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 312 毫秒
1.
Raychaudhury P  Basu AK 《Biochemistry》2011,50(12):2330-2338
γ-Radiation generates a variety of complex lesions in DNA, including the G[8,5-Me]T intrastrand cross-link in which C8 of guanine is covalently linked to the 5-methyl group of the 3'-thymine. We have investigated the toxicity and mutagenesis of this lesion by replicating a G[8,5-Me]T-modified plasmid in Escherichia coli with specific DNA polymerase knockouts. Viability was very low in a strain lacking pol II, pol IV, and pol V, the three SOS-inducible DNA polymerases, indicating that translesion synthesis is conducted primarily by these DNA polymerases. In the single-polymerase knockout strains, viability was the lowest in a pol V-deficient strain, which suggests that pol V is most efficient in bypassing this lesion. Most mutations were single-base substitutions or deletions, though a small population of mutants carrying two point mutations at or near the G[8,5-Me]T cross-link was also detected. Mutations in the progeny occurred at the cross-linked bases as well as at bases near the lesion site, but the mutational spectrum varied on the basis of the identity of the DNA polymerase that was knocked out. Mutation frequency was the lowest in a strain that lacked the three SOS DNA polymerases. We determined that pol V is required for most targeted G → T transversions, whereas pol IV is required for the targeted T deletions. Our results suggest that pol V and pol IV compete to carry out error-prone bypass of the G[8,5-Me]T cross-link.  相似文献   

2.
A hallmark of the Escherichia coli SOS response is the large increase in mutations caused by translesion synthesis (TLS). TLS requires DNA polymerase V (UmuD'2C) and RecA. Here, we show that pol V and RecA interact by two distinct mechanisms. First, pol V binds to RecA in the absence of DNA and ATP and second, through its UmuD' subunit, requiring DNA and ATP without ATP hydrolysis. TLS occurs in the absence of a RecA nucleoprotein filament but is inhibited in its presence. Therefore, a RecA nucleoprotein filament is unlikely to be required for SOS mutagenesis. Pol V activity is severely diminished in the absence of RecA or in the presence of RecA1730, a mutant defective for pol V mutagenesis in vivo. Pol V activity is strongly enhanced with RecA mutants constitutive for mutagenesis in vivo, suggesting that RecA is an obligate accessory factor that activates pol V for SOS mutagenesis.  相似文献   

3.
Escherichia coli plasmids containing the rpsL+ gene (Strs phenotype) as the target for mutation were treated in vitro with N-methyl-N-nitrosourea. Following fixation of mutations in E. coli MM294A cells (recA+ Strs), an unselected population of mutant and wild-type plasmids was isolated and transferred into a second host, E. coli 6451 (recA Strr). Strains carrying plasmid-encoded forward mutations were then selected as Strr isolates, while rpsL+ plasmids conferred the dominant Strs phenotype in the second host. Mutation induction and reduced survival of N-methyl-N-nitrosourea-treated plasmids were shown to be dose dependent. Because this system permitted analysis and manipulation of the levels of certain methylated bases produced in vitro by N-methyl-N-nitrosourea, it afforded the opportunity to assess directly the relative roles of these bases and of SOS functions in mutagenesis. The methylated plasmid DNA gave a mutation frequency of 6 X 10(-5) (a 40-fold increase over background) in physiologically normal cells. When the same methylated plasmid was repaired in vitro by using purified O6-methylguanine DNA methyltransferase (to correct O6-methylguanine and O4-methylthymine), no mutations were detected above background levels. In contrast, when the methylated plasmid DNA was introduced into host cells induced by UV light for the SOS functions, rpsL mutagenesis was enhanced eightfold over the level seen without SOS induction. This enhancement of mutagenesis by SOS was unaffected by prior treatment of the DNA with O6-methylguanine DNA methyltransferase. These results demonstrate a predominant mutagenic role for alkylation lesions other than O6-methylguanine or O4-methylthymine when SOS functions are induced. The mutation spectrum of N-methyl-N-nitrosourea under conditions of induced SOS functions revealed a majority of mutagenic events at A . T base pairs.  相似文献   

4.
Nucleotide incorporation fidelity, mismatch extension, and translesion DNA synthesis efficiencies were determined using SOS-induced Escherichia coli DNA polymerases (pol) II, IV, and V to copy 10R and 10S isomers of trans-opened benzo[a]pyrene-7,8-diol 9,10-epoxide (BaP DE) A and G adducts. A-BaP DE adducts were bypassed by pol V with moderate accuracy and considerably higher efficiency than by pol II or IV. Error-prone pol V copied G-BaP DE-adducted DNA poorly, forming A*G-BaP DE-S and -R mismatches over C*G-BaP DE-S and -R correct matches by factors of approximately 350- and 130-fold, respectively, even favoring G*G-BaP DE mismatches over correct matches by factors of 2-4-fold. In contrast, pol IV bypassed G-BaP DE adducts with the highest efficiency and fidelity, making misincorporations with a frequency of 10(-2) to 10(-4) depending on sequence context. G-BaP DE-S-adducted M13 DNA yielded 4-fold fewer plaques when transfected into SOS-induced DeltadinB (pol IV-deficient) mutant cells compared with the isogenic wild-type E. coli strain, consistent with the in vitro data showing that pol IV was most effective by far at copying the G-BaP DE-S adduct. SOS polymerases are adept at copying a variety of lesions, but the relative contribution of each SOS polymerase to copying damaged DNA appears to be determined by the lesion's identity.  相似文献   

5.
Cockayne syndrome (CS) is a human genetic disorder characterized by sensitivity to UV radiation, neurodegeneration, premature aging among other phenotypes. CS complementation group B (CS-B) gene (csb) encodes the CSB protein (CSB) that is involved in base excision repair of a number of oxidatively induced lesions in genomic DNA in vivo. We hypothesized that CSB may also play a role in cellular repair of the DNA helix-distorting tandem lesion (5'S)-8,5'-cyclo-2'-deoxyadenosine (S-cdA). Among many DNA lesions, S-cdA is unique in that it represents a concomitant damage to both the sugar and base moieties of the same nucleoside. Because of the presence of the C8-C5' covalent bond, S-cdA is repaired by nucleotide excision repair unlike most of other oxidatively induced lesions in DNA, which are subject to base excision repair. To test our hypothesis, we isolated genomic DNA from brain, kidney and liver of wild type and csb knockout (csb(-/-)) mice. Animals were not exposed to any exogenous oxidative stress before the experiment. DNA samples were analysed by liquid chromatography/mass spectrometry with isotope-dilution. Statistically greater background levels of S-cdA were observed in all three organs of csb(-/-) mice than in those of wild type mice. These results suggest the in vivo accumulation of S-cdA in genomic DNA due to lack of its repair in csb(-/-) mice. Thus, this study provides, for the first time, the evidence that CSB plays a role in the repair of the DNA helix-distorting tandem lesion S-cdA. Accumulation of unrepaired S-cdA in vivo may contribute to the pathology associated with CS.  相似文献   

6.
Reactive oxygen and nitrogen radicals produced during metabolic processes, such as respiration and inflammation, combine with DNA to form many lesions primarily at guanine sites. Understanding the roles of the polymerases responsible for the processing of these products to mutations could illuminate molecular mechanisms that correlate oxidative stress with cancer. Using M13 viral genomes engineered to contain single DNA lesions and Escherichia coli strains with specific polymerase (pol) knockouts, we show that pol V is required for efficient bypass of structurally diverse, highly mutagenic guanine oxidation products in vivo. We also find that pol IV participates in the bypass of two spiroiminodihydantoin lesions. Furthermore, we report that one lesion, 5-guanidino-4-nitroimidazole, is a substrate for multiple SOS polymerases, whereby pol II is necessary for error-free replication and pol V for error-prone replication past this lesion. The results spotlight a major role for pol V and minor roles for pol II and pol IV in the mechanism of guanine oxidation mutagenesis.  相似文献   

7.
Escherichia coli possesses five known DNA polymerases (pols). Pol III holoenzyme is the cell's main replicase, while pol I is responsible for the maturation of Okazaki fragments and filling gaps generated during nucleotide excision repair. Pols II, IV and V are significantly upregulated as part of the cell's global SOS response to DNA damage and under these conditions, may alter the fidelity of DNA replication by potentially interfering with the ability of pols I and III to complete their cellular functions. To test this hypothesis, we determined the spectrum of rpoB mutations arising in an isogenic set of mutL strains differentially expressing the chromosomally encoded pols. Interestingly, mutagenic hot spots in rpoB were identified that are susceptible to the actions of pols I–V. For example, in a recA730 lexA (Def) mutL background most transversions were dependent upon pols IV and V. In contrast, transitions were largely dependent upon pol I and to a lesser extent, pol III. Furthermore, the extent of pol I-dependent mutagenesis at one particular site was modulated by pols II and IV. Our observations suggest that there is considerable interplay among all five E. coli polymerases that either reduces or enhances the mutagenic load on the E. coli chromosome.  相似文献   

8.
We hypothesized that DNA damage products (5'R)-8,5'-cyclo-2'-deoxyadenosine (R-cdA) and (5'S)-8,5'-cyclo-2'-deoxyadenosine (S-cdA) may be well-suited biomarkers of risk and diagnosis for atherosclerosis. We tested this hypothesis by measuring the levels of R-cdA and S-cdA and another product, 8-hydroxy-2'-deoxyguanosine (8-OH-dG), in urine of atherosclerosis patients and healthy individuals using liquid chromatography-tandem mass spectrometry with isotope dilution. We showed the presence of these products at significantly greater concentrations in urine of atherosclerosis patients than in that of healthy individuals. Our data suggest that R-cdA and S-cdA can be accurately and reproducibly measured in human urine as potential biomarkers of risk and diagnosis for atherosclerosis.  相似文献   

9.
Reactive oxygen species induce oxidative damage in DNA precursors, i.e. dNTPs, leading to point mutations upon incorporation. Escherichia coli mutT strains, deficient in the activity hydrolysing 8‐oxo‐7,8‐dihydro‐2′‐deoxyguanosine 5′‐triphosphate (8‐oxo‐dGTP), display more than a 100‐fold higher spontaneous mutation frequency over the wild‐type strain. 8‐oxo‐dGTP induces A to C transversions when misincorporated opposite template A. Here, we report that DNA pol III incorporates 8‐oxo‐dGTP ≈ 20 times more efficiently opposite template A compared with template C. Single, double or triple deletions of pol I, pol II, pol IV or pol V had modest effects on the mutT mutator phenotype. Only the deletion of all four polymerases led to a 70% reduction of the mutator phenotype. While pol III may account for nearly all 8‐oxo‐dGTP incorporation opposite template A, it only extends ≈ 30% of them, the remaining 70% being extended by the combined action of pol I, pol II, pol IV or pol V. The unique property of pol III, a C‐family DNA polymerase present only in eubacteria, to preferentially incorporate 8‐oxo‐dGTP opposite template A during replication might explain the high spontaneous mutation frequency in E. coli mutT compared with the mammalian counterparts lacking the 8‐oxo‐dGTP hydrolysing activities.  相似文献   

10.
E. coli strains bearing the recA441 mutation and various mutations in the polA gene resulting in enzymatically well-defined deficiencies of DNA polymerase I have been constructed. It was found that the recA441 strains bearing either the polA1 or polA12 mutation causing deficiency of the polymerase activity of pol I are unable to grow at 42 degrees C on minimal medium supplemented with adenine, i.e., when the SOS response is continuously induced in strains bearing the recA441 mutation. Under these conditions the inhibition of DNA synthesis is followed in recA441 polA12 by DNA degradation and loss of cell viability. A similar lethal effect is observed with the recA730 polA12 mutant. The recA441 strain bearing the polA107 mutation resulting in the deficiency of the 5'-3' exonuclease activity of pol I shows normal growth under conditions of continuous SOS response. We postulate that constitutive expression of the SOS response leads to an altered requirement for the polymerase activity of pol I.  相似文献   

11.
The mutagenic potentials of DNAs containing site- and stereospecific intrastrand DNA crosslinks were evaluated in Escherichia coli cells that contained a full complement of DNA polymerases or were deficient in either polymerases II, IV, or V. Crosslinks were made between adjacent N(6)-N(6) adenines and consisted of R,R- and S,S-butadiene crosslinks and unfunctionalized 2-, 3-, and 4-carbon tethers. Although replication of single-stranded DNAs containing the unfunctionalized 3- and 4-carbon tethers were non-mutagenic in all strains tested, replication past all the other intrastrand crosslinks was mutagenic in all E. coli strains, except the one deficient in polymerase II in which no mutations were ever detected. However, when mutagenesis was analyzed in cells induced for SOS, mutations were not detected, suggesting a possible change in the overall fidelity of polymerase II under SOS conditions. These data suggest that DNA polymerase II is responsible for the in vivo mutagenic bypass of these lesions in wild-type E. coli.  相似文献   

12.
Forward mutations induced by the ultimate carcinogen N-acetoxy-N-2-acetylaminofluorene (N-Aco-AAF) in the tetracycline resistance gene carried on plasmid pBR322 are shown to be dependent upon the induction of the host SOS functions in wild-type and umuC Escherichia coli cells. The mutation frequency in the umuC strain is equal to about 40% of the mutation frequency observed in the umu+ background. In the excision-repair-deficient uvrA mutant strain the mutagenic response is the same as in SOS-induced wild-type cells whether or not the uvrA bacteria are SOS-induced. Equal mutation frequencies are obtained in both the wild-type and the uvrA strains for equal modification levels although the survival of AAF-modified plasmid DNA is greatly reduced in the uvrA strain as compared to the wild-type strain. Sequence analysis of the mutations reveals that more than 90% of the N-Aco-AAF-induced mutations are frameshift mutations. Two types of mutational hotspots are observed occurring either at repetitive sequences or at non-repetitive sequences. Both types of mutants appear at similar locations and frequencies in both the wild-type and the uvrA strains. On the other hand, only the non-repetitive sequence mutants are obtained in the umuC background. These non-repetitive sequence mutants preferentially occur within the sequence 5' G-G-C-G-C-C 3' (the NarI restriction enzyme recognition sequence). The analysis of the -AAF binding spectrum to the same DNA fragment shows that there is no direct correlation between the modification spectrum and the mutation spectrum. We suggest that certain sequences are "mutation-prone" in the sense that only these sequences can be efficiently mutated as the result of an active processing mediated by specific proteins. When a sequence is said to be mutation-prone it probably corresponds to a particular structure that is induced within this sequence as a result of the binding to the DNA of the mutagen. This sequence-specific conformational change is the substrate for the protein(s) that fixes the mutation. The mutagenic processing pathway(s) is part of the cellular response to DNA-damaging agents (the so-called SOS response). Two pathways for frameshift mutagenesis are suggested by the data: an umuC-dependent pathway, which is involved in the mutagenic processing of lesions within repetitive sequences; an umuC-independent pathway responsible for the fixation of mutations within specific non-repetitive sequences.  相似文献   

13.
Base excision repair (BER) is a very important repair mechanism to remove oxidative DNA damage. A major oxidative DNA damage after exposure to ionizing radiation is 7,8-dihydro-8-oxoguanine (8oxoG). 8oxoG is a strong mutagenic lesion, which may cause G:C to T:A transversions if not repaired correctly. Formamidopyrimidine-DNA glycosylase (Fpg), a repair enzyme which is part of BER, is the most important enzyme to repair 8oxoG. In the past years, evidence evolved that nucleotide excision repair (NER), a repair system originally thought to repair only bulky DNA lesions, can also repair some oxidative DNA damages. Examples of DNA damages which are recognized by NER are thymine glycol and abasic sites (AP sites). The main objective of this study is to determine if NER can act as a backup system for the repair of spontaneous and gamma-radiation-induced damages when Fpg is deficient. For that purpose, the effect of a NER-deficiency on the spontaneous and gamma-radiation-induced mutation spectrum in the lacZ gene was determined, using double-stranded (ds) M13 DNA, with the lacZalpha gene inserted as mutational target sequence. Subsequently the DNA was transfected into a fpg(-)uvrA(-) Escherichia coli strain (BH420) and the mutational spectra were compared with the spectra of a fpg(-) E. coli strain (BH410) and a wild type E. coli strain (JM105), which were determined in an earlier study. Furthermore, to examine effects which are caused by UvrA-deficiency, and not by Fpg-deficiency, the spontaneous and gamma-radiation-induced mutation spectra of an E. coli strain in which only UvrA is deficient (BH430) were also determined and compared with a wild type E. coli strain (JM105). The results of this study indicate that if only UvrA is deficient, there is an increase in spontaneous G:C to T:A transversions as compared to JM105 and a decrease in A:T to G:C transitions. The gamma-radiation-induced mutation spectrum of BH420 (fpg(-)uvrA(-)) shows a significant decrease in G:C to A:T and G:C to T:A mutations, as compared to BH410 where only Fpg is deficient. Based on these results, we conclude that in our experiments NER is not acting as a backup system if Fpg is deficient. Instead, NER seems to make mistakes, leading to the formation of mutations.  相似文献   

14.
In Escherichia coli, UV-irradiated cells resume DNA synthesis after a transient inhibition by a process called replication restart. To elucidate the role of several key proteins involved in this process, we have analysed the time dependence of replication restart in strains carrying a combination of mutations in lexA, recA, polB (pol II), umuDC (pol V), priA, dnaC, recF, recO or recR. We find that both pol II and the origin-independent primosome-assembling function of PriA are essential for the immediate recovery of DNA synthesis after UV irradiation. In their absence, translesion replication or 'replication readthrough' occurs approximately 50 min after UV and is pol V-dependent. In a wild-type, lexA+ background, mutations in recF, recO or recR block both pathways. Similar results were obtained with a lexA(Def) recF strain. However, lexA(Def) recO or lexA(Def) recR strains, although unable to facilitate PriA-pol II-dependent restart, were able to perform pol V-dependent readthrough. The defects in restart attributed to mutations in recF, recO or recR were suppressed in a recA730 lexA(Def) strain expressing constitutively activated RecA (RecA*). Our data suggest that in a wild-type background, RecF, O and R are important for the induction of the SOS response and the formation of RecA*-dependent recombination intermediates necessary for PriA/Pol II-dependent replication restart. In con-trast, only RecF is required for the activation of RecA that leads to the formation of pol V (UmuD'2C) and facilitates replication readthrough.  相似文献   

15.
Colis LC  Raychaudhury P  Basu AK 《Biochemistry》2008,47(31):8070-8079
Comparative mutagenesis of gamma- or X-ray-induced tandem DNA lesions G[8,5-Me]T and T[5-Me,8]G intrastrand cross-links was investigated in simian (COS-7) and human embryonic (293T) kidney cells. For G[8,5-Me]T in 293T cells, 5.8% of progeny contained targeted base substitutions, whereas 10.0% showed semitargeted single-base substitutions. Of the targeted mutations, the G --> T mutation occurred with the highest frequency. The semitargeted mutations were detected up to two bases 5' and three bases 3' to the cross-link. The most prevalent semitargeted mutation was a C --> T transition immediately 5' to the G[8,5-Me]T cross-link. Frameshifts (4.6%) (mostly small deletions) and multiple-base substitutions (2.7%) also were detected. For the T[5-Me,8]G cross-link, a similar pattern of mutations was noted, but the mutational frequency was significantly higher than that of G[8,5-Me]T. Both targeted and semitargeted mutations occurred with a frequency of approximately 16%, and both included a dominant G --> T transversion. As in 293T cells, more than twice as many targeted mutations in COS cells occurred in T[5-Me,8]G (11.4%) as in G[8,5-Me]T (4.7%). Also, the level of semitargeted single-base substitutions 5' to the lesion was increased and 3' to the lesion decreased in T[5-Me,8]G relative to G[8,5-Me]T in COS cells. It appeared that the majority of the base substitutions at or near the cross-links resulted from incorporation of dAMP opposite the template base, in agreement with the so-called "A-rule". To determine if human polymerase eta (hpol eta) might be involved in the mutagenic bypass, an in vitro bypass study of G[8,5-Me]T in the same sequence was carried out, which showed that hpol eta can bypass the cross-link incorporating the correct dNMP opposite each cross-linked base. For G[8,5-Me]T, nucleotide incorporation by hpol eta was significantly different from that by yeast pol eta in that the latter was more error-prone opposite the cross-linked Gua. The incorporation of the correct nucleotide, dAMP, by hpol eta opposite cross-linked T was 3-5-fold more efficient than that of a wrong nucleotide, whereas incorporation of dCMP opposite the cross-linked G was 10-fold more efficient than that with dTMP. Therefore, the nucleotide incorporation pattern by hpol eta was not consistent with the observed cellular mutations. Nevertheless, at and near the lesion, hpol eta was more error-prone compared to a control template. The in vitro data suggest that translesion synthesis by another Y-family DNA polymerase and/or flawed participation of an accessory protein is a more likely scenario in the mutagenesis of these lesions in mammalian cells. However, hpol eta may play a role in correct bypass of the cross-links.  相似文献   

16.
Recently, the Escherichia coli umuD and umuC genes have been shown to encode E. coli's fifth DNA polymerase, pol V (consisting of a heterotrimer of UmuD'(2)C). The main function of pol V appears to be the bypass of DNA lesions that would otherwise block replication by pols I-IV. This process is error-prone and leads to a striking increase in mutations at sites of DNA damage. While the enzymatic properties of pol V are now only beginning to be fully appreciated, a great deal is known about how E. coli regulates the intracellular levels of the Umu proteins so that the lesion-bypassing activity of pol V is available to help cells survive the deleterious consequences of DNA damage, yet keeps any unwarranted activity on undamaged templates to a minimum. Our review summarizes the multiple restrictions imposed upon pol V, so as to limit its activity in vivo and, in particular, highlights the pivotal role that the N-terminal tail of UmuD plays in regulating SOS mutagenesis.  相似文献   

17.
In the Ames Salmonella typhimurium reversion assay 1,6- and 1,8-dinitropyrenes (1,6- and 1,8-DNPs) are much more potent mutagens than 1-nitropyrene (1-NP). Genetic experiments established that certain differences in the metabolism of the DNPs, which in turn result in increased DNA adduction, play a role. It remained unclear, however, if the DNP adducts, N-(guanin-8-yl)-1-amino-6 ()-nitropyrene (Gua-C8-1,6-ANP and Gua-C8-1,8-ANP), which contain a nitro group on the pyrene ring covalently linked to the guanine C8, are more mutagenic than the major 1-NP adduct, N-(guanin-8-yl)-1-aminopyrene (Gua-C8-AP). In order to address this, we have compared the mutation frequency of the three guanine C8 adducts, Gua-C8-AP, Gua-C8-1,6-ANP, and Gua-C8-1,8-ANP in a CGCG*CG sequence. Single-stranded M13mp7L2 vectors containing these adducts and a control were constructed and replicated in Escherichia coli. A remarkable difference in the induced CpG deletion frequency between these adducts was noted. In repair-competent cells the 1-NP adduct induced 1.7% CpG deletions without SOS, whereas the 1,6- and 1,8-DNP adducts induced 6.8 and 10.0% two-base deletions, respectively. With SOS, CpG deletions increased up to 1.9, 11.1, and 15.1% by 1-NP, 1,6-, and 1,8-DNP adducts, respectively. This result unequivocally established that DNP adducts are more mutagenic than the 1-NP adduct in the repetitive CpG sequence. In each case the mutation frequency was significantly increased in a mutS strain, which is impaired in methyl-directed mismatch repair, and a dnaQ strain, which carries a defect in proofreading activity of the DNA polymerase III. Modeling studies showed that the nitro group on the pyrene ring at the 8-position can provide additional stabilization to the two-nucleotide extrahelical loop in the promutagenic slipped frameshift intermediate through its added hydrogen-bonding capability. This could account for the increase in CpG deletions in the M13 vector with the nitro-containing adducts compared with the Gua-C8-AP adduct itself.  相似文献   

18.
8-Oxo-7,8-dihydroguanine (8-hydroxyguanine) is oxidized more easily than normal nucleobases, which can produce spiroiminodihydantoin (Sp) and guanidinohydantoin (Gh). These secondary oxidation products of 8-oxo-7,8-dihydroguanine are highly mutagenic when formed within DNA. To evaluate the mutagenicity of the corresponding oxidation products of 8-oxo-7,8-dihydro-2'-deoxyguanosine 5'-triphosphate (8-hydroxy-2'- deoxyguanosine 5'-triphosphate) in the nucleotide pool, Escherichia coli cells deficient in the mutT gene were treated with H(2)O(2), and the induced mutations were analyzed. Moreover, the 2'-deoxyriboside 5'-triphosphate derivatives of Sp and Gh were also introduced into competent E. coli cells. The H(2)O(2) treatment of mutT E. coli cells resulted in increase of G:C → T:A and A:T → T:A mutations. However, the incorporation of exogenous Sp and Gh 2'-deoxyribonucleotides did not significantly increase the mutation frequency. These results suggested that the oxidation product(s) of 8-oxo-7,8-dihydro-2'-deoxyguanosine 5'-triphosphate induces G:C → T:A and A:T → T:A mutations, and that the 2'-deoxyriboside 5'-triphosphate derivatives of Sp and Gh exhibit quite weak mutagenicity, in contrast to the bases in DNA.  相似文献   

19.
Closed circular double stranded M13mp19 DNA containing a site-specifically placed HMT (4'-hydroxymethyl-4-5'-8-trimethylpsoralen) monoadduct or crosslink was synthesized in vitro. The damaged DNA were scored for loss of infectivity by transfection into repair proficient or deficient E. coli and into SOS induced E. coli. Mutant phages were detected by the loss of alpha-complementation between the viral and the host Lac Z genes or by the acquisition of resistance to kpn I digestion. Our results indicate that HMT mutagenesis is targeted and that deletion or transversion of the modified thymidine is the predominant sequence change elicited by a monoadduct or a crosslink. Transfection of the monoadducted DNA into a Uvr A deficient strain did not change the mutation pattern but did increase the respective mutation frequencies. Transfection of the crosslinked DNA into a SOS induced host resulted in the appearence of other types of mutations attributable to an increase in both targeted and untargeted mutations.  相似文献   

20.
To understand the role of POL30 in mutation suppression, 11 Saccharomyces cerevisiae pol30 mutator mutants were characterized. These mutants were grouped based on their mutagenic defects. Many pol30 mutants harbor multiple mutagenic defects and were placed in more than one group. Group A mutations (pol30-52, -104, -108, and -126) caused defects in mismatch repair (MMR). These mutants exhibited mutation rates and spectra reminiscent of MMR-defective mutants and were defective in an in vivo MMR assay. The mutation rates of group A mutants were enhanced by a msh2 or a msh6 mutation, indicating that MMR deficiency is not the only mutagenic defect present. Group B mutants (pol30-45, -103, -105, -126, and -114) exhibited increased accumulation of either deletions alone or a combination of deletions and duplications (4 to 60 bp). All deletion and duplication breakpoints were flanked by 3 to 7 bp of imperfect direct repeats. Genetic analysis of one representative group B mutant, pol30-126, suggested polymerase slippage as the likely mutagenic mechanism. Group C mutants (pol30-100, -103, -105, -108, and -114) accumulated base substitutions and exhibited synergistic increases in mutation rate when combined with msh6 mutations, suggesting increased DNA polymerase misincorporation as a mutagenic defect. The synthetic lethality between a group A mutant, pol30-104, and rad52 was almost completely suppressed by the inactivation of MSH2. Moreover, pol30-104 caused a hyperrecombination phenotype that was partially suppressed by a msh2 mutation. These results suggest that pol30-104 strains accumulate DNA breaks in a MSH2-dependent manner.  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