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 29 毫秒
1.
Prim‐pol is a recently identified DNA primase‐polymerase belonging to the archaeao‐eukaryotic primase (AEP) superfamily. Here, we characterize a previously unrecognized prim‐pol in human cells, which we designate hPrimpol1 (human primase‐polymerase 1). hPrimpol1 possesses primase and DNA polymerase activities in vitro, interacts directly with RPA1 and is recruited to sites of DNA damage and stalled replication forks in an RPA1‐dependent manner. Cells depleted of hPrimpol1 display increased spontaneous DNA damage and defects in the restart of stalled replication forks. Both RPA1 binding and the primase activity of hPrimpol1 are required for its cellular function during DNA replication. Our results indicate that hPrimpol1 is a novel factor involved in the response to DNA replication stress.  相似文献   

2.
3.
α-Accessory factor (AAF) stimulates the activity of DNA polymerase-α·primase, the only enzyme known to initiate DNA replication in eukaryotic cells (Goulian, M., Heard, C. J., and Grimm, S. L. (1990) J. Biol. Chem. 265 ,13221 -13230). We purified the AAF heterodimer composed of 44- and 132-kDa subunits from cultured cells and identified full-length cDNA clones using amino acid sequences from internal peptides. AAF-132 demonstrated no homologies to known proteins; AAF-44, however, is evolutionarily related to the 32-kDa subunit of replication protein A (RPA-32) and contains an oligonucleotide/oligosaccharide-binding (OB) fold domain similar to the OB fold domains of RPA involved in single-stranded DNA binding. Epitope-tagged versions of AAF-44 and -132 formed a complex in intact cells, and purified recombinant AAF-44 bound to single-stranded DNA and stimulated DNA primase activity only in the presence of AAF-132. Mutations in conserved residues within the OB fold of AAF-44 reduced DNA binding activity of the AAF-44·AAF-132 complex. Immunofluorescence staining of AAF-44 and AAF-132 in S phase-enriched HeLa cells demonstrated punctate nuclear staining, and AAF co-localized with proliferating cell nuclear antigen, a marker for replication foci containing DNA polymerase-α·primase and RPA. Small interfering RNA-mediated depletion of AAF-44 in tumor cell lines inhibited [methyl-3H]thymidine uptake into DNA but did not affect cell viability. We conclude that AAF shares structural and functional similarities with RPA-32 and regulates DNA replication, consistent with its ability to increase polymerase-α·primase template affinity and stimulate both DNA primase and polymerase-α activities in vitro.In eukaryotic cells, DNA replication is initiated at multiple origins internal to each chromosome; the origin recognition complex recruits cell division cycle and minichromosome maintenance proteins to form a preinitiation complex (1). At the G1-S phase transition, the latter complex is activated by cyclin-dependent protein kinases leading to formation of an initiation complex that alters local DNA structure through DNA helicase activity (1, 2). The replication protein A (RPA)2 is recruited to bind and stabilize single-stranded DNA (ssDNA) produced by the initiation complex (3, 4). RPA serves as an auxiliary factor for DNA polymerase-α (pol-α)·primase: it stabilizes the protein complex by direct interaction with both pol-α and primase subunits, and it reduces the misincorporation rate of pol-α, acting as a “fidelity clamp” (5, 6). The pol-α·primase complex consists of four subunits, including the catalytic pol-α subunit (p185), a regulatory B subunit (p70), and two primase subunits (p49 and p58). On an ssDNA template, the primase synthesizes short RNA primers from ribonucleoside triphosphates (rNTPs), which are elongated by pol-α in the presence of deoxyribonucleoside triphosphates (dNTPs) to form short DNA fragments. Through mechanisms requiring other replication factors, pol-α·primase is replaced by the more processive DNA polymerases pol-δ and pol-ε (7). Pol-ε synthesizes the leading strand, whereas pol-δ completes each Okazaki fragment initiated by pol-α·primase on the lagging strand and proofreads errors made by pol-α (7). The initiator RNA and DNA fragments are later removed by nucleases, and the Okazaki fragments are sealed by DNA ligase (7).The pol-α·primase complex is the only eukaryotic DNA polymerase able to initiate DNA synthesis de novo. In addition to initiating DNA replication and synthesizing Okazaki fragments, it appears to be one of the final targets of cell cycle checkpoint pathways that couple DNA replication to DNA damage response (2, 8). The role of RPA in initiation, elongation, and completion of lagging strand DNA synthesis has been thoroughly investigated (3, 9), but in vitro studies suggest that some additional factors that promote the rapidity of DNA replication in vivo are still lacking (2).In the course of purifying pol-α·primase from extracts of cultured mouse L1210 cells, we identified a factor we named α-accessory factor (AAF) that stimulates pol-α·primase activity in vitro (10, 11). The protein has a native molecular mass of ∼150 kDa as determined from its sedimentation coefficient and Stokes radius and is composed of two subunits of ∼132 and ∼44 kDa. AAF stimulates pol-α·primase activity with several different templates and types of reactions: (i) It stimulates selfprimed reactions with poly(dT), poly(dI·dT), or single-stranded circular DNA; (ii) it stimulates primed reactions with poly(dA)·oligo(dT) and multiply primed DNA in the absence of rNTPs, indicating that it affects pol-α activity when no primers are being made; and (iii) it stimulates primase activity on ssDNA in the absence of dNTPs, showing that it can enhance RNA primer synthesis in the absence of DNA synthesis (11). AAF increases the template affinity and processivity of pol-α·primase (12). AAF is highly specific for pol-α·primase and has no effect on the other mammalian DNA polymerases β, γ, or δ or on the DNA polymerase·primase complexes from Drosophila and Saccharomyces cerevisiae (11).The cloning of both AAF subunits based on peptide sequences obtained from the purified protein allowed us now to further characterize the AAF-44·AAF-132 complex structurally and functionally. Based on siRNA experiments in cancer cell lines, AAF appears to regulate DNA replication in vivo.  相似文献   

4.
Günes C  Rudolph KL 《The EMBO journal》2012,31(13):2833-2834
EMBO J 31 13, 2839–2851 (2012); published online May082012Senescence represents a major tumour suppressor checkpoint activated by telomere dysfunction or cellular stress factors such as oncogene activation. In this issue of The EMBO Journal, Suram et al (2012) reveal a surprising interconnection between oncogene activation and telomere dysfunction induced senescence. The study supports an alternative model of tumour suppression, indicating that oncogene-induced accumulation of telomeric DNA damage contributes to the induction of senescence in telomerase-negative tumours.Telomere shortening limits the proliferative capacity of primary human cells after 50–70 cell divisions by induction of replicative senescence activated by critically short, dysfunctional telomeres. Different mechanisms were thought to initiate senescence in response to oncogene activation, which occurs abruptly within a few cell doublings (Serrano et al, 1997). Oncogene-induced senescence (OIS) involves an activation of DNA damage signals at stalled replication forks induced by DNA replication stress (Bartkova et al, 2006; Di Micco et al, 2006). Replication fork stalling in response to oncogene activation preferentially affects common fragile sites of the DNA (Tsantoulis et al, 2008). The ends of eukaryotic chromosomes—the telomeres–represent common fragile sites that are sensitive to replication fork stalling (Sfeir et al, 2009). These data made it tempting to speculate whether replication fork stalling at telomeres was causatively involved in OIS. Studies on replicative senescence in human fibroblast also supported this possibility showing that mitogenic signals amplify DNA damage responses in senescent cells (Satyanarayana et al, 2004).Multiple studies revealed experimental evidences that senescence suppresses tumour progression in mouse models and early human tumours (for review see Collado and Serrano, 2010). The relative contribution of OIS and telomere dysfunction induced senescence (TDIS) to tumour suppression and possible interconnections between the two pathways at the level of checkpoint induction were not investigated in previous studies. In this issue of The EMBO Journal, Suram et al (2012) describe the presence of TDIS in human precursor lesions but not in the corresponding malignant tumours. Mechanistically, the study shows that oncogenic signals cause replication fork stalling, resulting in telomeric DNA damage accumulation and activation of DNA damage checkpoints reminiscent to TDIS. Telomerase expression does not rescue replication fork stalling but prevents the accumulation of DNA damage at telomeres allowing a bypass of OIS.The study has several important implications for molecular pathways and therapeutic approaches in cancer that need to be further explored (Figure 1):Open in a separate windowFigure 1Traditional and new models of senescence in tumour suppression. (A) Traditional model of replicative senescence: Telomerase-negative tumour cell clones experience telomere shortening as a consequence of cell division. After a lack period depending on the initial telomere length, tumour cells accumulate telomere dysfunction and activation of senescence impairs tumour growth. Telomerase activation represents a late event allowing tumour progression. (B) New model of oncogene induced, telomere-dependent senescence: Oncogene activation leads to abrupt accumulation of DNA damage at telomeres resulting in senescence and tumour suppression. Telomerase-positive stem cells could be resistant to OIS and may be selected as the cell type of origin of tumour development.(i) Telomere length independent roles of telomeres in tumour suppressionThe classical model of telomere-dependent tumour suppression indicates that proliferation-dependent telomere shortening leads to telomere dysfunction, activation of DNA damage checkpoints, and induction of senescence suppressing the growth of telomerase-negative tumour clones. Studies on mouse models supported this concept showing that telomere shortening impairs the progression of initiated tumours in a telomere length-dependent manner (Feldser and Greider, 2007). The new data from Suram et al (2012) indicate that oncogene-induced replication fork stalling activates a telomere-dependent senescence checkpoint, which is independent of telomere length. The study shows that replication forks stall in response to oncogene activation throughout the genome. However, stalled replication forks are resolved in non-telomeric regions, whereas fork stalling inside telomeres leads to un-repairable DNA damage in telomerase-negative cells. These findings are in line with recent publication showing accumulation of un-repairable DNA damage in telomeric DNA in response to aging and stress-induced DNA damage (Fumagalli et al, 2012).(ii) Telomere length independent roles of telomerase in tumour progressionFollowing the classical model telomeres in tumour suppression (Figure 1A), telomerase re-activation is required for tumour progression by limiting telomere dysfunction and the induction of DNA damage checkpoints in response to telomere shortening. The new data from Suram et al (2012) indicate that telomerase has an additional telomere length independent role in tumour progression. The study shows that catalytically active telomerase prevents the activation of DNA damage signals originating from stalled replication forks inside telomeres in response to oncogene activation (Figure 1B). The exact mechanisms of telomerase-dependent healing of stalled replication forks at telomeres remain to be elucidated. It is also unclear whether telomerase activity can prevent any type of DNA damage at telomeres as an over-expression of TERT could not suppress irradiation-induced cellular senescence or the persistence of telomeric DDR following irradiation, H2O2, or chemotherapy induced DNA damage (Hewitt et al, 2012).The data could provide a plausible explanation for the increased tumorigenesis in telomerase transgenic mice—a finding which is difficult to explain by telomere length dependent effects of telomerase given the long telomere reserves in mouse tissues (Gonzalez-Suarez et al, 2001). According to the findings of Suram et al (2012), anti-telomerase therapies could have immediate anti-cancer effects in tumours depending on telomerase-mediated healing of stalled replication forks at telomeres. Specific markers for this dependency could be of clinical value. In addition, the data support the concept that somatic stem cells could represent the cell type of origin of cancers. In contrast to differentiated somatic cells, tissues stem cells are often telomerase-positive, indicating that stem cells might be less sensitive to OIS.  相似文献   

5.
The Mcm10 protein is essential for chromosomal DNA replication in eukaryotic cells. We purified the Saccharomyces cerevisiae Mcm10 (ScMcm10) and characterized its DNA binding properties. Electrophoretic mobility shift assays and surface plasmon resonance analysis showed that ScMcm10 binds stably to both double strand (ds) DNA and single strand (ss) DNA. On short DNA templates of 25 or 50 bp, surface plasmon resonance analysis showed a ∼1:1 stoichiometry of ScMcm10 to dsDNA. On longer dsDNA templates, however, multiple copies of ScMcm10 cooperated in the rapid assembly of a large, stable nucleoprotein complex. The amount of protein bound was directly proportional to the length of the DNA, with an average occupancy spacing of 21–24 bp. This tight spacing is consistent with a nucleoprotein structure in which ScMcm10 is aligned along the helical axis of the dsDNA. In contrast, the stoichiometry of ScMcm10 bound to ssDNA of 20–50 nucleotides was ∼3:1 suggesting that interaction with ssDNA induces the assembly of a multisubunit ScMcm10 complex composed of at least three subunits. The tight packing of ScMcm10 on dsDNA and the assembly of a multisubunit complex on ssDNA suggests that, in addition to protein-DNA, protein-protein interactions may be involved in forming the nucleoprotein complex. We propose that these DNA binding properties have an important role in (i) initiation of DNA replication and (ii) formation and maintenance of a stable replication fork during the elongation phase of chromosomal DNA replication.MCM10 is a ubiquitous, conserved gene essential for DNA replication in eukaryotes. It was first discovered in yeast genetic screens designed to detect mutants defective in DNA synthesis and minichromosome maintenance (1, 2). In vivo, Mcm10 associates with chromatin and chromosomal replication origins in human cells (hMcm10), Xenopus laevis (XMcm10), Schizosaccharomyces pombe (SpMcm10), and Saccharomyces cerevisiae (ScMcm10) (36). In S. cerevisiae, initiation of chromosomal replication occurs at multiple specific sites known as autonomously replicating sequences (ARSs)2 (7). Mutations in MCM10 enhance the loss rate of plasmids bearing specific ARSs (8), suggesting a function for ScMcm10 in initiation.In eukaryotic systems replication initiation is a cell cycle-regulated process characterized by a multistep sequential loading of ORC, Cdc6, Cdt1, and the Mcm2–7 complex onto the origin in G1 to form the pre-RC complex. Binding of ORC, Cdc6p, and Cdt1p to chromatin is a prerequisite for the recruitment of Mcm2–7 (9, 10). The next step in the assembly of the initiation replication apparatus in S. cerevisiae involves protein kinases (Cdc28 and Cdc7/Dbf4), and recruitment of Mcm10, Cdc45, and the GINS complex. The mechanism for targeting Mcm10 to replications origins is unknown. However, recent studies in S. cerevisiae have shown that Mcm10 and Mcm2–7 physically interact (6, 11). It is now believed that in late G1, chromatin-bound Mcm2–7 is responsible for the recruitment of Mcm10 presumably via protein-protein interactions (12). Prior studies in the Xenopus laevis system reached similar conclusions (4). Additional interactions of Mcm10 with other components of the pre-RC cannot be excluded (13).A key step in the initiation of replication is the local melting of an origin DNA sequence, which occurs at the G1/S transition and throughout the S phase. The mechanism of this essential DNA-melting process is not understood. There is no evidence in S. cerevisiae that the assembled pre-RC complex leads to the melting of an origin DNA sequence. This unwinding may occur only following the recruitment of Mcm10, raising the possibility that Mcm10 is a key component of the initiation machinery responsible for this process. Results of a study in the Xenopus egg extract system (4), which showed that omission of XMcm10 blocks unwinding of plasmid DNA and initiation of DNA replication, are consistent with this notion. An additional function of Mcm10 in initiation is in the recruitment of Cdc45 to replication origins, presumably via Mcm10-Cdc45 physical interactions (5, 14). Cdc45 is believed to be important for the activation of replication origins and the assembly of the replication elongation complex (15). Upon initiation of DNA replication, ScMcm10 moves from the origin to origin-proximal sequences suggesting that ScMcm10 associates with moving replication forks (12) and is consistent with the observation that elevated temperatures cause pausing of replication forks in a mcm10-1 ts mutant (8). Both ScMcm10 and SpMcm10 interact with DNA polymerase α supporting the notion that replication fork movement requires Mcm10. ScMcm10 and polymerase α form a complex on and off the DNA in vivo (12). In S. pombe, SpMcm10 stimulates the activity of polymerase α in vitro and associates with a primase activity (16, 17) that has not been reported in other eukaryotes (18).Previous studies with Mcm10 in other systems showed that Mcm10 binds DNA. Using a filter binding assay Fien and Hurwitz (16) reported that SpMcm10 from S. pombe binds well to ssDNA but barely interacts (20-fold lower affinity) with dsDNA. It has been suggested that binding of SpMcm10 to ssDNA is important for the recruitment of polymerase α (16). Recently, it has been reported that a DNA binding activity is also associated with XMcm10 protein from X. laevis. Measurements of fluorescence anisotropy were used to show binding of XMcm10 to short, 25-nucleotide-long oligonucleotides (18). These studies have shown that XMcm10 has similar affinities for ssDNA and dsDNA. Unlike SpMcm10, which harbors a single DNA-binding domain in the N-terminal half of the protein, XMcm10 seems to contain two distinct domains for binding DNA. The biological implication of having two DNA-binding domains is not clear.It appears that there are differences in the quaternary structure of Mcm10 from different organisms. Although SpMcm10 and XMcm10 may be a homodimer in solution (17, 18), a recent electron microscopy study suggested that human hMcm10 has a hexameric ring structure (19). The same study reported that hMcm10 interacts with ssDNA but failed to bind dsDNA. The differences in structure and DNA binding properties may reflect differences in the function of Mcm10 in various organisms as well as in the protein preparations.Here we report, for the first time, the characterization of the DNA binding properties of purified Mcm10 from S. cerevisiae. We show that ScMcm10 forms a stable complex with dsDNA and ssDNA. In addition, we demonstrate that dsDNA longer than 50 bp sustains oligomerization of ScMcm10. The number of ScMcm10 molecules bound is directly proportional to the size of the dsDNA, suggesting that ScMcm10 is tightly packed on the dsDNA, perhaps in a head to tail oligomeric structure. In contrast to a 25-bp-long dsDNA, which supports the binding of a single copy of ScMcm10, ssDNA containing only 20 nucleotides may sustain binding of as many as three copies of ScMcm10, suggesting that a ScMcm10 complex composed of at least 3 subunits assembles on ssDNA. We believe that these distinct binding properties to dsDNA and ssDNA are important for the ScMcm10 functions in initiation, formation of replication forks, and the maintenance of replication fork progression during chromosomal DNA replication.  相似文献   

6.
7.
8.
Comment on: Leman AR, et al. Cell Cycle 2012; 11:2337-47.DNA replication is at the heart of the inheritance of genetic material. A single replication fork can progress through hundreds of kilobases of DNA, melting parental double-stranded DNA and leaving newly synthesized strands in its wake. A beautiful illustration showing how the replication machinery accomplishes this complex task is one of the triumphs of molecular biology. However, it is known that DNA replication is not always as processive as the textbooks suggest. Specifically, the rate of fork progression varies depending on the regions being replicated, and the replication fork even stalls in some circumstances, during replication of heterochromatin or damaged DNA, for example. A stalled replication fork has two fates. It may restart DNA replication, or it may collapse after prolonged stalling. A collapsed replication fork is particularly dangerous for the genome, because the DNA intermediate left by the collapsed fork may form a double-stranded break, a highly mutagenic lesion that can undergo illegitimate recombination. To circumvent replication fork collapse, cells are equipped with specialized proteins that stabilize the stalled replication fork. Timeless and Tipin are highly conserved in eukaryotes. from yeast to humans, and form a complex to protect stalled replication forks.In a paper published in Cell Cycle, Noguchi and his group investigated how Timeless plays a role in telomere replication in human cells.1 Telomeres consist of tandem arrays of short repetitive DNA (TTAGGG/CCCTAA in mammals) at the ends of chromosomes and numerous associated proteins. Telomeres are essential for the stable maintenance of genomic DNA, because they protect the DNA termini from undergoing accidental recombination and exonuclease attack. Dysfunctional telomeres lead to genetic instability that eventually results in senescence and cancer development. Because of the heterochromatic nature of telomeres, it has been recognized that telomere DNA is one of the genomic regions that impede replication fork progression. Indeed, in vitro DNA replication experiments using SV40 DNA, and cell extracts demonstrated that telomere DNA is replicated less efficiently and incurs more fork stalling than non-telomeric DNA.2 Moreover, overexpression of telomere-DNA binding protein TRF1 in HeLa cells led to an accumulation of replicating telomeres, consistent with a slower replication rate of telomeres under those circumstance. Furthermore, experiments using TRF1-deleted murine cells showed that TRF1 is essential for efficient telomere DNA replication.3 Collectively, these results confirm that the telomere is a difficult-to-replicate region.There is an apparent contradiction between two earlier studies, however, with TRF1 described as an anti-replication protein in one report2 and a pro-replication protein in the other.3 One potential explanation for the inconsistency might be that TRF1 requires other protein(s) to perform its pro-replication function, and the second factor was missing in the TRF1-overexpression experiments. Noguchi and his colleagues investigated this possibility by testing whether Timeless is required for proficient telomere DNA replication.1 They found that Timeless-knockdown cells displayed telomere length shortening and an increased frequency of dysfunctional telomeres. In vitro replication assays of SV40 DNA revealed that Timeless-depleted extracts supported non-telomere replication proficiently, while telomere replication was inefficient. They then demonstrated that addition of recombinant TRF1 to the replication system slowed telomere replication. Importantly, Timeless depletion and TRF1 addition did not produce additive effects on telomere replication, suggesting that Timeless and TRF1 function in the same pathway. These results suggest a model as described in Figure 1. A replication fork frequently stalls at telomeres because of the molecularly crowded nature of telomeric chromatin. Timeless presumably encounters TRF1 at telomeres and protects the stalled fork from undergoing collapse. In the absence of Timeless, the stalled forks easily collapse, leading to an abrupt shortening of telomeres. Several questions remain to be answered. Given that Timeless moves along the genomic DNA as a component of the replication machinery,4 it will be particularly interesting to see how Timeless (or the replication machinery) interacts with telomeric chromatin. In such studies, a dynamic transaction between the regional chromatin at telomeres and the replication machinery may be revealed.Open in a separate windowFigure 1. Hard life at telomeres. (A) Mammalian telomeres consist of repetitive DNA that potentially forms higher-ordered structures [G-quartet(G4)-DNA] and numerous proteins, including telomere DNA-binding protein TRF1. (B) Replication fork is frequently stalled at telomeres. Overexpressed TRF1 slows down fork progression at the telomere, while endogenous TRF1 together with Timeless protein facilitates it. Timeless protects the stalled replication fork from collapse. (C) Telomeres are unique in that the most distal replication fork is not coupled with another fork progressing inversely. (D) Prolonged fork stalling may lead to the formation of a DNA double-strand break. Because of the lack of another fork compensating the telomere replication (C), the break immediately results in the abrupt single-step shortening of telomere DNAs.  相似文献   

9.
10.
Chromosomal abnormalities are frequently caused by problems encountered during DNA replication. Although the ATR-Chk1 pathway has previously been implicated in preventing the collapse of stalled replication forks into double-strand breaks (DSB), the importance of the response to fork collapse in ATR-deficient cells has not been well characterized. Herein, we demonstrate that, upon stalled replication, ATR deficiency leads to the phosphorylation of H2AX by ATM and DNA-PKcs and to the focal accumulation of Rad51, a marker of homologous recombination and fork restart. Because H2AX has been shown to play a facilitative role in homologous recombination, we hypothesized that H2AX participates in Rad51-mediated suppression of DSBs generated in the absence of ATR. Consistent with this model, increased Rad51 focal accumulation in ATR-deficient cells is largely dependent on H2AX, and dual deficiencies in ATR and H2AX lead to synergistic increases in chromatid breaks and translocations. Importantly, the ATM and DNA-PK phosphorylation site on H2AX (Ser139) is required for genome stabilization in the absence of ATR; therefore, phosphorylation of H2AX by ATM and DNA-PKcs plays a pivotal role in suppressing DSBs during DNA synthesis in instances of ATR pathway failure. These results imply that ATR-dependent fork stabilization and H2AX/ATM/DNA-PKcs-dependent restart pathways cooperatively suppress double-strand breaks as a layered response network when replication stalls.Genome maintenance prevents mutations that lead to cancer and age-related diseases. A major challenge in preserving genome integrity occurs in the simple act of DNA replication, in which failures at numerous levels can occur. Besides the mis-incorporation of nucleotides, it is during this phase of the cell cycle that the relatively stable double-stranded nature of DNA is temporarily suspended at the replication fork, a structure that is susceptible to collapse into DSBs.2 Replication fork stability is maintained by a variety of mechanisms, including activation of the ATR-dependent checkpoint pathway.The ATR pathway is activated upon the generation and recognition of extended stretches of single-stranded DNA at stalled replication forks (1-4). Genome maintenance functions for ATR and orthologs in yeast were first indicated by increased chromatid breaks in ATR-/- cultured cells (5) and by the “cut” phenotype observed in Mec1 (Saccharomyces cerevisiae) and Rad3 (Schizosaccharomyces pombe) mutants (6-9). Importantly, subsequent studies in S. cerevisiae demonstrated that mutation of Mec1 or the downstream checkpoint kinase Rad53 led to increased chromosome breaks at regions of the genome that are inherently difficult to replicate (10), and a decreased ability to reinitiate replication fork progression following DNA damage or deoxyribonucleotide depletion (11-14).In vertebrates, similar replication fork stabilizing functions have been demonstrated for ATR and the downstream protein kinase Chk1 (15-20). Several possible mechanisms have been put forward to explain how ATR-Chk1 and orthologous pathways in yeast maintain replication fork stability, including maintenance of replicative polymerases (α, δ, and ε) at forks (17, 21), regulation of branch migrating helicases, such as Blm (22-25), and regulation of homologous recombination, either positively or negatively (26-29).Consistent with the role of the ATR-dependent checkpoint in replication fork stability, common fragile sites, located in late-replicating regions of the genome, are significantly more unstable (5-10-fold) in the absence of ATR or Chk1 (19, 20). Because these sites are favored regions of instability in oncogene-transformed cells and preneoplastic lesions (30, 31), it is possible that the increased tumor incidence observed in ATR haploinsufficient mice (5, 32) may be related to subtle increases in genomic instability. Together, these studies indicate that maintenance of replication fork stability may contribute to tumor suppression.It is important to note that prevention of fork collapse represents an early response to problems occurring during DNA replication. In the event of fork collapse into DSBs, homologous recombination (HR) has also been demonstrated to play a key role in genome stability during S phase by catalyzing recombination between sister chromatids as a means to re-establish replication forks (33). Importantly, a facilitator of homologous recombination, H2AX, has been shown to be phosphorylated under conditions that cause replication fork collapse (18, 34).Phosphorylation of H2AX occurs predominantly upon DSB formation (34-38) and has been reported to require ATM, DNA-PKcs, or ATR, depending on the context (37-42). Although H2AX is not essential for HR, studies have demonstrated that H2AX mutation leads to deficiencies in HR (43, 44), and suppresses events associated with homologous recombination, such as the focal accumulation of Rad51, BRCA1, BRCA2, ubiquitinated-FANCD2, and Ubc13-mediated chromatin ubiquitination (43, 45-51). Therefore, through its contribution to HR, it is possible that H2AX plays an important role in replication fork stability as part of a salvage pathway to reinitiate replication following collapse.If ATR prevents the collapse of stalled replication forks into DSBs, and H2AX facilitates HR-mediated restart, the combined deficiency in ATR and H2AX would be expected to dramatically enhance the accumulation of DSBs upon replication fork stalling. Herein, we utilize both partial and complete elimination of ATR and H2AX to demonstrate that these genes work cooperatively in non-redundant pathways to suppress DSBs during S phase. As discussed, these studies imply that the various components of replication fork protection and regeneration cooperate to maintain replication fork stability. Given the large number of genes involved in each of these processes, it is possible that combined deficiencies in these pathways may be relatively frequent in humans and may synergistically influence the onset of age-related diseases and cancer.  相似文献   

11.
Functional coordination between DNA replication helicases and DNA polymerases at replication forks, achieved through physical linkages, has been demonstrated in prokaryotes but not in eukaryotes. In Saccharomyces cerevisiae, we showed that mutations that compromise the activity of the MCM helicase enhance the physical stability of DNA polymerase α in the absence of their presumed linker, Mcm10. Mcm10 is an essential DNA replication protein implicated in the stable assembly of the replisome by virtue of its interaction with the MCM2-7 helicase and Polα. Dominant mcm2 suppressors of mcm10 mutants restore viability by restoring the stability of Polα without restoring the stability of Mcm10, in a Mec1-dependent manner. In this process, the single-stranded DNA accumulation observed in the mcm10 mutant is suppressed. The activities of key checkpoint regulators known to be important for replication fork stabilization contribute to the efficiency of suppression. These results suggest that Mcm10 plays two important roles as a linker of the MCM helicase and Polα at the elongating replication fork—first, to coordinate the activities of these two molecular motors, and second, to ensure their physical stability and the integrity of the replication fork.The key players of the replication machinery are the DNA polymerases that synthesize the leading and lagging daughter strands and the replicative helicase that unwinds the parental strands ahead of the polymerases. Coordination between the helicase and the polymerases is critical during replication. Uncoupling of these two molecular machines, especially during lagging strand synthesis, may result in an unrestrained helicase and the exposure of extensive single-stranded DNA (ssDNA), as observed in checkpoint mutants treated with hydroxyurea (HU) (37). Although there is no direct evidence, the implication is that the replicative helicase would be moving at a faster pace than would the DNA polymerase if synchrony were destroyed. In Escherichia coli, the replicative helicase (DnaB) and the primase (DnaG) are coupled by direct contact to form a tight complex (3). In T7, processivity of the gp5 polymerase in lagging strand synthesis requires coupling to the gp4 helicase (16). Recent studies of the budding yeast Saccharomyces cerevisiae suggest that Mrc1 may couple DNA polymerase ɛ and the MCM helicase on the leading strand as well as activate the checkpoint response under replication stress (1, 22, 28). A candidate for coupling DNA polymerase α primase and the MCM helicase on the lagging strand is Mcm10, because Mcm10 interacts with subunits of the Mcm2-7 helicase (26, 29) as well as Polα (14, 33) and the stability of Polα requires Mcm10 in both budding yeast and human cells (8, 33). Mcm10 is an essential protein known to be involved in various aspects of the replication process. It is required during both initiation and elongation steps of DNA replication and interacts with a wide range of replication factors, such as ORC (17, 23, 29), MCM helicase, DNA polymerases ɛ and δ (23), Cdc45 (34), and Polα (33). Therefore, Mcm10 is important for the overall stability of the elongation complex, but its essential function remains unknown.Accumulating evidence suggests that the major function of many checkpoint proteins is the stabilization of the replication machinery at the fork (9, 22, 39), in addition to regulation of the temporal and spatial firing of origins and prevention of premature mitosis (31, 35, 39). The main signal that leads to checkpoint activation is believed to be the exposure of RPA-coated ssDNA (42). In Xenopus, ssDNA exposure has been shown to be mediated by a functional uncoupling between the polymerase and the helicase (7), and it has been shown that the level of checkpoint activation depended on the extent of ssDNA accumulation. This observation suggests that uncoupling of the polymerase and the helicase activity would result in ssDNA accumulation that in turn would activate the checkpoint pathway to stabilize the fork.In our study, we carried out a random and a gene-targeted mutagenesis screen to identify mutations that suppress the conditional lethality of mcm10 caused by the lability of Mcm10 in budding yeast (27). We found suppressor mutations in MCM2, which encodes one of the six distinct subunits of the MCM helicase. These mcm2 mutations correct the fork defects of mcm10, particularly that which leads to Polα instability. The altered helicase activity and activation of the checkpoint pathway of the mcm2 mutants appeared to be required for viability of mcm10 mcm2. We showed that uncoupling the MCM helicase and DNA polymerase α by destabilizing Mcm10 leads to accumulation of ssDNA, which is suppressed by reducing the MCM helicase activity. Our findings suggest that the physical coupling of Polα and the helicase by Mcm10 may be replaced by an alternative stabilization mechanism that involves slowing down the helicase and activating the checkpoint proteins.  相似文献   

12.
13.
The genomes of mycobacteriophages of the L5 family, which includes the lytic phage D29, contain several genes putatively linked to DNA synthesis. One such gene is 65, which encodes a protein belonging to the RecA/DnaB helicase superfamily. In this study a recombinant version of the mycobacteriophage D29 gp65 was functionally characterized. The results indicated that it is not a helicase as predicted but an exonuclease that removes 3′ arms from forked structures in an ATP-dependent manner. The gp65 exonuclease acts progressively from the 3′ end, until the fork junction is reached. As it goes past, its progress is stalled over a stretch of seven to eight nucleotides immediately downstream of the junction. It efficiently acts on forked structures with single stranded arms. It also acts upon 5′ and 3′ flaps, though with somewhat relaxed specificity, but not on double-stranded forks. Sequence comparison revealed the presence of a KNRXG motif in the C-terminal half of the protein. This is a conserved element found in the RadA/Sms family of DNA repair proteins. A mutation (R203G) in this motif led to complete loss of nuclease activity. This indicated that KNRXG plays an important role in the nuclease function of not only gp65, but possibly other RadA/Sms family proteins as well. This is the first characterization of a bacteriophage-derived RadA/Sms class protein. Given its mode of action, it is very likely that gp65 is involved in processing branched replication intermediates formed during the replication of phage DNA.Fork structures are intricately associated with DNA replication. Such structures result due to unwinding of the DNA ahead of the replicating machinery. The unwound single strands are then used as templates for the synthesis of the new strands, either continuously (leading strand) or discontinuously (lagging strand). Repair of stalled forks involve complex mechanisms which may vary from one organism to another (5). However, in most cases the process requires nucleases that recognize stalled fork structures and cleave them specifically. Such nucleases are generally referred to as structure-specific nucleases (25). One such nuclease named FEN1 found in eukaryotes has been studied fairly extensively, and it is believed that this nuclease is involved in the removal of 5′ flaps from Okazaki fragments (11, 23). FEN1 belongs to a larger family of structure-specific nucleases, which includes human XPG (17), an endonuclease related to the disease xeroderma pigmentosa. Although the XPG family is associated with the removal of 5′ flaps the XPF type proteins are needed for removing the 3′ flaps (3). Similar proteins have been found in several Archaea (28). In Escherichia coli, the Holliday junction resolving enzyme system RuvABC is believed to be involved in resolving stalled forks by creating double-stranded breaks, which may be repaired through homologous recombination (29). Studies in E. coli have revealed that there are multiple redundant pathways that are capable of repairing stalled forks. One such pathway involves a protein named RadA/Sms, the absence of which results in partial increase in sensitivity to radiation in E. coli (2). Genes encoding RadA/Sms family proteins are present in many bacteria, including mycobacteria. Most of these members carry a conserved element KNRFG. It is believed (2) that RadA/Sms family of proteins may generate double-stranded breaks at fork junctions, although this has not been specifically demonstrated.Mycobacteriophages of the L5 family, which includes D29, BxB1, may be either temperate or potentially temperate (D29) (14, 15, 27). Despite their temperate character these phages share a strong resemblance with lytic phages. An important feature shared by lytic phages in general is their ability to synthesize DNA using phage-encoded DNA polymerases (13). They also possess many genes linked to nucleotide metabolism. It appears that as far as DNA replication is concerned, lytic phages prefer to be self-sufficient. This is apparently an important issue since lytic phages inactivate their host and therefore host-specific functions cannot be used to support phage growth.Following the availability of the genome sequence, many interesting aspects of mycobacteriophages have come to light. The central region of mycobacteriophage L5/D29 genome has been predicted to harbor several genes whose products may contribute directly or indirectly toward synthesis of new DNA strands. In a recent investigation from this laboratory it has been demonstrated (4) that at least some of the genes in this region are involved in the production of deoxyribonucleotide precursors which are probably needed at increased levels during phage replication. Apart from these genes there are several others which probably encode DNA polymerization related functions. One such gene that drew our interest was gene 65, which appears to encode a RecA/DnaB helicase superclass protein (22). The N-terminal region of this protein contains the Walker motifs A and B, which are characteristically present in the members of the RecA/DnaB superfamily. Walker motifs A and B (30) are found in proteins that hydrolyze ATP for executing their respective functions. To investigate the possible function of gp65, its gene was overexpressed in E. coli, and the recombinant protein was purified. Assays performed with the recombinant gp65 revealed that it is a structure-specific nuclease that acts exonucleolytically on fork structures, resulting in truncated forms lacking the 3′ arm. This function was demonstrated to require a particular motif KNRXG that is omnipresent in the RadA/Sms family of proteins (2). This characterization of D29 gp65 could give us better insight into how mycobacteriophages replicate their DNA within their hosts.  相似文献   

14.
Translesion DNA synthesis (TLS) is a process whereby specialized DNA polymerases are recruited to bypass DNA lesions that would otherwise stall high-fidelity polymerases. We provide evidence that TLS across cisplatin intrastrand cross-links is performed by multiple translesion DNA polymerases. First, we determined that PCNA monoubiquitination by RAD18 is necessary for efficient bypass of cisplatin adducts by the TLS polymerases eta (Polη), REV1, and zeta (Polζ) based on the observations that depletion of these proteins individually leads to decreased cell survival, cell cycle arrest in S phase, and activation of the DNA damage response. Second, we showed that in addition to PCNA monoubiquitination by RAD18, the Fanconi anemia core complex is also important for recruitment of REV1 to stalled replication forks in cisplatin treated cells. Third, we present evidence that REV1 and Polζ are uniquely associated with protection against cisplatin and mitomycin C-induced chromosomal aberrations, and both are necessary for the timely resolution of DNA double-strand breaks associated with repair of DNA interstrand cross-links. Together, our findings indicate that REV1 and Polζ facilitate repair of interstrand cross-links independently of PCNA monoubiquitination and Polη, whereas RAD18 plus Polη, REV1, and Polζ are all necessary for replicative bypass of cisplatin intrastrand DNA cross-links.Maintenance of genomic integrity involves the activation of cell cycle checkpoints coupled with DNA repair. Despite these sophisticated mechanisms to remove DNA lesions prior to DNA replication, replication forks may inevitably encounter nonrepaired lesions that block high fidelity polymerases, potentially leading to replication fork instability, gaps in replicated DNA, and the generation of DNA double-strand breaks (DSBs). In order to preserve replication fork stability by allowing replication through polymerase blocking lesions, template DNA containing a damaged base or abasic site can be replicated through the actions of specialized translesion DNA synthesis (TLS) polymerases (61). A key event in the regulation of TLS is the monoubiquitination of PCNA, a homotrimeric protein that functions as an auxiliary factor for DNA polymerases (28, 31, 57, 60). The RAD6 (E2)-RAD18 (E3) complex specifically monoubiquitinates PCNA on Lys-164 in response to replication fork stalling. This event is thought to operate as a molecular switch from normal DNA replication to the TLS pathway based on the observations that association of Y-family TLS polymerases with monoubiquitinated PCNA is strengthened through the cooperative binding of one or more ubiquitin-binding domains (UBM or UBZ) plus a PCNA-interacting domain (6, 25).Extensive biochemical evidence suggests that replication through a large variety of lesions requires the sequential action of two TLS polymerases (44). The Y-family polymerase eta (Polη) plays a key role in the efficient and error-free bypass of cyclobutane pyrimidine (TT) dimers, one of the major lesions resulting from exposure to UV radiation (45). In contrast, Polη can only insert a nucleotide directly opposite other lesions and requires an additional TLS polymerase, such as Polζ, to extend beyond the insertion (45). Polζ is comprised of the REV3 catalytic subunit that shares homology with B-family polymerases plus the REV7 accessory subunit (34). Polζ is unusual compared to other TLS polymerases due to the fact that it is relatively efficient at extending beyond mispaired primer termini and nucleotides inserted opposite a variety of DNA lesions, although this may occur in a potentially mutagenic manner (45). Genetic evidence in yeast suggest that Polζ activity is regulated by the Y family REV1 polymerase (21). In addition to a UBM domain that directly interacts with monoubiquitinated PCNA, REV1 possesses an N-terminal BRCT motif that directly contacts PCNA and potentially other proteins (24, 25). In addition, REV1 possesses a unique protein interaction domain in its carboxy terminus that interacts with the Polζ accessory subunit, REV7, and other TLS polymerases, including Polη and the Polζ catalytic subunit, REV3 (1, 18, 23, 40, 58). The characterization of these protein-protein interaction domains has led to the proposal that REV1 facilitates polymerase switching from a polymerase that directly inserts a nucleotide opposite a damaged base and Polζ, which subsequently performs the extension step beyond the inserted nucleotide opposite the damaged base (21).In addition to facilitating direct lesion bypass and filling in postreplicative gaps in DNA, REV1 and Polζ may also play an important role in the repair of interstrand cross-links (46, 63). Deletion of REV1, REV3, or REV7 in chicken DT40 cells leads to remarkable hypersensitivity to a wide variety of genotoxic stresses, most notably cisplatin and other DNA cross-linking agents such as mitomycin C (MMC) (38, 41, 55, 56). The genetic epistasis observed between REV1, REV3, and the Fanconi anemia (FA) complementation group C (FANCC) gene for cisplatin sensitivity further implicates TLS in the interstrand cross-link repair pathway (38). Current models suggest that when two replication forks converge upon an interstrand cross-link, the MUS81-EME1 endonuclease recognizes and cleaves the resulting branched DNA structure by making an incision at one side of the interstrand cross-link creating a replication-associated DSB (26). The XPF-ERCC1 endonuclease uncouples the cross-linked cDNA strands by making an incision on the other side of the interstrand cross-link (37). Recent biochemical evidence suggests that Polζ performs DNA synthesis opposite the DNA strand containing the residual cross-link and this process may be necessary to prepare the daughter strand for subsequent homologous recombination repair of the replication-associated DSB (46).Agents that introduce intra- and interstrand cross-links are widely used in cancer chemotherapy, and thus understanding the means by which cells repair or cope with these lesions will be instrumental in identifying novel mechanisms leading to drug resistance and designing new agents refractory to DNA damage tolerance mechanisms. Polη, REV1, and Polζ have all been implicated in mediating TLS past cisplatin intrastrand cross-links since lowering their expression increases sensitivity and reduces cisplatin-induced mutagenesis in human cancer cells (2, 5, 12, 42, 62). Furthermore, biochemical and structural analyses of Polη identified this polymerase as being capable of efficiently inserting dCTP opposite the 3′dG of a 1,2-d(GpG) cisplatin intrastrand cross-link (3). Here, we demonstrate that RAD18, Polη, and REV1 all localized to sites of replication stress marked by PCNA and γ-H2AX foci after treatment of cells with cisplatin. However, REV1 focus formation is specifically dependent upon both RAD18 and a functional FA core complex, suggesting FA core proteins are also necessary for directing REV1 to cisplatin-induced stalled replication forks. In addition, depletion of RAD18, Polη, REV1, or Polζ proteins lead to the induction of cellular responses indicative of inefficient lesion bypass of cisplatin adducts. Unexpectedly, we found that REV1- or Polζ-depleted cells displayed a greater loss in cell viability and the accumulation of chromosome aberrations and failed to resolve DSBs after cisplatin treatment. These results lead us to hypothesize that REV1 and Polζ may be necessary for the repair of cisplatin interstrand cross-links in addition to performing lesion bypass of cisplatin intrastrand cross-links. In agreement with this concept, we found that REV1 and Polζ-depleted cells were uniquely hypersensitive to MMC, accumulated greater numbers of chromosome aberrations, and failed to resolve replication-associated DSBs induced by MMC treatment.Together our findings support a model where replicative bypass of cisplatin intrastrand cross-links requires cooperation of multiple TLS polymerases in mammalian cells and is triggered by PCNA monoubiquitination. Our results also provide evidence that REV1 and Polζ facilitate repair of interstrand cross-links in human cells, and this process is likely independent of PCNA monoubiquitination.  相似文献   

15.
16.
Comment on: Kang YH, et al. Proc Natl Acad Sci USA 2012; 109:6042-7.In eukaryotes, the complex comprised of Mcm2–7, Cdc45 and GINS (CMG) is essential for DNA replication. Several lines of evidence indicate that the Mcm2–7 complex is the motor of the replicative helicase (reviewed in ref. 1), which is activated by its association with Cdc45 and GINS.2 Recently, we described the isolation and characterization of the human (h) CMG complex.3 In HeLa cells, this complex was formed only on chromatin and, following its isolation from cells, exhibited DNA helicase activity. Purified from Sf9 cells, hCMG possesses 3′→5′ DNA helicase activity, indicating that it moves ahead of the leading-strand DNA polymerase (pol). In contrast, the prokaryotic helicase DnaB, which unwinds DNA in the 5′→3′ direction, moves on the lagging strand. Detailed information about the progression of the prokaryotic replication fork was obtained using the rolling-circle method (ref. 4 and references therein). These studies permitted a detailed characterization of the joint action of the replicative pol and replicative helicase. In the rolling-circle reaction, the pol extends the 3′ end of a primer annealed to a minicircle that is then unwound simultaneously by the helicase (for a possible arrangement of proteins at the replication fork, see Fig. 1). The emerging single-stranded 5′-tail provides the template for lagging-strand synthesis. In most experiments, minicircles were engineered to contain only three nucleotides, allowing the distinction between leading- and lagging-strand nucleotide incorporation.Open in a separate windowFigure 1. Model of the human replication fork. The CMG complex unwinds DNA in the 3′→5′ direction. Polα/primase synthesizes primers to initiate leading- and lagging-strand synthesis. Polε and polδ are assigned as leading- and lagging-strand polymerases based on evidence in yeast.5,6 Both pols require the processivity factor PCNA. RPA binds to single-stranded DNA. Additional proteins are required for DNA replication, of which only Ctf4 and Mcm10 are shown for simplicity.We initiated experiments to develop a eukaryotic replication fork in order to investigate whether the hCMG helicase activity could be coupled with the replicative pols.3 We set up rolling circle reactions using a 200-nt minicircle, the putative leading strand pol ε5 and hCMG and showed that DNA chains longer than 10 kb were produced (representing > 50 turns of the circle). The putative lagging strand pol δ,6 however, did not replace hpol ε in this reaction, though both pols extended primers on single-stranded M13 to full-length products (about 7 kb). It is tempting to speculate that an interaction between hCMG and hpol ε, but not hpol δ, contributes to their different activities. Specific interaction between GINS, a component of the CMG complex, and hpol ε has been demonstrated.7 However, it is presently unclear whether this contributes to the observed preferential role of pol ε and thus requires further examination.The processivity of the CMG complex alone was about 500 bp, which was stimulated to about 1 kbp by the addition of a single-strand DNA binding protein, either E. coli SSB or hRPA. The rolling circle reaction is also dependent on E. coli SSB, presumably to sequester the emerging single-stranded 5′ tail. Surprisingly, hRPA did not replace E. coli SSB in the rolling circle reaction. This was attributed to its inhibitory effects on pol ε activity in vitro. The influence of hRPA on eukaryotic fork progression is presently unclear. In the in vitro SV40 viral DNA replication system, hRPA is essential for DNA synthesis and cannot be replaced by E. coli SSB (reviewed in ref. 8). In this system, the SV40 large T-antigen acts as the replicative helicase, and hRPA is essential for its interaction with the hpolα/primase complex, which positions primase to initiate RNA chains. In the SV40 replication reaction, hpol δ synthesizes both leading and lagging strands. Surprisingly, while prokaryotic pols (and their processivity factors) can replace hpol δ and its auxiliary proteins in the in vitro SV40 elongation reaction, hpol ε does not play a role,8 suggesting that, in this system, the action of hpol ε is preferentially excluded. Importantly, no rolling circle synthesis was detected when hpol δ was used in lieu of hpol ε.3 Whether a similar mechanism leading to the exclusion of hpol δ from leading-strand synthesis is operational with the CMG helicase remains to be investigated. Using an archaeal system consisting of Pol B, RFC, PCNA, the 3′→5′ DNA helicase Mcm and the DNA primase, we have performed both leading- and lagging-strand synthesis on a rolling circle substrate.9 Currently, our efforts are focused on the synthesis of the lagging-strand with human proteins.In cells, the replication machinery duplicates chromatinized DNA. Thus, it is likely that chromatin remodeling factors and nucleosome chaperones play roles in the progression of the replication fork. In support of this notion, FACT was identified as a component of the yeast replisome progression complex.10 Various other proteins associate with the replication fork, such as Mcm10, Ctf4, Tim-Tipin and Claspin. The effects of these proteins on the in vitro replication reaction in eukaryotes remain to be examined.  相似文献   

17.
Homologous recombination (HR) is a key pathway that repairs DNA double‐strand breaks (DSBs) and helps to restart stalled or collapsed replication forks. How HR supports replication upon genotoxic stress is not understood. Using in vivo and in vitro approaches, we show that the MMS22L–TONSL heterodimer localizes to replication forks under unperturbed conditions and its recruitment is increased during replication stress in human cells. MMS22L–TONSL associates with replication protein A (RPA)‐coated ssDNA, and the MMS22L subunit directly interacts with the strand exchange protein RAD51. MMS22L is required for proper RAD51 assembly at DNA damage sites in vivo, and HR‐mediated repair of stalled forks is abrogated in cells expressing a MMS22L mutant deficient in RAD51 interaction. Similar to the recombination mediator BRCA2, recombinant MMS22L–TONSL limits the assembly of RAD51 on dsDNA, which stimulates RAD51‐ssDNA nucleoprotein filament formation and RAD51‐dependent strand exchange activity in vitro. Thus, by specifically regulating RAD51 activity at uncoupled replication forks, MMS22L–TONSL stabilizes perturbed replication forks by promoting replication fork reversal and stimulating their HR‐mediated restart in vivo.  相似文献   

18.
19.
Replication Protein A is a single-stranded (ss) DNA-binding protein that is highly conserved in eukaryotes and plays essential roles in many aspects of nucleic acid metabolism, including replication, recombination, DNA repair and telomere maintenance. It is a heterotrimeric complex consisting of three subunits: RPA1, RPA2 and RPA3. It possesses four DNA-binding domains (DBD), DBD-A, DBD-B and DBD-C in RPA1 and DBD-D in RPA2, and it binds ssDNA via a multistep pathway. Unlike the RPA1 and RPA2 subunits, no ssDNA-RPA3 interaction has as yet been observed although RPA3 contains a structural motif found in the other DBDs. We show here using 4-thiothymine residues as photoaffinity probe that RPA3 interacts directly with ssDNA on the 3′-side on a 31 nt ssDNA.The replication protein A (RPA) is a single-stranded (ss) DNA-binding protein that is highly conserved in eukaryotes (1–3). RPA is one of the key players in various essential processes of DNA metabolism including replication, recombination, DNA repair and telomere maintenance (1,2,4–9). The functions of this protein are based on its DNA-binding activity and specific protein–protein interactions. Its ssDNA binding properties depend on DNA length and nucleotide sequence (6,10–13). RPA is a heterotrimeric protein, composed of 70-, 32- and 14-kDa subunits, commonly referred to as RPA1, RPA2 and RPA3, respectively. There are four DNA-binding domains (DBD) located in RPA1 (DBD A, DBD B, DBD C and DBD F), one located in RPA2 (DBD D) and one belongs to RPA3 (DBD E). RPA interacts with ssDNA via four DBD: DBD A, DBD B, DBD C and DBD D (14).It is now accepted (11) that RPA binds to ssDNA in a sequential pathway with a defined polarity (15–17). RPA binds ssDNA with three different binding modes. First, binding initially involves an unstable recognition site of 8–10 nt with the high-affinity DBD A and DBD B domains on the 5′-side of the occluded ssDNA; it is designated ‘compact conformation’ or 8–10 nt binding mode. Second, this step is followed by the weaker binding of DBD C, on the 3′-side, leading to an intermediate or ‘elongated contracted’ (13–22 nt) binding mode (18–19). Finally binding of DBD D on the 3′-side forms a stable ‘elongated extended’ complex characterized by a 30 nt long occluded binding site (30 nt binding mode). Although RPA3 contains an Oligonucleotide-Binding (OB)-fold motif found in the other DBDs, there is presently no biochemical evidence that this subunit directly contacts DNA. Thus positioning of the RPA3 subunit relative to the other domains is still speculative (11,20). It has been clearly demonstrated that RPA3 is crucial for RPA function (1,2): RPA3 is involved in heterotrimer formation and is responsible for the polarity of binding to DNA (11,21,22). The scope of the data indicates that either RPA3 participates only in protein–protein interactions or that putative interaction of RPA3 with ssDNA is unstable and too transient to be detected by standard biochemical experiments. This latter possibility is likely if such interaction is provided by the 3′-side of the ssDNA, since it has been suggested that this region might be transiently accessible to the RPA DBD domains (23,24).In the past few years, thionucleobases have been extensively used as intrinsic photolabels to probe the structure in solution of folded molecules and to identify transient contacts within nucleic acids and/or between nucleic acids and proteins, in nucleoprotein assemblies (25). Thio residues such as 4-thiothymine and 6-thioguanine absorb light at wavelengths longer than 320 nm, and thus can be selectively photo-activated. Owing to the high photo-reactivity of their triplet state, they exhibit high photo-cross-linking ability towards nucleic acid bases as well as towards amino acid residues. Here we used a combination of approaches including gel retardation assays, chemical cross-linking and cross-linking with photoreactive ssDNA probes containing 4-thiothymine, introduced at a defined site in the sequence of the ssDNA, to study interactions present in human RPA (hRPA): ssDNA complexes. These studies coupled with the identification of cross-linked targets using specific antibodies revealed that in the elongated extended hRPA:ssDNA complex RPA3 closely contacts the 3′-end positioned nucleotide and yields a covalent adduct with zero-length photolabel.  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