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 15 毫秒
1.
The base analog 6-N-hydroxylaminopurine (HAP) induces bidirectional GC -> AT and AT -> GC transitions that are enhanced in DNA polymerase ε and δ 3' -> 5' exonuclease-deficient yeast mutants, pol2-4 and pol3-01, respectively. We have constructed a set of isogenic strains to determine whether the DNA polymerases δ and ε contribute equally to proofreading of replication errors provoked by HAP during leading and lagging strand DNA synthesis. Site-specific GC -> AT and AT -> GC transitions in a Pol(+), pol2-4 or pol3-01 genetic background were scored as reversions of ura3 missense alleles. At each site, reversion was increased in only one proofreading-deficient mutant, either pol2-4 or pol3-01, depending on the DNA strand in which HAP incorporation presumably occurred. Measurement of the HAP-induced reversion frequency of the ura3 alleles placed into chromosome III near to the defined active replication origin ARS306 in two orientations indicated that DNA polymerases ε and δ correct HAP-induced DNA replication errors on opposite DNA strands.  相似文献   

2.
L. Giot  R. Chanet  M. Simon  C. Facca    G. Faye 《Genetics》1997,146(4):1239-1251
The POL3 encoded catalytic subunit of DNA polymerase δ possesses a highly conserved C-terminal cysteine-rich domain in Saccharomyces cerevisiae. Mutations in some of its cysteine codons display a lethal phenotype, which demonstrates an essential function of this domain. The thermosensitive mutant pol3-13, in which a serine replaces a cysteine of this domain, exhibits a range of defects in DNA repair, such as hypersensitivity to different DNA-damaging agents and deficiency for induced mutagenesis and for recombination. These phenotypes are observed at 24°, a temperature at which DNA replication is almost normal; this differentiates the functions of POL3 in DNA repair and DNA replication. Since spontaneous mutagenesis and spontaneous recombination are efficient in pol3-13, we propose that POL3 plays an important role in DNA repair after irradiation, particularly in the error-prone and recombinational pathways. Extragenic suppressors of pol3-13 are allelic to sdp5-1, previously identified as an extragenic suppressor of pol3-11. SDP5, which is identical to HYS2, encodes a protein homologous to the p50 subunit of bovine and human DNA polymerase δ. SDP5 is most probably the p55 subunit of Polδ of S. cerevisiae and seems to be associated with the catalytic subunit for both DNA replication and DNA repair.  相似文献   

3.
4.
5.
DNA polymerase δ (pol δ) is one of the two main replicative polymerases in eukaryotes; it synthesizes the lagging DNA strand and also functions in DNA repair. In previous work, we demonstrated that heterozygous expression of the pol δ L604G variant in mice results in normal life span and no apparent phenotype, whereas a different substitution at the same position, L604K, is associated with shortened life span and accelerated carcinogenesis. Here, we report in vitro analysis of the homologous mutations at position Leu-606 in human pol δ. Four-subunit human pol δ variants that harbor or lack 3′ → 5′-exonucleolytic proofreading activity were purified from Escherichia coli. The pol δ L606G and L606K holoenzymes retain catalytic activity and processivity similar to that of wild type pol δ. pol δ L606G is highly error prone, incorporating single noncomplementary nucleotides at a high frequency during DNA synthesis, whereas pol δ L606K is extremely accurate, with a higher fidelity of single nucleotide incorporation by the active site than that of wild type pol δ. However, pol δ L606K is impaired in the bypass of DNA adducts, and the homologous variant in mouse embryonic fibroblasts results in a decreased rate of replication fork progression in vivo. These results indicate that different substitutions at a single active site residue in a eukaryotic polymerase can either increase or decrease the accuracy of synthesis relative to wild type and suggest that enhanced fidelity of base selection by a polymerase active site can result in impaired lesion bypass and delayed replication fork progression.  相似文献   

6.
DNA polymerases (Pols) ε and δ perform the bulk of yeast leading- and lagging-strand DNA synthesis. Both Pols possess intrinsic proofreading exonucleases that edit errors during polymerization. Rare errors that elude proofreading are extended into duplex DNA and excised by the mismatch repair (MMR) system. Strains that lack Pol proofreading or MMR exhibit a 10- to 100-fold increase in spontaneous mutation rate (mutator phenotype), and inactivation of both Pol δ proofreading (pol3-01) and MMR is lethal due to replication error-induced extinction (EEX). It is unclear whether a similar synthetic lethal relationship exists between defects in Pol ε proofreading (pol2-4) and MMR. Using a plasmid-shuffling strategy in haploid Saccharomyces cerevisiae, we observed synthetic lethality of pol2-4 with alleles that completely abrogate MMR (msh2Δ, mlh1Δ, msh3Δ msh6Δ, or pms1Δ mlh3Δ) but not with partial MMR loss (msh3Δ, msh6Δ, pms1Δ, or mlh3Δ), indicating that high levels of unrepaired Pol ε errors drive extinction. However, variants that escape this error-induced extinction (eex mutants) frequently emerged. Five percent of pol2-4 msh2Δ eex mutants encoded second-site changes in Pol ε that reduced the pol2-4 mutator phenotype between 3- and 23-fold. The remaining eex alleles were extragenic to pol2-4. The locations of antimutator amino-acid changes in Pol ε and their effects on mutation spectra suggest multiple mechanisms of mutator suppression. Our data indicate that unrepaired leading- and lagging-strand polymerase errors drive extinction within a few cell divisions and suggest that there are polymerase-specific pathways of mutator suppression. The prevalence of suppressors extragenic to the Pol ε gene suggests that factors in addition to proofreading and MMR influence leading-strand DNA replication fidelity.  相似文献   

7.
DNA polymerase ε (Polε) is a multi-subunit polymerase that contributes to genomic stability via its roles in leading strand replication and the repair of damaged DNA. Here we report the ternary structure of the Polε catalytic subunit (Pol2) bound to a nascent G:C base pair (Pol2G:C). Pol2G:C has a typical B-family polymerase fold and embraces the template-primer duplex with the palm, fingers, thumb and exonuclease domains. The overall arrangement of domains is similar to the structure of Pol2T:A reported recently, but there are notable differences in their polymerase and exonuclease active sites. In particular, we observe Ca2+ ions at both positions A and B in the polymerase active site and also observe a Ca2+ at position B of the exonuclease site. We find that the contacts to the nascent G:C base pair in the Pol2G:C structure are maintained in the Pol2T:A structure and reflect the comparable fidelity of Pol2 for nascent purine-pyrimidine and pyrimidine-purine base pairs. We note that unlike that of Pol3, the shape of the nascent base pair binding pocket in Pol2 is modulated from the major grove side by the presence of Tyr431. Together with Pol2T:A, our results provide a framework for understanding the structural basis of high fidelity DNA synthesis by Pol2.  相似文献   

8.
Elucidating the sources of genetic variation within microsatellite alleles has important implications for understanding the etiology of human diseases. Mismatch repair is a well described pathway for the suppression of microsatellite instability. However, the cellular polymerases responsible for generating microsatellite errors have not been fully described. We address this gap in knowledge by measuring the fidelity of recombinant yeast polymerase δ (Pol δ) and ? (Pol ?) holoenzymes during synthesis of a [GT/CA] microsatellite. The in vitro HSV-tk forward assay was used to measure DNA polymerase errors generated during gap-filling of complementary GT(10) and CA(10)-containing substrates and ~90 nucleotides of HSV-tk coding sequence surrounding the microsatellites. The observed mutant frequencies within the microsatellites were 4 to 30-fold higher than the observed mutant frequencies within the coding sequence. More specifically, the rate of Pol δ and Pol ? misalignment-based insertion/deletion errors within the microsatellites was ~1000-fold higher than the rate of insertion/deletion errors within the HSV-tk gene. Although the most common microsatellite error was the deletion of a single repeat unit, ~ 20% of errors were deletions of two or more units for both polymerases. The differences in fidelity for wild type enzymes and their exonuclease-deficient derivatives were ~2-fold for unit-based microsatellite insertion/deletion errors. Interestingly, the exonucleases preferentially removed potentially stabilizing interruption errors within the microsatellites. Since Pol δ and Pol ? perform not only the bulk of DNA replication in eukaryotic cells but also are implicated in performing DNA synthesis associated with repair and recombination, these results indicate that microsatellite errors may be introduced into the genome during multiple DNA metabolic pathways.  相似文献   

9.
Impeded DNA replication or a deficiency of its control may critically threaten the genetic information of cells, possibly resulting in genome alterations, such as gross chromosomal translocations, microsatellite instabilities, or increased rates of homologous recombination (HR). We examined an Arabidopsis thaliana line derived from a forward genetic screen, which exhibits an elevated frequency of somatic HR. These HR events originate from replication stress in endoreduplicating cells caused by reduced expression of the gene coding for the catalytic subunit of the DNA polymerase δ (POLδ1). The analysis of recombination types induced by diverse alleles of polδ1 and by replication inhibitors allows the conclusion that two not mutually exclusive mechanisms lead to the generation of recombinogenic breaks at replication forks. In plants with weak polδ1 alleles, we observe genome instabilities predominantly at sites with inverted repeats, suggesting the formation and processing of aberrant secondary DNA structures as a result of the accumulation of unreplicated DNA. Stalled and collapsed replication forks account for the more drastic enhancement of HR in plants with strong polδ1 mutant alleles. Our data suggest that efficient progression of DNA replication, foremost on the lagging strand, relies on the physiological level of the polymerase δ complex and that even a minor disturbance of the replication process critically threatens genomic integrity of Arabidopsis cells.  相似文献   

10.
The tetrameric GINS complex, consisting of Sld5-Psf1-Psf2-Psf3, plays an essential role in the initiation and elongation steps of eukaryotic DNA replication, although its biochemical function is unclear. Here we investigate the function of GINS in fission yeast, using fusion of Psf1 and Psf2 subunits to a steroid hormone-binding domain (HBD) to make GINS function conditional on the presence of β-estradiol. We show that inactivation of Psf1-HBD causes a tight but rapidly reversible DNA replication arrest phenotype. Inactivation of Psf2-HBD similarly blocks premeiotic DNA replication and leads to loss of nuclear localization of another GINS subunit, Psf3. Inactivation of GINS has distinct effects on the replication origin association and chromatin binding of two of the replicative DNA polymerases. Inactivation of Psf1 leads to loss of chromatin binding of DNA polymerase ε, and Cdc45 is similarly affected. In contrast, chromatin association of the catalytic subunit of DNA polymerase α is not affected by defective GINS function. We suggest that GINS functions in a pathway that involves Cdc45 and is necessary for DNA polymerase ε chromatin binding, but that a separate pathway sets up the chromatin association of DNA polymerase α.  相似文献   

11.
The interaction of DNA polymerase purified from eggs of the teleost fish Misgurnus fossilis (loach) with DNA duplexes with single-strand gaps of 1-13 nucleotides was studied. In the absence of template-restricting DNA, the enzyme elongated primers on single-stranded DNA templates in a distributive manner. However, in the presence of the proximal 5"-terminus restricting the template, the enzyme activity significantly increased. In this case, the enzyme was capable of processive synthesis by filling gaps of 5-9 nucleotides in DNA duplexes. These data indicate that DNA polymerase can interact with both the 3"- and 5"-termini located upstream and downstream from the gap. Analysis of the complexes formed by DNA polymerase and different DNA substrates by electrophoretic mobility shift assay confirmed the assumption that this enzyme can interact with the proximal 5"-terminus restricting the gap. DNA polymerase displayed much higher affinity in duplexes with gaps of approximately 10 nucleotides compared to the standard template–primer complexes. Maximal affinity was observed in experiments with DNA substrates containing unpaired 3"-tails in primers. The results of this study suggest that DNA polymerase exerts high activity in the cell nuclei during repair of DNA intermediates with single strand gaps and unpaired 3"-termini.  相似文献   

12.
A key set of reactions for the initiation of new DNA strands during herpes simplex virus-1 replication consists of the primase-catalyzed synthesis of short RNA primers followed by polymerase-catalyzed DNA synthesis (i.e. primase-coupled polymerase activity). Herpes primase (UL5-UL52-UL8) synthesizes products from 2 to ∼13 nucleotides long. However, the herpes polymerase (UL30 or UL30-UL42) only elongates those at least 8 nucleotides long. Surprisingly, coupled activity was remarkably inefficient, even considering only those primers at least 8 nucleotides long, and herpes polymerase typically elongated <2% of the primase-synthesized primers. Of those primers elongated, only 4–26% of the primers were passed directly from the primase to the polymerase (UL30-UL42) without dissociating into solution. Comparing RNA primer-templates and DNA primer-templates of identical sequence showed that herpes polymerase greatly preferred to elongate the DNA primer by 650–26,000-fold, thus accounting for the extremely low efficiency with which herpes polymerase elongated primase-synthesized primers. Curiously, one of the DNA polymerases of the host cell, polymerase α (p70-p180 or p49-p58-p70-p180 complex), extended herpes primase-synthesized RNA primers much more efficiently than the viral polymerase, raising the possibility that the viral polymerase may not be the only one involved in herpes DNA replication.Herpes simplex virus 1 (HSV-1)2 encodes seven proteins essential for replicating its double-stranded DNA genome; five of these encode the heterotrimeric helicase-primase (UL5-UL52-UL8 gene products) and the heterodimeric polymerase (UL30-UL42 gene products) (1, 2). The helicase-primase unwinds the DNA at the replication fork and generates single-stranded DNA for both leading and lagging strand synthesis. Primase synthesizes short RNA primers on the lagging strand that the polymerase presumably elongates using dNTPs (i.e. primase-coupled polymerase activity). These two protein complexes are thought to replicate the viral genome on both the leading and lagging strands (1, 2).Previous studies have focused on the helicase-primase and polymerase separately. The helicase-primase contains three subunits, UL5, UL52, and UL8 in a 1:1:1 ratio (35). The UL5 subunit has helicase-like motifs and the UL52 subunit has primase-like motifs, yet the minimal active complex that demonstrates either helicase or primase activities contains both UL5 and UL52 (6, 7). Although the UL8 subunit has no known catalytic activity, several functions have been proposed, including enhancing helicase and primase activities, enhancing primer synthesis on ICP8 (the HSV-1 single-stranded binding protein)-coated DNA strands, and facilitating formation of the replisome (812). Although primase will synthesize short (23 nucleotides long) primers on a variety of template sequences, synthesis of longer primers up to 13 nucleotides long requires the template sequence, 3′-deoxyguanidine-pyrimidine-pyrimidine-5′ (13). Primase initiates synthesis at the first pyrimidine via the polymerization of two purine NTPs (13). Even after initiation at this sequence, however, the vast majority of products are only 2–3 nucleotides long (13, 14).The herpes polymerase consists of the UL30 subunit, which has polymerase and 3′ → 5′ exonuclease activities (1, 2), and the UL42 subunit, which serves as a processivity factor (1517). Unlike most processivity factors that encircle the DNA, the UL42 protein binds double-stranded DNA and thus directly tethers the polymerase to the DNA (18). Using pre-existing DNA primer-templates as the substrate, the heterodimeric polymerase (UL30-UL42) incorporates dNTPs at a rate of 150 s–1, a rate much faster than primer synthesis (for primers >7 nucleotides long, 0.0002–0.01 s–1) (19, 20).We examined primase-coupled polymerase activity by the herpes primase and polymerase complexes. Although herpes primase synthesizes RNA primers 2–13 nucleotides long, the polymerase only effectively elongates those at least 8 nucleotides long. Surprisingly, the polymerase elongated only a small fraction of the primase-synthesized primers (<1–2%), likely because of the polymerase elongating RNA primer-templates much less efficiently than DNA primer-templates. In contrast, human DNA polymerase α (pol α) elongated the herpes primase-synthesized primers very efficiently. The biological significance of these data is discussed.  相似文献   

13.
Mcm10 is an essential eukaryotic protein required for the initiation and elongation phases of chromosomal replication. Specifically, Mcm10 is required for the association of several replication proteins, including DNA polymerase α (pol α), with chromatin. We showed previously that the internal (ID) and C-terminal (CTD) domains of Mcm10 physically interact with both single-stranded (ss) DNA and the catalytic p180 subunit of pol α. However, the mechanism by which Mcm10 interacts with pol α on and off DNA is unclear. As a first step toward understanding the structural details for these critical intermolecular interactions, x-ray crystallography and NMR spectroscopy were used to map the binary interfaces between Mcm10-ID, ssDNA, and p180. The crystal structure of an Mcm10-ID·ssDNA complex confirmed and extended our previous evidence that ssDNA binds within the oligonucleotide/oligosaccharide binding-fold cleft of Mcm10-ID. We show using NMR chemical shift perturbation and fluorescence spectroscopy that p180 also binds to the OB-fold and that ssDNA and p180 compete for binding to this motif. In addition, we map a minimal Mcm10 binding site on p180 to a small region within the p180 N-terminal domain (residues 286–310). These findings, together with data for DNA and p180 binding to an Mcm10 construct that contains both the ID and CTD, provide the first mechanistic insight into how Mcm10 might use a handoff mechanism to load and stabilize pol α within the replication fork.To maintain their genomic integrity, cells must ensure complete and accurate DNA replication once per cell cycle. Consequently, DNA replication is a highly regulated and orchestrated series of molecular events. Multiprotein complexes assembled at origins of replication lead to assembly of additional proteins that unwind chromosomal DNA and synthesize nascent strands. The first event is the formation of a pre-replicative complex, which is composed of the origin recognition complex, Cdc6, Cdt1, and Mcm2–7 (for review, see Ref. 1). Initiation of replication at the onset of S-phase involves the activity of cyclin- and Dbf4-dependent kinases concurrent with recruitment of key factors to the origin. Among these, Mcm10 (2, 3) is recruited in early S-phase and is required for loading of Cdc45 (4). Mcm2–7, Cdc45, and the GINS complex form the replicative helicase (58). Origin unwinding is followed by loading of RPA,3 And-1/Ctf4, and pol α onto ssDNA (912). In addition, recruitment of Sld2, Sld3, and Dpb11/TopBP1 are essential for replication initiation (13, 14), and association of topoisomerase I, proliferating cellular nuclear antigen (PCNA), replication factor C, and the replicative DNA polymerases δ and ϵ completes the replisome (for review, see Ref. 15).Mcm10 is exclusive to eukaryotes and is essential to both initiation and elongation phases of chromosomal DNA replication (6, 8, 16). Mutations in Mcm10 in yeast result in stalled replication, cell cycle arrest, and cell death (2, 3, 1719). These defects can be explained by the number of genetic and physical interactions between Mcm10 and many essential replication proteins, including origin recognition complex, Mcm2–7, and PCNA (3, 12, 2024). In addition, Mcm10 has been shown to stimulate the phosphorylation of Mcm2–7 by Dbf4-dependent kinase in vitro (25). Thus, Mcm10 is an integral component of the replication machinery.Importantly, Mcm10 physically interacts with and stabilizes pol α and helps to maintain its association with chromatin (16, 26, 27). This is a critical interaction during replication because pol α is the only enzyme in eukaryotic cells that is capable of initiating DNA synthesis de novo. Indeed, Mcm10 stimulates the polymerase activity of pol α in vitro (28), and interestingly, the fission yeast Mcm10, but not Xenopus Mcm10, has been shown to exhibit primase activity (29, 30). Mcm10 is composed of three domains, the N-terminal (NTD), internal (ID), and C-terminal (CTD) domains (29). The NTD is presumably an oligomerization domain, whereas the ID and CTD both interact with DNA and pol α (29). The CTD is not found in yeast, whereas the ID is highly conserved among all eukaryotes. The crystal structure of Mcm10-ID showed that this domain is composed of an oligonucleotide/oligosaccharide binding (OB)-fold and a zinc finger motif, which form a unified DNA binding platform (31). An Hsp10-like motif important for the interaction with pol α has been identified in the sequence of Saccharomyces cerevisiae Mcm10-ID (16, 26).DNA pol α-primase is composed of four subunits: p180, p68, p58, and p48. The p180 subunit possesses the catalytic DNA polymerase activity, and disruption of this gene is lethal (32, 33). p58 and p48 form the DNA-dependent RNA polymerase (primase) activity (34, 35), whereas the p68 subunit has no known catalytic activity but serves a regulatory role (36, 37). Pol α plays an essential role in lagging strand synthesis by first creating short (7–12 nucleotide) RNA primers followed by DNA extension. At the critical length of ∼30 nucleotides, replication factor C binds to the nascent strand to displace pol α and loads PCNA with pols δ and ϵ (for review, see Ref. 38).The interaction between Mcm10 and pol α has led to the suggestion that Mcm10 may help recruit the polymerase to the emerging replisome. However, the molecular details of this interaction and the mechanism by which Mcm10 may recruit and stabilize the pol α complex on DNA has not been investigated. Presented here is the high resolution structure of the conserved Mcm10-ID bound to ssDNA together with NMR chemical shift perturbation competition data for pol α binding in the presence of ssDNA. Collectively, these data demonstrate a shared binding site for DNA and pol α in the OB-fold cleft of Mcm10-ID, with a preference for ssDNA over pol α. In addition, we have mapped the Mcm10-ID binding site on pol α to a 24-residue segment of the N-terminal domain of p180. Based on these results, we propose Mcm10 helps to recruit pol α to origins of replication by a molecular hand-off mechanism.  相似文献   

14.
DNA polymerases II (ε) and III(δ) are the only nuclear DNA polymerases known to possess an intrinsic 3′ → 5′ exonuclease in Saccharomyces cerevisiae. We have investigated the spontaneous mutator phenotypes of DNA polymerase δ and ε 3′ → 5′ exonuclease-deficient mutants, pol3-01 and pol2-4, respectively. pol3-01 and pol2-4 increased spontaneous mutation rates by factors of the order of 102 and 101, respectively, measured as URA3 forward mutation and his7-2 reversion. Surprisingly, a double mutant pol2-4 pol3-01 haploid was inviable. This was probably due to accumulation of unedited errors, since a pol2-4/pol2-4 pol3-01/pol3-01 diploid was viable, with the spontaneous his7-2 reversion rate increased by about 2 × 103-fold. Analysis of mutation rates of double mutants indicated that the 3′ → 5′ exonucleases of DNA polymerases δ and ε can act competitively and that, like the 3′ → 5′ exonuclease of DNA polymerase δ the 3′ → 5′ exonuclease of DNA polymerase ε acts in series with the PMS1 mismatch correction system. Mutational spectra at a URA3 gene placed in both orientations near to a defined replication origin provided evidence that the 3′ → 5′ exonucleases of DNA polymerases δ and ε act on opposite DNA strands, but were in sufficient to distinguish conclusively between different models of DNA replication.  相似文献   

15.
Abasic (AP) sites are very frequent and dangerous DNA lesions. Their ability to block the advancement of a replication fork has been always viewed as a consequence of their inhibitory effect on the DNA synthetic activity of replicative DNA polymerases (DNA pols). Here we show that AP sites can also affect the strand displacement activity of the lagging strand DNA pol δ, thus preventing proper Okazaki fragment maturation. This block can be overcome through a polymerase switch, involving the combined physical and functional interaction of DNA pol β and Flap endonuclease 1. Our data identify a previously unnoticed deleterious effect of the AP site lesion on normal cell metabolism and suggest the existence of a novel repair pathway that might be important in preventing replication fork stalling.Loss of purine and pyrimidine bases is a significant source of DNA damage in prokaryotic and eukaryotic organisms. Abasic (apurinic and apyrimidinic) lesions occur spontaneously in DNA; in eukaryotes it has been estimated that about 104 depurination and 102 depyrimidation events occur per genome per day. An equally important source of abasic DNA lesions results from the action of DNA glycosylases, such as uracil glycosylase, which excises uracil arising primarily from spontaneous deamination of cytosines (1). Although most AP sites are removed by the base excision repair (BER)5 pathway, a small fraction of lesions persists, and DNA with AP lesions presents a strong block to DNA synthesis by replicative DNA polymerases (DNA pols) (2, 3). Several studies have been performed to address the effects of AP sites on the template DNA strand on the synthetic activity of a variety of DNA pols. The major replicative enzyme of eukaryotic cells, DNA pol δ, was shown to be able to bypass an AP lesion, but only in the presence of the auxiliary factor proliferating cell nuclear antigen (PCNA) and at a very reduced catalytic efficiency if compared with an undamaged DNA template (4). On the other hand, the family X DNA pols β and λ were shown to bypass an AP site but in a very mutagenic way (5). Recent genetic evidence in Saccharomyces cerevisiae cells showed that DNA pol δ is the enzyme replicating the lagging strand (6). According to the current model for Okazaki fragment synthesis (79), the action of DNA pol δ is not only critical for the extension of the newly synthesized Okazaki fragment but also for the displacement of an RNA/DNA segment of about 30 nucleotides on the pre-existing downstream Okazaki fragment to create an intermediate Flap structure that is the target for the subsequent action of the Dna2 endonuclease and the Flap endonuclease 1 (Fen-1). This process has the advantage of removing the entire RNA/DNA hybrid fragment synthesized by the DNA pol α/primase, potentially containing nucleotide misincorporations caused by the lack of a proofreading exonuclease activity of DNA pol α/primase. This results in a more accurate copy synthesized by DNA pol δ. The intrinsic strand displacement activity of DNA pol δ, in conjunction with Fen-1, PCNA, and replication protein A (RP-A), has been also proposed to be essential for the S phase-specific long patch BER pathway (10, 11). Although it is clear that an AP site on the template strand is a strong block for DNA pol δ-dependent synthesis on single-stranded DNA, the functional consequences of such a lesion on the ability of DNA pol δ to carry on strand displacement synthesis have never been investigated so far. Given the high frequency of spontaneous hydrolysis and/or cytidine deamination events, any detrimental effect of an AP site on the strand displacement activity of DNA pol δ might have important consequences both for lagging strand DNA synthesis and for long patch BER. In this work, we addressed this issue by constructing a series of synthetic gapped DNA templates with a single AP site at different positions with respect to the downstream primer to be displaced by DNA pol δ (see Fig. 1A). We show that an AP site immediately upstream of a single- to double-strand DNA junction constitutes a strong block to the strand displacement activity of DNA pol δ, even in the presence of RP-A and PCNA. Such a block could be resolved only through a “polymerase switch” involving the concerted physical and functional interaction of DNA pol β and Fen-1. The closely related DNA pol λ could only partially substitute for DNA pol β. Based on our data, we propose that stalling of a replication fork by an AP site not only is a consequence of its ability to inhibit nucleotide incorporation by the replicative DNA pols but can also stem from its effects on strand displacement during Okazaki fragment maturation. In summary, our data suggest the existence of a novel repair pathway that might be important in preventing replication fork stalling and identify a previously unnoticed deleterious effect of the AP site lesion on normal cell metabolism.Open in a separate windowFIGURE 1.An abasic site immediately upstream of a double-stranded DNA region inhibits the strand displacement activity of DNA polymerase δ. The reactions were performed as described under “Experimental Procedures.” A, schematic representation of the various DNA templates used. The size of the resulting gaps is indicated in nt. The position of the AP site on the 100-mer template strand is indicated relative to the 3′ end. Base pairs in the vicinity of the lesion are indicated by dashes. The size of the gaps (35–38 nt) is consistent with the size of ssDNA covered by a single RP-A molecule, which has to be released during Okazaki fragment synthesis when the DNA pol is approaching the 5′-end of the downstream fragment. When the AP site is covered by the downstream terminator oligonucleotide (Gap-3 and Gap-1 templates) the nucleotide placed on the opposite strand is C to mimic the situation generated by spontaneous loss of a guanine or excision of an oxidized guanine, whereas when the AP site is covered by the primer (nicked AP template), the nucleotide placed on the opposite strand is A to mimic the most frequent incorporation event occurring opposite an AP site. B, human PCNA was titrated in the presence of 15 nm (lanes 2–4 and 10–12) or 30 nm (lanes 6–8 and 14–16) recombinant human four subunit DNA pol δ, on a linear control (lanes 1–8) or a 38-nt gap control (lanes 9–16) template. Lanes 1, 5, 9, and 13, control reactions in the absence of PCNA. C, human PCNA was titrated in the presence of 60 nm DNA pol δ, on a linear AP (lanes 2–4) or 38-nt gap AP (lanes 6–9) template. Lanes 1 and 5, control reactions in the absence of PCNA.  相似文献   

16.
The nuclear-encoded DNA polymerase γ (DNA POLγ) is the sole DNA polymerase required for the replication of the mitochondrial DNA. We have cloned the cDNA for human DNA POLγ and have mapped the gene to the chromosomal location 15q24. Additionally, the DNA POLγ gene fromDrosophila melanogasterand a partial cDNA for DNA POLγ fromGallus gallushave been cloned. The predicted human DNA POLγ polypeptide is 1239 amino acids, with a calculated molecular mass of 139.5 kDa. The human amino acid sequence is 41.6, 43.0, 48.7, and 77.6% identical to those ofSchizosaccharomyces pombe, Saccharomyces cerevisiae, Drosophila melanogaster,and the C-terminal half ofG. gallus,respectively. Polyclonal antibodies raised against the polymerase portion of the protein reacted specifically with a 140-kDa protein in mitochondrial extracts and immunoprecipitated a protein with DNA POLγ like activity from mitochondrial extracts. The human DNA POLγ is unique in that the first exon of the gene contains a CAG10trinucleotide repeat.  相似文献   

17.
Fidelity of DNA polymerases is vital for maintaining genomic integrity. Deficient DNA repair leads to age related disorders or cancer. If the age at which the decline in activity of predominant DNA repair enzymes starts is identified, and the deficient proteins supplemented, then the manifestation of these diseases can be delayed promoting healthy aging. DNA polymerase β (pol β) is a predominant repair enzyme in brain. DNA pol β activity declines with age in rat brain/neurons but the exact age during the life time of rat when this decline begins is not known, and comparison of this activity was not made between post mitotic and proliferating tissues therefore the pattern of pol β with age was studied in rat brain and tissues. The decline in pol β activity started between 30 and 45 days postnatal in all the tissues. Post mitotic tissues showed pronounced decline than the proliferating tissues.  相似文献   

18.
19.
20.
Homologous recombination is an error-free mechanism for the repair of DNA double-strand breaks (DSBs). Most DSB repair events occur by gene conversion limiting loss of heterozygosity (LOH) for markers downstream of the site of repair and restricting deleterious chromosome rearrangements. DSBs with only one end available for repair undergo strand invasion into a homologous duplex DNA, followed by replication to the chromosome end (break-induced replication [BIR]), leading to LOH for all markers downstream of the site of strand invasion. Using a transformation-based assay system, we show that most of the apparent BIR events that arise in diploid Saccharomyces cerevisiae rad51Δ mutants are due to half crossovers instead of BIR. These events lead to extensive LOH because one arm of chromosome III is deleted. This outcome is also observed in pol32Δ and pol3-ct mutants, defective for components of the DNA polymerase δ (Pol δ) complex. The half crossovers formed in Pol δ complex mutants show evidence of limited homology-dependent DNA synthesis and are partially Mus81 dependent, suggesting that strand invasion occurs and the stalled intermediate is subsequently cleaved. In contrast to rad51Δ mutants, the Pol δ complex mutants are proficient for repair of a 238-bp gap by gene conversion. Thus, the BIR defect observed for rad51 mutants is due to strand invasion failure, whereas the Pol δ complex mutants are proficient for strand invasion but unable to complete extensive tracts of recombination-initiated DNA synthesis.DNA double-strand breaks (DSBs) are potentially lethal lesions that can occur spontaneously during normal cell metabolism, by treatment of cells with DNA-damaging agents, or during programmed recombination processes (54). There are two major pathways to repair DSBs: nonhomologous end joining (NHEJ) and homologous recombination (HR). NHEJ involves the religation of the two ends of the broken chromosome and can occur with high fidelity or be accompanied by a gain or loss of nucleotides at the junction (9). Repair of two-ended DSBs by HR generally occurs by gene conversion resulting from a transfer of information from the intact donor duplex to the broken chromosome (Fig. (Fig.1).1). HR occurs preferentially during S and G2 when a sister chromatid is available to template repair (2, 19, 22). Sister-chromatid recombination events are genetically silent, whereas gene conversion between nonsister chromatids associated with an exchange of flanking markers can result in extensive loss of heterozygosity (LOH) or chromosome rearrangements (3, 21). One-ended DSBs that arise by replication fork collapse or by erosion of uncapped telomeres are thought to repair by strand invasion into homologous duplex DNA followed by replication to the end of the chromosome, a process referred to as break-induced replication (BIR) (35). BIR appears to be suppressed at two-ended breaks, presumably because it can lead to extensive LOH if it occurs between homologues or to chromosome translocations when strand invasion initiates within dispersed repeated sequences (5, 28, 31, 50, 52, 55).Open in a separate windowFIG. 1.Models for gene conversion and BIR. After formation of a DSB, the ends are resected to generate 3′ single-strand DNA tails. One end undergoes Rad51-dependent strand invasion to prime DNA synthesis from the invading 3′ end templated by the donor duplex. For gene conversion by the synthesis-dependent strand annealing model, the extended invading end is displaced and can anneal to the other side of the break; completion of repair requires DNA synthesis primed from the noninvading 3′ end. For a one-ended break, or if the other side of the break lacks homology to the donor duplex, DNA synthesis proceeds to the end of the chromosome. Centromeres are shown as solid ovals and a heterozygous marker centromere distal to the site of repair as A/a.The strand invasion step of BIR is assumed to be the same as that for gene conversion based on the requirement for the same HR proteins: Rad51, Rad52, Rad54, Rad55, and Rad57 (10). However, subsequent steps in BIR are less well defined. Recent studies of the fate of the invading end during BIR in diploid strains with polymorphic chromosome III homologues using a plasmid-based assay have shown that following strand invasion, the invading end is capable of dissociating from the initial homologous template. Following dissociation, the displaced end subsequently reinvades into the same or a different chromosome III homologue by a process termed template switching (52). One of the interesting features of the template switching events is that they occur over a region of about 10 kb downstream of the site of strand invasion and do not extend over the entire left arm of chromosome III. There are a number of possible mechanisms that could account for this apparent change in the processivity of BIR. First, it is possible that the strand invasion intermediate is cleaved by a structure-specific nuclease and once the invading strand is covalently joined to one of the template strands, the strand invasion process is irreversible. Recent studies of Schizosaccharomyces pombe have shown an essential role for Mus81, a structure-specific nuclease, in resolution of sister chromatid recombination intermediates during repair of collapsed replication forks (48). Another possibility is that there could be a switch between a translesion DNA polymerase and a highly processive DNA polymerase during BIR. The translesion polymerases in budding yeast, polymerase ζ (Pol ζ) and Pol η, are encoded by REV3-REV7 and RAD30, respectively (34, 40, 43). Deletion of REV3 has been shown to increase the fidelity of DNA synthesis associated with HR but has no effect on the overall frequency of DSB-induced HR (16). Deletion of POLη in chicken DT40 cells reduces the frequency of DSB-induced gene conversion, and human POL η has been shown to extend the invading 3′ end of D-loop intermediates in vitro (23, 36). However, this same preference for Pol η is not found for Saccharomyces cerevisiae. Instead, DNA synthesis during meiotic and mitotic recombination appears to be carried out by Pol δ, one of the three nuclear replicative polymerases, which normally functions with Pol α in Okazaki fragment synthesis (13, 32, 33, 44). Pol ɛ is thought to be the primary leading-strand polymerase (47), but in the absence of the Pol ɛ catalytic domain, Pol δ is presumed to carry out leading-strand synthesis (24). Recent studies by Lydeard et al. (30) have shown a requirement for the lagging-strand polymerases, Pol δ and Pol α, to form the initial primer extension product during BIR, and Pol ɛ is required to complete replication to the end of the chromosome. In contrast, repair of DSBs by gene conversion does not require Pol α, and there appears to be functional redundancy between Pol δ and Pol ɛ (56).To address the roles of Mus81, Pol δ, and Pol η in BIR and in particular template switching, we used the transformation-based BIR assay with diploids with polymorphic chromosome III homologues. Because the transformation assay can only be used with strains with viable mutations of replication factors, we used a null allele of POL32, encoding a nonessential subunit of the Pol δ complex (14), and a point mutation in the gene encoding the essential catalytic subunit, POL3. The pol3-ct allele results in a truncation removing the last four amino acids of the Pol3 protein; the C-terminal region of Pol3 is implicated in interaction with the other essential subunit of the Pol δ complex, Pol31 (15, 49). The interesting feature of the pol3-ct allele is that it decreases the length of gene conversion tracts during mitotic and meiotic recombination, presumably by affecting the processivity of Pol δ, but confers no apparent defect in normal DNA synthesis (32, 33). Because BIR requires more-extensive tracts of DNA synthesis than gene conversion, we expected the pol3-ct mutant to exhibit a BIR defect. We found that in the absence of a fully functional Pol δ complex, chromosome fragment (CF) formation proceeds by a half-crossover mechanism associated with loss of the template chromosome, an event with potentially catastrophic consequences (6, 57). This was also found to occur in rad51 mutants, suggesting nonreciprocal translocations arise by failure to undergo strand invasion or because replication following strand invasion is inefficient. In contrast to rad51 mutants, the Pol δ complex mutants are proficient for repair of a 238-bp gap by gene conversion and fully resistant to ionizing radiation, suggesting there is a unique requirement for Pol δ to complete BIR. Consistent with studies of gene conversion in S. cerevisiae (33), we found no role for Pol η in BIR or the process of template switching.  相似文献   

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