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.
The molecular mechanisms associated with rabies virus (RV) virulence are not fully understood. In this study, the RV Flury low-egg-passage (LEP) and high-egg-passage (HEP) strains were used as models to explore the attenuation mechanism of RV. The results of our studies confirmed that the R333Q mutation in the glycoprotein (GR333Q) is crucial for the attenuation of Flury RV in mice. The R333Q mutation is stably maintained in the HEP genome background but not in the LEP genome background during replication in mouse brain tissue or cell culture. Further investigation using chimeric viruses revealed that the polymerase L gene determines the genetic stability of the GR333Q mutation during replication. Moreover, a recombinant RV containing the LEP G protein with the R333Q mutation and the HEP L gene showed significant attenuation, genetic stability, enhancement of apoptosis, and immunogenicity. These results indicate that attenuation of the RV Flury strain results from the coevolution of G and L elements and provide important information for the generation of safer and more effective modified live rabies vaccine.Rabies virus (RV) belongs to the genus Lyssavirus of the family Rhabdoviridae and causes a fatal neurological disease in humans and animals (6). The RV genome is a nonsegmented negative-strand (NNS) RNA encoding five structural proteins: nucleoprotein (N), phosphoprotein (P), matrix protein (M), glycoprotein (G), and large polymerase (L). Among these, the G protein is a major contributor to RV pathogenicity (7, 31, 33). The G protein facilitates fast virus entry and transsynaptic spread and regulates the rate of virus replication, together with other viral elements (8, 30, 39). The G protein of nonpathogenic RV strains can trigger apoptosis, while the RV G of pathogenic strains induces less or no apoptosis (35, 59). The amino acid residue at position 333 of the G protein (G333) of some fixed strains has been shown to be an important determinant of virulence in adult mice (5). Strains that have arginine or lysine at position G333 kill adult mice, whereas mutants with other amino acids at this site cause a nonlethal infection (1, 5, 25, 36, 49, 53). However, the pathogenicity of RV strains is not solely determined by substitutions at the G333 position. Other substitutions in the G protein, such as N194K, have also been shown to affect viral pathogenicity in mice (10, 21, 50). In addition, other viral elements, such as the N, P, M, and L genes, the trailer sequence in the noncoding region, and the pseudogene, were also reported to modulate RV pathogenicity (12, 46, 57, 58). How these viral elements regulate the pathogenicity of RV remains to be fully explored, and further investigation is needed to understand the molecular basis of RV pathogenicity.Attenuated Flury RV low-egg-passage (LEP) and high-egg-passage (HEP) strains were established through serial passage in chicken brain, chicken embryos, and culture cells using a Flury RV isolated from a girl who died of rabies (23, 24). LEP has Arg at position G333 and kills adult mice after intracerebral (i.c.) inoculation, while HEP has Gln at G333 and causes only mild signs in adult mice. It has been demonstrated that HEP could regain lethality in adult mice by a single amino acid change at G333 from Gln to Arg (49), which indicated that Arg at position G333 is a key determinant of pathogenicity of Flury RV in adult mice. However, whether the Arg at G333 is indispensable for the lethal phenotype of LEP has not been demonstrated.In the current study, LEP and HEP Flury RV strains were used as models to investigate the mechanism of attenuation. We found that both G and L contribute to the attenuation of Flury RV. Substitution of Arg with Gln at G333 (GR333Q) eliminated LEP neuroinvasiveness but not the virus'' lethal phenotype in adult mice after i.c. inoculation. The GR333Q mutation could be kept stable only in the genome background of HEP but not in that of LEP during replication. The L gene contributes to the attenuation and enhanced immunogenicity of Flury RV by promoting the stabilization of the GR333Q mutation during virus replication in brain tissues or cells.  相似文献   

2.
3.
4.
Although few simian rotaviruses (RVs) have been isolated, such strains have been important for basic research and vaccine development. To explore the origins of simian RVs, the complete genome sequences of strains PTRV (G8P[1]), RRV (G3P[3]), and TUCH (G3P[24]) were determined. These data allowed the genotype constellations of each virus to be determined and the phylogenetic relationships of the simian strains with each other and with nonsimian RVs to be elucidated. The results indicate that PTRV was likely transmitted from a bovine or other ruminant into pig-tailed macaques (its host of origin), since its genes have genotypes and encode outer-capsid proteins similar to those of bovine RVs. In contrast, most of the genes of rhesus-macaque strains, RRV and TUCH, have genotypes more typical of canine-feline RVs. However, the sequences of the canine and/or feline (canine/feline)-like genes of RRV and TUCH are only distantly related to those of modern canine/feline RVs, indicating that any potential transmission of a progenitor of these viruses from a canine/feline host to a simian host was not recent. The remaining genes of RRV and TUCH appear to have originated through reassortment with bovine, human, or other RV strains. Finally, comparison of PTRV, RRV, and TUCH genes with those of the vervet-monkey RV SA11-H96 (G3P[2]) indicates that SA11-H96 shares little genetic similarity to other simian strains and likely has evolved independently. Collectively, our data indicate that simian RVs are of diverse ancestry with genome constellations that originated largely by interspecies transmission and reassortment with nonhuman animal RVs.Group A rotaviruses (RVs) are a major cause of acute dehydrating diarrhea in infants and children under the age of 5 years worldwide. These infections lead to approximately 527,000 deaths each year, the vast majority occurring in developing countries (33). RVs are also responsible for gastroenteritis in many other animal species, notably mammals and birds (16, 38). RVs are members of the family Reoviridae and possess a genome consisting of 11 segments of double-stranded RNA (dsRNA). The prototypic genome of a group A RV encodes six structural proteins (VP) and six nonstructural proteins (NSP) (5). The mature RV virion is a nonenveloped triple-layered icosahedral particle. The inner most protein layer is formed by the core lattice protein VP2. Attached to the interior surface of the VP2 layer near the fivefold axes are complexes of the viral RNA-dependent RNA polymerase VP1 and the RNA capping enzyme VP3. Collectively, VP1, VP2, VP3, and the dsRNA genome form the core of the virion (5, 11). The core is surrounded by VP6, the sole constituent of the intermediate protein layer of the virion. The antigenic properties of VP6 are used in classifying RV isolates into groups. The outer protein layer of the virion is composed of trimers of the VP7 glycoprotein penetrated by spikes of the VP4 attachment protein (50). The properties of VP7 and VP4 form the basis of a dual classification system defining RV G types (glycosylated) and P types (protease sensitive), respectively. At present, 23 G genotypes and 31 P genotypes have been recognized in the literature based on sequence analyses (17, 39, 42, 45, 47). Recently, a comprehensive sequence-based classification system was established for the RVs which, together with a uniform nomenclature, allows each genome segment of the virus to be assigned to a particular genotype. In the comprehensive classification system, the acronym Gx-P[x]-Ix-Rx-Cx-Mx-Ax-Nx-Tx-Ex-Hx defines the genotypes of VP7-VP4-VP6-VP1-VP2-VP3-NSP1-NSP2-NSP3-NSP4-NSP5 encoding genome segments (17, 18).Several years ago, Nakagomi et al. provided evidence by RNA-RNA hybridization assays that RVs originating from different animal species could be resolved into genogroups based upon the existence of unique species-specific genome constellations (29-31). More recently, the concept that RVs preferentially retain certain species-related genome constellations has been further supported by whole-genome sequencing (8, 24). For human RVs, two major genogroups (Wa-like genogroup 1 and DS-1-like genogroup 2) and one minor genogroup (AU-1-like genogroup 3) have been described (8, 17, 30). Although these genogroups are generally species specific, it is believed that the human AU-1 genogroup is of feline origin (31) and that the human Wa and DS-1 genogroups share common ancestor with porcine and bovine RVs, respectively (17). Another recent study based on full genome sequence data has indicated that the rarely seen human G3P[3] RVs are of feline or canine origin (46). Two additional sequence-based studies have indicated that human RVs with P[14] specificity may have originated after interspecies transmission from rabbit RVs and RVs from hosts belonging to the order Artiodactyla (i.e., hoofed mammals with even toes, including ruminants and pigs) (19, 20). These examples indicate that interspecies transmission of entire RV gene constellations from one host species to another may contribute significantly to viral evolution. In addition to interspecies transmission, complete genome sequencing of RVs have revealed multiple examples of naturally occurring inter- and intragenogroup reassortment (17, 19, 21-23, 37, 41).The simian RV strains, notably RRV and the SA11 derivatives (e.g., SA11-Cl3 and SA11-4F), have been used extensively as models in the study of all aspects of RV biology, including characterizing genome replication and virion assembly, delineating high-resolution structures of viral proteins and the virion capsid, and describing the functions of viral proteins. Moreover, the RRV strain was used to create a set of human-simian reassortant viruses that formed the basis of the first commercially licensed RV vaccine (Rotashield; Wyeth Laboratories) (10). Serological analyses have indicated that simian RVs are probably endemic in wild nonhuman primate (NHP) species in Africa (32). However, whether or not unique genogroups or preferred genome constellation exist for the simian RVs has not been determined, because of the lack of comprehensive genetic data. Most simian RVs isolated to date (e.g., rhesus macaque viruses RRV [43] and TUCH [25] and the pig-tailed macaque virus PTRV [9]) have been recovered from monkeys kept in captivity in the United States. An important exception is the SA11 isolate, which was recovered from a vervet monkey in South Africa (15). Simian RV infections occur mostly in young monkeys, similar to human RV infections in children (32, 40).To gain further insight into the origins and properties of simian RVs, we sequenced and contrasted the genomes of PTRV, RRV, and TUCH with other RVs, including SA11-H96 (G3P[2]), the only previously fully sequenced simian RV (41). Our results reveal that these four simian RVs are of divergent ancestry and have evolved by combinations of interspecies transmission and reassortment with RVs naturally occurring in other animal species. Thus, the simian RVs do not possess a common genome constellation nor define a unique genogroup. Although frequently used as disease models, the simian RVs show limited genetic similarity with the human RVs (genogroups 1 and 2) responsible for most human disease.  相似文献   

5.
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   

6.
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   

7.
8.
Recombinant rabies virus (RV)-based vectors have demonstrated their efficacy in generating long-term, antigen-specific immune responses in murine and monkey models. However, replication-competent viral vectors pose significant safety concerns due to vector pathogenicity. RV pathogenicity is largely attributed to its glycoprotein (RV-G), which facilitates the attachment and entry of RV into host cells. We have developed a live, single-cycle RV by deletion of the G gene from an RV vaccine vector expressing HIV-1 Gag (SPBN-ΔG-Gag). Passage of SPBN-ΔG-Gag on cells stably expressing RV-G allowed efficient propagation of the G-deleted RV. The in vivo immunogenicity data comparing single-cycle RV to a replication-competent control (BNSP-Gag) showed lower RV-specific antibodies; however, the overall isotype profiles (IgG2a/IgG1) were similar for the two vaccine vectors. Despite this difference, mice immunized with SPBN-ΔG-Gag and BNSP-Gag mounted similar levels of Gag-specific CD8+ T-cell responses as measured by major histocompatibility complex class I Gag-tetramer staining, gamma interferon-enzyme-linked immunospot assay, and cytotoxic T-cell assay. Moreover, these cellular responses were maintained equally at immunization titers as low as 103 focus-forming units for both RV vaccine vectors. CD8+ T-cell responses were significantly enhanced by a boost with a single-cycle RV complemented with a heterologous vesicular stomatitis virus glycoprotein. These findings demonstrate that single-cycle RV is an effective alternative to replication-competent RV vectors for future development of vaccines for HIV-1 and other infectious diseases.The global spread of HIV-1 represents one of the most significant pandemics to afflict humans (22). Despite tremendous efforts to increase HIV awareness in the general population, UNAIDS reports that fewer than one in five people has access to HIV prevention strategies and many are subject to cultural stigmas thwarting such efforts (43). As such, an HIV vaccine is paramount for preventing disease transmission. It is not yet clear precisely what characteristics are critical for an effective HIV vaccine, yet evidence suggests one would need to induce both antibody and CD8+ T-cell-mediated immunity (reviewed in reference 25). Live viruses are at the forefront of HIV vaccine development (7) because they are powerful inducers of both of these arms of immunity. We previously demonstrated that replication-competent rabies virus (RV)-based vectors can induce long-lasting antigen-specific immune responses in both murine and monkey models, as well as protect rhesus macaques from an AIDS-like disease (23, 24, 26-29, 42). However, there are safety concerns with the use of any replication-competent virus for widespread immunization. To address this, we sought to develop and evaluate the immunogenicity of a safer alternative: a single-cycle RV expressing HIV-1 Gag as a model antigen.Single-cycle viral vectors are defective in certain viral components that are required for infectious particle assembly (reviewed in reference 12). As such, the virus undergoes one complete cycle of replication in the primary infected cell and produces progeny virions that are unable to spread to a second round of cells. The progeny are noninfectious and provide inert antigen that may or may not be immunogenic (12). In contrast, so-called replication-deficient viruses do not complete that initial round of replication. These two attenuation strategies have been adopted for use with many different viruses including, but not limited to, adenovirus (Ad), vaccinia virus (VV), canarypox virus (CPV), herpes simplex virus (HSV), vesicular stomatitis virus (VSV), and, more recently, RV (4, 6, 9, 18, 21, 33, 35, 36, 38). However, the results regarding the immunogenicity of such vectors are mixed. For example, both the replication-deficient Ad5 vector and modified vaccinia Ankara (MVA) showed reduced humoral and cellular immunogenicity compared to their replication-competent counterparts, but the use of higher titers and multiple immunizations did increase such responses (18, 33, 35). In the case of CPV, the replication-deficient vector provided poor HIV-specific cellular responses, causing the termination of phase II HIV-1 vaccine trials (38). In contrast, single-cycle VSV, a rhabdovirus closely related to RV, has been shown to induce HIV-1 Env-specific CD8+ T-cell responses equivalent to full-length VSV when administered intramuscularly (36). However, protection of rhesus macaques against highly pathogenic simian immunodeficiency virus (SIV) challenge by both replication-competent and single-cycle VSV needs to be shown.In the study described here, we generated a single-cycle RV vector expressing HIV-1 Gag (SPBN-ΔG-Gag) by deletion of the entire RV glycoprotein (RV-G) from the RV genome. RV-G was chosen due to its critical role in the attachment and entry of RV into host cells, which makes RV-G one of the most important determinants of viral pathogenicity (10, 11, 37). RV particles lacking G are unable to spread, as evidenced by intracranial infection with a G-deleted RV that remains restricted to the primary infected neurons (13, 44). It must be noted that in the absence of RV-G, virions are still capable of budding though at a 30-fold lower efficiency (32). These virions, however, are incapable of attachment and entry into a secondary host cell. Because of this, SPBN-ΔG-Gag was propagated on a trans-complementing cell line induced to express RV-G (or VSV-RV-G), effectively facilitating virus spread. To evaluate the immunogenicity of the single-cycle vector, we immunized mice and compared the humoral and cellular responses to responses generated by replication-competent RV. Our results indicate that single-cycle RV generates reduced vector-specific antibody responses but similar HIV-1 Gag-specific CD8+ T-cell responses. Moreover, these responses can be significantly enhanced by a heterologous boost with a single-cycle RV complemented with a VSV glycoprotein. Taken together, the results presented here show evidence that single-cycle RV is a promising platform for a safe, live viral vaccine for use against HIV-1 and other applications.  相似文献   

9.
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   

10.
11.
12.
13.
14.
15.
Analysis of Lyme borreliosis (LB) spirochetes, using a novel multilocus sequence analysis scheme, revealed that OspA serotype 4 strains (a rodent-associated ecotype) of Borrelia garinii were sufficiently genetically distinct from bird-associated B. garinii strains to deserve species status. We suggest that OspA serotype 4 strains be raised to species status and named Borrelia bavariensis sp. nov. The rooted phylogenetic trees provide novel insights into the evolutionary history of LB spirochetes.Multilocus sequence typing (MLST) and multilocus sequence analysis (MLSA) have been shown to be powerful and pragmatic molecular methods for typing large numbers of microbial strains for population genetics studies, delineation of species, and assignment of strains to defined bacterial species (4, 13, 27, 40, 44). To date, MLST/MLSA schemes have been applied only to a few vector-borne microbial populations (1, 6, 30, 37, 40, 41, 47).Lyme borreliosis (LB) spirochetes comprise a diverse group of zoonotic bacteria which are transmitted among vertebrate hosts by ixodid (hard) ticks. The most common agents of human LB are Borrelia burgdorferi (sensu stricto), Borrelia afzelii, Borrelia garinii, Borrelia lusitaniae, and Borrelia spielmanii (7, 8, 12, 35). To date, 15 species have been named within the group of LB spirochetes (6, 31, 32, 37, 38, 41). While several of these LB species have been delineated using whole DNA-DNA hybridization (3, 20, 33), most ecological or epidemiological studies have been using single loci (5, 9-11, 29, 34, 36, 38, 42, 51, 53). Although some of these loci have been convenient for species assignment of strains or to address particular epidemiological questions, they may be unsuitable to resolve evolutionary relationships among LB species, because it is not possible to define any outgroup. For example, both the 5S-23S intergenic spacer (5S-23S IGS) and the gene encoding the outer surface protein A (ospA) are present only in LB spirochete genomes (36, 43). The advantage of using appropriate housekeeping genes of LB group spirochetes is that phylogenetic trees can be rooted with sequences of relapsing fever spirochetes. This renders the data amenable to detailed evolutionary studies of LB spirochetes.LB group spirochetes differ remarkably in their patterns and levels of host association, which are likely to affect their population structures (22, 24, 46, 48). Of the three main Eurasian Borrelia species, B. afzelii is adapted to rodents, whereas B. valaisiana and most strains of B. garinii are maintained by birds (12, 15, 16, 23, 26, 45). However, B. garinii OspA serotype 4 strains in Europe have been shown to be transmitted by rodents (17, 18) and, therefore, constitute a distinct ecotype within B. garinii. These strains have also been associated with high pathogenicity in humans, and their finer-scale geographical distribution seems highly focal (10, 34, 52, 53).In this study, we analyzed the intra- and interspecific phylogenetic relationships of B. burgdorferi, B. afzelii, B. garinii, B. valaisiana, B. lusitaniae, B. bissettii, and B. spielmanii by means of a novel MLSA scheme based on chromosomal housekeeping genes (30, 48).  相似文献   

16.
Clade B of the New World arenaviruses contains both pathogenic and nonpathogenic members, whose surface glycoproteins (GPs) are characterized by different abilities to use the human transferrin receptor type 1 (hTfR1) protein as a receptor. Using closely related pairs of pathogenic and nonpathogenic viruses, we investigated the determinants of the GP1 subunit that confer these different characteristics. We identified a central region (residues 85 to 221) in the Guanarito virus GP1 that was sufficient to interact with hTfR1, with residues 159 to 221 being essential. The recently solved structure of part of the Machupo virus GP1 suggests an explanation for these requirements.Arenaviruses are bisegmented, single-stranded RNA viruses that use an ambisense coding strategy to express four proteins: NP (nucleoprotein), Z (matrix protein), L (polymerase), and GP (glycoprotein). The viral GP is sufficient to direct entry into host cells, and retroviral vectors pseudotyped with GP recapitulate the entry pathway of these viruses (5, 13, 24, 31). GP is a class I fusion protein comprising two subunits, GP1 and GP2, cleaved from the precursor protein GPC (4, 14, 16, 18, 21). GP1 contains the receptor binding domain (19, 28), while GP2 contains structural elements characteristic of viral membrane fusion proteins (8, 18, 20, 38). The N-terminal stable signal peptide (SSP) remains associated with the mature glycoprotein after cleavage (2, 39) and plays a role in transport, maturation, and pH-dependent fusion (17, 35, 36, 37).The New World arenaviruses are divided into clades A, B, and C based on phylogenetic relatedness (7, 9, 11). Clade B contains the human pathogenic viruses Junin (JUNV), Machupo (MACV), Guanarito (GTOV), Sabia, and Chapare, which cause severe hemorrhagic fevers in South America (1, 10, 15, 26, 34). Clade B also contains the nonpathogenic viruses Amapari (AMAV), Cupixi, and Tacaribe (TCRV), although mild disease has been reported for a laboratory worker infected with TCRV (29).Studies with both viruses and GP-pseudotyped retroviral vectors have shown that the pathogenic clade B arenaviruses use the human transferrin receptor type 1 (hTfR1) to gain entry into human cells (19, 30). In contrast, GPs from nonpathogenic viruses, although capable of using TfR1 orthologs from other species (1), cannot use hTfR1 (1, 19) and instead enter human cells through as-yet-uncharacterized hTfR1-independent pathways (19). In addition, human T-cell lines serve as useful tools to distinguish these GPs, since JUNV, GTOV, and MACV pseudotyped vectors readily transduce CEM cells, while TCRV and AMAV GP vectors do not (27; also unpublished data). These properties of the GPs do not necessarily reflect a tropism of the pathogenic viruses for human T cells, since viral tropism is influenced by many factors and T cells are not a target for JUNV replication in vivo (3, 22, 25).  相似文献   

17.
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   

18.
In most cilia, the axoneme can be subdivided into three segments: proximal (the transition zone), middle (with outer doublet microtubules), and distal (with singlet extensions of outer doublet microtubules). How the functionally distinct segments of the axoneme are assembled and maintained is not well understood. DYF-1 is a highly conserved ciliary protein containing tetratricopeptide repeats. In Caenorhabditis elegans, DYF-1 is specifically needed for assembly of the distal segment (G. Ou, O. E. Blacque, J. J. Snow, M. R. Leroux, and J. M. Scholey. Nature. 436:583-587, 2005). We show that Tetrahymena cells lacking an ortholog of DYF-1, Dyf1p, can assemble only extremely short axoneme remnants that have structural defects of diverse natures, including the absence of central pair and outer doublet microtubules and incomplete or absent B tubules on the outer microtubules. Thus, in Tetrahymena, DYF-1 is needed for either assembly or stability of the entire axoneme. Our observations support the conserved function for DYF-1 in axoneme assembly or stability but also show that the consequences of loss of DYF-1 for axoneme segments are organism specific.Cilia are microtubule-rich cellular extensions that arise from basal bodies near the surfaces of most eukaryotic cell types. Defective cilia cause a wide variety of diseases, including polycystic kidney disease, primary ciliary dyskinesia, and retinal degeneration (3). A typical motile cilium has a microtubule-based framework, the axoneme, which contains nine outer (mostly doublet) microtubules and two central (singlet) microtubules. In most cilia, the axoneme can be subdivided into three segments: proximal (transition zone), middle (containing outer doublet microtubules), and distal (containing singlet extensions of peripheral microtubules). The outer doublet microtubules of the middle segment have a complete tubule A made of 13 protofilaments and an incomplete tubule B made of 11 protofilaments that is fused to the wall of the A tubule (36, 57). The outer microtubules in the distal segment lack the B tubule (32, 49). The distal segment also lacks dynein arms and radial spokes, and its microtubules are terminated by caps that are associated with the plasma membranes at the tips of cilia (11, 50). The distal segments are characterized by a high level of microtubule turnover, which could play a role in the regulation of the length of cilia (31).The mechanisms that establish the segmental subdivision of the axoneme are not well understood. Studies of Caenorhabditis elegans indicate that the distal segment is assembled using a mechanism that differs from the one utilized in the middle and proximal segments (54). In most cell types, ciliogenesis is dependent on the intraflagellar transport (IFT) pathway, a bidirectional motility of protein aggregates, known as IFT particles, that occurs along outer microtubules (10, 28, 29, 42). IFT particles are believed to provide platforms for transport of axonemal precursors (23, 44). The anterograde component of IFT that delivers cargo from the cell body to the tips of cilia is carried out by kinesin-2 motors (28, 63), whereas the cytoplasmic dynein DHC1b is responsible for the retrograde IFT (41, 43, 53). Importantly, in the well-studied amphid cilia of C. elegans, two distinct kinesin-2 complexes are involved in the anterograde IFT and differ in movement velocity: the “slow” heterotrimeric kinesin-II and the “fast” homodimeric OSM-3 kinesin (54). While kinesin-II and OSM-3 work redundantly to assemble the middle segment, OSM-3 alone functions in the assembly of the distal segment (39, 56).In C. elegans, DYF-1 is specifically required for assembly of the distal segment (39). In the DYF-1 mutant, the rate of IFT in the remaining middle segment is reduced to the level of the slow kinesin-II, suggesting that the Osm3 complex is nonfunctional and that kinesin-II functions alone in the middle segment. Thus, DYF-1 could either activate OSM-3 kinesin or dock OSM-3 to IFT particles (14, 39).However, a recent study of zebrafish has led to a different model for DYF-1 function. Zebrafish embryos that are homozygous for a loss of function of fleer, an ortholog of DYF-1, have shortened olfactory and pronephric cilia and ultrastructural defects in the axonemes. In the middle segment, the fleer axonemes have B tubules that are disconnected from the A tubule, indicating that DYF-1 functions in the middle segment and could play a role in the stability of doublet microtubules (40). Earlier, a similar mutant phenotype was reported in Tetrahymena for a mutation in the C-terminal tail domain of β-tubulin, at the glutamic acid residues that are used by posttranslational polymodifications (glycylation and glutamylation) (47). Glycylation (46) and glutamylation (12) are conserved polymeric posttranslational modifications that affect tubulin and are highly enriched on microtubules of axonemes and centrioles (reviewed in reference 20). Other studies have indicated that tubulin glutamylation contributes to the assembly and stability of axonemes and centrioles (4, 8). The fleer mutant zebrafish cilia have reduced levels of glutamylated tubulin (40). Pathak and colleagues proposed that the primary role of DYF-1/fleer is to serve as an IFT cargo adapter for a tubulin glutamic acid ligase (25) and that the effects of lack of function of DYF-1/fleer could be caused by deficiency in tubulin glutamylation in the axoneme (40). As an alternative hypothesis, the same authors proposed that DYF-1 is a structural component that stabilizes the doublet microtubules in the axoneme (40).Here, we evaluate the significance of a DYF-1 ortholog, Dyf1p, in Tetrahymena thermophila. Unexpectedly, we found that Tetrahymena cells lacking Dyf1p either fail to assemble an axoneme or can assemble an axoneme remnant. While our observations revealed major differences in the significance of DYF-1 for segmental differentiation in diverse models, it is clear that DYF-1 is a conserved and critical component that is required for assembly of the axoneme.  相似文献   

19.
Immunogold localization revealed that OmcS, a cytochrome that is required for Fe(III) oxide reduction by Geobacter sulfurreducens, was localized along the pili. The apparent spacing between OmcS molecules suggests that OmcS facilitates electron transfer from pili to Fe(III) oxides rather than promoting electron conduction along the length of the pili.There are multiple competing/complementary models for extracellular electron transfer in Fe(III)- and electrode-reducing microorganisms (8, 18, 20, 44). Which mechanisms prevail in different microorganisms or environmental conditions may greatly influence which microorganisms compete most successfully in sedimentary environments or on the surfaces of electrodes and can impact practical decisions on the best strategies to promote Fe(III) reduction for bioremediation applications (18, 19) or to enhance the power output of microbial fuel cells (18, 21).The three most commonly considered mechanisms for electron transfer to extracellular electron acceptors are (i) direct contact between redox-active proteins on the outer surfaces of the cells and the electron acceptor, (ii) electron transfer via soluble electron shuttling molecules, and (iii) the conduction of electrons along pili or other filamentous structures. Evidence for the first mechanism includes the necessity for direct cell-Fe(III) oxide contact in Geobacter species (34) and the finding that intensively studied Fe(III)- and electrode-reducing microorganisms, such as Geobacter sulfurreducens and Shewanella oneidensis MR-1, display redox-active proteins on their outer cell surfaces that could have access to extracellular electron acceptors (1, 2, 12, 15, 27, 28, 31-33). Deletion of the genes for these proteins often inhibits Fe(III) reduction (1, 4, 7, 15, 17, 28, 40) and electron transfer to electrodes (5, 7, 11, 33). In some instances, these proteins have been purified and shown to have the capacity to reduce Fe(III) and other potential electron acceptors in vitro (10, 13, 29, 38, 42, 43, 48, 49).Evidence for the second mechanism includes the ability of some microorganisms to reduce Fe(III) that they cannot directly contact, which can be associated with the accumulation of soluble substances that can promote electron shuttling (17, 22, 26, 35, 36, 47). In microbial fuel cell studies, an abundance of planktonic cells and/or the loss of current-producing capacity when the medium is replaced is consistent with the presence of an electron shuttle (3, 14, 26). Furthermore, a soluble electron shuttle is the most likely explanation for the electrochemical signatures of some microorganisms growing on an electrode surface (26, 46).Evidence for the third mechanism is more circumstantial (19). Filaments that have conductive properties have been identified in Shewanella (7) and Geobacter (41) species. To date, conductance has been measured only across the diameter of the filaments, not along the length. The evidence that the conductive filaments were involved in extracellular electron transfer in Shewanella was the finding that deletion of the genes for the c-type cytochromes OmcA and MtrC, which are necessary for extracellular electron transfer, resulted in nonconductive filaments, suggesting that the cytochromes were associated with the filaments (7). However, subsequent studies specifically designed to localize these cytochromes revealed that, although the cytochromes were extracellular, they were attached to the cells or in the exopolymeric matrix and not aligned along the pili (24, 25, 30, 40, 43). Subsequent reviews of electron transfer to Fe(III) in Shewanella oneidensis (44, 45) appear to have dropped the nanowire concept and focused on the first and second mechanisms.Geobacter sulfurreducens has a number of c-type cytochromes (15, 28) and multicopper proteins (12, 27) that have been demonstrated or proposed to be on the outer cell surface and are essential for extracellular electron transfer. Immunolocalization and proteolysis studies demonstrated that the cytochrome OmcB, which is essential for optimal Fe(III) reduction (15) and highly expressed during growth on electrodes (33), is embedded in the outer membrane (39), whereas the multicopper protein OmpB, which is also required for Fe(III) oxide reduction (27), is exposed on the outer cell surface (39).OmcS is one of the most abundant cytochromes that can readily be sheared from the outer surfaces of G. sulfurreducens cells (28). It is essential for the reduction of Fe(III) oxide (28) and for electron transfer to electrodes under some conditions (11). Therefore, the localization of this important protein was further investigated.  相似文献   

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