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 62 毫秒
1.
2.
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   

3.
During the search for haloarchaeal viruses, we isolated and characterized a new pleomorphic lipid-containing virus, Haloarcula hispanica pleomorphic virus 1 (HHPV-1), that infects the halophilic archaeon Haloarcula hispanica. The virus contains a circular double-stranded DNA genome of 8,082 bp in size. The organization of the genome shows remarkable synteny and amino acid sequence similarity to the genome and predicted proteins of the halovirus HRPV-1, a pleomorphic single-stranded DNA virus that infects a halophilic archaeon Halorubrum sp. Analysis of the two halovirus sequences, as well as the entire nucleotide sequence of the 10.8-kb pHK2-plasmid and a 12.6-kb chromosomal region in Haloferax volcanii, allows us to suggest a new group of closely related viruses with genomes of either single-stranded or double-stranded DNA. Currently, closely related viruses are considered to have the same genome type. Our observation clearly contradicts this categorization and indicates that we should reconsider the way we classify viruses. Our results also provide a new example of related viruses where the viral structural proteins have not diverged as much as the proteins associated with genome replication. This result further strengthens the proposal for higher-order classification to be based on virion architecture rather than on genome type or replication mechanism.Metagenomic studies have increased the amount of information on the nucleotide sequence space in our environment. It has also increased our awareness of the viral abundance and diversity not recognized before (16, 24, 26). Along with this new information, we have learned to acknowledge the significance of viruses in the evolution and behavior of other organisms (55). To reveal the dynamics and molecular interactions in the interplay between a particular virus and its host, however, isolation of single viruses and their hosts is needed. Even though a number of viruses pathogenic to humans, domestic animals, and plants, as well as some bacteriophages, have been studied in great detail, much of the diversity of the archaeal viruses has remained unknown. By the year 2007 only 44 archaeal viruses had been described (2). That embraces less than 1% of all reported viruses. Although the diversity among these few isolated archaeal viruses is considerable, a head-and-tail morphology is prevalent among isolated viruses infecting euryarchaeal cells. In contrast, viruses of Crenarchaeota are diverse and often unusual with no viruses having a head-tail morphology (53).Archaeal haloviruses infect euryarchaeal hosts living in environments up to saturated salt. This makes them an interesting group of viruses that reside in a very restricted habitat. In samples taken from high salt environments, the Dead Sea and Spanish solar salterns, viral morphotypes most often observed were spindle-shaped, head-and-tail or tailless icosahedral particles (25, 31, 47). Isolated haloviruses, however, do not seem to reflect the proportions of different morphotypes found in the nature as nearly all of the isolates possess a head-and-tail morphology (2). Molecular level studies on only two spindle-shaped (10, 11) and one tailless icosahedral particle have been carried out (37, 51). Virus-like particles of other morphologies have also been observed in high-salt environments (47), but only one additional morphotype has been described in detail (50). This recently isolated lipid containing halovirus, HRPV-1, is the first archaeal virus containing a single-stranded DNA (ssDNA) genome (50). It infects Halorubrum sp. and has a pleomorphic appearance with glycosylated spike structures protruding from its external membrane (49, 50).The evolution of prokaryotic viral genome sequences is very fast (18), and the assessment of viral relationships using homology of the genome sequences applies only to closely related viruses (17, 19). Current higher-order classification of viruses is based on the host organism, the nature of the genome (RNA/DNA, single stranded versus double stranded) and the virion morphology. Recently, a higher-order clustering of virus families has been proposed based on common principles of virion architectures as well as on the fold of the major capsid protein (1, 6, 12, 13, 42). Consequently, major capsid proteins most probably belong to the vertically inherited viral “self” (4), whereas proteins involved in replication of the viral genome can be swapped by horizontal exchange (21, 63). The proposal is based on observations that structurally related viruses have been found to infect organisms that reside in all three domains of life.We have isolated a new pleomorphic virus infecting Haloarcula hispanica (Har. hispanica pleomorphic virus 1 [HHPV-1]). Here, we determine the molecular constituents of HHPV-1 and its genetic relatedness to other archaeal viruses and putative proviruses. Sequence homology and gene order (synteny) shows distinct genomic regions shared between four genetic elements separating replication, virus assembly, and integration functions. Surprisingly, in spite of the close relatedness of HRPV-1 and HHPV-1, the genome types of these two viruses differ (ssDNA and dsDNA, respectively).  相似文献   

4.
5.
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   

6.
The emergence of natural isolates of human respiratory syncytial virus group B (HRSV-B) with a 60-nucleotide (nt) duplication in the G protein gene in Buenos Aires, Argentina, in 1999 (A. Trento et al., J. Gen. Virol. 84:3115-3120, 2003) and their dissemination worldwide allowed us to use the duplicated segment as a natural tag to examine in detail the evolution of HRSV during propagation in its natural host. Viruses with the duplicated segment were all clustered in a new genotype, named BA (A. Trento et al., J. Virol. 80:975-984, 2006). To obtain information about the prevalence of these viruses in Spain, we tested for the presence of the duplicated segment in positive HRSV-B clinical samples collected at the Severo Ochoa Hospital (Madrid) during 12 consecutive epidemics (1996-1997 to 2007-2008). Viruses with the 60-nt duplication were found in 61 samples, with a high prevalence relative to the rest of B genotypes in the most recent seasons. Global phylogenetic and demographic analysis of all G sequences containing the duplication, collected across five continents up until April 2009, revealed that the prevalence of the BA genotype increased gradually until 2004-2005, despite its rapid dissemination worldwide. After that date and coinciding with a bottleneck effect on the population size, a relatively new BA lineage (BA-IV) replaced all other group B viruses, suggesting further adaptation of the BA genotype to its natural host.Human respiratory syncytial virus (HRSV), a member of the Pneumovirus genus within the Paramyxoviridae family, is recognized as the leading agent responsible for severe respiratory infections in the pediatric population (31, 34, 35) and a pathogen of considerable importance in vulnerable adults (23, 24). The global respiratory syncytial virus (RSV) disease burden is estimated at 64 million cases and 160,000 deaths every year (70). This virus causes regular seasonal epidemics which take place during the winter months in temperate countries or during the rainy season in tropical areas (12). A peculiar aspect of HRSV is that the immune response produced by infection does not confer long-lasting protection, which is why reinfections are common throughout life (30).Neutralization tests performed with hyperimmune serum (16) and reactivity with specific monoclonal antibodies (4, 45) were used to classify HRSV isolates into two antigenic groups, A and B, which correlated with genetically distinct viruses (18). The main differences between these two groups are located in the major attachment G protein. This protein is a type II glycoprotein that shares neither sequence nor structural features with the attachment proteins (HN or H) of other paramyxoviruses (69), and it represents one of the targets of the immune response (27, 43). The full-length membrane-bound G protein (Gm) of 292 to 319 amino acids (depending on the viral strain) is also expressed in a secreted version (Gs) that lacks the transmembrane domain due to alternative initiation of translation at a second in-frame AUG codon in the G open reading frame (M48) (52). The G protein is the viral gene product with the highest degree of antigenic and genetic diversity among viral isolates (4, 18, 28, 45). Most changes are concentrated in two hypervariable regions that flank a highly conserved central region of the G protein ectodomain, which includes a cluster of four cysteines and the putative receptor binding site (43). It has been suggested that antigenic differences within this protein could facilitate repeated HRSV infections (37, 59). In addition, positive selection of amino acid changes was observed in the two hypervariable regions of the G protein ectodomain (7, 43, 71, 73, 74). One of the hypervariable regions, located in the C-terminal one-third of the G molecule, contains multiple epitopes recognized by monoclonal antibodies (43), suggesting that immune selection of new variants by antibodies may contribute to generation of HRSV diversity.Phylogenetic studies based on sequence analysis of the G protein have identified numerous genotypes in the antigenic groups A and B that show complex circulation patterns, since multiple genotypes of both antigenic groups may circulate within the same season and community, with one or two dominant genotypes being replaced in successive years (13, 14, 26, 27, 32, 49, 50). Each community shows a seasonal circulation pattern of genotypes, probably determined by local factors, such as the level of herd immunity to certain strains (3, 14, 49).The capacity of the G protein to accommodate drastic sequence changes was illustrated best by three antigenic group B viruses isolated in Buenos Aires, Argentina, in 1999 that contained a duplication of 60 nucleotides (nt) in the C-terminal third of the G protein gene (63). The global dissemination of these viruses allowed us to use the duplicated segment as a natural tag to reexamine the evolution of HRSV during propagation in its natural host. Phylogenetic analysis of G sequences revealed that all viruses with the duplicated segment clustered in a new genotype, named BA, and this finding supported the idea of a common ancestor for all viruses with the 60-nt duplication, dated about 1998 (64). The limited information about the molecular epidemiology of HRSV in Spain, together with an increase in G sequences with the duplicated segment reported worldwide, prompted us to conduct both a local search in Madrid for these viruses and a global phylogenetic analysis of HRSV with the 60-nt duplication from the time that these viruses were first detected, taking into account the geographic and temporal distribution of each isolate.  相似文献   

7.
In their vertebrate hosts, arboviruses such as Semliki Forest virus (SFV) (Togaviridae) generally counteract innate defenses and trigger cell death. In contrast, in mosquito cells, following an early phase of efficient virus production, a persistent infection with low levels of virus production is established. Whether arboviruses counteract RNA interference (RNAi), which provides an important antiviral defense system in mosquitoes, is an important question. Here we show that in Aedes albopictus-derived mosquito cells, SFV cannot prevent the establishment of an antiviral RNAi response or prevent the spread of protective antiviral double-stranded RNA/small interfering RNA (siRNA) from cell to cell, which can inhibit the replication of incoming virus. The expression of tombusvirus siRNA-binding protein p19 by SFV strongly enhanced virus spread between cultured cells rather than virus replication in initially infected cells. Our results indicate that the spread of the RNAi signal contributes to limiting virus dissemination.In animals, RNA interference (RNAi) was first described for Caenorhabditis elegans (27). The production or introduction of double-stranded RNA (dsRNA) in cells leads to the degradation of mRNAs containing homologous sequences by sequence-specific cleavage of mRNAs. Central to RNAi is the production of 21- to 26-nucleotide small interfering RNAs (siRNAs) from dsRNA and the assembly of an RNA-induced silencing complex (RISC), followed by the degradation of the target mRNA (23, 84). RNAi is a known antiviral strategy of plants (3, 53) and insects (21, 39, 51). Study of Drosophila melanogaster in particular has given important insights into RNAi responses against pathogenic viruses and viral RNAi inhibitors (31, 54, 83, 86, 91). RNAi is well characterized for Drosophila, and orthologs of antiviral RNAi genes have been found in Aedes and Culex spp. (13, 63).Arboviruses, or arthropod-borne viruses, are RNA viruses mainly of the families Bunyaviridae, Flaviviridae, and Togaviridae. The genus Alphavirus within the family Togaviridae contains several mosquito-borne pathogens: arboviruses such as Chikungunya virus (16) and equine encephalitis viruses (88). Replication of the prototype Sindbis virus and Semliki Forest virus (SFV) is well understood (44, 71, 74, 79). Their genome consists of a positive-stranded RNA with a 5′ cap and a 3′ poly(A) tail. The 5′ two-thirds encodes the nonstructural polyprotein P1234, which is cleaved into four replicase proteins, nsP1 to nsP4 (47, 58, 60). The structural polyprotein is encoded in the 3′ one-third of the genome and cleaved into capsid and glycoproteins after translation from a subgenomic mRNA (79). Cytoplasmic replication complexes are associated with cellular membranes (71). Viruses mature by budding at the plasma membrane (35).In nature, arboviruses are spread by arthropod vectors (predominantly mosquitoes, ticks, flies, and midges) to vertebrate hosts (87). Little is known about how arthropod cells react to arbovirus infection. In mosquito cell cultures, an acute phase with efficient virus production is generally followed by the establishment of a persistent infection with low levels of virus production (9). This is fundamentally different from the cytolytic events following arbovirus interactions with mammalian cells and pathogenic insect viruses with insect cells. Alphaviruses encode host response antagonists for mammalian cells (2, 7, 34, 38).RNAi has been described for mosquitoes (56) and, when induced before infection, antagonizes arboviruses and their replicons (1, 4, 14, 15, 29, 30, 32, 42, 64, 65). RNAi is also functional in various mosquito cell lines (1, 8, 43, 49, 52). In the absence of RNAi, alphavirus and flavivirus replication and/or dissemination is enhanced in both mosquitoes and Drosophila (14, 17, 31, 45, 72). RNAi inhibitors weakly enhance SFV replicon replication in tick and mosquito cells (5, 33), posing the questions of how, when, and where RNAi interferes with alphavirus infection in mosquito cells.Here we use an A. albopictus-derived mosquito cell line to study RNAi responses to SFV. Using reporter-based assays, we demonstrate that SFV cannot avoid or efficiently inhibit the establishment of an RNAi response. We also demonstrate that the RNAi signal can spread between mosquito cells. SFV cannot inhibit cell-to-cell spread of the RNAi signal, and spread of the virus-induced RNAi signal (dsRNA/siRNA) can inhibit the replication of incoming SFV in neighboring cells. Furthermore, we show that SFV expression of a siRNA-binding protein increases levels of virus replication mainly by enhancing virus spread between cells rather than replication in initially infected cells. Taken together, these findings suggest a novel mechanism, cell-to-cell spread of antiviral dsRNA/siRNA, by which RNAi limits SFV dissemination in mosquito cells.  相似文献   

8.
9.
Although enveloped virus assembly in the host cell is a crucial step in the virus life cycle, it remains poorly understood. One issue is how viruses include lipids in their membranes during budding from infected host cells. To analyze this issue, we took advantage of the fact that baby hamster kidney cells can be infected by two different viruses, namely, vesicular stomatitis virus and Semliki Forest virus, from the Rhabdoviridae and Togaviridae families, respectively. We purified the host plasma membrane and the two different viruses after exit from the host cells and analyzed the lipid compositions of the membranes by quantitative shotgun mass spectrometry. We observed that the lipid compositions of these otherwise structurally different viruses are virtually indistinguishable, and only slight differences were detected between the viral lipid composition and that of the plasma membrane. Taken together, the facts that the lipid compositions of the two viruses are so similar and that they strongly resemble the composition of the plasma membrane suggest that these viruses exert little selection in including lipids in their envelopes.Enveloped viruses acquire their lipid envelope from the membranes of host cells (43). In this process, the nucleocapsid or the nucleocapsid-matrix complex of the viruses buds out of the cell and becomes enveloped by a segment of the host membrane. This membrane segment is modified during the budding process, such that virally encoded membrane proteins are included in the viral envelope, while most host proteins are excluded. Since viruses usually do not carry lipid-synthesizing enzymes, the lipids in the viral envelope are derived from the host membrane. The lipid compositions of enveloped viruses have been studied for years (2, 15, 17, 18, 23, 25, 34, 36, 38, 40). One question that remains to be answered is whether the lipids are included passively, and thus the lipid composition of the envelope reflects the lipid composition of the host membrane, or whether lipid sorting occurs, leading to selective inclusion of some lipids and exclusion of others. This issue has been complicated by the fact that the lipid bilayer is no longer considered a homogenous liquid but contains fluctuating nanoscale assemblies of sphingolipids, saturated phospholipids, cholesterol (Chol), and proteins, called lipid rafts (13, 44). Lipid rafts can be induced to coalesce—usually by protein-protein interactions—into larger, dynamic platforms that function in signal transduction, intracellular membrane transport, and other membrane functions (45). It was also proposed that viruses make use of these membrane domains during their exit from cells (29, 32).A major complication in comparing viral envelopes with host cell membranes is the difficulty in obtaining host cell membranes of purity similar to that of the easily purified viruses. Many studies are faulted by the impurity of the cell membranes analyzed. Moreover, the early work in this field employed conventional analytical methods (such as thin-layer chromatography) that provide only semiquantitative estimates of the total abundance of the major lipid classes. Most importantly, lipid species diversity could not be analyzed. Recent developments in mass spectrometry (MS) have enabled comprehensive and quantitative analyses of lipidomes at the level of individual molecular species. The lipidomes of human immunodeficiency virus (HIV), murine leukemia virus (6, 7), and several bacteriophages (20, 21) were recently analyzed by these new methods.This paper focuses on two well-characterized enveloped viruses, Semliki Forest virus (SFV) and vesicular stomatitis virus (VSV). SFV is an RNA virus belonging to the Togaviridae family of the Alphaviridae that acquires its envelope by budding from the host cell plasma membrane (PM) (46). Early studies analyzed the lipid composition of the viral envelope and also that of the host cell PM (39, 40). These studies revealed strong similarity between the envelope of SFV and the host PM, but one important discrepancy was the higher Chol-to-phospholipid ratio in the virus.VSV is an RNA virus belonging to the Rhabdoviridae family and also hijacks its envelope from the host cell PM (35), but the lipid specificity of the budding process remains controversial. The most recent studies claim that VSV buds from localized regions that do not reflect the average composition of the PM (23, 36). It has also been claimed that lipid rafts are involved in VSV envelope assembly during budding (37).We used BHK-21 cells as host cells to purify SFV and VSV. The purposes of this study were (i) to establish a robust, comprehensive, and quantitative method to analyze lipidomes, including the full complement of glycerolipid, glycerophospholipid, and sphingolipid species as well as Chol; (ii) to establish a protocol for purification of PM suitable for MS analysis; and (iii) to analyze and compare the lipidomes of SFV, VSV, and the BHK-21 PM.We found that the lipidomes of SFV and VSV are similar in molecular composition and are closely related to that of the BHK-21 PM. The small differences observed could be explained by the high degrees of curvature generated during the viral budding process.  相似文献   

10.
Cytosolic chaperones are a diverse group of ubiquitous proteins that play central roles in multiple processes within the cell, including protein translation, folding, intracellular trafficking, and quality control. These cellular proteins have also been implicated in the replication of numerous viruses, although the full extent of their involvement in viral replication is unknown. We have previously shown that the heat shock protein 40 (hsp40) chaperone encoded by the yeast YDJ1 gene facilitates RNA replication of flock house virus (FHV), a well-studied and versatile positive-sense RNA model virus. To further explore the roles of chaperones in FHV replication, we examined a panel of 30 yeast strains with single deletions of cytosolic proteins that have known or hypothesized chaperone activity. We found that the majority of cytosolic chaperone deletions had no impact on FHV RNA accumulation, with the notable exception of J-domain-containing hsp40 chaperones, where deletion of APJ1 reduced FHV RNA accumulation by 60%, while deletion of ZUO1, JJJ1, or JJJ2 markedly increased FHV RNA accumulation, by 4- to 40-fold. Further studies using cross complementation and double-deletion strains revealed that the contrasting effects of J domain proteins were reproduced by altering expression of the major cytosolic hsp70s encoded by the SSA and SSB families and were mediated in part by divergent effects on FHV RNA polymerase synthesis. These results identify hsp70 chaperones as critical regulators of FHV RNA replication and indicate that cellular chaperones can have both positive and negative regulatory effects on virus replication.The compact genomes of viruses relative to those of other infectious agents restrict their ability to encode all proteins required to complete their replication cycles. To circumvent this limitation, viruses often utilize cellular factors or processes to complete essential steps in replication. One group of cellular proteins frequently targeted by viruses are cellular chaperones, which include a diverse set of heat shock proteins (hsps) that normally facilitate cellular protein translation, folding, trafficking, and degradation (18, 64). The connection between viruses and cellular chaperones was originally identified in bacteria, where the Escherichia coli hsp40 and hsp70 homologues, encoded by dnaJ and dnaK, respectively, were identified as bacterial genes essential for bacteriophage λ DNA replication (62). Research over the past 30 years has further revealed the importance of cellular chaperones in viral replication, such that the list of virus-hsp connections is now quite extensive and includes viruses from numerous families with diverse genome structures (4, 6, 7, 16, 19, 20, 23, 25, 40, 41, 44, 51, 54, 60). These studies have demonstrated the importance of cellular chaperones in multiple steps of the viral life cycle, including entry, viral protein translation, genome replication, encapsidation, and virion release. However, the list of virus-hsp connections is likely incomplete. Further studies to explore this particular host-pathogen interaction will shed light on virus replication mechanisms and pathogenesis, and potentially highlight targets for novel antiviral agents.To study the role of cellular chaperones in the genome replication of positive-sense RNA viruses, we use flock house virus (FHV), a natural insect pathogen and well-studied member of the Nodaviridae family. The FHV life cycle shares many common features with other positive-sense RNA viruses, including the membrane-specific targeting and assembly of functional RNA replication complexes (37, 38), the exploitation of various cellular processes and host factors for viral replication (5, 23, 60), and the induction of large-scale membrane rearrangements (24, 28, 38, 39). FHV virions contain a copackaged bipartite genome consisting of RNA1 (3.1 kb) and RNA2 (1.4 kb), which encode protein A, the viral RNA-dependent RNA polymerase, and the structural capsid protein precursor, respectively (1). During active genome replication, FHV produces a subgenomic RNA3 (0.4 kb), which encodes the RNA interference inhibitor protein B2 (12, 29, 32). These viral characteristics make FHV an excellent model system to study many aspects of positive-sense RNA virus biology.In addition to the benefits of a simple genome, FHV is able to establish robust RNA replication in a wide variety of genetically tractable eukaryotic hosts, including Drosophila melanogaster (38), Caenorhabditis elegans (32), and Saccharomyces cerevisiae (46). The budding yeast S. cerevisiae has been an exceptionally useful model host to study the mechanisms of viral RNA replication complex assembly and function with FHV (31, 37, 39, 45, 53, 55, 56, 60) as well as other positive-sense RNA viruses (11). The facile genetics of S. cerevisiae, along with the vast array of well-defined cellular and molecular tools and techniques, make it an ideal eukaryotic host for the identification of cellular factors required for positive-sense RNA virus replication. Furthermore, readily available yeast libraries with deletions and regulated expression of individual proteins have led to the completion of several high-throughput screens to provide a global survey of host factors that impact virus replication (26, 42, 52). An alternative approach with these yeast libraries that reduces the inherently high false-negative rates associated with high-throughput screens is to focus on a select set of host genes associated with a particular cellular pathway, process, or location previously implicated in virus replication.We have utilized such a targeted approach and focused on examining the impact of cytosolic chaperones on FHV RNA replication. Previously, we have shown that the cellular chaperone hsp90 facilitates protein A synthesis in Drosophila cells (5, 23), and the hsp40 encoded by the yeast YDJ1 gene facilitates FHV RNA replication in yeast, in part through effects on both protein A accumulation and function (60). In this report, we further extend these observations by examining FHV RNA accumulation in a panel of yeast strains with deletions of known or hypothesized cytosolic chaperones. We demonstrate that cytosolic chaperones can have either suppressive or enhancing effects on FHV RNA accumulation. In particular, related hsp70 members encoded by the SSA and SSB yeast chaperone families have marked and dramatically divergent effects on both genomic and subgenomic RNA accumulation and viral polymerase synthesis. These results highlight the complexities of the host-pathogen interactions that influence positive-sense RNA virus replication and identify the hsp70 family of cytosolic chaperones as key regulators of FHV replication.  相似文献   

11.
Antibodies against the extracellular virion (EV or EEV) form of vaccinia virus are an important component of protective immunity in animal models and likely contribute to the protection of immunized humans against poxviruses. Using fully human monoclonal antibodies (MAbs), we now have shown that the protective attributes of the human anti-B5 antibody response to the smallpox vaccine (vaccinia virus) are heavily dependent on effector functions. By switching Fc domains of a single MAb, we have definitively shown that neutralization in vitro—and protection in vivo in a mouse model—by the human anti-B5 immunoglobulin G MAbs is isotype dependent, thereby demonstrating that efficient protection by these antibodies is not simply dependent on binding an appropriate vaccinia virion antigen with high affinity but in fact requires antibody effector function. The complement components C3 and C1q, but not C5, were required for neutralization. We also have demonstrated that human MAbs against B5 can potently direct complement-dependent cytotoxicity of vaccinia virus-infected cells. Each of these results was then extended to the polyclonal human antibody response to the smallpox vaccine. A model is proposed to explain the mechanism of EV neutralization. Altogether these findings enhance our understanding of the central protective activities of smallpox vaccine-elicited antibodies in immunized humans.The smallpox vaccine, live vaccinia virus (VACV), is frequently considered the gold standard of human vaccines and has been enormously effective in preventing smallpox disease. The smallpox vaccine led to the worldwide eradication of the disease via massive vaccination campaigns in the 1960s and 1970s, one of the greatest successes of modern medicine (30). However, despite the efficacy of the smallpox vaccine, the mechanisms of protection remain unclear. Understanding those mechanisms is key for developing immunologically sound vaccinology principles that can be applied to the design of future vaccines for other infectious diseases (3, 101).Clinical studies of fatal human cases of smallpox disease (variola virus infection) have shown that neutralizing antibody titers were either low or absent in patient serum (24, 68). In contrast, neutralizing antibody titers for the VACV intracellular mature virion (MV or IMV) were correlated with protection of vaccinees against smallpox (68). VACV immune globulin (VIG) (human polyclonal antibodies) is a promising treatment against smallpox (47), since it was able to reduce the number of smallpox cases ∼80% among variola-exposed individuals in four case-controlled clinical studies (43, 47, 52, 53, 69). In animal studies, neutralizing antibodies are crucial for protecting primates and mice against pathogenic poxviruses (3, 7, 17, 21, 27, 35, 61, 66, 85).The specificities and the functions of protective antipoxvirus antibodies have been areas of intensive research, and the mechanics of poxvirus neutralization have been debated for years. There are several interesting features and problems associated with the antibody response to variola virus and related poxviruses, including the large size of the viral particles and the various abundances of many distinct surface proteins (18, 75, 91, 93). Furthermore, poxviruses have two distinct virion forms, intracellular MV and extracellular enveloped virions (EV or EEV), each with a unique biology. Most importantly, MV and EV virions share no surface proteins (18, 93), and therefore, there is no single neutralizing antibody that can neutralize both virion forms. As such, an understanding of virion structure is required to develop knowledge regarding the targets of protective antibodies.Neutralizing antibodies confer protection mainly through the recognition of antigens on the surface of a virus. A number of groups have discovered neutralizing antibody targets of poxviruses in animals and humans (3). The relative roles of antibodies against MV and EV in protective immunity still remain somewhat unclear. There are compelling data that antibodies against MV (21, 35, 39, 66, 85, 90, 91) or EV (7, 16, 17, 36, 66, 91) are sufficient for protection, and a combination of antibodies against both targets is most protective (66). It remains controversial whether antibodies to one virion form are more important than those to the other (3, 61, 66). The most abundant viral particles are MV, which accumulate in infected cells and are released as cells die (75). Neutralization of MV is relatively well characterized (3, 8, 21, 35). EV, while less abundant, are critical for viral spread and virulence in vivo (93, 108). Neutralization of EV has remained more enigmatic (3).B5R (also known as B5 or WR187), one of five known EV-specific proteins, is highly conserved among different strains of VACV and in other orthopoxviruses (28, 49). B5 was identified as a protective antigen by Galmiche et al., and the available evidence indicated that the protection was mediated by anti-B5 antibodies (36). Since then, a series of studies have examined B5 as a potential recombinant vaccine antigen or as a target of therapeutic monoclonal antibodies (MAbs) (1, 2, 7, 17, 40, 46, 66, 91, 110). It is known that humans immunized with the smallpox vaccine make antibodies against B5 (5, 22, 62, 82). It is also known that animals receiving the smallpox vaccine generate antibodies against B5 (7, 20, 27, 70). Furthermore, previous neutralization assays have indicated that antibodies generated against B5 are primarily responsible for neutralization of VACV EV (5, 83). Recently Chen at al. generated chimpanzee-human fusion MAbs against B5 and showed that the MAbs can protect mice from lethal challenge with virulent VACV (17). We recently reported, in connection with a study using murine monoclonal antibodies, that neutralization of EV is highly complement dependent and the ability of anti-B5 MAbs to protect in vivo correlated with their ability to neutralize EV in a complement-dependent manner (7).The focus of the study described here was to elucidate the mechanisms of EV neutralization, focusing on the human antibody response to B5. Our overall goal is to understand underlying immunobiological and virological parameters that determine the emergence of protective antiviral immune responses in humans.  相似文献   

12.
The multiplicity of infection (MOI), i.e., the number of virus genomes that infect a cell, is a key parameter in virus evolution, as it determines processes such as genetic exchange among genomes, selection intensity on viral genes, epistatic interactions, and the evolution of multipartite viruses. In fact, the MOI level is equivalent to the virus ploidy during genome expression. Nevertheless, there are few experimental estimates of MOI, particularly for viruses with eukaryotic hosts. Here we estimate the MOI of Tobacco mosaic virus (TMV) in its systemic host, Nicotiana benthamiana. The progress of infection of two TMV genotypes, differently tagged with the green or red fluorescent proteins GFP and RFP, was monitored by determining the number of leaf cell protoplasts that showed GFP, RFP, or GFP and RFP fluorescence at different times postinoculation. This approach allowed the quantitative analysis of the kinetics of infection and estimation of the generation time and the number of infection cycles required for leaf colonization. MOI levels were estimated from the frequency of cells infected by only TMV-GFP or TMV-RFP. The MOI was high, but it changed during the infection process, decreasing from an initial level of about 6 to a final one of 1 to 2, with most infection cycles occurring at the higher MOI levels. The decreasing MOI can be explained by mechanisms limiting superinfection and/or by genotype competition within double-infected cells, which was shown to occur in coinfected tobacco protoplasts. To our knowledge, this is the first estimate of MOI during virus colonization of a eukaryotic host.Virus evolution has been a very active area of research in the last few decades, as viruses are both important pathogens of humans, animals, and plants and good models to experimentally test hypotheses on parasite evolution or, more generally, central questions on evolutionary biology (11, 12, 21, 36). Considerable efforts have been devoted to modeling the evolution of viral populations. However, contrasting the theoretical models with reality may be hindered by limited experimental information on important parameters of the virus life cycle. The multiplicity of infection (MOI), i.e., the number of virus particles or genomes that may infect a cell, is a key parameter in many models of virus evolution (5, 6, 14, 15, 37, 38, 39, 52, 53, 57, 61) for which experimental estimates are scant.When a cell is coinfected by different viral genomes, competition may lead to decreased fitness of individual genotypes in comparison with their fitness in single infections (15, 31, 40). Thus, limiting coinfection may result in a selective advantage for viruses (58), which have developed mechanisms to prevent superinfection of previously infected cells (51, 60). On the other hand, infection of a cell by more than one virus genome is a prerequisite for two central phenomena in virus genetics to take place: recombination and complementation of defective mutants. Recombination between viral strains during replication in the same cell and complementation of defective mutants have been extensively documented for viruses infecting prokaryotes, animals, and plants (2, 25, 56), indicating that there must be some degree of coinfection and, hence, that the MOI must be higher than one in at least some infected cells. However, estimates of MOI in the natural hosts of viruses are surprisingly scarce in spite of this parameter''s relevance: values of about 2 to 3 have been reported for different DNA or RNA bacteriophages (26, 41, 51, 58), and a value of 4 to 5 was reported for Autographa californica nuclear polyhedrosis virus infecting larvae of the moth Tricoplusia ni (3), to our knowledge, the only estimate for a virus in its eukaryotic host. We are not aware of estimates reported for viruses infecting mammals or plants, although a MOI of about 3 can be inferred from the number of proviral copies of HIV in spleen cells of infected patients (29). This paucity of data may be due to the technical difficulty of directly measuring MOI, particularly within a eukaryotic host. Genetic approaches may provide valid alternatives for estimating MOI levels (3, 58), and here, the MOI of a plant virus is estimated through the analysis of the relative frequencies of two genotypes during the process of host colonization.Host colonization by plant-infecting viruses has been known for a long time to be a two-step phenomenon. First, colonization proceeds slowly from the initially infected cells to their neighbors by way of the cytoplasmic connections called plasmodesmata, a process known as cell-to-cell movement. After infection thus reaches the cells in the vasculature, the second step, known as long-distance or systemic movement, occurs as viruses move faster to distant organs through the vascular tissue, the phloem in most cases (59). As a result of these processes, the virus population within the infected plant may be strongly structured. Analyses of different viruses in different host plant species have shown that systemic movement causes population bottlenecks that may be severe (16, 28, 32, 34, 46), resulting in differences in the genetic composition of the virus subpopulations in different systemically infected organs. No analysis of population bottlenecks during cell-to-cell movement has been reported, but data indicate that the virus population within a leaf has a strong spatial structure with a separate distribution of different genotypes in different leaf areas. These reports derive from analyses of viruses that differ in genomic organization and gene expression strategies in different host plant species (9, 10, 23, 55); they indicate that a separate distribution of viral genotypes within the infected leaf is a general phenomenon and suggest limitation of coinfection. Data on the spatial exclusion of virus genotypes within the infected leaf are in apparent contradiction with the abundant evidence of recombination and complementation of defective mutants, which has been widely documented for plant viruses (19, 44, 50, 62). It should be pointed out that all reports on the spatial exclusion of virus genotypes in an infected leaf derive from microscopy observations, mostly at late times after infection of the tissue. No information is available on the kinetics of leaf colonization by viruses, and current data do not allow the estimation of MOI.In this report, we estimate the MOI of a plant RNA virus, Tobacco mosaic virus (TMV), in its systemic host, Nicotiana benthamiana. For this, we have reexamined the process of virus colonization by monitoring the progress of infection of two TMV genotypes in inoculated and in systemically infected leaves. The two TMV genotypes differed in the expression of fluorescent tags, either the green fluorescent protein (GFP) from Aequorea victoria (42, 43) or a red fluorescent protein (RFP) from Discosoma sp. (49). The expression of GFP and RFP allowed the precise quantification of the number of cells infected by either one or both TMV genotypes, and these data allowed the estimation of genotype frequencies and of MOI. The results show evidence of strong spatial structure of the virus population, with most cells being infected by either TMV-GFP or TMV-RFP alone and only a small fraction of cells being double infected. The kinetics of the single and double infections show that the MOI changes with time, decreasing as colonization progresses and therefore suggesting that exclusion mechanisms operate at later times after infection.  相似文献   

13.
Ubiquitin conjugation to lysine residues regulates a variety of protein functions, including endosomal trafficking and degradation. While ubiquitin plays an important role in the release of many viruses, the requirement for direct ubiquitin conjugation to viral structural proteins is less well understood. Some viral structural proteins require ubiquitin ligase activity, but not ubiquitin conjugation, for efficient release. Recent evidence has shown that, like other viruses, hepatitis B virus (HBV) requires a ubiquitin ligase for release from the infected cell. The HBV core protein contains two lysine residues (K7 and K96), and K96 has been suggested to function as a potential ubiquitin acceptor site based on the fact that previous studies have shown that mutation of this amino acid to alanine blocks HBV release. We therefore reexamined the potential connection between core lysine ubiquitination and HBV replication, protein trafficking, and virion release. In contrast to alanine substitution, we found that mutation of K96 to arginine, which compared to alanine is more conserved but also cannot mediate ubiquitin conjugation, does not affect either virus replication or virion release. We also found that the core lysine mutants display wild-type sensitivity to the antiviral activity of interferon, which demonstrates that ubiquitination of core lysines does not mediate the interferon-induced disruption of HBV capsids. However, mutation of K96 to arginine alters the nuclear-cytoplasmic distribution of core, leading to an accumulation in the nucleolus. In summary, these studies demonstrate that although ubiquitin may regulate the HBV replication cycle, these mechanisms function independently of direct lysine ubiquitination of core protein.The hepatitis B virus (HBV) particle consists of an enveloped nucleocapsid that contains the viral polymerase (Pol) and an incomplete 3.2-kb double-stranded DNA genome (9). In the cytoplasm, the viral core structural proteins interact to form homodimers, which further self-assemble into capsid particles that package Pol and the viral pregenomic RNA. Encapsidated Pol subsequently reverse transcribes pregenomic RNA to give rise to mature double-stranded relaxed circular DNA-containing capsids. HBV DNA-containing capsids are released from the cell as mature virions after acquiring an envelope consisting of cellular membrane lipids and the viral small, middle, and large envelope proteins (4, 9, 41). Due to the directed insertion of the envelope proteins in the endoplasmic reticulum and Golgi membrane, and the requirement of the large envelope protein for virion release, nucleocapsids are hypothesized to bud at intracellular membranes for release through the constitutive secretory pathway (5). Although the mechanism and site of HBV nucleocapsid envelopment and release remain poorly understood, emerging evidence indicates that the cellular ubiquitin pathway may play a role in this process.Structural proteins of some enveloped RNA viruses contain highly conserved sequences [PPXY, P(T/S)AP, and YPXL] termed late (L) domains that mediate interactions with proteins of the endocytic pathway to facilitate virus budding and release (1). The P(T/S)AP motif binds Tsg101 (8, 10, 19, 27, 47), a key ESCRT (for endosomal sorting complex required for transport) component for the recognition and sorting of ubiquitinated proteins to internal vesicles of the multivesicular body (MVB), while the YPXL motif binds Alix, an ESCRT-associated protein (26, 44, 48). The PPXY motif binds proteins of the Nedd4 family ubiquitin ligases, which are responsible for ubiquitination of proteins targeted for endocytosis and sorting to the MVB (20), suggesting a link between ubiquitin and viral budding (3, 16, 17, 22, 43, 55). The observation that proteasome inhibition, which depletes free cellular ubiquitin by interfering with ubiquitin recycling, results in a viral budding defect similar to that seen in virus L domain mutants further supports the implication that ubiquitin plays a role in mediating virion release (15, 31, 40, 43). Furthermore, fusion of ubiquitin to the Rous sarcoma virus (RSV) PPPY-containing Gag protein and the equine infectious anemia virus (EIAV) Gag protein containing a heterologous PTAP or PPPY motif rescues the virus-like particle release defect induced by proteasome inhibition (18, 31). While the role of L domains in mediating virion release is relatively well established, it remains unclear whether direct ubiquitination of viral structural proteins is generally required for virion release. Mutation of ubiquitin acceptor lysine residues in the RSV Gag protein inhibits virus budding, but such mutations in human immunodeficiency virus type 1 (HIV-1) or murine leukemia virus Gag protein exert no effect on virus release (29, 42). Recently, a retroviral (i.e., prototypic foamy virus) Gag protein engineered to lack ubiquitin acceptor lysines and encoding either the PSAP or PPXY motif of the L domain displayed no defect in viruslike particle release (58). Altogether, these results suggest that recruitment of host proteins to the L domain and ubiquitination of interacting proteins, but not the viral structural proteins, is required for ubiquitin-dependent virion release, at least for some viruses.The HBV core structural protein contains two potential ubiquitin acceptor lysine residues (K7 and K96) and an L-domain-like PPAY motif (Fig. (Fig.1A).1A). Structural studies indicate that residue K96 and the PPAY motif may be exposed on the surface of HBV capsid particles, at least transiently (4, 32, 37). Studies aimed at identifying interaction factors important for HBV particle release demonstrated a number of interesting findings. First, γ2-adaptin, a cellular trafficking adaptor that contains a ubiquitin-interacting motif (UIM), interacts with both the viral large envelope protein and HBV core, and disruption of the HBV/γ2-adaptin interaction inhibits virus secretion (14, 39). Second, core protein interacts with the Nedd4 ubiquitin ligase through the PPAY motif in core (39). Mutation of the tyrosine in the PPAY motif results in disrupted binding of Nedd4, and overexpression of a catalytically inactive Nedd4 mutant inhibits HBV particle secretion (39). Third, mutation of core K96, but not K7, to alanine results in a defective release phenotype, suggesting that K96 may serve as a ubiquitin conjugation site that aids virion release (32, 39). Recently, overexpression of dominant-negative proteins of the MVB machinery, such as the Vps4 ATPases and the ESCRT-III complex-forming CHMP proteins, were also shown to disrupt HBV budding and virion release, while subviral particles comprised only of envelope proteins were released efficiently (21, 24, 49). This suggests that nucleocapsids may release from the cell by a mechanism distinct from constitutive secretion. These studies show that similar to RNA viruses, HBV utilizes components of the cellular protein trafficking machinery to mediate virion release.Open in a separate windowFIG. 1.Generation of core lysine mutants. (A) The 21-kDa HBV core structural protein contains two lysine residues at positions 7 and 96 that serve as potential ubiquitin conjugation sites. These residues are highly conserved among the four major HBV genotypes (6). Core contains a late-domain-like PPXY motif that serves as a binding site for the Nedd4 E3 ubiquitin ligase. Core additionally contains a potential noncanonical SUMOylation motif at position 96. (B) Lysine mutations were generated by site-directed mutagenesis in the core gene contained within the HBV genome under the control of a CMV promoter. K7R contains a lysine-to-arginine mutation at position 7, K96R contains a lysine-to-arginine mutation at position 96, K96A contains a lysine-to-alanine mutation at position 96, and K7R/K96R contains arginine substituted at position 7 and position 96.Although these findings imply that core ubiquitination may be necessary for HBV particle release, direct evidence of core ubiquitination has been elusive (33, 39; unpublished results). As suggested by previous Gag lysine mutagenesis studies, however, ubiquitin may instead indirectly be required through conjugation to an interacting protein that is essential for mediating HBV release (29, 58). Although core K7 and K96 have been previously assayed in the context of virion release by mutation of the lysine residues to alanine (32, 39), we expanded these studies by assaying core mutants with an arginine substitution at position K7 (K7R) and K96 (K96R), as well as a double lysine-to-arginine mutation (K7R/K96R). Compared to alanine, arginine serves as a more conserved mutation for lysine while still abolishing the potential ubiquitin conjugation site. In the present study, we utilized these mutants to comprehensively examine the role of the core lysines in HBV virus release, the formation of replication intermediates, intracellular localization of core, and the interferon (IFN)-mediated antiviral response.  相似文献   

14.
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   

15.
16.
17.
18.
19.
Understanding the mechanisms underlying potential altered susceptibility to human immunodeficiency virus type 1 (HIV-1) infection in highly exposed seronegative (ES) individuals and the later clinical consequences of breakthrough infection can provide insight into strategies to control HIV-1 with an effective vaccine. From our Seattle ES cohort, we identified one individual (LSC63) who seroconverted after over 2 years of repeated unprotected sexual contact with his HIV-1-infected partner (P63) and other sexual partners of unknown HIV-1 serostatus. The HIV-1 variants infecting LSC63 were genetically unrelated to those sequenced from P63. This may not be surprising, since viral load measurements in P63 were repeatedly below 50 copies/ml, making him an unlikely transmitter. However, broad HIV-1-specific cytotoxic T-lymphocyte (CTL) responses were detected in LSC63 before seroconversion. Compared to those detected after seroconversion, these responses were of lower magnitude and half of them targeted different regions of the viral proteome. Strong HLA-B27-restricted CTLs, which have been associated with disease control, were detected in LSC63 after but not before seroconversion. Furthermore, for the majority of the protein-coding regions of the HIV-1 variants in LSC63 (except gp41, nef, and the 3′ half of pol), the genetic distances between the infecting viruses and the viruses to which he was exposed through P63 (termed the exposed virus) were comparable to the distances between random subtype B HIV-1 sequences and the exposed viruses. These results suggest that broad preinfection immune responses were not able to prevent the acquisition of HIV-1 infection in LSC63, even though the infecting viruses were not particularly distant from the viruses that may have elicited these responses.Understanding the mechanisms of altered susceptibility or control of human immunodeficiency virus type 1 (HIV-1) infection in highly exposed seronegative (ES) persons may provide invaluable information aiding the design of HIV-1 vaccines and therapy (9, 14, 15, 33, 45, 57, 58). In a cohort of female commercial sex workers in Nairobi, Kenya, a small proportion of individuals remained seronegative for over 3 years despite the continued practice of unprotected sex (12, 28, 55, 56). Similarly, resistance to HIV-1 infection has been reported in homosexual men who frequently practiced unprotected sex with infected partners (1, 15, 17, 21, 61). Multiple factors have been associated with the resistance to HIV-1 infection in ES individuals (32), including host genetic factors (8, 16, 20, 37-39, 44, 46, 47, 49, 59, 63), such as certain HLA class I and II alleles (41), as well as cellular (1, 15, 26, 55, 56), humoral (25, 29), and innate immune responses (22, 35).Seroconversion in previously HIV-resistant Nairobi female commercial sex workers, despite preexisting HIV-specific cytotoxic T-lymphocyte (CTL) responses, has been reported (27). Similarly, 13 of 125 ES enrollees in our Seattle ES cohort (1, 15, 17) have become late seroconverters (H. Zhu, T. Andrus, Y. Liu, and T. Zhu, unpublished observations). Here, we analyze the virology, genetics, and immune responses of HIV-1 infection in one of the later seroconverting subjects, LSC63, who had developed broad CTL responses before seroconversion.  相似文献   

20.
Polyomavirus JC (JCV) infects ∼60% of the general population, followed by asymptomatic urinary shedding in ∼20%. In patients with pronounced immunodeficiency, including HIV/AIDS, JCV can cause progressive multifocal leukoencephalopathy (PML), a devastating brain disease of high mortality. While JCV in the urine of healthy people has a linear noncoding control region called the archetype NCCR (at-NCCR), JCV in brain and cerebrospinal fluid (CSF) of PML patients bear rearranged NCCRs (rr-NCCRs). Although JCV NCCR rearrangements are deemed pathognomonic for PML, their role as a viral determinant is unclear. We sequenced JCV NCCRs found in CSF of eight HIV/AIDS patients newly diagnosed with PML and analyzed their effect on early and late gene expression using a bidirectional reporter vector recapitulating the circular polyomavirus early and late gene organization. The rr-NCCR sequences were highly diverse, but all increased viral early reporter gene expression in progenitor-derived astrocytes, glia-derived cells, and human kidney compared to the expression levels with the at-NCCR. The expression of simian virus 40 (SV40) large T antigen or HIV Tat expression in trans was associated with a strong increase of at-NCCR-controlled early gene expression, while rr-NCCRs were less responsive. The insertion of rr-NCCRs into the JCV genome backbone revealed higher viral replication rates for rr-NCCR compared to those of the at-NCCR JCV in human progenitor-derived astrocytes or glia cells, which was abrogated in SV40 large T-expressing COS-7 cells. We conclude that naturally occurring JCV rr-NCCR variants from PML patients confer increased early gene expression and higher replication rates compared to those of at-NCCR JCV and thereby increase cytopathology.Polyomavirus JC (JCV) infects approximately 60% of the general population, followed by asymptomatic urinary shedding in 20% of healthy individuals (20). Although JCV-associated nephropathy may occur in kidney transplant (14, 33) and HIV/AIDS patients (6, 27), the most prominent JCV disease is progressive multifocal leukoencephalopathy (PML) (44, 60). The pathology of PML was first described in 1958 as a rare complication of patients with chronic lymphocytic leukemia or Hodgkin''s lymphoma (3). Today, PML is recognized as a rare, virus-mediated demyelinating disease of the white brain matter in highly immunocompromised patients, including HIV/AIDS, transplantation, and chemotherapy patients and those exposed to immunomodulatory or depleting biologicals for the treatment of autoimmune diseases (29, 40). During the human immunodeficiency virus type 1 (HIV-1) pandemic, the incidence of PML rose significantly to rates of 1 to 8% prior to the use of highly active antiretroviral therapy (2, 5, 34). The definitive diagnosis requires brain tissue, but the detection of JCV by PCR in cerebrospinal fluid (CSF) is generally accepted for a laboratory-confirmed diagnosis in immunocompromised patients with (multi-)focal neurological deficits and corresponding radiological findings (8, 26). Due to the lack of effective antiviral therapy (13), the treatment of PML is based on improving overall immune functions. While this is difficult to achieve in cancer, chemotherapy, and transplantation, prompt antiretroviral therapy in HIV/AIDS patients has significantly improved PML survival, with increasing JCV-specific immune responses and declining intracerebral JCV replication (7, 15, 23, 35, 37). In patients diagnosed with PML after treatment with natalizumab for multiple sclerosis or inflammatory bowel disease, the removal of the monoclonal antibody by plasmapheresis has been tried to restore lymphocyte homing to, and the immune surveillance of, JCV replication sites in the central nervous system (38, 40, 52). However, the success of immune reconstitution in HIV/AIDS- and natalizumab-associated PML cases is limited by the fact that PML is typically diagnosed clinically by neurological deficits resulting from significant brain damage, where mounting antiviral immunity often may be too slow to modify the outcome. On the other hand, rapid recovery may cause immune reconstitution inflammatory syndrome with paradoxical clinical worsening and fatal outcomes (9, 16, 25, 38, 46). Although the etiologic role of JCV in PML is well documented, the pathogenesis and, in particular, the role of viral determinants is less clear. Virtually all JCV strains isolated from the brain or CSF of PML patients are characterized by highly variable genomic rearrangements of the noncoding control region (NCCR), which governs viral early and late genes in opposite directions of the circular polyomavirus DNA genome (1, 4, 31, 39, 41, 43, 49, 54, 59). In contrast, JCV detected in the urine of immunocompetent individuals show a consistent linear architecture called the archetype NCCR (at-NCCR). Thus, detecting rearranged NCCRs (rr-NCCRs) JCV in the central nervous system has been viewed as being derived from the archetype and closely linked to PML (4), but the functional consequences of rearrangements are unclear. To address the consequences of the rr-NCCR for JCV gene expression and replication, we characterized the sequences of JCV rr-NCCR from patients with PML and analyzed their effect on viral gene expression and replication with JCV at-NCCR in a bidirectional reporter assay and in recombinant JCV.  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