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 125 毫秒
1.
The replication of enteroviruses is sensitive to brefeldin A (BFA), an inhibitor of endoplasmic reticulum-to-Golgi network transport that blocks activation of guanine exchange factors (GEFs) of the Arf GTPases. Mammalian cells contain three BFA-sensitive Arf GEFs: GBF1, BIG1, and BIG2. Here, we show that coxsackievirus B3 (CVB3) RNA replication is insensitive to BFA in MDCK cells, which contain a BFA-resistant GBF1 due to mutation M832L. Further evidence for a critical role of GBF1 stems from the observations that viral RNA replication is inhibited upon knockdown of GBF1 by RNA interference and that replication in the presence of BFA is rescued upon overexpression of active, but not inactive, GBF1. Overexpression of Arf proteins or Rab1B, a GTPase that induces GBF1 recruitment to membranes, failed to rescue RNA replication in the presence of BFA. Additionally, the importance of the interaction between enterovirus protein 3A and GBF1 for viral RNA replication was investigated. For this, the rescue from BFA inhibition of wild-type (wt) replicons and that of mutant replicons of both CVB3 and poliovirus (PV) carrying a 3A protein that is impaired in binding GBF1 were compared. The BFA-resistant GBF1-M832L protein efficiently rescued RNA replication of both wt and mutant CVB3 and PV replicons in the presence of BFA. However, another BFA-resistant GBF1 protein, GBF1-A795E, also efficiently rescued RNA replication of the wt replicons, but not that of mutant replicons, in the presence of BFA. In conclusion, this study identifies a critical role for GBF1 in CVB3 RNA replication, but the importance of the 3A-GBF1 interaction requires further study.Enteroviruses are small, nonenveloped, positive-stranded RNA viruses that include many important pathogens, such as poliovirus (PV), coxsackievirus, echovirus, and human rhinovirus. Following virus entry and uncoating, the 7.5-kb enteroviral RNA genome is directly translated into a large polyprotein. This polyprotein is proteolytically processed by the virus-encoded proteases 2Apro, 3Cpro, and 3CDpro into the structural P1 region proteins and the nonstructural P2 and P3 region proteins that are involved in viral RNA replication.All RNA viruses with a positive-stranded genome induce the remodeling of cellular membranes to create a scaffold for genomic RNA replication. The organelle origin and morphology of these membranous replication sites, however, appear to vary for different viruses. Enteroviruses replicate their RNA genomes in nucleoprotein complexes that are associated with small vesicular membrane structures (6). The enteroviral proteins 2B, 2C, and 3A have been implicated in vesicle formation (4, 6, 27), but the mechanism and pathway of membrane reorganization are poorly understood. There are strong indications that these vesicular membranous structures, which are referred to here as “vesicles,” are derived from the early secretory pathway. Vesicles produced in PV-infected cells may form at the endoplasmic reticulum (ER) by the cellular COP-II budding machinery and may therefore share components with the membranous vesicles mediating ER-to-Golgi network transport (26). Further support for the involvement of the secretory pathway stems from the observation that brefeldin A (BFA), a well-known inhibitor of ER-to-Golgi network transport, completely inhibits enteroviral RNA replication (17, 20). In addition, the autophagocytic pathway appears to contribute to the formation of the membrane vesicles, many of which exhibit a double-membrane morphology characteristic of autophagosomes (18, 27). The utilization of individual components or reactions from different membrane metabolic pathways, rather than subversion of an entire pathway in toto, may represent a common strategy for building viral replication machinery.BFA inhibits activation of the small monomeric GTPase ADP ribosylation factor 1 (Arf1), a major regulator of intracellular protein transport (2). Arf1 cycles between an inactive, GDP-bound, cytosolic state and an active, GTP-bound, membrane-associated state, and this cycling is catalyzed by guanine nucleotide exchange factors (GEFs) and GTPase-activating proteins (13). BFA blocks the activities of the large GEFs GBF1, BIG1, and BIG2 by stabilizing an intermediate, abortive complex with inactive Arf1 (23), thus efficiently preventing activation of Arf1 and eventually formation of transport intermediates.Not only the fact that BFA blocks enteroviral replication suggests a role for Arf1 and/or its large GEFs in this process; recently, it was shown that Arf1 accumulates on membranes during PV infection (3). Arf1 translocation to membranes can be induced independently by enterovirus protein 3A or 3CD in vitro (5), but the underlying mechanisms seem to differ; the 3A protein specifically triggers the recruitment of GBF1 to membranes, most likely through a direct interaction with this GEF (32, 33), whereas 3CD recruits BIG1 and BIG2 to membranes (3). Here, we report the involvement of Arf1 and its large BFA-sensitive GEFs in coxsackievirus B3 (CVB3) replication.  相似文献   

2.
3.
In infected cells, hepatitis C virus (HCV) induces the formation of membrane alterations referred to as membranous webs, which are sites of RNA replication. In addition, HCV RNA replication also occurs in smaller membrane structures that are associated with the endoplasmic reticulum. However, cellular mechanisms involved in the formation of HCV replication complexes remain largely unknown. Here, we used brefeldin A (BFA) to investigate cellular mechanisms involved in HCV infection. BFA acts on cell membranes by interfering with the activation of several members of the family of ADP-ribosylation factors (ARF), which can lead to a wide range of inhibitory actions on membrane-associated mechanisms of the secretory and endocytic pathways. Our data show that HCV RNA replication is highly sensitive to BFA. Individual knockdown of the cellular targets of BFA using RNA interference and the use of a specific pharmacological inhibitor identified GBF1, a guanine nucleotide exchange factor for small GTPases of the ARF family, as a host factor critically involved in HCV replication. Furthermore, overexpression of a BFA-resistant GBF1 mutant rescued HCV replication in BFA-treated cells, indicating that GBF1 is the BFA-sensitive factor required for HCV replication. Finally, immunofluorescence and electron microscopy analyses indicated that BFA does not block the formation of membranous web-like structures induced by expression of HCV proteins in a nonreplicative context, suggesting that GBF1 is probably involved not in the formation of HCV replication complexes but, rather, in their activity. Altogether, our results highlight a functional connection between the early secretory pathway and HCV RNA replication.Hepatitis C virus (HCV) is an important human pathogen. It mainly infects human hepatocytes, and this often leads to chronic hepatitis, cirrhosis, or hepatocarcinoma. HCV studies have been hampered for many years by the difficulty in propagating this virus in vitro. Things have recently changed with the development of a cell culture model referred to as HCVcc (34, 60, 65), which allows the study of the HCV life cycle in cell culture and facilitates studies of the interactions between HCV and the host cell.HCV is an enveloped positive-strand RNA virus belonging to the family Flaviviridae (35). The viral genome contains a single open reading frame, which is flanked by two noncoding regions that are required for translation and replication. All viral proteins that are produced after proteolytic processing of the initially synthesized polyprotein are membrane associated (15, 43). This reflects the fact that virtually all steps of the viral life cycle occur in close association with cellular membranes.Interactions of HCV with cell membranes begin during entry. Several receptors, coreceptors, and other entry factors have been discovered over the years, which link HCV entry to specialized domains of the plasma membrane, such as tetraspanin-enriched microdomains and tight junctions (8, 16, 59). The internalization of the viral particle occurs by clathrin-mediated endocytosis (5, 40). The fusion of the viral envelope with the membrane of an acidic endosome likely mediates the transfer of the viral genome to the cytosol of the cell (5, 40, 57). However, little is known regarding the pre- and postfusion intracellular transport steps of entering viruses in the endocytic pathway.HCV RNA replication is also associated with cellular membranes. Replication begins with the translation of the genomic RNA of an incoming virus. This leads to the production of viral proteins, which in turn initiate the actual replication of the viral RNA. Mechanisms regulating the transition from the translation of the genomic RNA to its replication are not yet known. All viral proteins are not involved in RNA replication. Studies performed with subgenomic replicons demonstrated that proteins NS3-4A, NS4B, NS5A, and NS5B are necessary and sufficient for replication (6, 27, 37). RNA replication proceeds through the synthesis of a cRNA strand (negative strand), catalyzed by the RNA-dependent RNA polymerase activity of NS5B, which is then used as a template for the synthesis of new positive strands.Electron microscopy studies using a subgenomic replicon model suggested that replication takes place in membrane structures made of small vesicles, referred to as “membranous webs,” which are induced by the virus (26). Membranous webs are detectable not only in cells carrying subgenomic replicons but also in infected cells (50). They appear to be associated with the endoplasmic reticulum (ER) (26). In addition to the membranous webs, a second type of ER-associated replicase that is smaller and more mobile has recently been described (63). Cellular mechanisms leading to these membrane alterations are still poorly understood. In cells replicating and secreting infectious viruses effectively, the situation appears to be even more complex, since replicase components appear to be, at least in part, associated with cytoplasmic lipid droplets (41, 50, 56). This association depends on the capsid protein (41) and may reflect a coupling between replication and assembly. Indeed, HCV assembly and secretion show some similarities with very-low-density lipoprotein (VLDL) maturation and secretion (24, 64).Our knowledge of the cellular membrane mechanisms involved in the HCV life cycle is still limited. The expression of NS4B alone induces membrane alterations that are reminiscent of membranous webs (19). However, cellular factors that participate in this process are still unknown. On the other hand, several cellular proteins potentially involved in the HCV life cycle have been identified through their interactions with viral proteins. For some of these proteins, a functional role in infection was recently confirmed using RNA interference (48). It is very likely that other cellular factors critical to HCV infection have yet to be identified.To gain more insight into cellular mechanisms underlying HCV infection, we made use of brefeldin A (BFA), a macrocyclic lactone of fungal origin that exhibits a wide range of inhibitory actions on membrane-associated mechanisms of the secretory and endocytic pathways (30). BFA acts on cell membranes by interfering with the activation of several members of the family of ADP-ribosylation factors (ARFs). ARFs are small GTP-binding proteins of the Ras superfamily. They function as regulators of vesicular traffic, actin remodeling, and phospholipid metabolism by recruiting effectors to membranes. BFA does not actually interfere directly with ARF GTPases but rather interferes with their activation by regulators known as guanine nucleotide exchange factors (GEFs) (14, 25). We now report the identification of an ARF GEF as a cellular BFA-sensitive factor that is required for HCV replication.  相似文献   

4.
The genome of the severe acute respiratory syndrome-associated coronavirus (SARS-CoV) contains eight open reading frames (ORFs) that encode novel proteins. These accessory proteins are dispensable for in vitro and in vivo replication and thus may be important for other aspects of virus-host interactions. We investigated the functions of the largest of the accessory proteins, the ORF 3a protein, using a 3a-deficient strain of SARS-CoV. Cell death of Vero cells after infection with SARS-CoV was reduced upon deletion of ORF 3a. Electron microscopy of infected cells revealed a role for ORF 3a in SARS-CoV induced vesicle formation, a prominent feature of cells from SARS patients. In addition, we report that ORF 3a is both necessary and sufficient for SARS-CoV-induced Golgi fragmentation and that the 3a protein accumulates and localizes to vesicles containing markers for late endosomes. Finally, overexpression of ADP-ribosylation factor 1 (Arf1), a small GTPase essential for the maintenance of the Golgi apparatus, restored Golgi morphology during infection. These results establish an important role for ORF 3a in SARS-CoV-induced cell death, Golgi fragmentation, and the accumulation of intracellular vesicles.The severe acute respiratory syndrome-associated coronavirus (SARS-CoV) genome encodes several smaller open reading frames (ORFs) located in the 3′ region of the genome that are predicted to express eight novel proteins termed accessory proteins. The accessory proteins are designated ORFs 3a, 3b, 6, 7a, 7b, 8a, 8b, and 9b and range in size from 39 to 274 amino acids (35, 50). These SARS-CoV-specific ORFs are not present in other coronaviruses and do not display significant homology with any known proteins in the NCBI database. Five of these are predicted to code for polypeptides of greater than 50 amino acids (35, 50). Antibodies reactive against all of the SARS-CoV proteins have been detected in sera isolated from SARS patients, indicating that these proteins are expressed by the virus in vivo (7, 9, 17-19, 45, 59). Expression of three of the ORF proteins has been demonstrated during infection using protein-specific antibodies and include the ORFs 3a, 6, and 7a (12, 37, 41, 60). Six of the eight group-specific ORFs, including ORFs 3a, 3b, 6, 7a, 7b, and 9b, were deleted from recombinant SARS-CoV and shown to be dispensable for in vitro and in vivo replication (66).Related coronaviruses also encode unique accessory proteins in the 3′ region of the genome, often referred to as group-specific ORFs. Similar to SARS-CoV, several of these proteins are dispensable for viral replication. Murine hepatitis virus (MHV) expresses accessory proteins ORFs 2a, 4, and 5a. A recombinant virus in which ORF 2a was deleted replicated normally in vitro but caused attenuated disease in vivo (55). Deletion of the group-specific ORF 7 in porcine coronavirus TGEV also results in reduced replication and virulence in vivo despite normal replication in vitro (38). Similarly, in feline infectious peritonitis virus (FIPV), group-specific proteins are dispensable for replication in cell culture but contribute to pathogenesis in vivo (20). Thus, while the SARS-CoV group specific proteins are unnecessary for in vitro and in vivo replication, their expression may underlie the devastating pathology associated with SARS disease. Detailed characterization of these novel proteins may contribute to a better understanding of SARS pathogenesis and host-virus interactions.The ORF 3a protein is expressed from subgenomic RNA3, which contains the 3a and 3b ORFs (35, 50). The 3a protein, which is the largest group-specific SARS-CoV accessory protein at 274 amino acids, has been reported to localize to the Golgi apparatus, the plasma membrane, and intracellular vesicles of unknown origin (67, 68). The protein is efficiently transported to the cell surface and is also internalized during the process of endocytosis (60).The mechanism of SARS-CoV-induced cell death has been investigated by several groups. Studies to date have used overexpression of individual SARS-CoV ORFs to evaluate their intrinsic cytotoxicity. Using this approach, the following proteins have been reported to cause apoptosis: the 3CL-like protease; spike; ORFs 3a, 3b, and 7a; and the envelope (E), membrane (M), and nucleocapsid (N) proteins (23, 31, 32, 36, 46, 58, 61, 65, 69). However, since all of these reports utilize overexpression of individual proteins, it is unclear whether these effects may be attributable to high, nonphysiological levels of protein and whether they occur during infection. Analysis of recombinant viruses with specific mutations or deletions is necessary to determine the relative contribution of these proteins to the cytotoxicity of SARS-CoV during infection (63). Therefore, the cytotoxic component(s) of SARS-CoV have not been fully defined.Here, we have investigated the function of the ORF 3a protein in the context of SARS-CoV infection and by overexpression. We confirm that ORF 3a contributes to SARS-CoV cytotoxicity using a recombinant strain deficient for expression of ORF 3a. While characterizing this deficient strain, we observed that SARS-CoV-induced vesicle formation, a feature that has been documented in cells from infected SARS patients, is dependent on ORF 3a. Furthermore, we observed that SARS-CoV infection causes Golgi fragmentation by ORF 3a. Additional characterization of 3a in transfected cells revealed that the protein colocalizes with markers of the trans-Golgi network (TGN) and late endosomal pathways and causes an accumulation of these vesicles. Finally, we report that Arf1 overexpression rescued SARS-CoV or 3a-induced Golgi fragmentation, suggesting that the ORF 3a protein may perturb Arf1-mediated vesicle trafficking.  相似文献   

5.
6.
Golgi fragmentation is a process that is necessary to allow its redistribution into daughter cells during mitosis, a process controlled by serine-threonine kinases. This Golgi fragmentation is activated by MEK1 and Plk3. Plk3 is a kinase that is a downstream target in the Golgi fragmentation pathway induced by MEK1 or by nocodazole. In this work, we have identified that Plk3 and VRK1 are two consecutive steps in this signaling pathway. Plk3 interacts with VRK1, forming a stable complex detected by reciprocal immunoprecipitations and pull-down assays; VRK1 colocalizes with giantin in the Golgi apparatus, as Plk3 also does, forming clearly detectable granules. VRK1 does not phosphorylate Plk3, but Plk3 phosphorylates the C-terminal region of VRK1 in Ser342. VRK1 with substitutions in S342 is catalytically active but blocks Golgi fragmentation, indicating that its specific phosphorylation is necessary for this process. The induction of Golgi fragmentation by MEK1 and Plk3 can be inhibited by kinase-dead VRK1, the knockdown of VRK1 by siVRK1, kinase-dead Plk3, or PD98059, a MEK1 inhibitor. The Plk3-VRK1 kinase module might represent two consecutive steps of a signaling cascade that participates in the regulation of Golgi fragmentation.The Golgi apparatus in mammalian cells is formed by cistern stacks, tubules, and small vesicles, which undergo extensive and sequential fragmentation in mitosis (33). The reorganization of the Golgi apparatus, involving fragmentation, dispersal, and reassembly, is tightly regulated during mitosis (1, 27, 30), and reversible phosphorylation plays a critical role (1, 21), although the components and their sequential organization in the context of the initiation or execution of the signal required for Golgi fragmentation are only partially known.Many signaling pathways are composed of consecutive kinases. Characterization of new signaling pathways requires the identification of their components, the connections between them, and the order in which they are organized. Human VRK1 is a novel serine-threonine kinase that phosphorylates several proteins implicated in cellular responses to stress and DNA damage, such as p53 (5, 20, 40), c-Jun (31), and ATF2 (32), as well as proteins needed for nuclear envelope assembly required at the end of mitosis, such as Baf (25). In addition, VRK1 kinase activity is inhibited by interaction with RanGDP, and this inhibition is relieved by RanGTP, suggesting an asymmetric distribution of its activity within the nucleus and in mitosis (29). These properties suggest that the VRK1 gene plays a role in the regulation of cell cycle initiation and/or progression, consistent with its requirement for entry into the cell cycle, where it behaves as an immediate-early response gene like c-MYC and FOS (36). The loss of VRK1 by use of small interfering RNA (siRNA) induces an early G1 block, before cyclin D1 expression (36), which is accompanied by a reduction in the phospho-retinoblastoma level and an accumulation of cycle inhibitors, such as p27 (36), resulting in a stop in cell cycle progression (36, 40).Several kinases are implicated in the control of cell proliferation and in different mitotic checkpoints; among them are the polo-like kinase (Plk) family, which is a group composed of four proteins (14, 39, 46). One of them, Plk3, contributes as a mediator of DNA damage checkpoint responses, since its kinase activity increases after oxidative stress (43) and induction of DNA damage by ionizing radiomimetic drugs (45). Plk3 physically interacts with and phosphorylates p53 in Ser20, and this interaction increases in response to DNA damage and induces either cell cycle arrest or apoptosis (44) so that genetic stability can be maintained by the prevention of the accumulation of genetic damage. Furthermore, Plk3 interacts with Chk2 (2, 45), an important mediator of DNA damage responses (6, 16), and there is a functional connection between them since Plk3 phosphorylates Chk2 in Ser62 and Ser73, which are necessary for full Chk2 activation by ATM (4). In mitotic cells, Plk3 is localized associated with the spindle poles and mitotic spindles, and deregulated expression of Plk3 induces cell cycle arrest and apoptosis by the perturbation of microtubule integrity (41). In addition, Plk3 expression is induced after mitogenic stimulation, and it is required for mitotic (28) and S-phase (48) entry. Plk3 also regulates Cdc25C (3, 23, 26) and the NF-κB signaling pathway (19). VRK1 phosphorylates p53 in Thr18 (20, 40), a residue phosphorylated in response to taxol, an inhibitor of microtubule polymerization (34).There is a possibility that VRK1 and Plk3 might be connected in some way, since subpopulations of both VRK1 (37) and Plk3 (28) have been detected in the Golgi apparatus near the centrosome, where they colocalize with Golgi markers such as giantin or GM130 (33). Golgi fragmentation can be induced by MEK1 (1, 15), and this signal is partly mediated by Plk3 (28, 42). Moreover, Golgi fragmentation is a required step during mitosis, occurring late in the G2/M phase of the cell cycle (11), and MEK1 is implicated in the activation of this process (1, 15).The common biological aspects of VRK1 and Plk3 proteins and the association of VRK1 and Plk3 subpopulations in the Golgi apparatus led us to think that there might be a functional connection between these two kinases and thus that they might be components in a common signaling pathway. In this work, we explored the possible connection between VRK1 and Plk3 and determined if they were functionally related in a biological process, Golgi fragmentation, in which one of them, Plk3, is already known to participate. This work demonstrates that Plk3 and VRK1 are consecutive components in the signaling pathway that induces Golgi fragmentation in mitosis.  相似文献   

7.
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.
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   

10.
Hepatitis C virus (HCV) RNA genome replicates within the ribonucleoprotein (RNP) complex in the modified membranous structures extended from endoplasmic reticulum. A proteomic analysis of HCV RNP complexes revealed the association of oxysterol binding protein (OSBP) as one of the components of these complexes. OSBP interacted with the N-terminal domain I of the HCV NS5A protein and colocalized to the Golgi compartment with NS5A. An OSBP-specific short hairpin RNA that partially downregulated OSBP expression resulted in a decrease of the HCV particle release in culture supernatant with little effect on viral RNA replication. The pleckstrin homology (PH) domain located in the N-terminal region of OSBP targeted this protein to the Golgi apparatus. OSBP deletion mutation in the PH (ΔPH) domain failed to localize to the Golgi apparatus and inhibited the HCV particle release. These studies suggest a possible functional role of OSBP in the HCV maturation process.Hepatitis C virus (HCV) infection is one of the leading causes of chronic hepatitis. HCV infection is associated with cirrhosis, steatosis, and hepatocellular carcinoma (33). The HCV RNA genome of ∼9.6 kb is translated via an internal ribosome entry site element on the rough endoplasmic reticulum (ER) as a polyprotein precursor of about 3,010 amino acids that is co- and posttranslationally processed by cellular and viral proteases into mature structural and nonstructural (NS) proteins (33). HCV replicates within ribonucleoprotein (RNP) complexes associated with modified ER membranous structures (15). Recent work implicated lipid droplets that emanate from the ER as sites of RNA replication (28, 44). Almost all of the HCV NS proteins along with a variety of cellular factors are associated with the RNP complexes engaged in viral RNA replication (37). It is likely that these NS proteins not only participate in replication process but also are involved in the various steps of virion morphogenesis and assembly. Membrane-associated RNP complexes are generally composed of viral proteins, replicating RNA, host proteins, and altered cellular membranes (1). In this respect, a growing body of evidence implicates the functional role of NS5A in early steps of virion assembly and morphogenesis (3, 27, 45). NS5A is a phosphoprotein that migrates in sodium dodecyl sulfate gels as 56-kDa (basally phosphorylated) and 58-kDa (hyperphosphorylated) forms of proteins. The C-terminal domain III region of NS5A and the phosphorylated residue (Ser457) are important for virion maturation (3, 27, 45). NS5A domain III contains the binding site for viral core protein, indicating the possible involvement of NS5A protein in virus assembly (27). NS5A anchors to the ER membrane by an N-terminal hydrophobic α-helix, and this attachment is needed for its key role(s) in viral replication (10). Studies suggest that phosphorylation of NS5A plays a functional role in viral replication (12). The hyperphosphorylated NS5A reduces its interaction with the human vesicle-associated membrane protein-associated protein A (VAP-A) (12). VAP-A binds both NS5A and NS5B (13, 17). These associations are important for RNA replication (13, 17).HCV alters lipid homeostasis to benefit its infectious processes. Host lipids and their synthesis affect viral infectious process (21, 40, 51, 57). HCV RNA replication can be induced by saturated and monounsaturated fatty acids and inhibited by polyunsaturated fatty acids (18, 21). HCV gene expression induces lipogenesis by stimulating the activation of the sterol regulatory element binding proteins, the master regulators of lipid/fatty acid biosynthetic pathways (51). Reagents that interfere with host lipid biosynthetic pathways abrogate viral replication (21, 57). It has been suggested that HCV utilizes the very-low-density lipoprotein (VLDL) secretion pathway for its viral particle release (14, 19). These studies collectively suggest that host lipid metabolism plays a key role in the viral life cycle including replication, virion assembly, and secretion (56).In the present study, we focus on the functional role of oxysterol binding protein (OSBP) that was identified by proteomic analysis as one of the host factors associated with the HCV RNP complexes. OSBP belongs to a family of the OSBP-related proteins. Originally discovered as a major cytosolic receptor for oxidized cholesterols, it undergoes translocation from the cytosolic/vesicular compartment to the Golgi apparatus upon ligand (hydroxycholesterol) binding (38). OSBP also binds to VAP-A via its FFAT motif (53). Golgi apparatus translocation of OSBP is regulated by the pleckstrin homology (PH) domain. This domain also harbors binding sites for phosphatidylinositol 4-phosphate (PI4P) and phosphatidylinositol 4,5-biphosphate (PI4,5P2) (25). OSBP and OSBP-related proteins are implicated in cholesterol homeostasis, phospholipid metabolism, vesicular transport, and cell signaling (55). OSBP functions as sterol sensor that regulates the transport of ceramide from the ER to the Golgi apparatus for de novo synthesis of sphingomyelin by coordinated action with ceramide transport protein (CERT) (36). OSBP also functions as a scaffolding protein for two phosphatases (phosphatase 2A/HePTP) (49). This complex regulates the activity of extracellular signal-regulate kinase. This cytosolic 440-kDa complex disassembles by the addition of 25-hydroxycholesterol (25-HC) or depletion of cholesterol, both of which cause OSBP translocation to the Golgi compartment (49). Thus, in addition to its role in intracellular trafficking, OSBP appears to regulate cell signaling. We investigated the functional significance of OSBP association with HCV RNP complexes. RNA interference studies support a functional role of OSBP in virion morphogenesis and release process. The OSBP PH domain deletion mutant (ΔPH) failed to localize to the Golgi apparatus and caused an inhibition of the HCV particle release. Our work described herein also demonstrates that the association of OSBP with NS5A may also contribute to the overall HCV maturation process.  相似文献   

11.
12.
13.
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   

14.
15.
Coxsackievirus A9 (CAV9) is a member of the human enterovirus B species within the Enterovirus genus of the family Picornaviridae. It has been shown to utilize αV integrins, particularly αVβ6, as its receptors. The endocytic pathway by which CAV9 enters human cells after the initial attachment to the cell surface has so far been unknown. Here, we present a systematic study concerning the internalization mechanism of CAV9 to A549 human lung carcinoma cells. The small interfering RNA (siRNA) silencing of integrin β6 subunit inhibited virus proliferation, confirming that αVβ6 mediates the CAV9 infection. However, siRNAs against integrin-linked signaling molecules, such as Src, Fyn, RhoA, phosphatidylinositol 3-kinase, and Akt1, did not reduce CAV9 proliferation, suggesting that the internalization of the virus does not involve integrin-linked signaling events. CAV9 endocytosis was independent of clathrin or caveolin-1 but was restrained by dynasore, an inhibitor of dynamin. The RNA interference silencing of β2-microglobulin efficiently inhibited virus infection and caused CAV9 to accumulate on the cell surface. Furthermore, CAV9 infection was found to depend on Arf6 as both silencing of this molecule by siRNA and the expression of a dominant negative construct resulted in decreased virus infection. In conclusion, the internalization of CAV9 to A549 cells follows an endocytic pathway that is dependent on integrin αVβ6, β2-microglobulin, dynamin, and Arf6 but independent of clathrin and caveolin-1.Coxsackievirus A9 (CAV9), a member of the human enterovirus B species in the family Picornaviridae, is a significant human pathogen. It causes infections of the central nervous system, myocarditis, and respiratory diseases and may occasionally cause fatal generalized infections in newborns (6, 22, 26). The CAV9 particle is about 30 nm in diameter and consists of a naked capsid with an icosahedral symmetry, surrounding a positive-sense RNA genome of approximately 7,400 nucleotides (30). The capsid is made up of 60 copies of each of the four proteins VP1 to VP4 and interacts with cell surface integrins during the early stages of infection via arginine-glycine-aspartic acid (RGD) motif that resides in the C terminus of the VP1 protein (11). While CAV9 binds to both integrin αVβ3 and αVβ6 in vitro (53, 61), our recent data show that integrin αVβ6 is the primary receptor of the virus (29).Viruses can utilize several endocytic pathways to enter mammalian cells: macropinocytosis and clathrin-mediated, caveolin-mediated, and clathrin- and caveolin-independent routes (14, 40-41, 50). Recent studies have shown that some of these pathways differ only slightly from each other, and certain endocytic components can participate in more than just one pathway (35, 41, 55). Most of the research carried out on enterovirus endocytosis has been done with echovirus 1 (EV1), coxsackievirus B3 (CBV3), and poliovirus (PV). Recently, Karjalainen et al. showed that EV1 enters SAOS cells via tubulovesicular structures in a dynamin-independent manner that resembles fluid-phase endocytosis and macropinocytosis and that at later stages of infection is targeted to caveosomes (33). EV1 entry to CV-1 cells, on the other hand, was shown to be strictly dynamin dependent (49). PV is endocytosed to HeLa cells by a rapid clathrin- and caveolin-independent pathway, whereas in brain microvascular endothelial cells it uses slower, caveolin- and dynamin-dependent endocytosis (4, 7, 17). CBV3 enters HeLa cells by clathrin-mediated endocytosis (13) and polarized epithelial CaCo-2 cells by a process that combines features of caveolar endocytosis and macropinocytosis (16, 18). Foot-and-mouth-disease virus (FMDV), a member of the Aphthovirus genus of the family Picornaviridae, binds to several αV-integrins, including αVβ6, and is internalized through the clathrin-mediated pathway (5, 19, 31). In the light of these examples, it is evident that enterovirus internalization to human cells is a complex phenomenon wherein a virus may use different mechanisms to enter different cell types.In the CAV9 infection cycle, the steps that follow the initial attachment of the virus to the cell surface integrins are still poorly characterized. An early electron microscopic work by Hecker et al. has shown that single CAV9 particles enter monkey kidney cells in vesicles, which then occasionally fuse and form larger structures (28). Interestingly, they found that most internalized virus particles became eventually trapped in large vacuoles, presumably lysosomes, where they were confined without proceeding to capsid uncoating and RNA release. More recently, a number of cell surface molecules have been proposed to contribute to CAV9 internalization. A subunit of major histocompatibility complex class I (MHC-I) complex, β2-microglobulin (β2M), has been shown to be essential for the infection of several picornaviruses, including CAV9, and it is supposed to have a postattachment role (12, 59, 61). In addition, heat shock 70-kDa protein 5 (HSPA5 protein, also known as glucose-regulated protein 78-kDa, or GRP78) has been suggested to function as a coreceptor for the virus and to mediate CAV9 infection by its interaction with β2M on the cell surface (57). CAV9 entry has been proposed to occur through lipid microdomains, where a number of signaling events takes place (58).The aim of this study was to elucidate the internalization mechanism of CAV9 in A549 human lung carcinoma cells. We used chemical inhibitors, RNA interference (RNAi) silencing, and the expression of dominant negative constructs combined to virus infectivity assays and confocal imaging to examine which cellular molecules are involved in the entry process. The results indicate that CAV9 internalization is dependent on integrin αVβ6, β2M, dynamin 2, and Arf6 (ADP-ribosylation factor 6) but not clathrin or caveolin-1.  相似文献   

16.
17.
18.
19.
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   

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