首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Homologous recombination is an important biological process that facilitates genome rearrangement and repair of DNA double-strand breaks (DSBs). The induction of Epstein-Barr virus (EBV) lytic replication induces ataxia telangiectasia-mutated (ATM)-dependent DNA damage checkpoint signaling, leading to the clustering of phosphorylated ATM and Mre11/Rad50/Nbs1 (MRN) complexes to sites of viral genome synthesis in nuclei. Here we report that homologous recombinational repair (HRR) factors such as replication protein A (RPA), Rad51, and Rad52 as well as MRN complexes are recruited and loaded onto the newly synthesized viral genome in replication compartments. The 32-kDa subunit of RPA is extensively phosphorylated at sites in accordance with those with ATM. The hyperphosphorylation of RPA32 causes a change in RPA conformation, resulting in a switch from the catalysis of DNA replication to the participation in DNA repair. The levels of Rad51 and phosphorylated RPA were found to increase with the progression of viral productive replication, while that of Rad52 proved constant. Furthermore, biochemical fractionation revealed increases in levels of DNA-bound forms of these HRRs. Bromodeoxyuridine-labeled chromatin immunoprecipitation and PCR analyses confirmed the loading of RPA, Rad 51, Rad52, and Mre11 onto newly synthesized viral DNA, and terminal deoxynucleotidyltransferase-mediated dUTP nick end labeling analysis demonstrated DSBs in the EBV replication compartments. HRR factors might be recruited to repair DSBs on the viral genome in viral replication compartments. RNA interference knockdown of RPA32 and Rad51 prevented viral DNA synthesis remarkably, suggesting that homologous recombination and/or repair of viral DNA genome might occur, coupled with DNA replication to facilitate viral genome synthesis.Replication protein A (RPA), the eukaryotic single-stranded DNA (ssDNA)-binding protein, is a heterotrimeric complex composed of three tightly associated subunits of 70, 32, and 14 kDa (referred as to RPA70, RPA32, and RPA14, respectively) that is essential for DNA replication, recombination, and all major types of DNA repair (4). RPA participates in such diverse pathways through its ability to interact with DNA and numerous proteins involved in its processing. During DNA replication, RPA associates with ssDNA at forks and facilitates nascent-strand DNA synthesis by replicative DNA polymerases localized at replication foci during S phase. Under DNA-damaging conditions, RPA binds to ssDNA at damaged sites and interacts with repair and recombination components to process double-strand DNA breaks (DSBs) and other lesions (6, 14, 21, 32, 38, 41).RPA undergoes both DNA damage-independent and -dependent phosphorylation on the N-terminal 33 residues of RPA32. Unstressed cell cycle-dependent phosphorylation occurs during the G1/S-phase transition and in M phase, primarily at the conserved cyclin-CDK phosphorylation sites of Ser-23 and Ser-29 in the N terminus of the RPA32 subunit (13, 15). In contrast, stress-induced hyperphosphorylation of RPA is much more extensive. Nine potential phosphorylation sites within the N-terminal domain of RPA32, Ser-4, Ser-8, Ser-11/Ser-12/Ser-13, Thr-21, Ser-23, Ser-29, and Ser-33, in response to DNA-damaging agents, have been suggested (33, 54). Although this region of RPA32 is not required for the ssDNA-binding activity of RPA (5, 22), a phosphorylation-induced subtle conformation change in RPA, resulting from altered intersubunit interactions, regulates the interaction of RPA with both interacting proteins and DNA (30). The hyperphosphorylated form of RPA32 is unable to localize to replication centers in normal cells, while binding to DNA damage foci is unaffected (46). Therefore, RPA phosphorylation following damage is thought to both prevent RPA from catalyzing DNA replication and potentially serve as a marker to recruit repair factors to sites of DNA damage. RPA localizes to nuclear foci where DNA repair is occurring after DNA damage and is essential for multiple DNA repair pathways, participating in damage recognition, excision, and resynthesis reactions (4, 56).Mammalian cells can repair DSBs by homologous recombination (HR) or by nonhomologous end joining. HR is an accurate repair process, the first step of which is the resection of the 5′ ends of the DSB to generate 3′ ssDNA overhangs. This reaction is carried out by the Mre11/Rad50/Nbs1 (MRN) complex, which not only functions as a damage sensor upstream of ataxia telangiectasia-mutated (ATM)/ATM-Rad3-related (ATR) activation but also plays a role in DSB repair (4). RPA and members of the RAD52 epistasis group of gene products, such as Rad51, Rad52, and Rad54, bind to the resulting 3′ ssDNA strands and form a helical, nucleoprotein filament that facilitates the invasion of a damaged DNA strand into the homologous double-stranded DNA partner. The human Rad51 protein is a structural and functional homolog of the Escherichia coli RecA protein, which promotes homologous pairing and strand transfer reactions in vitro. Both Rad51 and Rad52 bind specifically to the terminal regions of tailed duplex DNA, the substrate thought to initiate recombination in vivo. Furthermore, nucleoprotein filaments of Rad51, formed on tailed DNA, catalyze strand invasion of homologous duplex DNA in a reaction that is stimulated by Rad52 and RPA (3).Epstein-Barr virus (EBV) is a human herpesvirus that infects B lymphocytes, inducing their continuous proliferation. In B-lymphoblastoid cell lines, there is no production of virus particles, which is termed latent infection (52). Reactivation from latency is characterized by the expression of lytic genes, and one of the first detectable changes is the expression of the BZLF1 immediate-early gene product, which trans-activates viral promoters (16), leading to an ordered cascade of viral early and late gene expression. This lytic EBV DNA replication occurs in discrete sites in nuclei, called replication compartments, in which seven viral replication proteins are assembled (44). The viral genome is amplified several hundredfold by the viral replication machinery and is thought to generate highly branched replication intermediates through HR coupled with viral DNA replication (48). With the progression of lytic replication, the replication compartments become larger and appeared to fuse to form large globular structures that eventually filled the nucleus at late stages of infection (8, 45).We previously isolated latently EBV-infected Tet-BZLF1/B95-8 cells in which the exogenous BZLF1 protein is conditionally expressed under the control of a tetracycline-regulated promoter, leading to a highly efficient induction of lytic replication (28). Using this system, we have demonstrated that the induction of the EBV lytic program results in the inhibition of replication of cellular DNA in spite of the replication of viral DNA (28) and elicits a cellular DNA damage response, with the activation of the ATM-Chk2-p53 DNA damage transduction pathway (29). The DNA damage sensor MRN complex and phosphorylated ATM are recruited and retained in viral replication compartments (29).Here we report that RPA32 is extensively phosphorylated after EBV lytic replication is induced, with the phosphorylation sites in accordance with those for ATM. Phosphorylated RPA, Rad51, and Rad52, which are involved in HR repair (HRR), are recruited and retained in viral replication compartments as well as the MRN complex. Furthermore, DSBs could be demonstrated to occur during viral genome synthesis in the EBV replication compartments. HRR factors might be recruited to repair DSBs on the viral genome in viral replication compartments. RNA interference (RNAi) knockdown of RPA32 and Rad51 prevented viral DNA synthesis remarkably, suggesting that HR and/or repair of viral DNA genome might occur, coupled with DNA replication, to facilitate viral genome synthesis.  相似文献   

2.
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   

3.
Although the herpes simplex virus type 1 (HSV-1) genome might be expected to induce a DNA damage response, the ATR kinase is not activated in infected cells. We previously proposed that spatial uncoupling of ATR from its interaction partner, ATRIP, could be the basis for inactivation of the ATR kinase in infected cells; however, we now show that ATR and ATRIP are in fact both recruited to HSV-1 replication compartments and can be coimmunoprecipitated from infected-cell lysates. ATRIP and replication protein A (RPA) are recruited to the earliest detectable prereplicative sites, stage II microfoci. In a normal cellular DNA damage response, ATR/ATRIP are recruited to stretches of RPA-coated single-stranded DNA in an RPA- and kinase-dependent manner, resulting in the phosphorylation of RPA by ATR in damage foci. In contrast, in HSV-1-infected cells, RPA is not phosphorylated, and endogenous phosphorylated RPA is excluded from stage II microfoci; in addition, the recruitment of ATR/ATRIP is independent of RPA and the kinase activity of ATR. Furthermore, we show that ATR/ATRIP play a beneficial role in viral gene expression and virus production. Although ICP0 has been shown to be important for partial inactivation of other cellular DNA repair pathways, we show that ICP0 is not responsible for the inactivation of ATR signaling and, furthermore, that neither ATR nor ATRIP is a target of ICP0 degradation. Thus, ATR and ATRIP may function outside the context of the canonical ATR damage signaling pathway during HSV-1 infection to participate in the viral life cycle.Herpes simplex virus type 1 (HSV-1) is a large linear double-stranded DNA virus that replicates in the nucleus of the host cell. The incoming viral genome contains nicks and gaps (42), and cellular DNA repair machinery might be expected to recognize it as damaged, resulting in the activation of one or more cellular DNA damage pathways. Activation of DNA damage response pathways can result not only in repair of the damaged DNA but also in cell cycle arrest, gene silencing, and apoptosis (9). The later outcomes could result in suppression of viral gene expression and DNA replication and thus have negative consequences for lytic infection. Activation of a cellular DNA damage response during viral infection could, therefore, represent a form of intrinsic antiviral immunity (14, 15). On the other hand, HSV-1 and other DNA viruses which replicate in the nucleus have also been shown to utilize cellular DNA repair machinery to promote productive infection (28). Thus, HSV-1 has apparently evolved to manipulate the host DNA damage response by utilizing some components and inactivating others in an attempt to create an environment conducive to lytic viral infection.The cellular DNA damage response is regulated by the three phosphoinositide 3-kinase-related kinases (PIKKs), DNA-PK (DNA-dependent protein kinase), ATM (ataxia-telangiectasia-mutated), and ATR (ATM and Rad3-related) (1, 9). DNA-PK and ATM respond predominantly to double-strand breaks, and ATR responds to stalled replication forks and long stretches of single-stranded DNA (ssDNA). DNA-PK is required for nonhomologous end joining (NHEJ), while ATM activation promotes homologous recombination. Interestingly, in some cell types, the catalytic subunit of DNA-PK (DNA-PKcs) is proteolytically degraded during infection by the immediate-early (IE) protein ICP0, a viral E3 ubiquitin ligase (25, 37), thereby resulting in the probable inactivation of the NHEJ pathway. ATM kinase activity, on the other hand, is activated during HSV-1 infection once viral DNA replication is initiated (26, 47, 56). Despite phosphorylation of several ATM targets, ATM signaling is also modulated by ICP0, which degrades the ubiquitin ligases RNF8 and RNF168. The function of these ubiquitin ligases is to promote the tethering of ATM pathway proteins at sites of cellular DNA damage (27). Thus, ICP0 functions to partially inactivate portions of both the DNA-PK- and ATM-mediated repair pathways.During a cellular DNA damage response, ATM activation and processing of DNA ends generate ssDNA adjacent to double-stranded DNA (dsDNA), a structure that is known to activate ATR (9, 38). The ssDNA is coated by the cellular ssDNA binding protein, replication protein A (RPA), which then serves to recruit ATR through a direct interaction with ATR-interacting protein (ATRIP) (4, 12, 58). ATR signaling results in the phosphorylation of many substrates, including RPA and Chk1. During HSV-1 infection, the ATR substrates RPA and Chk1 are not phosphorylated (47, 54-56), indicating that ATR signaling may be disabled.A hallmark of HSV-1 infection is the reorganization of the infected-cell nucleus, resulting in the formation of large globular replication compartments as well as the rearrangement of cellular proteins involved in several homeostatic pathways. In addition to cellular DNA repair proteins, HSV-1 infection also causes the reorganization of components of the cellular protein quality control pathways, resulting in the formation of virus-induced chaperone-enriched (VICE) domains, which act to maintain nuclear protein quality control during infection (31). Viral gene expression, DNA replication, and encapsidation of viral genomes occur in replication compartments (24, 39, 41). In this work we revisit the study of proteins recruited to and restricted from replication compartments in an attempt to better understand how HSV-1 manipulates components of the cellular DNA damage response for its own benefit.  相似文献   

4.
Adenovirus infection activates cellular DNA damage response and repair pathways. Viral proteins that are synthesized before viral DNA replication prevent recognition of viral genomes as a substrate for DNA repair by targeting members of the sensor complex composed of Mre11/Rad50/NBS1 for degradation and relocalization, as well as targeting the effector protein DNA ligase IV. Despite inactivation of these cellular sensor and effector proteins, infection results in high levels of histone 2AX phosphorylation, or γH2AX. Although phosphorylated H2AX is a characteristic marker of double-stranded DNA breaks, this modification was widely distributed throughout the nucleus of infected cells and was coincident with the bulk of cellular DNA. H2AX phosphorylation occurred after the onset of viral DNA replication and after the degradation of Mre11. Experiments with inhibitors of the serine-threonine kinases ataxia telangiectasia mutated (ATM), AT- and Rad3-related (ATR), and DNA protein kinase (DNA-PK), the kinases responsible for H2AX phosphorylation, indicate that H2AX may be phosphorylated by ATR during a wild-type adenovirus infection, with some contribution from ATM and DNA-PK. Viral DNA replication appears to be the stimulus for this phosphorylation event, since infection with a nonreplicating virus did not elicit phosphorylation of H2AX. Infected cells also responded to high levels of input viral DNA by localized phosphorylation of H2AX. These results are consistent with a model in which adenovirus-infected cells sense and respond to both incoming viral DNA and viral DNA replication.Cellular DNA damage response pathways protect and preserve the integrity of the genome. These pathways, which are activated in response to various forms of DNA damage, involve a number of proteins that participate in both DNA repair and cell cycle progression (62). The serine-threonine kinases ataxia telangiectasia mutated (ATM), AT- and Rad3-related (ATR), and DNA protein kinase (DNA-PK) are activated in response to distinct types of damage. The ATM pathway is activated primarily by double-stranded DNA breaks (4, 30). DNA-PK acts in conjunction with the DNA ligase IV/XRCC4 complex to mediate the ligation of double-stranded breaks through nonhomologous end joining (34). The ATR pathway can be activated in response to a wide range of genotoxic stresses, such as base or nucleotide excision, double-stranded breaks, or single-stranded breaks. Activation of ATR is generally thought to occur via the recognition of single-stranded tracks of DNA (63). Each of these pathways leads to the phosphorylation and activation of a number of cellular proteins such as the variant histone H2AX, checkpoint kinases 1 and 2 (Chk1 and Chk2), and Nijmegen break syndrome protein 1 (NBS1), among others (62). Signals transmitted by a cascade of phosphorylation events result in cell cycle arrest and the accumulation of repair protein complexes at sites of DNA damage.Upon recognition of a double-stranded DNA break by the cell, H2AX is phosphorylated on an extended C-terminal tail at serine 139 by the phosphatidylinositol 3-kinase (PI3K)-related kinases ATM, ATR, and DNA-PK (9, 41, 44, 58). Considered one of the earliest indications of a double-stranded DNA break, phosphorylated H2AX (γH2AX) acts as a scaffolding protein to which a number of DNA repair factors can dock to facilitate repair of the damaged DNA (36, 42, 53). Areas of phosphorylated H2AX, termed γH2AX foci, are enriched for proteins involved in both homologous recombination and nonhomologous end joining, such as NBS1, BRCA1 (42), and Mdc1 (24, 50).Although adenovirus is able to activate both ATM and ATR pathways (11), adenoviral proteins limit the extent and consequences of signaling through these pathways. The E1B-55K and E4orf6 proteins form an E3 ubiquitin ligase with the cellular proteins Cullin-5, elongins B and C, and Rbx1 (28, 43). This complex targets key cellular proteins involved in cellular response to DNA damage, including p53 (28, 43), Mre11 (51), and DNA ligase IV (3). The E4orf3 gene product targets cellular proteins central to both the cellular DNA damage response and the antiviral response. The E4orf3 protein of species C adenoviruses alters the localization of Mre11/Rad50/NBS1 (MRN) complex members within the nucleus to prevent association with centers of viral DNA replication and to ensure efficient viral DNA replication (17, 18, 52). In addition, these three viral early proteins direct members of the MRN complex (2, 35) and the single-stranded DNA-binding protein 2 (20) to cytoplasmic aggresomes, where these sequestered proteins are effectively inactivated. These viral activities, along with the inactivation of DNA-PK by E4orf3 and E4orf6 gene products (7), appear to prevent recognition of viral genomes by the MRN complex and prevent ligation of these genomes through nonhomologous end joining. In cells infected with a virus with E4 deleted, Mre11 physically binds to viral DNA in an NBS1-dependent manner and may prevent efficient genome replication (37). The overlapping means by which adenovirus disables the MRN complex and prevents DNA damage repair serves to illustrate the importance of this activity for a productive adenovirus infection. However, despite having DNA damage signaling and DNA repair pathways dismantled, adenovirus-infected cells exhibit some characteristic changes associated with DNA damage signaling events, such as the phosphorylation of H2AX (6, 15). Thus, it appears that adenovirus effectively inhibits DNA repair activity but may not fully suppress the early events of DNA damage signaling.The focus of the present study was to elucidate the activation of DNA damage signaling pathways revealed by phosphorylation of the variant histone H2AX during wild-type adenovirus infection and to determine what stage of the virus life cycle leads to this activation. We demonstrate that infected cells respond to viral genome replication with high levels of H2AX phosphorylation throughout the cell nucleus. This phosphorylation event is not localized to viral replication centers and does not appear to be concurrent with cellular double-stranded DNA breaks; rather, H2AX phosphorylation occurs coincident with the bulk of cellular chromatin. H2AX phosphorylation follows viral DNA replication and reaches peak levels after the degradation of the Mre11. In addition, we observed that infected cells can respond to both the presence of incoming viral genomes and genome replication by initiating H2AX phosphorylation.  相似文献   

5.
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   

6.
7.
8.
9.
10.
11.
12.
13.
14.
Dna2 is a highly conserved helicase/nuclease that in yeast participates in Okazaki fragment processing, DNA repair, and telomere maintenance. Here, we investigated the biological function of human Dna2 (hDna2). Immunofluorescence and biochemical fractionation studies demonstrated that hDna2 was present in both the nucleus and the mitochondria. Analysis of mitochondrial hDna2 revealed that it colocalized with a subfraction of DNA-containing mitochondrial nucleoids in unperturbed cells. Upon the expression of disease-associated mutant forms of the mitochondrial Twinkle helicase which induce DNA replication pausing/stalling, hDna2 accumulated within nucleoids. RNA interference-mediated depletion of hDna2 led to a modest decrease in mitochondrial DNA replication intermediates and inefficient repair of damaged mitochondrial DNA. Importantly, hDna2 depletion also resulted in the appearance of aneuploid cells and the formation of internuclear chromatin bridges, indicating that nuclear hDna2 plays a role in genomic DNA stability. Together, our data indicate that hDna2 is similar to its yeast counterpart and is a new addition to the growing list of proteins that participate in both nuclear and mitochondrial DNA maintenance.DNA damage arises from errors in the replication process, as well as a myriad of intrinsic and extrinsic DNA-damaging agents that continually assault cells. Failure to efficiently repair DNA lesions leads to accumulation of mutations that contribute to numerous pathologies, including carcinogenesis. In addition to genomic DNA, mitochondrial DNA (mtDNA) is subject to damage that requires repair to maintain integrity. For these reasons, it is not surprising that DNA replication and repair proteins display significant plasticity that allows participation in several divergent replication and repair processes. In addition, numerous mechanisms, including alternative splicing, posttranslational modifications, or utilization of alternative translation initiation start sites, allow DNA replication and repair proteins such as Pif1, DNA ligase III, and APE1 to localize to the nucleus and the mitochondrion and participate in DNA replication and/or repair (9, 17, 25), thus ensuring genomic DNA and mtDNA integrity.Dna2 is an evolutionarily conserved helicase/nuclease enzyme. Originally discovered in Saccharomyces cerevisiae, Dna2 orthologs are found throughout the animal kingdom, including humans (5, 22, 28). Early studies demonstrated that Dna2 functions in concert with Flap endonuclease 1 (FEN1) to remove long DNA flaps that form upon lagging-strand DNA replication (6). However, in contrast to FEN1, Dna2 is an essential gene in yeast, suggesting that other proteins, including FEN1, cannot compensate for its loss in DNA replication or that it possesses functions beyond its role in Okazaki fragment processing. In agreement with this, genetic and biochemical studies have implicated Dna2 in DNA double-strand break (DSB) repair, telomere regulation, and mitochondrial function (8, 10, 15, 26, 38, 44, 45).Analysis of Dna2 in yeast revealed that it undergoes dynamic cell cycle localization. Dna2 localizes to telomeres during G1, relocalizes throughout the genome in S phase, and moves back to the telomere during late S/G2, where it participates in telomere replication and telomerase-dependent telomere elongation (10). Dna2 also leaves the telomere following treatment with bleomycin and localizes to sites of DNA DSBs (10). In addition, dna2 mutants are sensitive to DNA damage induced by gamma radiation and methanesulfonic acid methyl ester (7, 15). These phenotypes may be explained by recent work demonstrating that Dna2 plays an important role in 5′-end resection following DSBs. Indeed, upon induction of DSBs and initiation of 5′-end resection by the Mre11-Rad50-Xrs2 complex, Dna2 and Sgs1 cooperate to further degrade the 5′ end, creating long 3′ strands essential for homologous recombination (26, 45). Finally, while dna2Δ mutations are lethal in budding yeast, the dna2Δ pif1-m2 (nuclear PIF1) double mutations rescue dna2Δ lethality but produce a petite phenotype, suggesting that Dna2 is also involved in mtDNA maintenance (8).Recently, the human ortholog of Dna2 was cloned and characterized (23, 29). Biochemical analysis revealed that, similar to its yeast counterpart, the human Dna2 (hDna2) protein possesses nuclease, ATPase, and limited helicase activities (23, 29), suggesting that it carries out analogous functions in yeast and mammalian cells. However, hDna2''s putative role in genomic DNA repair and replication was called into question by a recent study suggesting that hDna2 is absent from the nucleus and found exclusively within the mitochondria, where it participates in mtDNA repair (44). Further in vitro biochemical studies suggested that hDna2 also participates in mtDNA replication (44). Here, we confirm that hDna2 localizes to the mitochondria and demonstrate that hDna2 participates in mtDNA replication and repair. However, our studies go further by uncovering a nuclear form of hDna2 that plays an important role in genomic stability. Indeed, we demonstrate that depletion of hDna2 leads to the appearance of aneuploid cells and the formation of internuclear chromatin bridges, indicating that hDna2, like its yeast counterpart, is essential to maintain nuclear DNA stability.  相似文献   

15.
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   

16.
17.
18.
19.
Cells from complementation groups A through G of the heritable sun-sensitive disorder xeroderma pigmentosum (XP) show defects in nucleotide excision repair of damaged DNA. Proteins representing groups A, B, C, D, F, and G are subunits of the core recognition and incision machinery of repair. XP group E (XP-E) is the mildest form of the disorder, and cells generally show about 50% of the normal repair level. We investigated two protein factors previously implicated in the XP-E defect, UV-damaged DNA binding protein (UV-DDB) and replication protein A (RPA). Three newly identified XP-E cell lines (XP23PV, XP25PV, and a line formerly classified as an XP variant) were defective in UV-DDB binding activity but had levels of RPA in the normal range. The XP-E cell extracts did not display a significant nucleotide excision repair defect in vitro, with either UV-irradiated DNA or a uniquely placed cisplatin lesion used as a substrate. Purified UV-DDB protein did not stimulate repair of naked DNA by DDB XP-E cell extracts, but microinjection of the protein into DDB XP-E cells could partially correct the repair defect. RPA stimulated repair in normal, XP-E, or complemented extracts from other XP groups, and so the effect of RPA was not specific for XP-E cell extracts. These data strengthen the connection between XP-E and UV-DDB. Coupled with previous results, the findings suggest that UV-DDB has a role in the repair of DNA in chromatin.The heritable human disorder xeroderma pigmentosum (XP) is chiefly characterized by an increased incidence of benign and malignant skin lesions after exposure to sunlight. Affected individuals fall into one of eight different genetic complementation groups. Cells from the seven complementation groups A through G have reduced nucleotide excision repair (NER) of damaged DNA, while cells from the variant, or V, group are defective in a less-defined process of cellular recovery after DNA damage (11). Genes and proteins representing XP groups A (XP-A) B, C, D, F, and G have all been isolated and found to represent some of the subunits of the core NER recognition and incision machinery. XP-E is the mildest form of the disorder, and cells of this group generally have 40 to 60% of the normal repair level, as shown by autoradiographic measurement of unscheduled DNA synthesis (UDS) after UV irradiation. Cell fusion studies have assigned at least 16 individuals to this form of the disorder (6, 19, 23, 40).There are several indications that a DNA damage binding protein denoted UV-DDB (or DDB) is involved in the primary XP-E defect. The protein has been detected in extracts of vertebrate cells as an activity that preferentially binds damaged oligonucleotides in electrophoretic mobility shift or filter binding assays. The protein has a particular affinity for (6-4) photoproducts in UV-irradiated DNA (10, 15, 16, 34, 41, 43), but UV-DDB also binds to DNA damaged by other agents, including cisplatin and nitrogen mustard (32). The activity has been purified as a single 127-kDa protein (2) and as a complex with two subunits of 127 and 48 kDa (21). Damage-binding activity is missing from some cells in the XP-E group, designated DDB, but is present in other XP-E cell lines, designated DDB+ (3, 15, 19, 23). The genes encoding the p127 protein (7, 17, 39) and the p48 protein (7) have been isolated, but DNA sequence features have not yet yielded firm clues about their functions. Microinjection of purified UV-DDB into XP-E cells lacking UV-DDB activity substantially corrects the NER defect, as measured by UDS after UV irradiation, but UV-DDB+ cells are not corrected (22). Sequence alterations in the gene for p48 have been reported for several XP-E cell lines (29), and it is possible that these are causative mutations for XP-E.There are also suggestions that the single-stranded DNA binding activity of replication protein A (RPA) is involved in the XP-E defect. RPA is a heterotrimer of three subunits with sizes of 70, 34, and 14 kDa that plays key roles in DNA replication, recombination, and DNA repair (44). It is one of the core components of the eukaryotic nucleotide excision-incision system (1, 12, 28). With regard to XP, it was recently reported that XP-E cell extracts are severely defective in NER in vitro and that RPA can specifically correct the repair defect of these extracts, but not those of extracts of other complementation groups (20). Moreover, it has been found that RPA copurifies to some extent with UV-DDB protein and that the two proteins interact, showing a tighter association with chromatin after UV irradiation of cells (31).The availability of lymphoblastoid cell lines derived from three newly identified XP-E individuals has given us the opportunity to further investigate the possible relationships of UV-DDB and RPA to the molecular defect in XP-E and the influence of these proteins on NER.  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