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 78 毫秒
1.
The replication of plus-strand RNA viruses depends on subcellular membranes. Recent genome-wide screens have revealed that the sterol biosynthesis genes ERG25 and ERG4 affected the replication of Tomato bushy stunt virus (TBSV) in a yeast model host. To further our understanding of the role of sterols in TBSV replication, we demonstrate that the downregulation of ERG25 or the inhibition of the activity of Erg25p with an inhibitor (6-amino-2-n-pentylthiobenzothiazole; APB) leads to a 3- to 5-fold reduction in TBSV replication in yeast. In addition, the sterol biosynthesis inhibitor lovastatin reduced TBSV replication by 4-fold, confirming the importance of sterols in viral replication. We also show reduced stability for the p92pol viral replication protein as well as a decrease in the in vitro activity of the tombusvirus replicase when isolated from APB-treated yeast. Moreover, APB treatment inhibits TBSV RNA accumulation in plant protoplasts and in Nicotiana benthamiana leaves. The inhibitory effect of APB on TBSV replication can be complemented by exogenous stigmasterol, the main plant sterol, suggesting that sterols are required for TBSV replication. The silencing of SMO1 and SMO2 genes, which are orthologs of ERG25, in N. benthamiana reduced TBSV RNA accumulation but had a lesser inhibitory effect on the unrelated Tobacco mosaic virus, suggesting that various viruses show different levels of dependence on sterol biosynthesis for their replication.Plus-stranded RNA [(+)RNA] viruses usurp various intracellular/organellar membranes for their replication. These cellular membranes are thought to facilitate the building of viral factories, promote a high concentration of membrane-bound viral proteins, and provide protection against cellular nucleases and proteases (1, 12, 35, 44). The membrane lipids and proteins may serve as scaffolds for targeting the viral replication proteins or for the assembly of the viral replicase complex. The subcellular membrane also may provide critical lipid or protein cofactors to activate/modulate the function of the viral replicase. Indeed, the formation of spherules, consisting of lipid membranes bended inward and viral replication proteins as well as recruited host proteins, has been demonstrated for several (+)RNA viruses (20, 30, 48). These virus-induced spherules serve as sites of viral replication. Importantly, (+)RNA viruses also induce membrane proliferation that requires new lipid biosynthesis. Therefore, it is not surprising that several genome-wide screens for the identification of host factors affecting (+)RNA virus replication unraveled lipid biosynthesis/metabolism genes (8, 23, 38, 50). However, in spite of these intensive efforts, understanding the roles of various lipids and lipid biosynthesis enzymes and pathways in (+)RNA virus replication is limited.Tomato bushy stunt virus (TBSV) is among the most advanced model systems regarding the identification of host factors affecting (+)RNA virus replication (32). Among the five proteins encoded by the TBSV genome, only the p33 replication cofactor and the p92pol RNA-dependent RNA polymerase (RdRp) are essential for TBSV RNA replication (55). p33 and p92pol are integral membrane proteins, and they are present on the cytosolic surface of the peroxisomes, the site of replicase complex formation and viral RNA replication (30, 42). Electron microscopic images of cells actively replicating tombusviruses have revealed the extensive remodeling of membranes and indicated active lipid biosynthesis (30, 34).Additional support for the critical roles of various lipids in TBSV replication comes from a list of 14 host genes involved in lipid biosynthesis/metabolism, which affected tombusvirus replication and recombination based on systematic genome-wide screens in yeast, a model host. These screens covered 95% of the host genes (16, 38, 50, 51). The 14 identified host genes involved in lipid biosynthesis/metabolism included 8 genes affecting phospholipid biosynthesis, 4 genes affecting fatty acid biosynthesis/metabolism, and 2 genes affecting ergosterol synthesis. These findings suggest that these lipids likely are involved, directly or indirectly, in TBSV replication in yeast.To further understand the roles of cellular membranes, lipids, and host factors in viral (+)RNA replication, we analyzed the importance of sterol biosynthesis in tombusvirus replication. Sterols are ubiquitous and essential membrane components in all eukaryotes, affecting many membrane functions. Sterols regulate membrane rigidity, fluidity, and permeability by interacting with other lipids and proteins within the membranes (4, 5). They also are important for the organization of detergent-resistant microdomains, called lipid rafts (45). The sterol biosynthesis differs in several steps in animals, fungi, and plants, but the removal of two methyl groups at the C-4 position is critical and rate limiting. The C-4 demethylation steps are performed by SMO1 (sterol4α-methyl-oxidase) and SMO2 in plants and by the orthologous ERG25 gene in yeast (10). Accordingly, erg25 mutant yeast accumulates 4,4-dimethylzymosterol, an intermediate in the sterol biosynthesis pathway (3). However, sterol molecules become functional structural components of membranes only after the removal of the two methyl groups at C-4. Therefore, ERG25 is an essential gene for yeast growth.Our previous genome-wide screens for factors affecting tombusvirus replication have identified two sterol synthesis genes, ERG25 and ERG4, that participate in different steps in the sterol biosynthesis pathway (11). In this work, we further characterized the importance of ERG25 in TBSV replication in yeast. The downregulation or pharmacological inhibition of ERG25 in yeast led to a 4- to 5-fold decreased TBSV RNA accumulation. The in vitro activity of the tombusvirus replicase was reduced when isolated from the yeast cells described above. We also found that the stability of p92pol viral replication protein decreased by 3-fold in yeast treated with a chemical inhibitor of ERG25. The inhibition of sterol biosynthesis in plant protoplasts or in plant leaves with a chemical inhibitor or the silencing of SMO1 and SMO2 genes also resulted in a reduction in TBSV RNA accumulation, supporting the roles of sterols in tombusvirus replication in plants as well.  相似文献   

2.
3.
The replication of positive-strand RNA viruses occurs in cytoplasmic membrane-bound virus replication complexes (VRCs). Depending on the virus, distinct cellular organelles such as the endoplasmic reticulum (ER), chloroplast, mitochondrion, endosome, and peroxisome are recruited for the formation of VRC-associated membranous structures. Previously, the 6,000-molecular-weight protein (6K) of plant potyviruses was shown to be an integral membrane protein that induces the formation of 6K-containing membranous vesicles at endoplasmic reticulum (ER) exit sites for potyvirus genome replication. Here, we present evidence that the 6K-induced vesicles predominantly target chloroplasts, where they amalgamate and induce chloroplast membrane invaginations. The vesicular transport pathway and actomyosin motility system are involved in the trafficking of the 6K vesicles from the ER to chloroplasts. Viral RNA, double-stranded RNA, and viral replicase components are concentrated at the 6K vesicles that associate with chloroplasts in infected cells, suggesting that these chloroplast-bound 6K vesicles are the site for potyvirus replication. Taken together, these results suggest that plant potyviruses sequentially recruit the ER and chloroplasts for their genome replication.The replication of eukaryotic positive-strand RNA viruses in infected cells is closely associated with unique virus-induced intracellular membranous vesicles (22). These membranous vesicles have been proposed to provide a scaffold for anchoring the virus replication complex (VRC), confine the process of RNA replication to a specific safeguarded cytoplasmic location, and prevent the activation of certain host defense mechanisms that can be triggered by double-stranded RNA (dsRNA) intermediates during virus replication (33, 47). Depending on the type of virus, the virus-induced membranous vesicles are derived from various intracellular organelles in the host. Many plant and animal viruses remodel and utilize the endoplasmic reticulum (ER) in VRCs (1, 6, 17, 33, 34, 36, 38, 39, 46). Other cellular organelles such as endosomes, lysomes, chloroplasts, peroxisomes, and mitochondria have also been suggest to be the replication site for togaviruses, tymoviruses, and tombusviruses, respectively (25, 27, 31). Given that the ER appears to be the site where the host cell translation machinery is hijacked for the biosynthesis of the first set of viral proteins, the subcellular location of virus replication (either in the vicinity of the ER or elsewhere) and the mechanism of transport to locations other than the ER are poorly understood.Plant potyviruses, accounting for ∼30% of known plant viruses including many agriculturally important viruses, e.g., Turnip mosaic virus (TuMV), Maize dwarf mosaic virus (MDMV), Tobacco etch virus (TEV), and Potato virus Y (PVY), are related to picornaviruses and picorna-like viruses (20, 21, 43). The potyviral genome is a single-stranded positive-sense RNA of about 10 kb in length and encodes at least 11 mature viral proteins (8, 43). Of these 11 proteins, the 6-kDa protein (designated 6K or 6K2) contains a central hydrophobic domain (35). In seminal work, Carrington and colleagues determined that 6K induces the formation of the ER-derived vesicles for TEV replication (35, 38). More recently, viral proteins required for replication and several host factors, namely, eukaryotic initiation factor (isoform) 4E, poly(A)-binding protein, eukaryotic elongation factor 1A, and heat shock cognate 70-3 protein, have been shown to associate with the TuMV 6K-induced vesicles (9, 41), raising the possibility that the potyviral 6K vesicles represent sites of viral genome replication. Furthermore, we have demonstrated that the biogenesis of the potyviral 6K vesicles occurs at COPII-accumulating ER exit sites (ERES) on the ER membrane (45). In this study, we further studied the trafficking of 6K-induced vesicles and found that the 6K-induced mobile vesicles trafficked predominantly from the ER to the periphery of chloroplasts. We show that these 6K vesicles docked on the outer chloroplast envelope and induced chloroplast invaginations. The chloroplast-associated 6K vesicles contained viral replicase components and dsRNA and were concentrated with viral RNA. We provide evidence that the early secretory pathway and actomyosin motility system were required for the trafficking of 6K vesicles from the ER to chloroplasts. These results suggest that plant potyviruses sequentially recruit the ER and chloroplasts for their genome replication.  相似文献   

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.
7.
8.
9.
10.
11.
12.
The mechanisms underlying HIV-1 control by protective HLA class I alleles are not fully understood and could involve selection of escape mutations in functionally important Gag epitopes resulting in fitness costs. This study was undertaken to investigate, at the population level, the impact of HLA-mediated immune pressure in Gag on viral fitness and its influence on HIV-1 pathogenesis. Replication capacities of 406 recombinant viruses encoding plasma-derived Gag-protease from patients chronically infected with HIV-1 subtype C were assayed in an HIV-1-inducible green fluorescent protein reporter cell line. Viral replication capacities varied significantly with respect to the specific HLA-B alleles expressed by the patient, and protective HLA-B alleles, most notably HLA-B*81, were associated with lower replication capacities. HLA-associated mutations at low-entropy sites, especially the HLA-B*81-associated 186S mutation in the TL9 epitope, were associated with lower replication capacities. Most mutations linked to alterations in replication capacity in the conserved p24 region decreased replication capacity, while most in the highly variable p17 region increased replication capacity. Replication capacity also correlated positively with baseline viral load and negatively with baseline CD4 count but did not correlate with the subsequent rate of CD4 decline. In conclusion, there is evidence that protective HLA alleles, in particular HLA-B*81, significantly influence Gag-protease function by driving sequence changes in Gag and that conserved regions of Gag should be included in a vaccine aiming to drive HIV-1 toward a less fit state. However, the long-term clinical benefit of immune-driven fitness costs is uncertain given the lack of correlation with longitudinal markers of disease progression.There is broad heterogeneity in the ability of HIV-infected individuals to control virus replication, ranging from elite controllers, who maintain undetectable viral loads without treatment, to rapid progressors, who progress to AIDS within 2 years of infection (9, 22, 32). Many interrelated factors, including host and viral genetic factors involved in antiviral immunity and the viral life cycle, may partially account for the differences in the course of disease progression (10, 11, 30, 41). The complex interplay between host genetic factors and viral factors is exemplified by human leukocyte antigen (HLA) class I-restricted cytotoxic T-lymphocyte (CTL) responses, which exert considerable immune pressure on the virus, resulting in escape mutations that affect the interaction of viral and host proteins, thereby influencing infection outcome.The exact mechanisms by which some HLA class I alleles, such as HLA-B*57 and HLA-B*27, are associated with slower progression to AIDS, while others, such as B*5802 and B*18, are associated with accelerated disease progression (6, 20, 42), are unclear. The magnitude and/or breadth of HLA-restricted CTL responses to the conserved Gag protein has been correlated inversely with disease progression or markers of disease progression in several studies (12, 21, 28, 31, 35, 43, 46), although there are some exceptions (4, 16, 37), while preferential targeting of the highly variable envelope protein (as occurs in HLA-B*5802-positive individuals) correlates with higher viral loads (21, 29). Protective HLA alleles restrict CTL responses that impose a strong selection pressure on a few specific Gag p24 epitopes, resulting in escape mutations (14) for which fitness costs have been demonstrated either through site-directed mutations introduced into a reference strain background (2, 8, 25, 38) or through in vivo reversion of these mutations after transmission to an HLA-mismatched individual (8, 24). Recent evidence suggests that Gag escape mutations with a fitness cost, particularly those in p24, are a significant determinant of disease progression: the transmitted number of HLA-B-associated polymorphisms in Gag was found to significantly impact the viral set point in recipients (although an associated fitness cost was not shown) (7, 15), and in a small number of infants, decreased fitness of the transmitted virus with HLA-B*5703/5801-selected mutations in Gag p24 epitopes resulted in slower disease progression (33, 39). Also, the number of reverting Gag mutations (thought to revert as a consequence of fitness costs) associated with individual HLA-B alleles was strongly correlated with the HLA-linked viral set point in chronically infected patients (26). A recent in vitro study showed that HLA-associated variation in Gag-protease, with resulting reduced replication capacity, may contribute to viral control in HIV-1 subtype B-infected elite controllers (27). Taken together, these studies suggest that CTL responses restricted by favorable HLA alleles select for escape mutations in conserved epitopes, particularly those in Gag, resulting in a fitness cost to HIV and therefore at least partly explaining the slower disease progression in individuals carrying these alleles.To date, many of the studies investigating the fitness cost of Gag escape mutations and their clinical relevance have concentrated on escape mutations associated with protective HLA alleles, have not assessed fitness consequences in the natural sequence background (in the presence of other escape and compensatory mutations), and/or have focused on a limited number of patients. Most importantly, the majority of studies have focused on HIV-1 subtype B. The present study is the first to use a large population-based approach and clinically derived Gag-protease sequences to investigate comprehensively the relationships between immune-driven sequence variation in Gag, viral replication capacity, and markers of disease progression in chronic infection with HIV-1 subtype C, the most predominant subtype in the epidemic. We assayed the replication capacity of recombinant viruses encoding patient Gag-protease in an HIV-1-inducible green fluorescent protein (GFP) reporter cell line and found associations between lower replication capacities, protective HLA alleles, protective HLA-associated mutations, lower baseline viral loads, and higher baseline CD4 counts. However, Gag-protease replication capacity did not correlate with the subsequent rate of CD4 decline.  相似文献   

13.
The parvovirus adeno-associated virus (AAV) contains a small single-stranded DNA genome with inverted terminal repeats that form hairpin structures. In order to propagate, AAV relies on the cellular replication machinery together with functions supplied by coinfecting helper viruses such as adenovirus (Ad). Here, we examined the host cell response to AAV replication in the context of Ad or Ad helper proteins. We show that AAV and Ad coinfection activates a DNA damage response (DDR) that is distinct from that seen during Ad or AAV infection alone. The DDR was also triggered when AAV replicated in the presence of minimal Ad helper proteins. We detected autophosphorylation of the kinases ataxia telangiectasia mutated (ATM) and DNA-dependent protein kinase catalytic subunit (DNA-PKcs) and signaling to downstream targets SMC1, Chk1, Chk2, H2AX, and XRCC4 and multiple sites on RPA32. The Mre11 complex was not required for activation of the DDR to AAV infection. Additionally, we found that DNA-PKcs was the primary mediator of damage signaling in response to AAV replication. Immunofluorescence revealed that some activated damage proteins were found in a pan-nuclear pattern (phosphorylated ATM, SMC1, and H2AX), while others such as DNA-PK components (DNA-PKcs, Ku70, and Ku86) and RPA32 accumulated at AAV replication centers. Although expression of the large viral Rep proteins contributed to some damage signaling, we observed that the full response required replication of the AAV genome. Our results demonstrate that AAV replication in the presence of Ad helper functions elicits a unique damage response controlled by DNA-PK.Replication of viral genomes produces a large amount of extrachromosomal DNA that may be recognized by the cellular DNA damage machinery. This is often accompanied by activation of DNA damage response (DDR) signaling pathways and recruitment of cellular repair proteins to sites of viral replication. Viruses therefore provide good model systems to study the recognition and response to DNA damage (reviewed in reference 48). The Mre11/Rad50/Nbs1 (MRN) complex functions as a sensor of chromosomal DNA double-strand breaks (DSBs) and is involved in activation of damage signaling (reviewed in reference 41). The MRN complex also localizes to DNA DSBs and is found at viral replication compartments during infection with a number of DNA viruses (6, 40, 47, 70, 75, 77, 87, 93). The phosphatidylinositol 3-kinase-like kinases (PIKKs) ataxia telangiectasia mutated (ATM), ATM and Rad3-related kinase (ATR), and the catalytic subunit of the DNA-dependent protein kinase (DNA-PKcs) are involved in the signal transduction cascades activated by DNA damage (reviewed in references 43, 51, and 71). These kinases respond to distinct types of damage and regulate DSB repair during different phases of the cell cycle (5), either through nonhomologous end-joining (NHEJ) or homologous recombination pathways (reviewed in references 63, 81, and 86). The DNA-PK holoenzyme is composed of DNA-PKcs and two regulatory subunits, the Ku70 and Ku86 heterodimer. DNA-PK functions with XRCC4/DNA ligase IV to repair breaks during NHEJ, and works with Artemis to process DNA hairpin structures during VDJ recombination and during a subset of DNA DSB events (46, 50, 86). While the kinase activity of DNA-PKcs leads to phosphorylation of a large number of substrates in vitro as well as autophosphorylation of specific residues (reviewed in references 16 and 85), it is currently unclear how DNA-PKcs contributes to signaling in cells upon different types of damage.The adeno-associated virus (AAV) genome consists of a molecule of single-stranded DNA with inverted terminal repeats (ITRs) at both ends that form double-hairpin structures due to their palindromic sequences (reviewed in reference 52). The ITRs are important for replication and packaging of the viral genome and for integration into the host genome. Four viral Rep proteins (Rep78, Rep68, Rep52, and Rep40) are also required for replication and packaging of the AAV genome into virions assembled from the Cap proteins. Although the Rep and Cap genes are replaced in recombinant AAV vectors (rAAV) that retain only the ITRs flanking the gene of interest, these vectors can be replicated by providing Rep in trans (reviewed in reference 7). Productive AAV infection requires helper functions supplied by adenovirus (Ad) or other viruses such as herpes simplex virus (HSV) (reviewed in reference 27), together with components of the host cell DNA replication machinery (54, 55, 58). In the presence of helper viruses or minimal helper proteins from Ad or HSV, AAV replicates in the nucleus at centers where the viral DNA and Rep proteins accumulate (35, 76, 84, 89). Cellular and viral proteins involved in AAV replication, including replication protein A (RPA), Ad DNA-binding protein (DBP), and HSV ICP8, localize with Rep proteins at these viral centers (29, 33, 76).A number of published reports suggest associations between AAV and the cellular DNA damage machinery. For example, transduction by rAAV vectors is increased by genotoxic agents and DNA damaging treatments (1, 62, 91) although the mechanisms involved remain unclear. Additionally, the ATM kinase negatively regulates rAAV transduction (64, 92), and we have shown that the MRN complex poses a barrier to both rAAV transduction and wild-type AAV replication (11, 67). UV-inactivated AAV particles also appear to activate a DDR involving ATM and ATR kinases that perturbs cell cycle progression (39, 60, 88). It has been suggested that this response is provoked by the AAV ITRs (60) and that UV-treated particles mimic stalled replication forks in infected cells (39). In addition to AAV genome components, the viral Rep proteins have been observed to exhibit cytotoxicity and induce S-phase arrest (3, 65).The role of cellular repair proteins in AAV genome processing has also been explored by examining the molecular fate of rAAV vectors, which are converted into circular and concatemeric forms that persist episomally (18, 19, 66). Proteins shown to regulate circularization in cell culture include ATM and the MRN complex (14, 64), while in vivo experiments using mouse models have implicated ATM and DNA-PK in this process (14, 20, 72). Additionally, DNA-PKcs and Artemis have recently been shown to cleave the ITR hairpins of rAAV vectors in vivo in a tissue-dependent manner (36). Despite these studies, it is not clear how damage response factors function together and how they impact AAV transduction and replication in human cells.In this study we examined the cellular response to AAV replication in the context of Ad infection or helper proteins. We show that coinfection with AAV and Ad activates a DDR that is distinct from that seen during infection with Ad alone. The ATM and DNA-PKcs damage kinases are activated and signal to downstream substrates, but the response does not require the MRN complex and is primarily mediated by DNA-PKcs. Although expression of the large Rep proteins induced some DDR events, full signaling appeared to require AAV replication and was accompanied by accumulation of DNA-PK at viral replication compartments. Our results demonstrate that AAV replication induces a unique DNA damage signal transduction response and provides a model system for studying DNA-PK.  相似文献   

14.
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   

15.
16.
The ectopic overexpression of Bcl-2 restricts both influenza A virus-induced apoptosis and influenza A virus replication in MDCK cells, thus suggesting a role for Bcl-2 family members during infection. Here we report that influenza A virus cannot establish an apoptotic response without functional Bax, a downstream target of Bcl-2, and that both Bax and Bak are directly involved in influenza A virus replication and virus-induced cell death. Bak is substantially downregulated during influenza A virus infection in MDCK cells, and the knockout of Bak in mouse embryonic fibroblasts yields a dramatic rise in the rate of apoptotic death and a corresponding increase in levels of virus replication, suggesting that Bak suppresses both apoptosis and the replication of virus and that the virus suppresses Bak. Bax, however, is activated and translocates from the cytosol to the mitochondria; this activation is required for the efficient induction of apoptosis and virus replication. The knockout of Bax in mouse embryonic fibroblasts blocks the induction of apoptosis, restricts the infection-mediated activation of executioner caspases, and inhibits virus propagation. Bax knockout cells still die but by an alternative death pathway displaying characteristics of autophagy, similarly to our previous observation that influenza A virus infection in the presence of a pancaspase inhibitor leads to an increase in levels of autophagy. The knockout of Bax causes a retention of influenza A virus NP within the nucleus. We conclude that the cell and virus struggle to control apoptosis and autophagy, as appropriately timed apoptosis is important for the replication of influenza A virus.The pathology of influenza A virus infection usually arises from acute lymphopenia and inflammation of the lungs and airway columnar epithelial cells (23, 38). Influenza A virus induces apoptotic death in infected epithelial, lymphocyte, and phagocytic cells, and apoptosis is a source of tissue damage during infection (3, 22, 33) and increased susceptibility to bacterial pathogens postinfection (31). While the induction of apoptosis by influenza A virus has been well documented (4, 19-21, 28, 33, 37), the mechanisms of this interaction are not well understood. Two viral proteins, NS1 and PB1-F2, have been associated with viral killing of cells. NS1, originally characterized as being proapoptotic (34), was later identified as being an interferon antagonist, inhibiting the activation of several key antiviral responses and restricting the apoptotic response to infection (1, 10, 15, 18, 35, 39, 46). In contrast, PB1-F2 induces apoptosis primarily by localizing to the outer mitochondrial membrane, promoting cytochrome c release, and triggering the apoptotic cascade (43). This effect, however, is typically restricted to infected monocytes, leading to the hypothesis that PB1-F2 induces apoptosis specifically to clear the landscape of immune responders (5, 44). Although PB1-F2 activity does not directly manipulate virus replication or virus-induced apoptosis, PB1-F2 localization to the mitochondrial membrane during infection potentiates the apoptotic response in epithelial and fibroblastic cells through tBID signaling with proapoptotic Bcl-2 family protein members Bax and Bak (22, 43, 44).The Bcl-2 protein family consists of both pro- and antiapoptotic members that regulate cytochrome c release during mitochondrion-mediated apoptosis through the formation of pore-like channels in the outer mitochondrial membrane (12, 16). During the initiation of mitochondrion-mediated apoptosis, cytoplasmic Bid is cleaved to form tBID. This, in turn, activates proapoptotic Bax and Bak (40), which drive cytochrome c release and subsequent caspase activation. Bak is constitutively associated with the mitochondrial membrane, whereas inactive Bax is primarily cytosolic, translocating to the outer mitochondrial membrane only after activation (6). The activation of Bax and Bak results in homo- and heterodimer formation at the outer mitochondrial membrane, generating pores that facilitate mitochondrial membrane permeabilization and cytochrome c release (14, 17), leading to caspase activation and the apoptotic cascade (8). Antiapoptotic members of the Bcl-2 protein family, including Bcl-2, inhibit the activation of proapoptotic Bax and Bak primarily by sequestering inactive Bax and Bak monomers via interactions between their BH3 homology domains (7).Bcl-2 expression has been linked to decreased viral replication rates (26). Bcl-2 overexpression inhibits influenza A virus-induced cell death and reduces the titer and spread of newly formed virions (29). The activation of caspase-3 in the absence of sufficient Bcl-2 is critical to the influenza A virus life cycle. Both Bcl-2 expression and the lack of caspase activation during infection lead to the nuclear accumulation of influenza virus ribonucleoprotein (RNP) complexes, thereby leading to the improper assembly of progeny virions and a marked reduction in titers of infectious virus (26, 41, 42, 45).Here we show that influenza A virus induces mitochondrion-mediated (intrinsic-pathway) apoptosis signaled specifically through Bax and that this Bax signaling is essential for the maximum efficiency of virus propagation. In contrast, Bak expression is strongly downregulated during infection. Cells lacking Bak (while expressing Bax) display a much more severe apoptotic phenotype in response to infection and produce infectious virions at a higher rate than the wild type (WT), suggesting that Bak, which can suppress viral replication, is potentially downregulated by the virus. Our results indicate essential and opposing roles for Bax and Bak in both the response of cells to influenza A virus infection and the ability of the virus to maximize its own replicative potential.  相似文献   

17.
18.
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   

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