首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
2.
3.
4.
The positive-strand RNA genome of Japanese encephalitis virus (JEV) terminates in a highly conserved 3′-noncoding region (3′NCR) of six domains (V, X, I, II-1, II-2, and III in the 5′-to-3′ direction). By manipulating the JEV genomic RNA, we have identified important roles for RNA elements present within the 574-nucleotide 3′NCR in viral replication. The two 3′-proximal domains (II-2 and III) were sufficient for RNA replication and virus production, whereas the remaining four (V, X, I, and II-1) were dispensable for RNA replication competence but required for maximal replication efficiency. Surprisingly, a lethal mutant lacking all of the 3′NCR except domain III regained viability through pseudoreversion by duplicating an 83-nucleotide sequence from the 3′-terminal region of the viral open reading frame. Also, two viable mutants displayed severe genetic instability; these two mutants rapidly developed 12 point mutations in domain II-2 in the mutant lacking domains V, X, I, and II-1 and showed the duplication of seven upstream sequences of various sizes at the junction between domains II-1 and II-2 in the mutant lacking domains V, X, and I. In all cases, the introduction of these spontaneous mutations led to an increase in RNA production that paralleled the level of protein accumulation and virus yield. Interestingly, the mutant lacking domains V, X, I, and II-1 was able to replicate in hamster BHK-21 and human neuroblastoma SH-SY5Y cells but not in mosquito C6/36 cells, indicating a cell type-specific restriction of its viral replication. Thus, our findings provide the basis for a detailed map of the 3′ cis-acting elements in JEV genomic RNA, which play an essential role in viral replication. They also provide experimental evidence for the function of 3′ direct repeat sequences and suggest possible mechanisms for the emergence of these sequences in the 3′NCR of JEV and perhaps in other flaviviruses.Japanese encephalitis virus (JEV), a mosquito-borne flavivirus of the family Flaviviridae, is serologically related to several significant human pathogens, including West Nile virus (WNV), Kunjin virus (KUNV), St. Louis encephalitis virus, and Murray Valley encephalitis virus. It is also phylogenetically close to other clinically important human pathogens, including yellow fever virus (YFV) and dengue virus (DENV) (11, 67). JEV is the leading cause of viral encephalitis in Southeast Asia, including China, Japan, Korea, the Philippines, Thailand, and India, and it has begun to expand throughout the Indonesian archipelago and as far as Australia (21, 43). Despite the fact that JEV is generally asymptomatic, ∼50,000 cases are reported annually, and the disease has a mortality rate of ∼25%, mainly in children and young adults (29, 63). Thus, the geographic expansion and clinical importance of JEV infection have drawn increasing attention from the international public health community (44, 71).Like other flaviviruses, JEV is a spherical enveloped virus (∼50 nm diameter) with a single-stranded positive-sense RNA genome that contains a 5′ cap structure but lacks a 3′ polyadenylated tail. Its genomic RNA of ∼11,000 nucleotides (nt) consists of a single long open reading frame (ORF) with two noncoding regions (NCRs) at the 5′ and 3′ ends (41, 84). The ORF is translated into an ∼3,400-amino acid polyprotein precursor, which is co- or posttranslationally cleaved by a cellular protease(s) or a viral protease complex into 10 mature proteins: (i) three structural proteins, the capsid (C), premembrane (prM; which is further processed into pr and M), and envelope (E) proteins; and (ii) seven nonstructural proteins, NS1, NS2A, NS2B, NS3, NS4A, NS4B, and NS5, as arranged in the genome (13, 41, 84). The nonstructural proteins, together with cellular factors, form a viral replicase complex that directs the replication of the genomic RNA in the cytoplasm of the host cell in association with perinuclear membranes (40, 74). For the synthesis of the genomic RNA to take place, this replicase complex must specifically recognize viral cis-acting RNA elements, defined by primary sequences or secondary/tertiary structures. These RNA elements are found in various locations within the genome but most frequently are located in the 5′- and 3′NCRs (23, 47). The identification and characterization of these cis-acting RNA elements is critical for understanding the complete cycle of JEV genome replication.The availability of the complete nucleotide sequence of YFV genomic RNA (57) has led to the identification of three major conserved elements in the 5′- and 3′-terminal regions of the genomic RNA that contain the short primary sequences and secondary structures required for flavivirus RNA replication. (i) Both ends of the genomic RNA terminate with the conserved dinucleotides 5′-AG and CU-3′ (9, 10, 32, 45, 57, 72, 73) in all flaviviruses except an insect cell fusing agent virus (12). Mutations substituting another nucleotide for one of these four nucleotides in KUNV or WNV replicon RNA are known to abolish or compromise RNA replication (35, 69). (ii) A 3′ stem-loop structure (3′SL) has been recognized in all flaviviruses within the ∼90-nt 3′-terminal region of the genomic RNA (9, 45, 57). The structural and functional importance of this 3′SL in RNA replication has been demonstrated in several flaviviruses (9, 18, 49, 50, 61, 70, 82, 86). (iii) The presence of short 5′ and 3′ cyclization sequences (5′CYC and 3′CYC, respectively) in all mosquito-borne flaviviruses suggests that flavivirus genomes can cyclize via 5′-3′ long-range base-pairing interaction, since the 3′CYC upstream of the 3′SL is complementary to the 5′CYC in the 5′ coding region of the C protein (30). The role of these CYC motifs in RNA replication has been well characterized via cell-based assays in many mosquito-borne flaviviruses, including KUNV (34), WNV (42), YFV (8, 14), and DENV (2, 22, 49), and in cell-free systems in the case of WNV (51) and DENV (1, 3, 79, 80). Other RNA elements that have recently been shown to be important for RNA replication in DENV and WNV include an additional pair of complementary sequences (designated 5′- and 3′UARs) that participate in genome cyclization (3, 4, 17, 87) and a 5′ stem-loop structure (designated 5′SLA) present within the 5′NCR that promotes RNA synthesis in association with the 3′NCR (22).In all flaviviruses, the 3′NCR of the genomic RNA is relatively long (∼400 to ∼800 nt), with an array of conserved primary sequences and secondary structures. Although significant progress has been made in identifying cis-acting elements within the 3′NCRs that are essential for RNA replication, most of these elements (i.e., the 3′CYC, 3′SL, and CU-3′) are limited to the ∼100-nt 3′-terminal region that is highly conserved in these viruses (see recent reviews in references 23 and 47). However, the functional importance of the remaining 5′-proximal region of the 3′NCR, which differs in sequence between the various serological groups, is poorly understood. In particular, comparative sequence analyses and genetic algorithm-based computer modeling have suggested that in addition to the well-studied ∼100-nt 3′-proximal region, the remaining ∼474-nt 5′-proximal region of the 574-nt JEV 3′NCR also contains several RNA elements that may play critical roles in the viral life cycle (52, 55, 56, 68). To date, however, experimental evidence for the functional importance of these potential RNA elements in JEV genomic RNA replication is lacking.In the present study, we have identified and characterized the 3′ cis-acting RNA elements within the JEV 3′NCR and shown that they play an essential and/or regulatory role in genomic RNA replication. In particular, we have constructed and functionally characterized genome-length JEV mutant cDNAs with a series of 5′-to-3′ or 3′-to-5′ progressive deletions within the 3′NCR. In addition to identifying particular mutations within this region that affect either the competence or efficiency of genomic RNA replication, we found that the serial passaging of these mutants in susceptible BHK-21 cells produced a large number of pseudorevertants bearing a wide variety of spontaneous point mutations and sequence duplications, some of which were capable of restoring the replication competence of the defective mutants or enhancing replication efficiency. In addition, we assessed the replication of these mutants in three different cell types (BHK-21, SH-SY5Y, and C6/36 cells). Collectively, these data offer new insights into the functional importance of 3′ cis-acting RNA elements that regulate the cell type-dependent replication of JEV and perhaps other closely related mosquito-borne flaviviruses. Our findings also provide experimental evidence for the emergence of functional 3′ direct repeat sequences that are duplicated from the coding region and 3′NCR of JEV genomic RNA.  相似文献   

5.
6.
The herpes simplex virus (HSV) virion host shutoff protein (vhs) encoded by gene UL41 is an mRNA-specific RNase that triggers accelerated degradation of host and viral mRNAs in infected cells. We report here that vhs is also able to modulate reporter gene expression without greatly altering the levels of the target mRNA in transient-transfection assays conducted in HeLa cells. We monitored the effects of vhs on a panel of bicistronic reporter constructs bearing a variety of internal ribosome entry sites (IRESs) located between two test cistrons. As expected, vhs inhibited the expression of the 5′ cistrons of all of these constructs; however, the response of the 3′ cistron varied with the IRES: expression driven from the wild-type EMCV IRES was strongly suppressed, while expression controlled by a mutant EMCV IRES and the cellular ApaF1, BiP, and DAP5 IRES elements was strongly activated. In addition, several HSV type 1 (HSV-1) 5′ untranslated region (5′ UTR) sequences also served as positive vhs response elements in this assay. IRES activation was also observed in 293 and HepG2 cells, but no such response was observed in Vero cells. Mutational analysis has yet to uncouple the ability of vhs to activate 3′ cistron expression from its shutoff activity. Remarkably, repression of 5′ cistron expression could be observed under conditions where the levels of the reporter RNA were not correspondingly reduced. These data provide strong evidence that vhs can modulate gene expression at the level of translation and that it is able to activate cap-independent translation through specific cis-acting elements.The virion host shutoff protein (vhs) encoded by herpes simplex virus (HSV) gene UL41 is an endoribonuclease that is packaged into the tegument of mature HSV virions. Once delivered into the cytoplasm of newly infected cells, vhs triggers shutoff of host protein synthesis, disruption of preexisting polysomes, and degradation of host mRNAs (reviewed in reference 62). The vhs-dependent shutoff system destabilizes many cellular and viral mRNAs (36, 46, 67). The rapid decline in host mRNA levels presumably helps viral mRNAs gain access to the cellular translational apparatus. In addition, the relatively short half-lives of viral mRNAs contribute to the sharp transitions between the successive phases of viral protein synthesis by tightly coupling changes in the rates of synthesis of viral mRNAs to altered mRNA levels (46). These effects enhance virus replication and may account for the modest reduction in virus yield displayed by vhs mutants in cultured Vero cells (55, 61).vhs also plays a critical role in HSV pathogenesis: vhs mutants are severely impaired for replication in the corneas and central nervous systems of mice and cannot efficiently establish or reactivate from latency (63, 65, 66). Mounting evidence indicates that this attenuation stems at least in part from an impaired ability to disarm elements of the innate and adaptive host immune responses (reviewed in reference 62). For example, vhs suppresses certain innate cellular antiviral responses, including production of proinflammatory cytokines and chemokines (68); dampens the type I interferon system (11, 45, 49, 78); and blocks activation of dendritic cells (58). Moreover, vhs mutants display enhanced virulence in knockout mice lacking type I interferon (IFN) receptors (37, 45) or Stat1 (48) and are hypersensitive to the antiviral effects of IFN in some cells in tissue culture (11, 49, 68). Thus, vhs is arguably a bona fide virulence factor.vhs present in extracts of HSV virions or purified from bacteria has nonspecific RNase activity capable of degrading all RNA substrates (15, 70, 71, 79). However, vhs is highly selective in vivo, targeting mRNAs and sparing other cytoplasmic RNAs (36, 46). In vivo and in mammalian whole-cell extracts, vhs-induced decay of at least some mRNAs initiates near regions of translation initiation and proceeds in an overall 5′-to-3′ direction (12, 13, 29, 52). Moreover, vhs binds to the translation initiation factors eIF4H, eIF4B, and eIF4A II, all components of the cap recognition factor eIF4F (10, 16, 17). Thus, it has been proposed that vhs selectively targets actively translated mRNAs through interactions with eIF4F components (17). Consistent with this hypothesis, recent data document that eIF4H is required for vhs activity in vivo (59).A previous report from this laboratory documented that the internal ribosome entry sites (IRESs) of the picornaviruses poliovirus and encephalomyocarditis virus (EMCV) strongly target vhs-induced RNA cleavage events to sequences immediately 3′ to the IRES in an in vitro translation system derived from rabbit reticulocyte lysates (RRL) (13). IRES elements are highly structured RNA sequences that are able to direct cap-independent translational initiation (reviewed in references 21, 25, 30, and 64). In the case of the poliovirus and EMCV elements, this is achieved by directly recruiting the eIF4F scaffolding protein eIF4G, thus bypassing the requirement for the cap-binding eIF4F subunit, eIF4E (reviewed in reference 30). Based on these data, we suggested that vhs is strongly targeted to the picornavirus IRES elements via interactions with eIF4 factors.A growing number of cellular mRNAs have been proposed to bear IRES elements in their 5′ untranslated regions (5′ UTRs). These include many that are involved in cellular stress responses, apoptosis, and cell cycle progression (24, 64, 74). Given the striking ability of picornavirus IRES elements to target vhs RNase activity in vitro, we asked whether viral and cellular IRES elements are able to modify the susceptibility of mRNAs to vhs in vivo. During the course of preliminary experiments designed to test this hypothesis, we unexpectedly discovered that vhs is able to strongly activate gene expression controlled by some cellular IRES elements and HSV 5′ UTR sequences in in vivo bicistronic reporter assays. These observations are the subject of the present report.  相似文献   

7.
8.
9.
10.
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   

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

12.
13.
The human immunodeficiency virus type 1 (HIV-1) matrix (MA) protein targets HIV-1 precursor Gag (PrGag) proteins to assembly sites at plasma membrane (PM) sites that are enriched in cholesterol and phosphatidylinositol-(4,5)-bisphosphate [PI(4,5)P2]. MA is myristoylated, which enhances membrane binding, and specifically binds PI(4,5)P2 through headgroup and 2′ acyl chain contacts. MA also binds nucleic acids, although the significance of this association with regard to the viral life cycle is unclear. We have devised a novel MA binding assay and used it to examine MA interactions with membranes and nucleic acids. Our results indicate that cholesterol increases the selectivity of MA for PI(4,5)P2-containing membranes, that PI(4,5)P2 binding tolerates 2′ acyl chain variation, and that the MA myristate enhances membrane binding efficiency but not selectivity. We also observed that soluble PI(4,5)P2 analogues do not compete effectively with PI(4,5)P2-containing liposomes for MA binding but surprisingly do increase nonspecific binding to liposomes. Finally, we have demonstrated that PI(4,5)P2-containing liposomes successfully outcompete nucleic acids for MA binding, whereas other liposomes do not. These results support a model in which RNA binding protects MA from associating with inappropriate cellular membranes prior to PrGag delivery to PM assembly sites.The matrix (MA) domain of the human immunodeficiency virus type 1 (HIV-1) precursor Gag (PrGag) protein serves several functions in the viral replication cycle. One essential function is to target PrGag proteins to their assembly sites at the plasma membranes (PMs) of infected cells (4, 5, 11, 16, 25, 29, 30, 33, 35, 39, 43-45, 47, 50, 54, 56, 57). A second function is the recruitment of the viral surface/transmembrane (SU/TM; also referred to as gp120/gp41) envelope (Env) protein complex into virions (14, 15, 18, 19, 27, 51-53). In addition to these activities, numerous reports have attributed nucleic acid binding properties to retroviral MAs (24, 38, 47), and with some viruses MA appears to serve in an encapsidation capacity (24). While no encapsidation role has been assigned for HIV-1 MA, experiments have shown that MA can substitute for the HIV-1 nucleocapsid (NC) protein assembly function (38) under some circumstances, presumably by virtue of its facility to concentrate PrGag proteins by binding them to RNAs (38).A number of structural studies have been conducted on HIV-1 MA (1, 22, 41, 42, 49). The protein is N terminally myristoylated and composed of six α helices, capped by a three-strand β sheet (7, 22, 41, 42, 49). The protein trimerizes in solution and in crystals (22, 28, 49) and recently has been shown to organize as hexamers of trimers on lipid membranes (1). The membrane binding face of HIV-1 MA is basic, fostering its ability to associate with negatively charged phospholipid headgroups (1, 22, 30, 41, 42, 49). The importance of such an interaction has been underscored in molecular genetic experiments which demonstrated that depletion of PM phosphatidylinositol-(4,5)-bisphosphate [PI(4,5)P2] reduced the assembly efficiency of HIV-1 (9, 36). Consistent with these observations, HIV-1 MA preferentially binds to soluble PI(4,5)P2 mimics through contacts with the headgroup and 2′ acyl chain, and binding promotes exposure of the MA myristate group and protein oligomerization (17, 21, 40-43, 46). However, PI(4,5)P2 is not the only lipid to demonstrate an association with HIV-1. In particular, HIV-1 appears to assemble at cholesterol-rich PM sites, cholesterol is highly enriched in HIV-1 virions, and cholesterol depletion reduces viral infectivity (2, 6, 8, 20, 23, 26, 31, 34, 37). The HIV-1 lipidome shows additional differences from the PM lipids of infected cells (2, 5, 8), suggesting that other lipids could affect PrGag-membrane binding or virus assembly site selection.To gain a better understanding of the functions and interactions of HIV-1 MA, we have examined the liposome and nucleic acid binding properties of purified myristoylated MA. Using liposome flotation assays and a novel liposome bead binding assay, we have demonstrated that the PI(4,5)P2 binding specificity of MA is enhanced by cholesterol, that protein myristoylation increases membrane binding efficiency but not specificity, and that 2′ acyl chain variation is compatible with PI(4,5)P2 binding. We also examined whether soluble PI(4,5)P2 mimics could compete with liposomes for MA binding. Surprisingly, we found that soluble mimics not only failed to compete with PI(4,5)P2 liposomes but also increased MA binding to membranes that do not contain acidic phospholipids. Finally, we have observed that while MA does bind nucleic acids, nucleic acid binding is outcompeted by PI(4,5)P2-containing liposomes. Our results suggest models for PrGag-membrane and RNA association and the HIV-1 assembly pathway.  相似文献   

14.
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   

15.
16.
Borna disease virus (BDV), the prototypic member of the family Bornaviridae within the order Mononegavirales, provides an important model for the investigation of viral persistence within the central nervous system (CNS) and of associated brain disorders. BDV is highly neurotropic and enters its target cell via receptor-mediated endocytosis, a process mediated by the virus surface glycoprotein (G), but the cellular factors and pathways determining BDV cell tropism within the CNS remain mostly unknown. Cholesterol has been shown to influence viral infections via its effects on different viral processes, including replication, budding, and cell entry. In this work, we show that cell entry, but not replication and gene expression, of BDV was drastically inhibited by depletion of cellular cholesterol levels. BDV G-mediated attachment to BDV-susceptible cells was cholesterol independent, but G localized to lipid rafts (LR) at the plasma membrane. LR structure and function critically depend on cholesterol, and hence, compromised structural integrity and function of LR caused by cholesterol depletion likely inhibited the initial stages of BDV cell internalization. Furthermore, we also show that viral-envelope cholesterol is required for BDV infectivity.Borna disease virus (BDV) is an enveloped virus with a nonsegmented negative-strand RNA genome whose organization (3′-N-p10/P-M-G-L-5′) is characteristic of mononegaviruses (6, 28, 46, 48). However, based on its unique genetics and biological features, BDV is considered to be the prototypic member of a new virus family, Bornaviridae, within the order Mononegavirales (8, 28, 46, 49).BDV can infect a variety of cell types in cell culture but in vivo exhibits exquisite neurotropism and causes central nervous system (CNS) disease in different vertebrate species, which is frequently manifested in behavioral abnormalities (19, 33, 44, 53). Both host and viral factors contribute to a variable period of incubation and heterogeneity in the symptoms and pathology associated with BDV infection (14, 16, 29, 42, 44). BDV provides an important model for the investigation of both immune-mediated pathological events associated with virus-induced neurological disease and mechanisms whereby noncytolytic viruses induce neurodevelopmental and behavioral disturbances in the absence of inflammation (15, 18, 41). Moreover, serological data and molecular epidemiological studies suggest that BDV, or a BDV-like virus, can infect humans and that it might be associated with certain neuropsychiatric disorders (17, 24), which further underscores the interest in understanding the mechanisms underlying BDV persistence in the CNS and its effect on brain cell functions. The achievement of these goals will require the elucidation of the determinants of BDV cell tropism within the CNS.BDV enters its target cell via receptor-mediated endocytosis, a process in which the BDV G protein plays a central role (1, 5, 13, 14, 39). Cleavage of BDV G by the cellular protease furin generates two functional subunits: GP1 (GPN), involved in virus interaction with a yet-unidentified cell surface receptor (1, 39), and GP2 (GPC), which mediates a pH-dependent fusion event between viral and cellular membranes (13). However, a detailed characterization of cellular factors and pathways involved in BDV cell entry remains to be done.Besides cell surface molecules that serve as viral receptors, many other cell factors, including nonproteinaceous molecules, can influence cell entry by virus (52). In this regard, cholesterol, which plays a critical role in cellular homeostasis (55), has also been identified as a key factor required for productive infection by different viruses. Accordingly, cholesterol participates in a variety of processes in virus-infected cells, including fusion events between viral and cellular membranes (3), viral replication (23), and budding (35, 37), as well as maintenance of lipid rafts (LR) (12) as scaffold structures where the viral receptor and coreceptor associate (11, 26, 32, 36). LR are specialized microdomains within cellular membranes constituted principally of proteins, sphingolipids, and cholesterol. LR facilitate the close proximity and interaction of specific sets of proteins and contribute to different processes associated with virus multiplication (38). Cholesterol can also influence virus infection by contributing to the maintenance of the properties of the viral envelope required for virus particle infectivity (21, 54). Here, we show for the first time that cholesterol plays a critical role in BDV infection. Depletion of cellular cholesterol prior to, but not after, BDV cell entry prevented productive BDV infection, likely due to disruption of plasma membrane LR that appear to be the cell entry point for BDV. In addition, we document that cholesterol also plays an essential role in the properties of the BDV envelope required for virus particle infectivity.  相似文献   

17.
18.
19.
20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

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