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.
Cellular and biochemical studies support a role for all five human RecQ helicases in DNA replication; however, their specific functions during this process are unclear. Here we investigate the in vivo association of the five human RecQ helicases with three well-characterized human replication origins. We show that only RECQ1 (also called RECQL or RECQL1) and RECQ4 (also called RECQL4) associate with replication origins in a cell cycle-regulated fashion in unperturbed cells. RECQ4 is recruited to origins at late G1, after ORC and MCM complex assembly, while RECQ1 and additional RECQ4 are loaded at origins at the onset of S phase, when licensed origins begin firing. Both proteins are lost from origins after DNA replication initiation, indicating either disassembly or tracking with the newly formed replisome. Nascent-origin DNA synthesis and the frequency of origin firing are reduced after RECQ1 depletion and, to a greater extent, after RECQ4 depletion. Depletion of RECQ1, though not that of RECQ4, also suppresses replication fork rates in otherwise unperturbed cells. These results indicate that RECQ1 and RECQ4 are integral components of the human replication complex and play distinct roles in DNA replication initiation and replication fork progression in vivo.The RecQ helicases are a family of DNA-unwinding enzymes essential for the maintenance of genome integrity in all kingdoms of life. Five RecQ enzymes have been found in human cells: RECQ1 (also called RECQL or RECQL1), BLM (RECQ2 or RECQL3), WRN (RECQ3 or RECQL2), RECQ4 (RECQL4), and RECQ5 (RECQL5) (3, 7). Here we refer to these helicases as RECQ1, RECQ4, and RECQ5, without the “L” that is present in the official gene names. Mutations in the BLM, WRN, and RECQ4 genes are linked to Bloom syndrome (BS), Werner syndrome (WS), and the subset of Rothmund-Thomson syndrome (RTS) patients at high risk of developing osteosarcomas, respectively (19, 31, 71). RECQ4 mutations have also been associated with RAPADILINO and Baller-Gerold syndrome (56, 61). Although these disorders are all associated with inherent genomic instability and cancer predisposition, they show distinct clinical features, suggesting that BLM, WRN, and RECQ4 are involved in different aspects of DNA metabolism. However, the molecular events underlying the pathogenesis of BS, WS, and RTS remain obscure. Mutations in the remaining two human RecQ helicase genes, RECQ1 and RECQ5, have not as yet been identified as causes of either genomic instability or heritable cancer predisposition disorders.Several lines of evidence suggest that RecQ helicases play an important role in DNA replication control (3, 10). In particular, RecQ helicases are thought to facilitate replication by preserving the integrity of stalled replication forks and by remodeling or repairing damaged or collapsed forks to allow the resumption of replication. Consistent with these ideas, several investigators have shown that primary fibroblasts from BS, WS, and RTS patients and RecQ5-deficient mouse embryonic fibroblasts all show differential hypersensitivity to agents that perturb DNA replication (12, 14, 26, 29). Moreover, BLM and WRN are recruited to DNA replication forks after replicative stress, and DNA fiber track analyses have shown that both BLM and WRN are required for normal fork progression after DNA damage or replication arrest (11-13, 47, 54). In particular, BLM in conjunction with DNA topoisomerase III and two other accessory proteins, RMI-1 and RMI-2, has been shown to catalyze the resolution of double-Holliday-junction recombination intermediates to generate noncrossover products. This dissolution reaction could play an important role in the error-free recombinational repair of damaged or stalled forks during S phase (57, 67). WRN also appears to promote error-free repair by contributing to the resolution of gene conversion events to generate noncrossover products (46). In line with the above observations, WRN and BLM can be found associated with replication foci or other DNA damage response proteins in damaged cells. In contrast, in unperturbed cells, a majority of each protein is found in the nucleolus (WRN) or associated with PML bodies (BLM) (5, 37, 62).RECQ4 has also been implicated in DNA replication. Recent studies have shown that hypomorphic mutants of the Drosophila melanogaster homolog of human RECQ4, DmRECQ4, have reduced DNA replication-dependent chorion gene amplification (65). These findings are thus consistent with a postulated role for Xenopus laevis RECQ4 (XRECQ4) in the initiation of DNA replication (39, 48). The N terminus of XRECQ4 bears homology to the N termini of the yeast proteins Sld2 (Saccharomyces cerevisiae [budding yeast]) and DRC1 (Schizosaccharomyces pombe [fission yeast]), which play a central role, in association with budding yeast Dpb11 and the fission yeast homolog Cut5/Rad4, in the establishment of DNA replication forks (38, 41, 63). Consistently, the N terminus of XRECQ4 has been shown to interact with the X. laevis variant of Cut5, and XRECQ4 depletion severely perturbs DNA replication initiation in X. laevis egg extracts (39, 48). The notion that the function of XRECQ4 is evolutionarily conserved in mammals is supported by the observations that the human protein can complement its Xenopus counterpart in cell-free assays for replication initiation and that depletion of human RECQ4 inhibits cellular proliferation and DNA synthesis (39, 48). Moreover, deletion of the N-terminal region of mouse RECQ4 has been shown to be an embryonic lethal mutation (27). These observations suggest that vertebrate RECQ4 might be a functional homolog of Sld2/DRC11, although its precise function during replication initiation and progression is not known. Recent results, published while this work was in progress, indicate that human RECQ4 interacts with the MCM replicative complex during replication initiation and that this interaction is regulated by CDK phosphorylation of RECQ4 (69). These findings, together with our results below, provide clues to the mechanism regulating RECQ4 interaction with the replication machinery.RECQ1 is the most abundant of the human RecQ helicases and was the first of the human RecQ proteins to be discovered on the basis of its potent ATPase activity (50). Despite this, little is known about the cellular functions of RECQ1, and no human disease associations have been identified to date. Recent studies have shown that RECQ1 is involved in the maintenance of genome integrity and that RECQ1 depletion affects cellular proliferation (51). Moreover, biochemical studies have shown that RECQ1 and BLM display distinct substrate specificities, indicating that these helicases are likely to perform nonoverlapping functions (43). These results suggest an important—though as yet mechanistically ill-defined—role for RECQ1 in cell cycle progression and/or DNA repair (52).In order to better delineate the role of human RecQ helicases in DNA replication, we investigated the in vivo interactions of all five human RecQ enzymes with three well-characterized human DNA replication origins in quantitative chromatin immunoprecipitation (ChIP) assays. We also determined how nascent-origin-dependent DNA synthesis, chromatin binding of replication proteins, origin firing frequency, and replication fork rates were altered by depleting specific human RecQ helicase proteins. We found that only two of the five human RecQ helicases, RECQ1 and RECQ4, specifically interact with origins in unperturbed cells. Our results provide new mechanistic insight into the distinct roles of human RECQ1 and RECQ4 in DNA replication initiation and in replication fork progression.  相似文献   

2.
3.
4.
Tegument is a unique structure of herpesvirus, which surrounds the capsid and interacts with the envelope. Morphogenesis of gammaherpesvirus is poorly understood due to lack of efficient lytic replication for Epstein-Barr virus and Kaposi''s sarcoma-associated herpesvirus/human herpesvirus 8, which are etiologically associated with several types of human malignancies. Murine gammaherpesvirus 68 (MHV-68) is genetically related to the human gammaherpesviruses and presents an excellent model for studying de novo lytic replication of gammaherpesviruses. MHV-68 open reading frame 33 (ORF33) is conserved among Alpha-, Beta-, and Gammaherpesvirinae subfamilies. However, the specific role of ORF33 in gammaherpesvirus replication has not yet been characterized. We describe here that ORF33 is a true late gene and encodes a tegument protein. By constructing an ORF33-null MHV-68 mutant, we demonstrated that ORF33 is not required for viral DNA replication, early and late gene expression, viral DNA packaging or capsid assembly but is required for virion morphogenesis and egress. Although the ORF33-null virus was deficient in release of infectious virions, partially tegumented capsids produced by the ORF33-null mutant accumulated in the cytoplasm, containing conserved capsid proteins, ORF52 tegument protein, but virtually no ORF45 tegument protein and the 65-kDa glycoprotein B. Finally, we found that the defect of ORF33-null MHV-68 could be rescued by providing ORF33 in trans or in an ORF33-null revertant virus. Taken together, our results indicate that ORF33 is a tegument protein required for viral lytic replication and functions in virion morphogenesis and egress.Gammaherpesviruses are associated with tumorigenesis. Like other herpesviruses, they are characterized as having two distinct stages in their life cycle: lytic replication and latency (15, 16, 18, 21, 54). Latency provides the viruses with advantages to escape host immune surveillance and to establish lifelong persistent infection and contributes to transformation and development of malignancies. However, it is through lytic replication that viruses propagate and transmit among hosts to maintain viral reservoirs. Both viral latency and lytic replication play important roles in tumorigenesis. The gammaherpesvirus subfamily includes Epstein-Barr virus (EBV), Kaposi''s sarcoma-associated herpesvirus (KSHV)/human herpesvirus 8 and murine gammaherpesvirus 68 (MHV-68), among others. EBV is associated with Burkitt''s lymphoma, nasopharyngeal carcinoma, Hodgkin''s disease, and lymphoproliferative diseases in immunodeficient patients (28). KSHV is etiologically linked with Kaposi''s sarcoma, primary effusion lymphoma, and multicentric Castleman''s disease (11-13, 22, 52). Neither in vivo nor in vitro studies of EBV and KSHV are convenient due to their propensity to establish latency in cell culture and their limited host ranges.MHV-68 is genetically related to these two human gammaherpesviruses, especially to KSHV, based on the alignment of their genomic sequences and other biological properties (55). As a natural pathogen of wild rodents, MHV-68 also infects laboratory mice (6, 40, 46) and replicates to a high titer in a variety of fibroblast and epithelial cell lines. These advantages make MHV-68 an excellent model for studying the lytic replication of gammaherpesviruses in vitro and certain aspects of virus-host interactions in vivo. In addition, the MHV-68 genome has been cloned as a bacterial artificial chromosome (BAC) that can propagate in Escherichia coli (1, 2, 36, 51), making it convenient to study the function of each open reading frame (ORF) by genetic methods. Exploring the functions of MHV-68 ORFs will likely shed light on the functions of their homologues in human gammaherpesviruses.Gammaherpesviral particles have a characteristic multilayered architecture. An infectious virion contains a double-stranded DNA genome, an icosahedral capsid shell, a thick, proteinaceous tegument compartment, and a lipid bilayer envelope spiked with glycoproteins (14, 30, 47, 49). As a unique structure of herpesviruses, the tegument plays important roles in multiple aspects of the viral life cycle, including virion assembly and egress (38, 48, 53), translocation of nucleocapsids into the nucleus, transactivation of viral immediate-early genes, and modulation of host cell gene expression, innate immunity, and signal transduction (9, 10, 23, 60). Some components of MHV-68 tegument have been identified by a mass spectrometric study (8), and the functions of some tegument proteins have been revealed, such as ORF45, ORF52, and ORF75c (7, 24, 29).MHV-68 ORF33 is conserved among Alpha-, Beta-, and Gammaherpesvirinae subfamilies. Its homologues include human herpes simplex virus type 1 (HSV-1) UL16, human herpes simplex virus type 2 (HSV-2) UL16, human cytomegalovirus (HCMV) UL94, EBV BGLF2, KSHV ORF33, and rhesus monkey rhadinovirus (RRV) ORF33. HSV-1 UL16 has been identified as a tegument protein and may function in viral DNA packaging, virion assembly, budding, and egress (5, 32, 35, 41, 44). HCMV UL94 is a virion associated protein and might function in virion assembly and budding (31, 57). EBV BGLF2, KSHV ORF33, and RRV ORF33 are also virion-associated proteins, but their functions are not clear (26, 43, 59). The mass spectrometric study of MHV-68 did not identify ORF33 as a virion component (8), although ORF33 is found to be essential for viral lytic replication by transposon mutagenesis of the MHV-68 genome cloned as a BAC (51). However, insertion of the 1.2-kbp Mu transposon in that study may influence the expression of ORFs approximate to ORF33. Consequently, the role ORF33 plays in viral replication needs to be confirmed, preferably through site-directed mutagenesis. Whether ORF33 is a tegument protein and the exact viral replication stage in which it functions also need to be investigated.We determined that MHV-68 ORF33 encodes a tegument protein and is expressed with true late kinetics. To explore the function of ORF33 in viral lytic phase, we used site-directed mutagenesis and generated an ORF33-null mutant, taking advantage of the MHV-68 BAC system. We showed that the ORF33-null mutant is capable of viral DNA replication, early and late gene expression, capsid assembly, and DNA packaging, but incapable of virion release. The defect of ORF33-null mutant can be rescued in trans by an ORF33 expression plasmid.  相似文献   

5.
6.
7.
8.
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   

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

10.
Nitrofurazone is reduced by cellular nitroreductases to form N2-deoxyguanine (N2-dG) adducts that are associated with mutagenesis and lethality. Much attention recently has been given to the role that the highly conserved polymerase IV (Pol IV) family of polymerases plays in tolerating adducts induced by nitrofurazone and other N2-dG-generating agents, yet little is known about how nitrofurazone-induced DNA damage is processed by the cell. In this study, we characterized the genetic repair pathways that contribute to survival and mutagenesis in Escherichia coli cultures grown in the presence of nitrofurazone. We find that nucleotide excision repair is a primary mechanism for processing damage induced by nitrofurazone. The contribution of translesion synthesis to survival was minor compared to that of nucleotide excision repair and depended upon Pol IV. In addition, survival also depended on both the RecF and RecBCD pathways. We also found that nitrofurazone acts as a direct inhibitor of DNA replication at higher concentrations. We show that the direct inhibition of replication by nitrofurazone occurs independently of DNA damage and is reversible once the nitrofurazone is removed. Previous studies that reported nucleotide excision repair mutants that were fully resistant to nitrofurazone used high concentrations of the drug (200 μM) and short exposure times. We demonstrate here that these conditions inhibit replication but are insufficient in duration to induce significant levels of DNA damage.Replication in the presence of DNA damage is thought to produce most of the mutagenesis, genomic rearrangements, and lethality that occur in all cells. UV-induced photoproducts, X-ray-induced strand breaks, psoralen- or cis-platin-interstrand cross-links, oxidized bases from reactive oxygen species, and base depurination are just a few of the structurally distinct challenges that the replication machinery must overcome. It seems likely that the mechanisms that process these lesions will vary depending on the nature of the impediment.While a number of the lesions described above are known to block replication, the events associated with UV-induced damage have been the most extensively characterized. UV irradiation causes the formation of cyclobutane pyrimidine dimers and 6-4 photoproducts in DNA that block the progression of the replication fork (16, 29, 30, 37). Following the arrest of replication at UV-induced damage, RecA and several RecF pathway proteins are required to process the replication fork such that the blocking lesion is removed or bypassed (2, 5, 6, 8-10). Cells lacking either RecA or any of several RecF pathway proteins are hypersensitive to UV-induced damage and fail to recover replication following disruption by the lesions (2, 6, 10). RecBCD is an exonuclease/helicase complex that is involved in repairing double-strand breaks (38). It also is required for resistance to UV-induced damage, although it is not required to process or restore disrupted replication forks, and the substrates it acts upon after UV irradiation currently remain unclear (3, 10, 19).Survival and the ability to resume DNA synthesis following UV-induced damage depend predominantly on the removal of the lesions by nucleotide excision repair (5, 7, 36). Cells deficient in nucleotide excision repair are unable to remove UV-induced DNA lesions and exhibit elevated levels of mutagenesis, strand exchanges, rearrangements, and cell lethality (16, 33, 34). In cases where replication fork processing or lesion repair is prevented, the recovery of replication and survival become entirely dependent on translesion synthesis by DNA polymerase V (Pol V) (6). However, in repair-proficient cells, the contribution of translesion synthesis to recovery and survival is minor and is detected only following UV doses that exceed the repair capacity of the cell (5, 6).Less is known about how replication recovers from other forms of DNA damage. We chose to characterize nitrofurazone, because a number of studies suggested that N2-deoxyguanine (N2-dG) adducts induced by this and other agents would be processed differently than UV-induced lesions. Nitrofurazone is a topical antibacterial agent that historically has been used for treating burns and skin grafts in patients and animals (14, 15, 32). Nitrofurazone toxicity is known to require activation by cellular nitroreductases (25, 42). However, the mechanism and targets of its antimicrobial properties have yet to be fully elucidated. In addition to its antimicrobial properties, the reduced nitrofurazone metabolites also target DNA and have been shown to induce free radical damage, strand breaks, and N2-dG adducts (26, 40, 42, 45), and they are mutagenic and carcinogenic in rodent models (1, 15, 24, 39).Whereas nucleotide excision repair is the predominant mechanism required for survival after UV-induced damage, a number of studies suggest that translesion synthesis plays a larger role in survival after nitrofurazone-induced DNA damage. dinB mutants lacking Pol IV were shown to be hypersensitive to nitrofurazone compared to cells that constitutively express the polymerase (17). Biochemically, Pol IV and a number of Pol IV homologs from other organisms have been shown to efficiently replicate over a range of N2-dG adducts in vitro (17, 35, 44). In addition, several studies have reported that uvrA mutants, which are defective in nucleotide excision repair, do not exhibit any hypersensitivity to nitrofurazone or other agents that induce similar adducts in vivo (12, 21, 27). Early studies also observed a direct correlation between nitrofurazone-induced mutations and lethality, suggesting that mutagenic lesions persist in the DNA to cause toxicity (21, 23, 27, 43). Consistent with these observations, nitrofuran-induced lesions were found to be poor substrates for nucleotide excision repair in vitro (46).Taken together, these observations suggest to us that the cellular response to nitrofurazone will be distinct from its response to UV irradiation. However, no study has examined the relative contributions that nucleotide excision repair, translesion synthesis, or recombination has in recovering from nitrofurazone-induced damage. In this study, we characterized the mechanism by which nitrofurazone inhibits DNA replication and identified the genes that contribute to the recovery, survival, and mutagenesis of Escherichia coli treated with nitrofurazone. In contrast to previous studies, we found that survival following nitrofurazone-induced damage depends predominantly on nucleotide excision repair. Similarly to UV-induced DNA damage, both the RecF and RecBC pathways contribute to survival following nitrofurazone-induced DNA damage. The contribution of translesion polymerases to survival was minor and was mediated by Pol IV. In addition, we found that nitrofurazone can act to inhibit DNA replication directly when used at higher concentrations. The direct inhibition of replication is reversible and occurs independently of DNA damage, suggesting that DNA is not the primary target of its antimicrobial properties.  相似文献   

11.
12.
13.
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.
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   

16.
The combinatorial nature of genetic recombination can potentially provide organisms with immediate access to many more positions in sequence space than can be reached by mutation alone. Recombination features particularly prominently in the evolution of a diverse range of viruses. Despite rapid progress having been made in the characterization of discrete recombination events for many species, little is currently known about either gross patterns of recombination across related virus families or the underlying processes that determine genome-wide recombination breakpoint distributions observable in nature. It has been hypothesized that the networks of coevolved molecular interactions that define the epistatic architectures of virus genomes might be damaged by recombination and therefore that selection strongly influences observable recombination patterns. For recombinants to thrive in nature, it is probably important that the portions of their genomes that they have inherited from different parents work well together. Here we describe a comparative analysis of recombination breakpoint distributions within the genomes of diverse single-stranded DNA (ssDNA) virus families. We show that whereas nonrandom breakpoint distributions in ssDNA virus genomes are partially attributable to mechanistic aspects of the recombination process, there is also a significant tendency for recombination breakpoints to fall either outside or on the peripheries of genes. In particular, we found significantly fewer recombination breakpoints within structural protein genes than within other gene types. Collectively, these results imply that natural selection acting against viruses expressing recombinant proteins is a major determinant of nonrandom recombination breakpoint distributions observable in most ssDNA virus families.Genetic recombination is a ubiquitous biological process that is both central to DNA repair pathways (10, 57) and an important evolutionary mechanism. By generating novel combinations of preexisting nucleotide polymorphisms, recombination can potentially accelerate evolution by increasing the population-wide genetic diversity upon which adaptive selection relies. Recombination can paradoxically also prevent the progressive accumulation of harmful mutations within individual genomes (18, 35, 53). Whereas its ability to defend high-fitness genomes from mutational decay possibly underlies the evolutionary value of sexuality in higher organisms, in many microbial species where pseudosexual genetic exchange is permissible among even highly divergent genomes, recombination can enable access to evolutionary innovations that would otherwise be inaccessible by mutation alone.Such interspecies recombination is fairly common in many virus families (8, 17, 27, 44, 82). It is becoming clear, however, that as with mutation events, most recombination events between distantly related genomes are maladaptive (5, 13, 38, 50, 63, 80). As genetic distances between parental genomes increase, so too does the probability of fitness defects in their recombinant offspring (16, 51). The viability of recombinants is apparently largely dependent on how severely recombination disrupts coevolved intragenome interaction networks (16, 32, 51). These networks include interacting nucleotide sequences that form secondary structures, sequence-specific protein-DNA interactions, interprotein interactions, and amino acid-amino acid interactions within protein three-dimensional folds.One virus family where such interaction networks appear to have a large impact on patterns of natural interspecies recombination are the single-stranded DNA (ssDNA) geminiviruses. As with other ssDNA viruses, recombination is very common among the species of this family (62, 84). Partially conserved recombination hot and cold spots have been detected in different genera (39, 81) and are apparently caused by both differential mechanistic predispositions of genome regions to recombination and natural selection disfavoring the survival of recombinants with disrupted intragenome interaction networks (38, 51).Genome organization and rolling circle replication (RCR)—the mechanism by which geminiviruses and many other ssDNA viruses replicate (9, 67, 79; see reference 24 for a review)—seem to have a large influence on basal recombination rates in different parts of geminivirus genomes (20, 33, 39, 61, 81). To initiate RCR, virion-strand ssDNA molecules are converted by host-mediated pathways into double-stranded “replicative-form” (RF) DNAs (34, 67). Initiated by a virus-encoded replication-associated protein (Rep) at a well-defined virion-strand replication origin (v-ori), new virion strands are synthesized on the complementary strand of RF DNAs (28, 73, 74) by host DNA polymerases. Virion-strand replication is concomitant with the displacement of old virion strands, which, once complete, yields covalently closed ssDNA molecules which are either encapsidated or converted into additional RF DNAs. Genome-wide basal recombination rates in ssDNA viruses are probably strongly influenced by the specific characteristics of host DNA polymerases that enable RCR. Interruption of RCR has been implicated directly in geminivirus recombination (40) and is most likely responsible for increased basal recombination rates both within genes transcribed in the opposite direction from that of virion-strand replication (40, 71) and at the v-ori (1, 9, 20, 69, 74).Whereas most ssDNA virus families replicate via either a rolling circle mechanism (the Nanoviridae, Microviridae, and Geminiviridae) (3, 23, 24, 31, 59, 67, 74) or a related rolling hairpin mechanism (the Parvoviridae) (25, 76), among the Circoviridae only the Circovirus genus is known to use RCR (45). Although the Gyrovirus genus (the other member of the Circoviridae) and the anelloviruses (a currently unclassified ssDNA virus group) might also use RCR, it is currently unknown whether they do or not (78). Additionally, some members of the Begomovirus genus of the Geminiviridae either have a second genome component, called DNA-B, or are associated with satellite ssDNA molecules called DNA-1 and DNA-Beta, all of which also replicate by RCR (1, 47, 68).Recombination is known to occur in the parvoviruses (19, 43, 70), microviruses (66), anelloviruses (40, 46), circoviruses (11, 26, 60), nanoviruses (30), geminivirus DNA-B components, and geminivirus satellite molecules (2, 62). Given that most, if not all, of these ssDNA replicons are evolutionarily related to and share many biological features with the geminiviruses (22, 31, 36), it is of interest to determine whether conserved recombination patterns observed in the geminiviruses (61, 81) are evident in these other groups. To date, no comparative analyses have ever been performed with different ssDNA virus families to identify, for example, possible influences of genome organization on recombination breakpoint distributions found in these viruses.Here we compare recombination frequencies and recombination breakpoint distributions in most currently described ssDNA viruses and satellite molecules and identify a number of sequence exchange patterns that are broadly conserved across this entire group.  相似文献   

17.
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   

18.
19.
The processing of lagging-strand intermediates has not been demonstrated in vitro for herpes simplex virus type 1 (HSV-1). Human flap endonuclease-1 (Fen-1) was examined for its ability to produce ligatable products with model lagging-strand intermediates in the presence of the wild-type or exonuclease-deficient (exo) HSV-1 DNA polymerase (pol). Primer/templates were composed of a minicircle single-stranded DNA template annealed to primers that contained 5′ DNA flaps or 5′ annealed DNA or RNA sequences. Gapped DNA primer/templates were extended but not significantly strand displaced by the wild-type HSV-1 pol, although significant strand displacement was observed with exo HSV-1 pol. Nevertheless, the incubation of primer/templates containing 5′ flaps with either wild-type or exo HSV-1 pol and Fen-1 led to the efficient production of nicks that could be sealed with DNA ligase I. Both polymerases stimulated the nick translation activity of Fen-1 on DNA- or RNA-containing primer/templates, indicating that the activities were coordinated. Further evidence for Fen-1 involvement in HSV-1 DNA synthesis is suggested by the ability of a transiently expressed green fluorescent protein fusion with Fen-1 to accumulate in viral DNA replication compartments in infected cells and by the ability of endogenous Fen-1 to coimmunoprecipitate with an essential viral DNA replication protein in HSV-1-infected cells.Herpes simplex virus type 1 (HSV-1), the prototypic member of the family of Herpesviridae and that of the alphaherpesviridae subfamily, has served as the model for understanding the replication of herpesvirus genomes during lytic virus replication (29). The 152-kbp genome of herpes simplex virus type 1 (HSV-1) possesses approximately 85 genes, 7 of which have been shown to be necessary and sufficient for viral DNA replication within host cells (reviewed in references 5 and 38). These seven genes encode a DNA polymerase (pol) and its processivity factor (UL42), a heterotrimeric complex containing a DNA helicase (UL5), primase (UL52), and noncatalytic accessory protein (UL8), a single-stranded DNA binding protein (infected cell protein 8 [ICP-8]), and an origin binding protein with DNA helicase activity (UL9). There is strong evidence in support of the circularization of the linear virion DNA shortly after entry, and DNA replication then is thought to initiate at one or more of the three redundant origins of replication (29, 38). At least in the earliest stages of viral DNA replication, UL9 protein is required, presumably to bind to and unwind the DNA and to attract the other DNA replication proteins (29, 38). The electron microscopic examination of pulse-labeled replicating HSV-1 DNA indicates the presence of lariats, eye-forms, and D-forms (21), which is consistent with bidirectional theta-like replication from origins. To date, however, no biochemical assay has demonstrated origin-dependent DNA replication in vitro. However, in the absence of UL9, the other six HSV DNA replication proteins can support initiation and replication from a circular single-stranded DNA (ssDNA) template in an origin-independent fashion (15, 26), resembling the rolling-circle mode of replication thought to occur during the later stages of viral replication.Although nicks and small gaps have been observed in isolated replicating and virion DNA (38), the evidence for bidirectional duplex synthesis, the rapid rate of viral DNA replication, and the absence of long stretches of ssDNA in replicating and mature DNA isolated from HSV-1-infected cells suggest that leading- and lagging-strand synthesis are closely coordinated in vivo. Falkenberg et al. (15) used a minicircle DNA template with a strand bias and the six essential HSV-1 DNA replication proteins needed for rolling circle replication to demonstrate lagging-strand synthesis in vitro. However, replication from the parental strand template (leading-strand synthesis) was more efficient than synthesis from the complementary-strand template (lagging-strand synthesis). These results suggest the possibility that one or more host functions required for efficient lagging-strand synthesis or for its close coordination with leading-strand synthesis is missing in such in vitro systems.Although leading- and lagging-strand syntheses share many of the same requirements for bulk DNA synthesis, lagging-strand synthesis is a more complex process. Because the direction of polymerization of lagging-strand intermediates is opposite the direction of replication fork movement, lagging-strand synthesis requires that priming and extension occur many times to produce discontinuous segments called Okazaki fragments (reviewed in reference 25). Okazaki fragments need to be processed to remove the RNA primer, to fill in the area previously occupied by the RNA, and to seal the remaining nick between fragments, all of which must occur efficiently, accurately, and completely. Failure to do so would result in the accumulation of DNA breaks, multiple mutations, delayed DNA replication, and/or cell death (16, 61).In eukaryotes, what is currently known regarding the process of lagging-strand synthesis is based on genetic and biochemical studies with Saccharomyces cerevisiae and on in vitro reconstitution studies to define the mammalian enzymes required for simian virus 40 (SV40) T-antigen-dependent DNA replication (17, 37, 44, 57, 58). These studies have revealed that the extension of a newly synthesized Okazaki fragment DNA with pol δ causes the strand displacement of the preceding fragment to produce a 5′ flap (25). Results suggest that flap endonuclease 1 (Fen-1) is the activity responsible for the removal of the bulk of the 5′ flaps generated (1, 44, 48), although dna2 protein may facilitate the removal of longer flaps coated with the ssDNA binding protein complex (2, 44). In addition, the overexpression of exonuclease I can partially compensate for the loss of Fen-1 function in yeast (24, 51). For the proper processing of lagging-strand intermediates, the entire 5′ flap and all of the RNA primer need to be removed, and the gap must be filled to achieve a ligatable nick. DNA ligase I has been shown to be the enzyme involved in sealing Okazaki fragments in yeast and in humans (3, 31, 50, 56, 57). DNA ligase I requires a nick in which there is a 5′ phosphate on one end and a 3′ hydroxyl linked to a deoxyribose sugar entity on the other, and it works poorly in the presence of mismatches (54). The close coordination of Fen-1 and DNA ligase I activities for Okazaki fragment processing is facilitated by the interactions of these proteins with proliferating cell nuclear antigen (PCNA), the processivity factor for pol δ and ɛ (6, 30, 32, 46, 52, 53).HSV-1 does not appear to encode a protein with DNA ligase activity or one that can specifically cleave 5′ flaps, although it does encode a 5′-to-3′ exonuclease activity (UL12 [10, 20]) and a 3′-to-5′ exonuclease activity that is part of the HSV-1 pol catalytic subunit (27). As for most eukaryotes, RNA primers are essential for HSV-1 DNA synthesis, as demonstrated by the presence of oligoribonucleotides in replicating DNA in vivo (4), by the well-characterized ability of the UL52 protein in complex with the UL5 helicase activity to synthesize oligoribonucleotide primers on ssDNA in vitro (11, 13), and by the requirement of the conserved catalytic residues in the UL52 primase in vitro and in HSV-1-infected cells (14, 26). It is the strand displacement activity of pol δ that produces the 5′ flaps that are key to the removal of RNA primers during Okazaki fragment processing (6, 25). However, we previously demonstrated that wild-type HSV-1 DNA polymerase possesses poor strand displacement activity (62), in contrast to mammalian DNA pol δ (25). Thus, it is not apparent how RNA primers would be removed when encountered by HSV-1 pol during HSV-1 lagging-strand synthesis or how such intermediates would be processed.We wished to test the hypothesis that the nick translation activity of mammalian Fen-1 could function in collaboration with HSV-1 pol to facilitate the proper removal of RNA primers and/or short flaps to produce the ligatable products required for Okazaki fragment processing. In this report, we have examined the ability of wild-type and exonuclease-deficient (exo) HSV-1 pol, which differ in their respective strand displacement activities, to extend model lagging-strand substrates in the presence or absence of mammalian Fen-1. Our results demonstrate that both wild-type and exo HSV-1 pol can cooperate with and enhance Fen-1 activity to achieve a ligatable nick in vitro. Moreover, colocalization and coimmunoprecipitation studies reveal a physical association of Fen-1 with HSV-1 DNA replication proteins, supporting a model for the involvement of Fen-1 in HSV-1 DNA replication.  相似文献   

20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号