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 147 毫秒
1.
2.
Using a pair of plasmids carrying the rpsL target sequence in different orientations to the replication origin, we analyzed a large number of forward mutations generated in wild-type and mismatch-repair deficient (MMR(-)) Escherichia coli cells to assess the effects of directionality of replication-fork movement on spontaneous mutagenesis and the generation of replication error. All classes of the mutations found in wild-type cells but not MMR(-) cells were strongly affected by the directionality of replication fork movement. It also appeared that the directionality of replication-fork movement governs the directionality of sequence substitution mutagenesis, which occurred in wild-type cells at a frequency comparable to base substitutions and single-base frameshift mutations. A very strong orientation-dependent hot-spot site for single-base frameshift mutations was discovered and demonstrated to be caused by the same process involved in sequence substitution mutagenesis. It is surprising that dnaE173, a potent mutator mutation specific for sequence substitution as well as single-base frameshift, did not enhance the frequency of the hot-spot frameshift mutation. Furthermore, the frequency of the hot-spot frameshift mutation was unchanged in the MMR(-) strain, whereas the mutHLS-dependent mismatch repair system efficiently suppressed the generation of single-base frameshift mutations. These results suggested that the hot-spot frameshift mutagenesis might be initiated at a particular location containing a DNA lesion, and thereby produce a premutagenic replication intermediate resistant to MMR. Significant numbers of spontaneous single-base frameshift mutations are probably caused by similar mechanisms.  相似文献   

3.
Non-targeted mutagenesis of lambda phage by ultraviolet light is the increase over background mutagenesis when non-irradiated phage are grown in irradiated Escherichia coli host cells. Such mutagenesis is caused by different processes from targeted mutagenesis, in which mutations in irradiated phage are correlated with photoproducts in the phage DNA. Non-irradiated phage grown in heavily irradiated uvr+ host cells showed non-targeted mutations, which were 3/4 frameshifts, whereas targeted mutations were 2/3 transitions. For non-targeted mutagenesis in heavily irradiated host cells, there were one to two mutant phage per mutant burst. From this and the pathways of lambda DNA synthesis, it can be argued that non-targeted mutagenesis involves a loss of fidelity in semiconservative DNA replication. A series of experiments with various mutant host cells showed a major pathway of non-targeted mutagenesis by ultraviolet light, which acts in addition to "SOS induction" (where cleavage of the LexA repressor by RecA protease leads to din gene induction): (1) the induction of mutants has the same dependence on irradiation for wild-type and for umuC host cells; (2) a strain in which the SOS pathway is constitutively induced requires irradiation to the same level as wild-type cells in order to fully activate non-targeted mutagenesis; (3) non-targeted mutagenesis occurs to some extent in irradiated recA recB cells. In cells with very low levels of PolI, the induction of non-targeted mutagenesis by ultraviolet light is enhanced. We propose that the major pathway for non-targeted mutagenesis in irradiated host cells involves binding of the enzyme DNA polymerase I to damaged genomic DNA, and that the low polymerase activity leads to frameshift mutations during semiconservative DNA replication. The data suggest that this process will play a much smaller role in ultraviolet mutagenesis of the bacterial genome than it does in the mutagenesis of lambda phage.  相似文献   

4.
We have recently shown that single-base frameshifts were predominant among mutations induced within the rpsL target sequence upon oriC plasmid DNA replication in vitro. We found that the occurrence of +1 frameshifts at a run of 6 residues of dA/dT could be increased proportionally by increasing the concentration of dATP present in the in vitro replication. Using single-stranded circular DNA containing either the coding sequence of the rpsL gene or its complementary sequence, the +1 frameshift mutagenesis by DNA polymerase III holoenzyme of Escherichia coli was extensively examined. A(6) --> A(7) frameshifts occurred 30 to 90 times more frequently during DNA synthesis with the noncoding sequence (dT tract) template than with the coding sequence (dA tract). Excess dATP enhanced the occurrence of +1 frameshifts during DNA synthesis with the dT tract template, but no other dNTPs showed such an effect. In the presence of 0.1 mM dATP, the A(6) --> A(7) mutagenesis with the dT tract template was not inhibited by 1.5 mM dCTP, which is complementary to the residue immediately upstream of the dT tract. These results strongly suggested that the A(6) --> A(7) frameshift mutagenesis possesses an asymmetric strand nature and that slippage errors leading to the +1 frameshift are made during chain elongation within the tract rather than by misincorporation of nucleotides opposite residues next to the tract.  相似文献   

5.
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   

6.
RecA is important in recombination, DNA repair and repair of replication forks. It functions through the production of a protein-DNA filament. To study the localization of RecA in live Escherichia coli cells, the RecA protein was fused to the green fluorescence protein (GFP). Strains with this gene have recombination/DNA repair activities three- to tenfold below wild type (or about 1000-fold above that of a recA null mutant). RecA-GFP cells have a background of green fluorescence punctuated with up to five foci per cell. Two types of foci have been defined: 4,6-diamidino-2-phenylindole (DAPI)-sensitive foci that are bound to DNA and DAPI-insensitive foci that are DNA-less aggregates/storage structures. In log phase cells, foci were not localized to any particular region. After UV irradiation, the number of foci increased and they localized to the cell centre. This suggested colocalization with the DNA replication factory. recA, recB and recF strains showed phenotypes and distributions of foci consistent with the predicted effects of these mutations.  相似文献   

7.
Antimutator alleles indentify genes whose normal products are involved in spontaneous mutagenesis pathways. Mutant alleles of the recA and umuC genes of Escherichia coli, whose wild-type alleles are components of the inducible SOS response, were shown to cause a decrease in the level of spontaneous mutagenesis. Using a series of chromosomal mutant trp alleles, which detect point mutations, as a reversion assay, it was shown that the reduction in mutagenesis is limited to base-pair substitutions. Within the limited number of sites than could be examined, transversions at AT sites were the favored substitutions. Frameshift mutagenesis was slightly enhanced by a mutant recA allele and unchanged by a mutant umuC allele. The wild-type recA and umuC genes are involved in the same mutagenic base-pair substitution pathway, designated "SOS-dependent spontaneous mutagenesis" (SDSM), since a recAumuC strain showed the same degree and specificity of antimutator activity as either single mutant strain. The SDSM pathway is active only in the presence of oxygen, since wild-type, recA, and umuC strains all show the same levels of reduced spontaneous mutagenesis anaerobically. The SDSM pathway can function in starving/stationary cells and may, or may not, be operative in actively dividing cultures. We suggest that, in wild-type cells, SDSM results from basal levels of SOS activity during DNA synthesis. Mutations may result from synthesis past cryptic DNA lesions (targeted mutagenesis) and/or from mispairings during synthesis with a normal DNA template (untargeted mutagenesis). Since it occurs in chromosomal genes of wild-type cells, SDSM may be biologically significant for isolates of natural enteric bacterial populations where extended starvation is often a common mode of existence.  相似文献   

8.
The DNA damage-inducible SOS response of Escherichia coli includes an error-prone translesion DNA replication activity responsible for SOS mutagenesis. In certain recA mutant strains, in which the SOS response is expressed constitutively, SOS mutagenesis is manifested as a mutator activity. Like UV mutagenesis, SOS mutator activity requires the products of the umuDC operon and depends on RecA protein for at least two essential activities: facilitating cleavage of LexA repressor to derepress SOS genes and processing UmuD protein to produce a fragment (UmuD') that is active in mutagenesis. To determine whether RecA has an additional role in SOS mutator activity, spontaneous mutability (tryptophan dependence to independence) was measured in a family of nine lexA-defective strains, each having a different recA allele, transformed or not with a plasmid that overproduces either UmuD' alone or both UmuD' and UmuC. The magnitude of SOS mutator activity in these strains, which require neither of the two known roles of RecA protein, was strongly dependent on the particular recA allele that was present. We conclude that UmuD'C does not determine the mutation rate independently of RecA and that RecA has a third essential role in SOS mutator activity.  相似文献   

9.
Y. Cao  T. Kogoma 《Genetics》1995,139(4):1483-1494
The mechanism of recA polA lethality in Escherichia coli has been studied. Complementation tests have indicated that both the 5' -> 3' exonuclease and the polymerization activities of DNA polymerase I are essential for viability in the absence of RecA protein, whereas the viability and DNA replication of DNA polymerase I-defective cells depend on the recombinase activity of RecA. An alkaline sucrose gradient sedimentation analysis has indicated that RecA has only a minor role in Okazaki fragment processing. Double-strand break repair is proposed for the major role of RecA in the absence of DNA polymerase I. The lexA(Def)::Tn5 mutation has previously been shown to suppress the temperature-sensitive growth of recA200(Ts) polA25::spc mutants. The lexA(Def) mutation can alleviate impaired DNA synthesis in the recA200(Ts) polA25::spc mutant cells at the restrictive temperature. recF(+) is essential for this suppression pathway. recJ and recQ mutations have minor but significant adverse effects on the suppression. The recA200(Ts) allele in the recA200(Ts) polA25::spc lexA(Def) mutant can be replaced by δrecA, indicating that the lexA(Def)-induced suppression is RecA independent. lexA(Def) reduces the sensitivity of δrecA polA25::spc cells to UV damage by ~10(4)-fold. lexA(Def) also restores P1 transduction proficiency to the δrecA polA25::spc mutant to a level that is 7.3% of the recA(+) wild type. These results suggest that lexA(Def) activates a RecA-independent, RecF-dependent recombination repair pathway that suppresses the defect in DNA replication in recA polA double mutants.  相似文献   

10.
The role of Tyr264 in nucleotide binding and hydrolysis catalyzed by the RecA protein of Escherichia coli was investigated by constructing Gly, Ser, and Phe substitution mutations using oligonucleotide-directed mutagenesis. The corresponding mutant recA genes neither restored resistance to killing by ultraviolet irradiation nor increased homologous recombination in a recA strain. The purified RecA(Gly264) protein was unable to bind nucleotide, hydrolyze ATP, or form stable ternary complexes with adenosine 5'-O-thiotriphosphate and DNA although the mutant protein bound DNA normally in the absence of nucleotide. The RecA (Phe264) and RecA(Ser264) proteins hydrolyzed ATP poorly and the rates were reduced approximately 8- and 18-fold, respectively. Although capable of low levels of ATP hydrolysis, neither the RecA(Phe264) nor the RecA(Ser264) protein promoted DNA pairing or strand exchange reactions in vitro. Furthermore, these mutant RecA proteins were impaired in their ability to form salt-resistant ternary complexes with adenosine 5'-O-thiotriphosphate) and DNA as judged by filter binding. Nevertheless, nucleoprotein complexes formed with either RecA(Phe264) or RecA(Ser264) protein directed efficient cleavage of LexA repressor in vitro. These results demonstrate that Tyr264 is required for efficient ATP hydrolysis and for homologous pairing of DNA but does not participate in activating RecA protein for LexA repressor autodigestion.  相似文献   

11.
To better understand the mechanisms of SOS mutagenesis in the bacterium Escherichia coli, we have undertaken a genetic analysis of the SOS mutator activity. The SOS mutator activity results from constitutive expression of the SOS system in strains carrying a constitutively activated RecA protein (RecA730). We show that the SOS mutator activity is not enhanced in strains containing deficiencies in the uvrABC nucleotide excision-repair system or the xth and nfo base excision-repair systems. Further, recA730-induced errors are shown to be corrected by the MutHLS-dependent mismatch-repair system as efficiently as the corresponding errors in the rec+ background. These results suggest that the SOS mutator activity does not reflect mutagenesis at so-called cryptic lesions but instead represents an amplification of normally occurring DNA polymerase errors. Analysis of the base-pair-substitution mutations induced by recA730 in a mismatch repair-deficient background shows that both transition and transversion errors are amplified, although the effect is much larger for transversions than for transitions. Analysis of the mutator effect in various dnaE strains, including dnaE antimutators, as well as in proofreading-deficient dnaQ (mutD) strains suggests that in recA730 strains, two types of replication errors occur in parallel: (i) normal replication errors that are subject to both exonucleolytic proofreading and dnaE antimutator effects and (ii) recA730-specific errors that are not susceptible to either proofreading or dnaE antimutator effects. The combined data are consistent with a model suggesting that in recA730 cells error-prone replication complexes are assembled at sites where DNA polymerization is temporarily stalled, most likely when a normal polymerase insertion error has created a poorly extendable terminal mismatch. The modified complex forces extension of the mismatch largely at the exclusion of proofreading and polymerase dissociation pathways. SOS mutagenesis targeted at replication-blocking DNA lesions likely proceeds in the same manner.  相似文献   

12.
The RecA protein plays a key role in Escherichia coli recombination and DNA repair. We have created new recA mutants with mutations in the vicinity of the recA430 mutation (Gly-204----Ser) which is known to affect RecA coprotease activity. Mutants carrying recA659 or recA611, located 3 and 7 amino acids downstream of residue 204, respectively, lose all RecA activities, while the mutant carrying recA616, which is located at 12 amino acids from this residue, keeps the coprotease activity but is unable to promote recombination. Complementation experiments show that both mutations recA611 and recA659 are dominant over the wild-type or recA430 allele while recA616 seems to be recessive to recA+ and dominant over recA430. It is suggested that these mutations are located in RecA domains which direct conformational modifications.  相似文献   

13.
It has been hard to detect forward mutations generated during DNA synthesis in vitro by replicative DNA polymerases, because of their extremely high fidelity and a high background level of pre-existing mutations in the single-stranded template DNA used. Using the oriC plasmid DNA replication in vitro system and the rpsL forward mutation assay, we examined the fidelity of DNA replication catalyzed by the replicative apparatus of Escherichia coli. Upon DNA synthesis by the fully reconstituted system, the frequency of rpsL-mutations in the product DNA was increased to 1.9x10(-4), 50-fold higher than the background level of the template DNA. Among the mutations generated in vitro, single-base frameshifts predominated and occurred with a pattern similar to those induced in mismatch-repair deficient E. coli cells, indicating that the major replication error was slippage at runs of the same nucleotide. Large deletions and other structural alterations of DNA appeared to be induced also during the action of the replicative apparatus.  相似文献   

14.
RecA protein in the SOS response: milestones and mysteries   总被引:8,自引:0,他引:8  
E M Witkin 《Biochimie》1991,73(2-3):133-141
The role of RecA protein in the SOS response of Escherichia coli is traced from the isolation of the first recA mutant to our current understanding of the scope and regulation of this DNA damage-inducible system. In addition, possible RecA protein activities that may be essential in the expression of several SOS phenotypes (stable DNA replication, DNA replication recovery, SOS mutagenesis and RecA association with the cell membrane) are discussed.  相似文献   

15.
16.
After unsuccessful attempts to recover a viable RecA-deficient mutant of the Lyme borreliosis agent Borrelia burgdorferi, we characterized the functional activities of RecA of B. burgdorferi, as well as RecA of the relapsing fever spirochete Borrelia hermsii and the free-living spirochete Leptospira biflexa, in a recA mutant of Escherichia coli. As a control, E. coli RecA was expressed from the same plasmid vector. DNA damage repair activity was assessed after exposure of the transgenic cells to UV light or the radiomimetic chemicals methyl methanesulfonate and mitomycin C. Recombination activity in the cells was assessed by using an assay for homologous recombination between repeats in the chromosome and by measuring the ability of the cells to foster lytic growth by red gam mutant bacteriophage lambda. Overall, we found that transgenic cells with recA genes of B. burgdorferi, B. hermsii, and L. biflexa had approximately equivalent activities in promoting homologous recombination in the lacZ duplication assay, but cells with B. burgdorferi recA and, most notably, B. hermsii recA were significantly less capable than cells with L. biflexa recA or E. coli recA in responding to DNA damage or in facilitating plaque formation in the phage assay. The comparatively poor function of Borrelia recA in the latter set of assays may be the consequence of impaired coordination in the loading of the transgenic RecA by RecBCD and/or RecFOR in E. coli.  相似文献   

17.
We developed a system to examine forward mutations that occurred in the rpsL gene of Escherichia coli placed on a multicopy plasmid. Using this system we determined the mutational specificity for a dnaE173 mutator strain in which the editing function of DNA polymerase III is impeded. The frequency of rpsL- mutations increased 32,000-fold, due to the dnaE173 mutator, and 87 independent rpsL- mutations in the mutator strain were analyzed by DNA sequencing, together with 100 mutants recovered from dnaE+ strain, as the control. While half the number of mutations that occurred in the wild-type strain were caused by insertion elements, no such mutations were recovered from the mutator strain. A novel class of mutation, named "sequence substitution" was present in mutants raised in the dnaE173 strain; seven sequence substitutions induced in the mutator strain occurred at six sites, and all were located in quasipalindromic sequences, carrying the GTG or CAC sequence at one or both endpoints. While other types of mutation were found in both strains, single-base frameshifts were the most frequent events in the mutator strain. Thus, the mutator effect on this class of mutation was 175,000-fold. A total of 95% of the single-base frameshifts in the mutator strain were additions, most of which occurred at runs of A or C bases so as to increase the number of identical residues. Base substitutions, the frequency of which was enhanced 25,000-fold by the mutator effect, occurred primarily at several hotspots in the mutator strain, whereas those induced in the wild-type strain were more randomly distributed throughout the rpsL sequence. The dnaE173 mutator also increased the frequency of duplications 28,000-fold. Of the three duplications recovered from the mutator strain, one was a simple duplication, the region of which was flanked by direct repeats. The other duplications were complex, one half part of which was in the inverted orientation of a region containing two sets of inverted repeats. The same duplications were also recovered from the wild-type strain. The present data suggest that dnaE173 is a novel class of mutator that sharply induces sequence-directed mutagenesis, yielding high frequencies of single base frameshifts, duplications with inversions, sequence substitutions and base substitutions at hotspots.  相似文献   

18.
The effect of gaseous nitrogen dioxide (NO2) on cytotoxicity, induction of synthesis of UmuC and RecA proteins, and mutagenesis was studied in Escherichia coli strains with different capacities of DNA repair. Gaseous NO2 (90, 180 microliter/l) killed Escherichia coli. The recA mutant was most sensitive, the lexA mutant moderately sensitive, and the uvrA mutant and the wild-type the least sensitive. When 90 microliter/l NO2 gas was bubbled into bacterial suspensions for 30 min at a flow rate of 100 ml/min, the induction of umuC gene expression increased in the wild-type strain. NO2 also induced the recA gene expression in the wild-type strain. The synthesis of neither RecA nor UmuC proteins was induced in the recA and lexA mutants. We further investigated the NO2 mutagenesis in the cells treated with bubbling of NO2 gas. NO2 caused mutation to Trp+ of WP2.  相似文献   

19.
Oxidative damage of nucleotides within DNA or precursor pools caused by oxygen radicals is thought to play an important role in spontaneous mutagenesis, as well as carcinogenesis and aging. In particular, 8-oxodGTP and 2-OHdATP are potent mutagenic substrate for DNA synthesis. Mammalian MTH1 catalyzes hydrolysis of these mutagenic substrates, suggesting that it functions to prevent mutagenesis caused by these oxidized nucleotides. We have established MTH1(-/-) mice lacking the 8-oxodGTPase activity, which were shown to be susceptible to lung, liver and stomach cancers. To examine in vivo mutation events due to the MTH1-deficiency, a reporter gene, rpsL of Escherichia coli, was introduced into MTH1(-/-) mice. Interestingly, the net frequency of rpsL(-) forward mutants showed no apparent increase in MTH1(-/-) mice as compared to MTH1(+/+) mice. However, we found differences between these two genotypes in the class- and site-distributions of the rpsL(-) mutations recovered from the mice. Unlike MutT-deficient E. coli showing 1000-fold higher frequency of A:T-->C:G transversion than the wild type cells, an increase in frequency of A:T-->C:G transversion was not evident in MTH1 nullizygous mice. Nevertheless, the frequency of single-base frameshifts at mononucleotide runs was 5.7-fold higher in spleens of MTH1(-/-) mice than in those of wild type mice. Since the elevated incidence of single-base frameshifts at mononucleotide runs is a hallmark of the defect in MSH2-dependent mismatch repair system, this weak site-specific mutator effect of MTH1(-/-) mice could be attributed to a partial sequestration of the mismatch repair function that may act to correct mispairs with the oxidized nucleotides. Consistent with this hypothesis, a significant increase in the frequency of G:C-->T:A transversions was observed with MTH1(-/-) MSH2(-/-) mice over MSH2(-/-) mice alone. These results suggest a possible involvement of multiple anti-mutagenic pathways, including the MTH1 protein and other repair system(s), in mutagenesis caused by the oxidized nucleotides.  相似文献   

20.
Expression of the Bacillus subtilis gene coding for SspC, a small, acid-soluble protein, caused both killing and mutation in a number of Escherichia coli B and K-12 strains. SspC was previously shown to bind E. coli DNA in vivo, and in vitro this protein binds DNA and converts it into an A-like conformation. Analysis of revertants of nonsense mutations showed that SspC caused single-base changes, and a greater proportion of these were at A-T base pairs. Mutation in the recA gene abolished the induction of mutations upon synthesis of SspC, but the killing was only slightly greater than in RecA+ cells. Mutations in the umuC and umuD genes eliminated most of the mutagenic effect of SspC but not the killing, while the lexA mutation increased mutagenesis but did not appreciably affect the killing. Since there was neither killing nor mutation of E. coli after synthesis of a mutant SspC which does not bind DNA, it appears likely that the binding of wild-type SspC to DNA, with the attendant conformational change, was responsible for the killing and mutation. A strain containing the B. subtilis gene that is constitutive for the RecA protein at 42 degrees C showed a lower frequency of mutation when that temperature was used to induce the RecA protein than when the temperature was 30 degrees C, where the RecA level is low, suggesting that at the elevated temperature the high RecA level could be inhibiting binding of the B. subtilis protein to DNA.  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