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 46 毫秒
1.
The Escherichia coli RecA protein is the prototype of the RecA/RAD51/DMC1 family of strand transferases acting in genetic recombination. The E96D mutant was previously isolated in a screen for toxic recA mutants and was found to constitutively derepress the SOS genes and inhibit chromosome segregation in E. coli. Here, we have found that the E96D mutation lowers the RecA kcat value for ATP hydrolysis 100-fold. Use of this mutant reveals that the ATPase and branch migration activities of RecA are not necessarily required for catalyzing in vivo recombinational pairing and LexA cleavage. In addition to its effect on ATP hydrolysis, the mutation causes ATP to more strongly promote the transition to the biologically active, extended conformation of the RecA enzyme. The enhanced ATP binding is apparently the cause for a broader nucleic acid ligand specificity. The use of RNA and double-stranded DNA as cofactors for LexA cleavage could give rise to the inappropriate, constitutive derepression of the SOS genes. This underscores the need for the ATP affinity to be optimized so that RecA becomes selectively activated only during DNA repair and recombination through binding single-stranded DNA.  相似文献   

2.
RecA protein of E.coli is a multifunctional protein participating in genetic recombination, recombinational repair and mutagenesis. We used E.coli recA gene as a probe for complementation of repair defects after treatment of N-methyl-N'-nitro-N-nitrosoguanidine in the pso4-1 and rad51::URA3 mutants of S. cerevisiae. We tried to find the role of the RecA protein in S. cerevisiae mutants defective in different repair pathways. The RecA protein had no effect on survival of haploid and diploid pso4-1 mutants, but it had a significant effect on MNNG induced mutagenesis, which was increased to the wild type level. No effect of the RecA protein on survival was observed in rad51::URA3 mutant after MNNG treatment. However, in this case the RecA protein decreased the induced mutagenesis. We can suppose that the RecA protein, with its multifunctional enzymatic activity, can bind to special intermediates and initiate function of different repair pathways depending on repair defects of the mutants studied.  相似文献   

3.
The RecA, UmuC, and UmuD' proteins are essential for error-prone, replicative bypass of DNA lesions. Normally, RecA protein mediates homologous pairing of DNA. We show that purified Umu(D')2C blocks this recombination function. Biosensor measurements establish that the mutagenic complex binds to the RecA nucleoprotein filament with a stoichiometry of one Umu(D')2C complex for every two RecA monomers. Furthermore, Umu(D')2C competitively inhibits LexA repressor cleavage but not ATPase activity, implying that Umu(D')2C binds in or proximal to the helical groove of the RecA nucleoprotein filament. This binding reduces joint molecule formation and even more severely impedes DNA heteroduplex formation by RecA protein, ultimately blocking all DNA pairing activity and thereby abridging participation in recombination function. Thus, Umu(D')2C restricts the activities of the RecA nucleoprotein filament and presumably, in this manner, recruits it for mutagenic repair function. This modulation by Umu(D')2C is envisioned as a key event in the transition from a normal mode of genomic maintenance by "error-free" recombinational repair, to one of "error-prone" DNA replication.  相似文献   

4.
BACKGROUND: The RAD51 gene of Saccharomyces cerevisiae is homologous to the Escherichia coli recA gene and plays a key role in genetic recombination and DNA double-strand break repair. To construct an improved experimental system of homologous recombination in higher eukaryotes, we have chosen the South African clawed frog, Xenopus laevis, whose egg extracts might be useful for the in vitro studies. We identified and characterized a Xenopus homologue of RAD51 gene, the XRAD51.1. RESULTS: Recombinant XRad51.1 was expressed in E. coli. The purified XRad51.1 protein showed ssDNA-dependent ATPase activity and promoted the DNA strand exchange reaction between two 55-mer oligonucleotides. The binding stoichiometry of XRad51.1 to ssDNA was determined by fluorescence of poly(d epsilonA), a chemically modified poly(dA), and was found to be about six bases/XRad51.1 monomer in a nucleoprotein filament, a similar value to E. coli RecA protein. The kinetics of the fluorescence change of poly(d epsilonA) after XRad51.1 binding in the presence of ATP was significantly different from that observed with RecA protein. The affinity of XRad51.1 to ssDNA in the presence of ATP was higher than that of RecA protein, and the dissociation of the XRad51.1-ssDNA complex was slower than the RecA-ssDNA complex. CONCLUSIONS: Purified recombinant XRad51.1 protein promoted the strand exchange between short DNA molecules. While the binding stoichiometry of XRad51.1 protein to ssDNA was identical to that of the RecA protein, XRad51.1 has a significantly higher affinity and binding stability to ssDNA than did the RecA protein in the presence of ATP.  相似文献   

5.
Escherichia coli K12 assay-system is designed in order to detect bioantimutagens, agents preventing mutagenesis by modulation of DNA repair and replication. The assay is composed of four tests aimed at the detection of inhibition of spontaneous and induced mutations (Tests A and B) and at the estimation whether the anti-mutagenic agent acts by increasing the fidelity of DNA replication (Test B), by inhibition of SOS error prone repair (Test C), or by favoring error-free recombinational repair (Test D). In Test A, repair proficient strain and its uvrA counterpart are used for detection of spontaneous and UV-induced mutations, while in Test B mismatch repair deficient strains (mutH, mutS, mutL and uvrD) are used for amplified detection of spontaneous mutations caused by replication errors. In Test C, repair proficient strain carrying sfiA::lacZ fusion is used for measuring the level of SOS induction by monitoring the level of beta-galactosidase. In Test D, the strains carrying different recA alleles (recA+, recA730 and DeltarecA) are used for measuring intrachromosomal recombination between nonoverlapping deletions in duplicated lac operon, by monitoring Lac+ recombinants. The assay-system is validated with model bioantimutagens and used for detection of anti-mutagenic potential of different terpenoid fractions from sage (Salvia officinalis L.). Extract E1/3 of cultivated sage, distinguished from others by its high content of monoterpenoid camphor, reduces UV-induced mutagenesis in Test A, while it has no effect in Tests B and C. In Test D, it enhances intrachromosomal recombination in untreated and UV-irradiated recA+ and recA730 strains. The results suggest that the protective effect is due to stimulation of recombinational repair, similarly to coumarin. We speculate that monoterpenoids from sage enhance genetic recombination by intervening in a formation of RecA-DNA complex and channeling it into recombination reaction.  相似文献   

6.
The recA locus of pathogenic mycobacteria differs from that of non-pathogenic species in that it contains large intervening sequences termed protein introns or inteins that are excised by an unusual protein-splicing reaction. In addition, a high degree of illegitimate recombination has been observed in the pathogenic Mycobacterium tuberculosis complex. Homologous recombination is the main mechanism of integration of exogenous nucleic acids in M. smegmatis, a non-pathogenic mycobacterium species that carries an inteinless RecA and is amenable to genetic manipulations. To investigate the function of recA in mycobacteria, recA- strains of M. smegmatis were generated by allelic exchange techniques. These strains are characterized (i) by increased sensitivity towards DNA-damaging agents [ethylmethylsulphonate (EMS), mitomycin C, UV irradiation] and (ii) by the inability to integrate nucleic acids by homologous recombination. Transformation efficiencies using integrative or replicative vectors were not affected in recA- mutants, indicating that in mycobacteria RecA does not affect plasmid uptake or replication. Complementation of the recA- mutants with the recA from M. tuberculosis restored resistance towards EMS, mitomycin C and UV irradiation. Transformation of the complemented strains with suicide vectors targeting the pyrF gene resulted in numerous allelic exchange mutants. From these data, we conclude that the intein apparently does not interfere with RecA function, i.e. with respect to competency for homologous recombination, the RecAs from pathogenic and non-pathogenic mycobacteria are indistinguishable.  相似文献   

7.
RecQ helicase is important to homologous recombination and DNA repair in Escherichia coli. We demonstrate that RecQ helicase, in conjunction with RecA and SSB proteins, can initiate recombination events in vitro. In addition, RecQ protein is capable of unwinding a wide variety of DNA substrates, including joint molecules formed by RecA protein. These data are consistent with RecQ helicase assuming two roles in the cell; it can be (1) an initiator of homologous recombination, or (2) a disrupter of joint molecules formed by aberrant recombination. These findings also shed light on the function of the eukaryotic homologs of RecQ helicase, the Sgs1, Blm, and Wrn helicases.  相似文献   

8.
A DNA fragment containing the recA gene of Gluconobacter oxydans was isolated and further characterized for its nucleotide sequence and ability to functionally complement various recA mutations. When expressed in an Escherichia coli recA host, the G. oxydans recA protein could efficiently function in homologous recombination and DNA damage repair. The recA gene's nucleotide sequence analysis revealed a protein of 344 amino acids with a molecular mass of 38 kDa. We observed an E. coli-like LexA repressor-binding site in the G. oxydans recA gene promoter region, suggesting that a LexA-like mediated response system may exist in G. oxydans. The expression of G. oxydans recA in E. coli RR1, a recA+ strain, surprisingly caused a remarkable reduction of the host wild-type recA gene function, whereas the expression of both Serratia marcescens recA and Pseudomonas aeruginosa recA gene caused only a slight inhibitory effect on function of the host wild-type recA gene product. Compared with the E. coli RecA protein, the identity of the amino acid sequence of G. oxydans RecA protein is much lower than those RecA proteins of both S. marcescens and Pseudomonas aeruginosa. This result suggests that the expression of another wild-type RecA could interfere with host wild-type recA gene's function, and the extent of such an interference is possibly correlated to the identity of the amino acid sequence between the two classes of RecA protein.  相似文献   

9.
10.
The Escherichia coli RecA protein promotes DNA strand exchange in homologous recombination and recombinational DNA repair. Stopped-flow kinetics and fluorescence resonance energy transfer (FRET) were used to study RecA-mediated strand exchange between a 30-bp duplex DNA and a homologous single-stranded 50mer. In our standard assay, one end of the dsDNA helix was labeled at apposing 5' and 3' ends with hexachlorofluorescein and fluorescein, respectively. Strand exchange was monitored by the increase in fluorescence emission resulting upon displacement of the fluorescein-labeled strand from the initial duplex. The potential advantages of FRET in study of strand exchange are that it noninvasively measures real-time kinetics in the previously inaccessible millisecond time regime and offers great sensitivity. The oligonucleotide substrates model short-range mechanistic effects that might occur within a localized region of the ternary complex formed between RecA and long DNA molecules during strand exchange. Reactions in the presence of ATP with 0.1 microM duplex and 0.1-1.0 microM ss50mer showed triphasic kinetics in 600 s time courses, implying the existence of three mechanistic steps subsequent to presynaptic filament formation. The observed rate constants for the intermediate phase were independent of the concentration of ss50mer and most likely characterize a unimolecular isomerization of the ternary complex. The observed rate constants for the first and third phases decreased with increasing ss50mer concentration. Kinetic experiments performed with the nonhydrolyzable analogue ATPgammaS showed overall changes in fluorescence emission identical to those observed in the presence of ATP. In addition, the observed rate constants for the two fastest reaction phases were identical in ATP or ATPgammaS. The observed rate constant for the slowest phase showed a 4-fold reduction in the presence of ATPgammaS. Results in ATPgammaS using an alternate fluorophore labeling pattern suggest a third ternary intermediate may form prior to ssDNA product release. The existence of two or three ternary intermediates in strand exchange with a 30 bp duplex suggests the possibility that the step size for base pair switching may be 10-15 bp. Products of reactions in the presence of ATP and ATPgammaS, with and without proteinase K treatment, were analyzed on native polyacrylamide gels. In reactions in which only short-range RecA-DNA interactions were important, ATP hydrolysis was not required for recycling of RecA from both oligonucleotide products. Hydrolysis or deproteinization was required for RecA to release the heteroduplex product, but not the outgoing single strand.  相似文献   

11.
We have developed a new assay to characterize the double-stranded DNA (dsDNA) binding properties of RecA protein. This assay is based on measurement of changes in the fluorescence of a 4',6-diamidino-2-phenylindole (DAPI)-dsDNA complex upon RecA protein binding. The binding of RecA protein to a complex of DAPI and dsDNA results in displacement of the bound DAPI, producing a decrease in the observed fluorescence. DAPI displacement is dependent on both RecA protein and ATP; dATP and, to a lesser extent, UTP and dCTP also support the DAPI displacement reaction, but dGTP, GTP, dITP and TTP do not. Binding stoichiometry for the RecA protein-dsDNA complex measured by DAPI displacement is 3 bp per RecA protein monomer in the presence of ATP. These results, taken together with data for mutant RecA proteins, suggest that this DAPI displacement assay monitors formation of the high affinity DNA binding state of RecA protein. Since this state of RecA protein defines the form of the nucleoprotein filament that is active in DNA strand exchange, these findings raise the possibility that the RecA protein-dsDNA filament may possess a homologous pairing capacity.  相似文献   

12.
The recA gene of a thermophilic eubacterial strain, Thermus thermophilus (T.th.) HB8, was cloned from a genomic DNA library by Southern hybridization using a gene-internal fragment amplified by the polymerase chain reaction (PCR) method as the probe. The gene encoded a 36 kDa polypeptide whose amino acid sequence showed 61% identity with that of the Escherichia coli RecA protein. Characteristic amino acid changes between the two RecA proteins were found. In the amino acid composition of the T.th. RecA protein, the number of Pro residues was increased, the number of Cys residues was decreased, and Lys residues were replaced by Arg, Asp by Glu, Thr by Val, and Ile by Val or Leu. These changes are supposed to stabilize the native protein conformation against heat denaturation. The amino acid residues in the nucleotide binding site of the protein and in the protein-protein interaction site responsible for the oligomer formation were well conserved. The T.th. recA gene has the ability to complement the ultraviolet light (UV) sensitivity of a E. coli recA deletion mutant. Thus, the thermophilic bacterium has a RecA protein whose function will be common to the E. coli RecA protein.  相似文献   

13.
In addition to its critical role in genetic recombination, the Escherichia coli RecA protein plays a pivotal role in SOS-induced mutagenesis. This role can be separated genetically into three steps: (i) depression of the SOS regulon by mediating the posttranslational cleavage of the LexA repressor, (ii) activation of UmuD'-like proteins by mediating cleavage of the UmuD-like proteins, and (iii) a direct step, possibly to interact with and to target the Umu-like mutagenesis proteins to lesions in DNA. We have analyzed RecA's third role biochemically using protein affinity chromatography and an agarose-based DNA mobility-shift assay. RecA730 protein from a crude cell extract was specifically retained on UmuD and UmuD' protein affinity columns, suggesting that these proteins physically interact. Normally, neither UmuD nor UmuD' shows any affinity for DNA. In the presence of RecA protein, however, UmuD and UmuD' were targeted to DNA. RecA1730 protein, which is defective for UmuD' but proficient for MucA'-promoted mutagenesis, showed a dramatically reduced capacity to target UmuD' to DNA but was able to target a significant portion of MucA' to DNA. These data support the suggestion that the direct role of RecA protein in SOS-induced mutagenesis is to interact with and target the Umu-like mutagenesis proteins to DNA.  相似文献   

14.
The Rad51 protein of Saccharomyces cerevisiae is a eukaryotic homolog of the RecA protein, the prototypic DNA strand-exchange protein of Escherichia coli. RAD51 gene function is required for efficient genetic recombination and for DNA double-strand break repair. Recently, we demonstrated that RecA protein has a preferential affinity for GT-rich DNA sequences-several of which exhibit enhanced RecA protein-promoted homologous pairing activity. The fundamental similarity between the RecA and Rad51 proteins suggests that Rad51 might display an analogous bias. Using in vitro selection, here we show that the yeast Rad51 protein shares the same preference for GT-rich sequences as its prokaryotic counterpart. This bias is also manifest as an increased ability of Rad51 protein to promote the invasion of supercoiled DNA by homologous GT-rich single-stranded DNA, an activity not previously described for the eukaryotic pairing protein. We propose that the preferred utilization of GT-rich sequences is a conserved feature among all homologs of RecA protein, and that GT-rich regions are loci for increased genetic exchange in both prokaryotes and eukaryotes.  相似文献   

15.
The genetic requirements for adaptive mutation in Escherichia coli parallel those for homologous recombination in the RecBCD pathway. Recombination-deficient recA and recB null mutant strains are deficient in adaptive reversion. A hyper-recombinagenic recD strain is hypermutable, and its hypermutation depends on functional recA and recB genes. Genes of subsidiary recombination systems are not required. These results indicate that the molecular mechanism by which adaptive mutation occurs includes recombination. No such association is seen for spontaneous mutation in growing cells.  相似文献   

16.
The role of the RAD57 gene in double-strand gap (DSG) repair has been examined. The repair of a linearized plasmid, bearing a DSG, has been analyzed in a rad57-1 mutant of Saccharomyces cerevisiae. For effective rejoining of the ends of plasmid DNA in the rad57 mutant the sequence of chromosomal DNA homologous to the DSG region is required. However, DSG repair (restoration of plasmid circularity) in rad57 cells is not accompanied by the recovery of DSGs. The DSG repair, which depends on an homologous chromosomal DNA sequence, requires the cohesive ends of DSGs. The non-cohesive-ended DSGs are repaired in rad57 cells by a pathway independent of the homologous recombination between chromosomal and plasmid DNA. We presume that the rad57-1 mutation is connected with the inhibition of DNA repair synthesis, required for filling the DSG. This situation produces a condition of "homology-dependent ligation", the alternative minor mechanism of recombinational DSG repair, that takes place in mutant cells. A molecular model for "homology-dependent ligation" in rad57 cells is proposed.  相似文献   

17.
The Escherichia coli RecA protein has served as a model for understanding protein-catalyzed homologous recombination, both in vitro and in vivo. Although RecA proteins have now been sequenced from over 60 different bacteria, almost all of our structural knowledge about RecA has come from studies of the E. coli protein. We have used electron microscopy and image analysis to examine three different structures formed by the RecA protein from the thermophilic bacterium Thermus aquaticus. This protein has previously been shown to catalyze an in vitro strand exchange reaction at an optimal temperature of about 60 degrees C. We show that the active filament formed by the T. aquaticus RecA on DNA in the presence of a nucleotide cofactor is extremely similar to the filament formed by the E. coli protein, including the extension of DNA to a 5.1-A rise per base pair within this filament. This parameter appears highly conserved through evolution, as it has been observed for the eukaryotic RecA analogs as well. We have also characterized bundles of filaments formed by the T. aquaticus RecA in the absence of both DNA and nucleotide cofactor, as well as hexameric rings of the protein formed under all conditions examined. The bundles display a very large plasticity of mass within the RecA filament, as well as showing a polymorphism in filament-filament contacts that may be important to understanding mutations that affect surface residues on the RecA filament.  相似文献   

18.
A 1.8-kb NsiI-StuI fragment containing the recA gene of Xanthomonas campestris pv. campestris was cloned by a PCR-based approach and complementation of Escherichia coli HB 101. Sequence analysis of this fragment revealed an ORF (orf343) of 1,032 bp able to encode a protein of 343 amino acids with a calculated MW of 37,021 Da, a size similar to the values detected by in vitro system and Western blotting. It showed 69.6% identity to the E. coli RecA in amino acid sequence. Amino acid residues of the E coli RecA associated with functional activities are conserved in this Xc17 RecA. The recA mutant, L1, constructed by gene replacement, was sensitive to ultraviolet irradiation and methyl methanesulfonate, and deficient in homologous recombination.  相似文献   

19.
20.
RecA protein of Escherichia coli plays an essential role in homologous recombination of DNA strands. To analyze the interaction of RecA with single-stranded DNA (ssDNA), we performed a fluorescence competition assay employing 1-anilinonaphthalene-8-sulfonic acid (ANS) as an extrinsic fluorescent probe. ANS bound to RecA at three sites, leading to enhancement of ANS fluorescence. Addition of synthetic polynucleotides to the RecA-ANS complex in the absence of a nucleotide quenched the ANS fluorescence, indicating displacement of ANS molecules by ssDNA. Less effective quenching by poly(dA) suggests that the nucleoprotein filament on poly(dA) may differ from those on poly(dT) and poly(dC). A titration experiment with poly(dT) and poly(dA) showed clear stoichiometric binding of 3.5 nucleotides per protein. The site size for poly(dC) was 7.0, which could be explained by the formation of a double helix of poly(dC). ATP and other nucleotides also displaced the ANS. To identify ANS-binding sites, ANS was incorporated into RecA by UV irradiation, and fluorescent peptides were isolated from the proteolytic digest. Sequence analysis suggested that ANS binds to or near the ATP-binding region. These results suggest that the fluorescence quenching and photoincorporation assay using ANS may be useful for the analysis of the interaction of a protein and its ligand.  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