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.
The toxin MazF in Escherichia coli cleaves single-stranded RNAs specifically at ACA sequences. MazF overexpression virtually eliminates all cellular mRNAs to completely block protein synthesis. However, protein synthesis can continue on an mRNA that is devoid of ACA triplets. The finding that ribosomal RNAs remain intact in the face of complete translation arrest suggested a purpose for such preservation. We therefore examined the sequences of all transcribed RNAs to determine if there was any statistically significant bias against ACA. While ACA motifs are absent from tmRNA, 4.5S RNA, and seven of the eight 5S rRNAs, statistical analysis revealed that only for tmRNA was the absence nonrandom. The introduction of single-strand ACAs makes tmRNA highly susceptible to MazF cleavage. Furthermore, analysis of tmRNA sequences from 442 bacteria showed that the discrimination against ACA in tmRNAs was seen mostly in enterobacteria. We propose that the unusual bias against ACA in tmRNA may have coevolved with the acquisition of MazF.Toxin-antitoxin (TA) modules, originally found in plasmids, are believed to specifically prevent the growth of dividing cells that do not inherit the plasmid (13, 20). This type of selection process was given the term “postsegregational killing.” New findings demonstrate that the prevalence of these genes on plasmids is due to their ability to inhibit the replication of TA-negative plasmids after within-host competition (7, 8).Surprisingly, many species of bacteria and archaea contain homologues of these TA pairs in their genomes, often containing more than one copy (35, 51). Current research is revealing how diverse the targets of these toxins are, from topoisomerase to mRNAs (21, 36, 37, 44). Whatever its target, the defining characteristic of a toxin is its ability to slow the growth of the organism in which the toxin is overexpressed or in which its cognate antitoxin is degraded. Considering the apparent loss of fitness during toxin expression, the functional importance of the many TA modules found throughout the bacterial and archaeal kingdoms, often in the same organism, is still in dispute. Theories involving quorum sensing (28) and programmed cell death have been proposed, where the death of one population may benefit the survival of the species (32). Others have proposed that rather than being irreversible, toxin-induced growth inhibition may be the cell''s last resort to temporarily conserve its metabolic stores under extreme environmental conditions (5, 6, 11, 12). The radically different theory put forth by Saavedra De Bast et al. (38) argues that postsegregational killing was the selective pressure for the fixation of TA systems in the genome and that these genes may no longer have any biological role once the plasmid-borne TA modules evolve enough that they cannot be neutralized. Those authors showed that the possession of a cognate antitoxin in its genome does indeed allow the bacterium to survive the toxic effects of losing a TA-containing plasmid. However, that work did not address the fact that most antitoxins in genomes are found together with their toxins, the presence of which would be unnecessary, according to their theory.One of the best-studied toxins, MazF, is a sequence-specific endoribonuclease. This toxin cleaves single-strand ACA sequences and, when overexpressed, rapidly blocks protein synthesis (53). However, we demonstrated that translation itself is unaffected during MazF induction by overexpressing a protein whose mRNA is ACA-less (41). As both soluble and membrane-bound proteins can be expressed upon MazF induction (41), we inferred that the noncoding RNAs involved in protein synthesis and translocation must also be protected from MazF cleavage. This finding prompted us to analyze all RNAs involved in translation for the presence of ACAs in their secondary structures.In bacteria, the repertoire of RNAs involved in protein synthesis includes not only rRNAs, tRNAs, and mRNAs but also 4.5S and tmRNA. 4.5S RNA, in a complex with the Ffh protein, plays a pivotal role in cotranslational protein targeting and translocation (31). The ssrA gene encodes another small RNA (tmRNA) that is required for normal growth because it mediates the release of stalled ribosomes from defective mRNAs (48). All eubacterial genomes examined so far contain the ssrA gene (1).Here we examined the occurrence of every possible nucleotide triplet in the RNAs of Escherichia coli K-12. By ranking ACA among the 64 triplets possible, we found the frequency of ACAs to be generally low in the RNAs involved in translation. However, statistical analysis showed that only tmRNA had a significantly low number of ACAs in its single-strand region. Furthermore, examination of over 400 tmRNA sequences from different bacteria revealed a significant bias against single-strand ACAs most often in enterobacteria. As MazF cleavage of ACAs in tmRNA was biochemically confirmed, we propose that tmRNA function may be critical for the MazF-related stress response and that the unusual bias against ACA in tmRNA may have coevolved with the acquisition of a sequence-specific endoribonuclease.  相似文献   

3.
4.
5.
Porcine circovirus type 1 (PCV1), originally isolated as a contaminant of PK-15 cells, is nonpathogenic, whereas porcine circovirus type 2 (PCV2) causes an economically important disease in pigs. To determine the factors affecting virus replication, we constructed chimeric viruses by swapping open reading frame 1 (ORF1) (rep) or the origin of replication (Ori) between PCV1 and PCV2 and compared the replication efficiencies of the chimeric viruses in PK-15 cells. The results showed that the replication factors of PCV1 and PCV2 are fully exchangeable and, most importantly, that both the Ori and rep of PCV1 enhance the virus replication efficiencies of the chimeric viruses with the PCV2 backbone.Porcine circovirus (PCV) is a single-stranded DNA virus in the family Circoviridae (34). Type 1 PCV (PCV1) was discovered in 1974 as a contaminant of porcine kidney cell line PK-15 and is nonpathogenic in pigs (31-33). Type 2 PCV (PCV2) was discovered in piglets with postweaning multisystemic wasting syndrome (PMWS) in the mid-1990s and causes porcine circovirus-associated disease (PCVAD) (1, 9, 10, 25). PCV1 and PCV2 have similar genomic organizations, with two major ambisense open reading frames (ORFs) (16). ORF1 (rep) encodes two viral replication-associated proteins, Rep and Rep′, by differential splicing (4, 6, 21, 22). The Rep and Rep′ proteins bind to specific sequences within the origin of replication (Ori) located in the intergenic region, and both are responsible for viral replication (5, 7, 8, 21, 23, 28, 29). ORF2 (cap) encodes the immunogenic capsid protein (Cap) (26). PCV1 and PCV2 share approximately 80%, 82%, and 62% nucleotide sequence identity in the Ori, rep, and cap, respectively (19).In vitro studies using a reporter gene-based assay system showed that the replication factors of PCV1 and PCV2 are functionally interchangeable (2-6, 22), although this finding has not yet been validated in a live infectious-virus system. We have previously shown that chimeras of PCV in which cap has been exchanged between PCV1 and PCV2 are infectious both in vitro and in vivo (15), and an inactivated vaccine based on the PCV1-PCV2 cap (PCV1-cap2) chimera is used in the vaccination program against PCVAD (13, 15, 18, 27).PCV1 replicates more efficiently than PCV2 in PK-15 cells (14, 15); thus, we hypothesized that the Ori or rep is directly responsible for the differences in replication efficiencies. The objectives of this study were to demonstrate that the Ori and rep are interchangeable between PCV1 and PCV2 in a live-virus system and to determine the effects of swapped heterologous replication factors on virus replication efficiency in vitro.  相似文献   

6.
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   

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

8.
Mrc1 plays a role in mediating the DNA replication checkpoint. We surveyed replication elongation proteins that interact directly with Mrc1 and identified a replicative helicase, Mcm6, as a specific Mrc1-binding protein. The central portion of Mrc1, containing a conserved coiled-coil region, was found to be essential for interaction with the 168-amino-acid C-terminal region of Mcm6, and introduction of two amino acid substitutions in this C-terminal region abolished the interaction with Mrc1 in vivo. An mcm6 mutant bearing these substitutions showed a severe defect in DNA replication checkpoint activation in response to stress caused by methyl methanesulfonate. Interestingly, the mutant did not show any defect in DNA replication checkpoint activation in response to hydroxyurea treatment. The phenotype of the mcm6 mutant was suppressed when the mutant protein was physically fused with Mrc1. These results strongly suggest for the first time that an Mcm helicase acts as a checkpoint sensor for methyl methanesulfonate-induced DNA damage through direct binding to the replication checkpoint mediator Mrc1.Progression of the DNA replication machinery along chromosomes is a complex process. Replication forks pause occasionally when they encounter genomic regions that are difficult to replicate, such as highly transcribed regions, tRNA genes, and regions with specialized chromatin structure, like centromeric and heterochromatic regions (17). Replication forks also stall when treated with chemicals like methyl methanesulfonate (MMS), which causes DNA damage, or hydroxyurea (HU), which limits the cellular concentration of the deoxynucleoside triphosphate pool (17). Because de novo assembly and programming of the replisome do not occur after the onset of S phase (18), DNA replication forks must be protected from replicative stresses. The DNA replication checkpoint constitutes a surveillance mechanism for S-phase progression that safeguards replication forks from various replicative stresses (22, 38, 40), and malfunction of this checkpoint leads to chromosome instability and cancer development in higher organisms (4, 9).The Saccharomyces cerevisiae DNA replication checkpoint mediator Mrc1 is functionally conserved and is involved directly in DNA replication as a component of the replisome (1, 8, 16, 19, 29, 30). Mrc1, together with Tof1 and Csm3, is required for forming a replication pausing complex when the fork is exposed to replicative stress by HU (16). The pausing complex subsequently triggers events leading to DNA replication checkpoint activation and hence stable replicative arrest. A sensor kinase complex, Mec1-Ddc2 (ATR-ATRIP homolog of higher eukaryotes), is then recruited to the complex (14, 16). Mec1-Ddc2-mediated phosphorylation of Mrc1 activates the pausing complex, and phosphorylated Mrc1 likely recruits Rad53 (a putative homolog of CHK2 of higher eukaryotes), which is then activated via phosphorylation by Mec1-Ddc2 (1, 16, 20, 30). Activated Rad53 subsequently elicits a stress responses, i.e., stabilization of replication forks, induction of repair genes, and suppression of late-firing origins (24). It remains unclear, however, whether DNA replication checkpoint activation is induced in response to DNA damage by MMS, a reagent commonly used to study the DNA replication stress response. Several lines of evidence have suggested that MMS-induced damage is also sensed directly by the replication machinery (38, 40).Although biochemical and genetic interaction data have placed Mrc1 at the center of the replication checkpoint signal transduction cascade, its molecular function remains largely unknown. The proteins Mrc1, Tof1, and Csm3 associate with the Mcm complex (8, 27), a heterohexameric DNA helicase consisting of Mcm2 to Mcm7 proteins which unwinds the parental DNA duplex to allow replisome progression (3, 12, 18, 31, 32, 35). The Mcm complex associates with a specific set of regulatory proteins at forks to form replisome progression complexes (8). In addition to Mcm, Tof1, Csm3, and Mrc1, replisome progression complexes include factors such as Cdc45 and the GINS complex that are also required for fork progression (13, 26, 31, 32, 39). Claspin, a putative Xenopus laevis homolog of Mrc1, is also reported to associate with Cdc45, DNA polymerase ɛ (Polɛ), replication protein A, and two of the replication factor C complexes in aphidicolin-treated Xenopus egg extracts (19). Recently, Mrc1 was reported to interact directly with Polɛ (23).The aim of this study was to provide mechanistic insight into Mrc1 function in the DNA replication checkpoint. For this purpose, it was essential to identify, among all the essential proteins in the replication machinery, a specific protein that interacts with Mrc1 and to examine the role of this interaction in the DNA replication checkpoint. We found that Mrc1 interacts with Mcm6 directly and specifically. When the interaction between Mrc1 and Mcm6 was impaired, cells no longer activated the DNA replication checkpoint in response to MMS-induced replicative stress. Interestingly and unexpectedly, this interaction was not required for DNA replication checkpoint activation in response to HU-induced replicative stress. Our results provide the first mechanistic evidence that cells use separate mechanisms to transmit replicative stresses caused by MMS and HU for DNA replication checkpoint activation.  相似文献   

9.
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.
Eukaryotic mRNA translation begins with recruitment of the 40S ribosome complex to the mRNA 5′ end through the eIF4F initiation complex binding to the 5′ m7G-mRNA cap. Spliced leader (SL) RNA trans splicing adds a trimethylguanosine (TMG) cap and a sequence, the SL, to the 5′ end of mRNAs. Efficient translation of TMG-capped mRNAs in nematodes requires the SL sequence. Here we define a core set of nucleotides and a stem-loop within the 22-nucleotide nematode SL that stimulate translation of mRNAs with a TMG cap. The structure and core nucleotides are conserved in other nematode SLs and correspond to regions of SL1 required for early Caenorhabditis elegans development. These SL elements do not facilitate translation of m7G-capped RNAs in nematodes or TMG-capped mRNAs in mammalian or plant translation systems. Similar stem-loop structures in phylogenetically diverse SLs are predicted. We show that the nematode eukaryotic translation initiation factor 4E/G (eIF4E/G) complex enables efficient translation of the TMG-SL RNAs in diverse in vitro translation systems. TMG-capped mRNA translation is determined by eIF4E/G interaction with the cap and the SL RNA, although the SL does not increase the affinity of eIF4E/G for capped RNA. These results suggest that the mRNA 5′ untranslated region (UTR) can play a positive and novel role in translation initiation through interaction with the eIF4E/G complex in nematodes and raise the issue of whether eIF4E/G-RNA interactions play a role in the translation of other eukaryotic mRNAs.Cap-dependent translation initiation in eukaryotes is a complex process involving many factors and serves as the primary mechanism for eukaryotic translation (37, 44). The first step in the initiation process, recruitment of the m7G (7-methylguanosine)-capped mRNA to the ribosome, is widely considered the rate-limiting step. It begins with recognition of and binding to the m7G cap at the 5′ end of the mRNA by the eukaryotic translation initiation factor 4F (eIF4F) complex, which contains three proteins: eIF4E (a cap-binding protein), eIF4G (a scaffold protein with RNA binding sites), and eIF4A (an RNA helicase). eIF4G''s interaction with eIF3, itself a multisubunit complex that interacts with the 40S ribosome, facilitates the actual recruitment of capped RNA to the ribosome. With the help of several other initiation factors, the small ribosomal subunit scans the mRNA from 5′ to 3′ until a translation initiation codon (AUG) in appropriate context is identified and an 80S ribosomal complex is formed, after which the first peptide bond is formed, thus ending the initiation process (37, 44). The AUG context can play an important role in the efficiency of translation initiation (23, 44). The length, structure, and presence of AUGs or open reading frames in the mRNA 5′ untranslated region (UTR) can negatively affect cap-dependent translation and ribosomal scanning. In general, long and highly structured 5′ UTRs, as well as upstream AUGs leading to short open reading frames, can impede ribosome scanning and lead to reduced translation (23, 44). In addition, 5′ UTRs less than 10 nucleotides (nt) in length are thought to be too short to enable preinitiation complex assembly and scanning (24). Thus, several attributes of the mRNA 5′ UTR are known to negatively affect translation initiation, whereas only the AUG context and the absence of negative elements are known to have a positive effect on translation initiation (44).Two of the important mRNA features associated with cap-dependent translation, the cap and the 5′ UTR, are significantly altered by an RNA processing event known as spliced leader (SL) trans splicing (3, 8, 17, 26, 36, 47). This takes place in members of a diverse group of eukaryotic organisms, including some protozoa, sponges, cnidarians, chaetognaths, flatworms, nematodes, rotifers, crustaceans, and tunicates (17, 28, 39, 55, 56). In SL trans splicing, a separately transcribed small exon (16 to 51 nucleotides [nt]) with its own cap gets added to the 5′ end of pre-mRNAs. This produces mature mRNAs with a unique cap and a conserved sequence in the 5′ UTR. In metazoa, the m7G cap is replaced with a trimethylguanosine (TMG) cap (m2,2,7GpppN) (27, 30, 46, 49). In nematodes, ∼70% of all mRNAs are trans spliced and therefore have a TMG cap and an SL (2). In general, eukaryotic eIF4E proteins do not effectively recognize the TMG cap (35). This raises the issues of how the translation machinery in trans-splicing metazoa effectively recognizes TMG-capped trans-spliced mRNAs, what role the SL sequence plays in translation initiation, and how the conserved translation initiation machinery has adapted to effectively translate trans-spliced mRNAs.Previous work has shown that efficient translation of TMG-capped messages in nematodes requires the SL sequence (22 nt) immediately downstream of the cap (5, 25, 29). In the current studies, we sought to understand the manner in which the SL enhanced the translation of TMG-capped mRNAs. Using a cell-free nematode in vitro translation system, we carried out mutational analyses that define the specific sequences in the SL that are required and sufficient for efficient translation of TMG-capped mRNAs. These analyses led to the discovery of a small, discrete stem-loop immediately adjacent to the TMG cap in trans-spliced messages required for efficient translation. Notably, the sequences involved in the base pairing of the stem are highly conserved in alternative SL sequences found in nematodes. We further show that the nematode eIF4E/G complex plays a major role in facilitating the SL enhancement of TMG-capped mRNA that likely occurs after the initial cap-binding step. The results demonstrate the importance of specific enhancing elements in the 5′ UTR and adaptation in the eIF4F complex necessary for optimal cap-dependent translation.  相似文献   

12.
13.
14.
15.
16.
Naturally occurring hepatitis C virus (HCV) subgenomic RNAs have been found in several HCV patients. These subgenomic deletion mutants, mostly lacking the genes encoding envelope glycoproteins, were found in both liver and serum, where their relatively high abundance suggests that they are capable of autonomous replication and can be packaged and secreted in viral particles, presumably harboring the envelope proteins from wild type virus coinfecting the same cell. We recapitulated some of these natural subgenomic deletions in the context of the isolate JFH-1 and confirmed these hypotheses in vitro. In Huh-7.5 cells, these deletion-containing genomes show robust replication and can be efficiently trans-packaged and infect naïve Huh-7.5 cells when cotransfected with the full-length wild-type J6/JFH genome. The genome structure of these natural subgenomic deletion mutants was dissected, and the maintenance of both core and NS2 regions was proven to be significant for efficient replication and trans-packaging. To further explore the requirements needed to achieve trans-complementation, we provided different combinations of structural proteins in trans. Optimal trans-complementation was obtained when fragments of the polyprotein encompassing core to p7 or E1 to NS2 were expressed. Finally, we generated a stable helper cell line, constitutively expressing the structural proteins from core to p7, which efficiently supports trans-complementation of a subgenomic deletion encompassing amino acids 284 to 732. This cell line can produce and be infected by defective particles, thus representing a powerful tool to investigate the life cycle and relevance of natural HCV subgenomic deletion mutants in vivo.Hepatitis C virus (HCV) is an enveloped virus belonging to the family Flaviviridae. The virus genome is a positive-stranded RNA of about 9,600 nucleotides, which contains a single open reading frame (ORF) encoding both structural (core, E1, E2, and p7) and nonstructural (NS2, NS3, NS4A, NS4B, NS5A, and NS5B) proteins (47). Two highly conserved untranslated regions (UTRs) are found at the 5′ and 3′ ends, which play critical roles in both viral translation and replication (13, 14, 23).HCV is estimated to infect 170 million people worldwide (1, 43) and in a high percentage of individuals causes a chronic liver infection that frequently evolves into an array of diseases, including cirrhosis (12, 15, 38) and hepatocellular carcinoma (4, 7, 17, 30, 38, 49).The HCV RNA-dependent RNA polymerase (NS5B) has a high frequency of incorrect nucleotide insertions, in the range of 10−4 to 10−5 base substitutions per site, which can result in the rapid generation of HCV quasispecies (37, 45). Because of this huge genetic diversity, HCV is currently classified into six major genotypes and more than 80 subtypes (44). Recombination may be another mechanism exploited by HCV to increase genetic diversity: naturally occurring intergenotypic recombinant viruses that often have their recombination points in the trans-membrane domains of NS2 were recently identified (20, 21, 26, 27, 33, 35).Recent publications have reported the presence of natural HCV subgenomic RNAs in serum and liver of infected patients, mostly containing large in-frame deletions from E1 up to NS2, always found together with the full-length wild-type (wt) RNAs (5, 16, 36, 54). These mutant viral genomes persist for a long time (at least 2 years), and sequence analysis suggests that subgenomic (the predominant species during this period) and full-length HCV evolve independently (54). The relative abundance and persistence of such subgenomic RNAs in vivo suggests that (i) they are capable of autonomous replication and (ii) they can be packaged and secreted in infectious viral particles, presumably harboring the envelope proteins from wt virus coinfecting the same cell.Analysis of the genetic structure of 18 independent subgenomic deletion-containing RNAs (5, 16, 36, 54) strongly suggests that the possibility of recombination and/or deletion is restricted to specific regions.As expected, the 5′ UTR, the 3′ UTR, and the region coding from NS3 to NS5B are always conserved, in line with the notion that these regions are the minimal requirements for RNA replication. In addition to the regions required for RNA replication, however, naturally occurring subgenomic HCV RNA invariably contains an intact core region and the protease domain of NS2. In fact, the 5′ deletion boundary falls between the codons for amino acid 189 of core and amino acid 4 of E1 in 33% of cases and between the codons for amino acids 21 and 29 of E1 in 39% of cases, such that in 72% of cases overall, the 5′ boundary is between the codons for amino acid 189 of core and amino acid 29 of E1 (see Fig. S1 in the supplemental material). Likewise, the 3′ boundary shows a distinct localization, occurring between the codons for amino acids 51 and 79 of NS2 in 56% of cases.In the present study, we modified the infectious isolate JFH-1 in order to recapitulate in vitro the genetic structure of two of the most representative in-frame natural subgenomic deletion-containing RNAs found circulating in patients (28, 51, 57). Using this system, we analyzed natural subgenomic variants for their ability to replicate autonomously and demonstrated that the natural subgenomic deletion mutants are replication competent and are trans-packaged into infectious virions when coexpressed together with wt virus. Furthermore, our data suggest that the presence of the NS2 protease domain is required in order to generate the correct NS3 N terminus, required for RNA replication. Unexpectedly, the presence of NS2 generates, in turn, a strict cis requirement for the core region in order to allow efficient trans-packaging of the subgenomic RNA, revealing a complex interplay between the NS2 and the core viral genes. Finally, we performed trans-complementation studies of the subgenomic deletions with different HCV structural proteins to gain insight into the minimal requirements for the assembly and release of HCV virions.  相似文献   

17.
The highly conserved bacterial ybeY gene is a heat shock gene whose function is not fully understood. Previously, we showed that the YbeY protein is involved in protein synthesis, as Escherichia coli mutants with ybeY deleted exhibit severe translational defects in vivo. Here we show that the in vitro activity of the translation machinery of ybeY deletion mutants is significantly lower than that of the wild type. We also show that the lower efficiency of the translation machinery is due to impaired 30S small ribosomal subunits.Many heat shock proteins are chaperones and proteases that constitute the protein quality control system (4, 5, 13, 18). Recent studies demonstrated that beyond protein quality control, the heat shock response includes proteins implemented in the translation machinery (16, 17), such as FtsJ (2, 3) and Hsp15 (11).FtsJ catalyzes methylation of U2552 in 23S rRNA (3). This modification occurs during the final steps of 50S biogenesis and is important for the structural stability of the 50S subunit (2). ftsJ deletion mutants accumulate ribosomal subunits at the expense of polysomes (2). Consequently, crude ribosome extracts prepared from ftsJ deletion mutants are far less active than wild-type preparations (3). Hsp15 recognizes and binds with high affinity to the aberrant state of the 50S subunit in complex with peptidyl tRNA positioned at the A site (10), which is more frequent at high temperatures (10). It has been proposed that Hsp15 participates in releasing the bound peptide and thereby helps recycle the 50S subunit (8, 10). Thus, heat shock proteins play a significant role both in the biogenesis of ribosomes and in the translation process.YbeY is a 17-kDa heat shock protein, highly conserved among bacteria, that belongs to the UPF0054 family of metal-dependent hydrolases, suggesting that it may have a potential hydrolytic function (14, 21). In Aquifex aeolicus, analysis of YbeY structure homology showed similarity to eukaryotic extracellular proteinases such as collagenase and gelatinase. However, in vitro experiments could not detect collagenase, gelatinase, or other hydrolase activity in YbeY (14).Recently, we showed that ybeY deletion mutants exhibit severe translational defects manifested by a very low level of polysomes and accumulation of free ribosomes and ribosomal subunits, indicating that most ribosomes in the cell are not engaged in translation. This translational defect intensifies at elevated temperatures (42°C) and results in growth arrest (17).Here we present in vitro studies indicating that the activity of the translation machinery prepared from ybeY deletion mutants is lower than in the wild type. In addition, we show that this lower activity stems specifically from a defective 30S ribosomal subunit.  相似文献   

18.
The Asf1 and Rad6 pathways have been implicated in a number of common processes such as suppression of gross chromosomal rearrangements (GCRs), DNA repair, modification of chromatin, and proper checkpoint functions. We examined the relationship between Asf1 and different gene products implicated in postreplication repair (PRR) pathways in the suppression of GCRs, checkpoint function, sensitivity to hydroxyurea (HU) and methyl methanesulfonate (MMS), and ubiquitination of proliferating cell nuclear antigen (PCNA). We found that defects in Rad6 PRR pathway and Siz1/Srs2 homologous recombination suppression (HRS) pathway genes suppressed the increased GCR rates seen in asf1 mutants, which was independent of translesion bypass polymerases but showed an increased dependency on Dun1. Combining an asf1 deletion with different PRR mutations resulted in a synergistic increase in sensitivity to chronic HU and MMS treatment; however, these double mutants were not checkpoint defective, since they were capable of recovering from acute treatment with HU. Interestingly, we found that Asf1 and Rad6 cooperate in ubiquitination of PCNA, indicating that Rad6 and Asf1 function in parallel pathways that ubiquitinate PCNA. Our results show that ASF1 probably contributes to the maintenance of genome stability through multiple mechanisms, some of which involve the PRR and HRS pathways.DNA replication must be highly coordinated with chromatin assembly and cell division for correct propagation of genetic information and cell survival. Errors arising during DNA replication are corrected through the functions of numerous pathways including checkpoints and a diversity of DNA repair mechanisms (32, 33, 35). However, in the absence of these critical cellular responses, replication errors can lead to the accumulation of mutations and gross chromosomal rearrangements (GCRs) as well as chromosome loss, a condition generally termed genomic instability (33). Genome instability is a hallmark of many cancers as well as other human diseases (24). There are many mechanisms by which GCRs can arise, and over the last few years numerous genes and pathways have been implicated in playing a role in the suppression of GCRs in Saccharomyces cerevisiae and in some cases in the etiology of cancer (27, 28, 33, 39-47, 51, 53, 56, 58, 60), including S. cerevisiae ASF1, which encodes the main subunit of the replication coupling assembly factor (37, 62).Asf1 is involved in the deposition of histones H3 and H4 onto newly synthesized DNA during DNA replication and repair (62), and correspondingly, asf1 mutants are sensitive to chronic treatment with DNA-damaging agents (2, 30, 62). However, asf1 mutants do not appear to be repair defective and can recover from acute treatment with at least some DNA-damaging agents (2, 8, 30, 31, 54), properties similar to those described for rad9 mutants (68). In the absence of Asf1, both the DNA damage and replication checkpoints become activated during normal cell growth, and in the absence of checkpoint execution, there is a further increase in checkpoint activation in asf1 mutants (30, 46, 54). It has been suggested that asf1 mutants are defective for checkpoint shutoff and that this might account for the increased steady-state levels of checkpoint activation seen in asf1 mutants (8); however, another study has shown that asf1 mutants are not defective for checkpoint shutoff and that in fact Asf1 and the chromatin assembly factor I (CAF-I) complex act redundantly or cooperate in checkpoint shutoff (31). Furthermore, Asf1 might be involved in proper activation of the Rad53 checkpoint protein, as Asf1 physically interacts with Rad53 and this interaction is abrogated in response to exogenous DNA damage (15, 26); however, the physiological relevance of this interaction is unclear. Asf1 is also required for K56 acetylation of histone H3 by Rtt109, and both rtt109 mutants and histone H3 variants that cannot be acetylated (38) share many of the properties of asf1 mutants, suggesting that at least some of the requirement for Asf1 in response to DNA damage is mediated through Rtt109 (11, 14, 22, 61). Subsequent studies of checkpoint activation in asf1 mutants have led to the hypothesis that replication coupling assembly factor defects result in destabilization of replication forks which are then recognized by the replication checkpoint and stabilized, suggesting that the destabilized replication forks account for both the increased GCRs and increased checkpoint activation seen in asf1 mutants (30). This hypothesis is supported by other recent studies implicating Asf1 in the processing of stalled replication forks (16, 57). This role appears to be independent of CAF-I, which can cooperate with Asf1 in chromatin assembly (63). Asf1 has also been shown to function in disassembly of chromatin, suggesting other possibilities for the mechanism of action of Asf1 at the replication fork (1, 2, 34). Thus, while Asf1 is thought to be involved in progression of the replication fork, both the mechanism of action and the factors that cooperate with Asf1 in this process remain obscure.Stalled replication forks, particularly those that stall at sites of DNA damage, can be processed by homologous recombination (HR) (6) or by a mechanism known as postreplication repair (PRR) (reviewed in reference 67). There are two PRR pathways, an error-prone pathway involving translesion synthesis (TLS) by lower-fidelity polymerases and an error-free pathway thought to involve template switching (TS) (67). In S. cerevisiae, the PRR pathways are under the control of the RAD6 epistasis group (64). The error-prone pathway depends on monoubiquitination of proliferating cell nuclear antigen (PCNA) on K164 by Rad6 (an E2 ubiquitin-conjugating enzyme) by Rad18 (E3 ubiquitin ligase) (23). This results in replacement of the replicative DNA polymerase with nonessential TLS DNA polymerases, such as REV3/REV7-encoded DNA polymerase ζ (polζ) and RAD30-encoded DNA polη, which can bypass different types of replication-blocking damage (67). The error-free pathway is controlled by Rad5 (E3) and a complex consisting of Ubc13 and Mms2 (E2 and E2 variant, respectively), which add a K63-linked polyubiquitin chain to monoubiquitinated PCNA, leading to TS to the undamaged nascent sister chromatid (4, 25, 65). Furthermore, in addition to modification with ubiquitin, K164 of PCNA can also be sumoylated by Siz1, resulting in subsequent recruitment of the Srs2 helicase and inhibition of deleterious Rad51-dependent recombination events (50, 52, 55), although it is currently unclear if these are competing PCNA modifications or if both can exist on different subunits in the same PCNA trimer. A separate branch of the Rad6 pathway involving the E3 ligase Bre1 monoubiquitinates the histone H2B (29, 69) as well as Swd2 (66), which stimulates Set1-dependent methylation of K4 and Dot1-dependent methylation of K79 of histone H3 (48, 49, 66). Subsequently, K79-methylated H3 recruits Rad9 and activates the Rad53 checkpoint (19, 70). Activation of Rad53 is also bolstered by Rad6-Rad18-dependent ubiquitination of Rad17, which is part of the 9-1-1 complex that functions upstream in the checkpoint pathway (17). Finally, Rad6 complexes with the E3 Ubr1, which mediates protein degradation by the N-end rule pathway (13).Due to the role of the PRR pathways at stalled replication forks and a recent study implicating the Rad6 pathway in the suppression of GCRs (39), we examined the relationship between these ubiquitination and sumoylation pathways and the Asf1 pathway in order to gain additional insights into the function of Asf1 during DNA replication and repair. Our findings suggest that Asf1 has multiple functions that prevent replication damage or act in the cellular responses to replication damage and that these functions are modified by and interact with the PRR pathways. The TLS PRR pathway does not appear to be involved, and both a Dun1-dependent replication checkpoint and HR are important for preventing the deleterious effects of PRR and Asf1 pathway defects. We hypothesize that this newly observed cooperation between Asf1 and the PRR pathways may be required for resolving stalled replication forks, leading to suppression of GCRs and successful DNA replication.  相似文献   

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