection that frequently leads to serious liver disease (1), and current models indicate that both morbidity and mortality as a consequence of HCV infection will continue to rise for about the next 20 years (10, 11, 29).HCV is the only assigned species of the Hepacivirus genus within the family Flaviviridae. The virus can be classified into six genetic groups or clades (numbered 1 to 6) and then further separated into subtypes (e.g., 1a, 1b, 2a, 2b, etc.) (53, 55). HCV has a single-stranded, positive-sense RNA genome that is approximately 9.6 kb in length (reviewed in reference 46). Genomic RNA carries a single open reading frame flanked by 5′ and 3′ nontranslated regions, which are important for both replication and translation (19, 20, 34, 47, 56). Viral RNA is translated by the host ribosomal machinery, and the resultant polyprotein is co- and posttranslationally cleaved to generate the mature viral proteins. The structural proteins (core, E1, and E2) and a small hydrophobic polypeptide called p7 are produced by the cellular proteases signal peptidase and signal peptide peptidase (28, 45, 54). Two virus-encoded proteases, the NS2-3 autoprotease and the NS3 serine protease (5, 13, 26), are responsible for maturation of the nonstructural (NS) proteins (NS2, NS3, NS4A, NS4B, NS5A, and NS5B). With the exception of NS2, the NS proteins are necessary for genome replication (8, 40) and form replication complexes (RCs), which are located at the endoplasmic reticulum (ER) membrane (14, 24, 52, 57, 59). The functions of all viral constituents of RCs have not been characterized in detail. It is known that NS5B is the RNA-dependent RNA polymerase (6), while NS3 possesses helicase and nucleoside triphosphatase activities in addition to acting as a protease (32, 58). However, the precise roles of the other proteins remain to be firmly established.Expression of NS4B, one of the replicase proteins, generates rearrangements at the ER membrane that have been termed the “membranous web” (14, 24) and “membrane-associated foci” (MAFs) (25). Detection of viral RNA at such foci suggests that NS4B is involved in creating the sites where genome synthesis occurs (18, 24, 59). It is predicted that NS4B has an amphipathic α-helix within its N-terminal region, which is followed by four transmembrane domains (TMDs) in the central portion of the protein (17, 42). As a result, the majority of NS4B is likely to be tightly anchored to membranes, and experimental evidence indicates that it has characteristics consistent with an integral membrane protein (27). It is thought that after membrane association, NS4B rearranges membranes into a network, thereby generating foci which act as a “scaffold” to facilitate RNA replication. The mechanisms engaged in formation of foci are not known but include the notion that the NS4B N terminus can translocate into the ER lumen, resulting in rearrangement of cellular membranes (41, 42). Alternatively, palmitoylation, a lipid modification, might facilitate polymerization of NS4B, in turn promoting formation of RCs on the ER membrane (68).Apart from inducing membranous changes required for replication, NS4B may perform other tasks in HCV RNA synthesis. For example, studies of cell culture adaptive mutations in subgenomic replicons (SGRs) have identified amino acid changes that can stimulate RNA production (39), suggesting that NS4B may exert a regulatory role in determining replication efficiency. In support of a regulatory function, replacement of NS4B sequences in an SGR from strain H77 (a genotype 1a strain) with those from strain Con-1 (a genotype 1b strain) gave higher levels of replication than for a wild-type (wt) strain H77 SGR (7). The corresponding replacement of strain Con-1 NS4B sequences with those from strain H77 reduced the replication efficiency of a Con-1 SGR (7). Moreover, interactions of NS4B with the RC can affect the behavior of other replicase proteins. For example, NS4B is needed for hyperphosphorylation of NS5A (35, 48) and restricts its intracellular movement (30).To try to gain greater insight into the functional organization of the components that constitute RCs, trans-complementation assays using defective and helper SGRs have been established (2, 64). Such studies reveal that the only protein capable of trans-complementation is NS5A, while active replication cannot be restored for replicons harboring deleterious mutations in NS3, NS4B, and NS5B. These data led to the conclusion that functional NS5A may be able to exchange between RCs (2), whereas, by inference, such exchange would not be possible for other HCV replicase proteins. In transient-replication assays, complementation by NS5A also relied on its expression as part of a polyprotein (minimally NS3-NS5A), and production of the protein alone failed to restore replication for an inactive SGR (2). However, in a separate study, stable expression of wt NS5A was capable of complementing a defective replicon (64). Thus, different assay systems can give dissimilar results for complementation by NS5A.In this study, we have created a series of mutations in the NS4B gene of HCV strain JFH1 (31) to explore the function of the protein in the HCV life cycle. We focused our attention on the C-terminal portion of NS4B, downstream from the predicted TMD regions, since it is relatively well conserved and is predicted to lie on the cytosolic side of the ER membrane (15, 42). Our analysis examines the impact of mutations on replication efficiency and the intracellular characteristics of the mutants compared to the behavior of the wt protein. In addition, we have utilized this series of mutants to reassess trans-complementation of NS4B in replication assays. Finally, we also analyze the impact of mutations which do not affect replication on the production of infectious virus to determine whether NS4B plays a role in virus assembly and release.  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