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.
Purification of an SOS repressor from Bacillus subtilis.   总被引:6,自引:5,他引:1       下载免费PDF全文
C M Lovett  Jr  K C Cho    T M O'Gara 《Journal of bacteriology》1993,175(21):6842-6849
We have identified in Bacillus subtilis a DNA-binding protein that is functionally analogous to the Escherichia coli LexA protein. We show that the 23-kDa B. subtilis protein binds specifically to the consensus sequence 5'-GAACN4GTTC-3' located within the putative promoter regions of four distinct B. subtilis DNA damage-inducible genes: dinA, dinB, dinC, and recA. In RecA+ strains, the protein's specific DNA binding activity was abolished following treatment with mitomycin C; the decrease in DNA binding activity after DNA damage had a half-life of about 5 min and was followed by an increase in SOS gene expression. There was no detectable decrease in DNA binding activity in B. subtilis strains deficient in RecA (recA1, recA4) or otherwise deficient in SOS induction (recM13) following mitomycin C treatment. The addition of purified B. subtilis RecA protein, activated by single-stranded DNA and dATP, abolished the specific DNA binding activity in crude extracts of RecA+ strains and strains deficient in SOS induction. We purified the B. subtilis DNA-binding protein more than 4,000-fold, using an affinity resin in which a 199-bp DNA fragment containing the dinC promoter region was coupled to cellulose. We show that B. subtilis RecA inactivates the DNA binding activity of the purified B. subtilis protein in a reaction that requires single-stranded DNA and nucleoside triphosphate. By analogy with E. coli, our results indicate that the DNA-binding protein is the repressor of the B. subtilis SOS DNA repair system.  相似文献   

2.
3.
Neisseria gonorrhoeae is a human-specific organism that is not usually exposed to UV light or chemicals but is likely to encounter reactive oxygen species during infection. Exposure of N. gonorrhoeae to sublethal hydrogen peroxide revealed that the ng1427 gene was upregulated sixfold. N. gonorrhoeae was thought to lack an SOS system, although NG1427 shows amino acid sequence similarity to the SOS response regulator LexA from Escherichia coli. Similar to LexA and other S24 peptidases, NG1427 undergoes autoproteolysis in vitro, which is facilitated by either the gonococcal or E. coli RecA proteins or high pH, and autoproteolysis requires the active and cleavage site residues conserved between LexA and NG1427. NG1427 controls a three gene regulon: itself; ng1428, a Neisseria-specific, putative integral membrane protein; and recN, a DNA repair gene known to be required for oxidative damage survival. Full NG1427 regulon de-repression requires RecA following methyl methanesulphonate or mitomycin C treatment, but is largely RecA-independent following hydrogen peroxide treatment. NG1427 binds specifically to the operator regions of the genes it controls, and DNA binding is abolished by oxidation of the single cysteine residue encoded in NG1427. We propose that NG1427 is inactivated independently of RecA by oxidation.  相似文献   

4.
Coxiella burnetii , a Gram-negative obligate intracellular pathogen, replicates within an parasitophorous vacuole with lysosomal characteristics. To understand how C. burnetii maintains genomic integrity in this environment, a database search for genes involved in DNA repair was performed. Major components of repair, SOS response and recombination were identified, including recA and ruvABC , but lexA and recBCD were absent. Instead, C. burnetii possesses addAB orthologous genes, functional equivalents to recBCD . Survival after treatment with UV, mitomycin C (MC) or methyl methanesulfonate (MMS), as well as homologous recombination in Hfr mating was restored in Escherichia coli deletion strains by C. burnetii recA or addAB . Despite the absence of LexA, co-protease activity for C. burnetii RecA was demonstrated. Dominant-negative inhibition of C. burnetii RecA by recA mutant alleles, modelled after E. coli recA1 and recA56 , was observed and more apparent with expression of C. burnetii RecAG159D mutant protein. Expression of a subset of repair genes in C. burnetii was monitored and, in contrast to the non-inducible E. coli recBCD , addAB expression was strongly upregulated under oxidative stress. Constitutive SOS gene expression due to the lack of LexA and induction of AddAB likely reflect a unique repair adaptation of C. burnetii to its hostile niche.  相似文献   

5.
Purification of a RecA protein analogue from Bacillus subtilis   总被引:29,自引:0,他引:29  
We have identified in Bacillus subtilis an analogue of the Escherichia coli RecA protein. Its activities suggest that it has a corresponding role in general genetic recombination and in regulation of SOS (DNA repair) functions. The B. subtilis protein (B. subtilis Rec) has a Mr of 42,000 and cross-reacts with antisera raised against E. coli RecA protein. Its level is significantly reduced in the recombination-deficient recE4 mutant. B. subtilis Rec is induced 10- to 20-fold in rec+ strains following treatment with mitomycin C, whereas it is not induced in the recombination-deficient mutants recE4, recE45, and recA1. We have purified B. subtilis Rec about 2000-fold to near homogeneity and we describe its activities. It catalyzes DNA-dependent hydrolysis of dATP at a rate comparable to that of E. coli RecA protein. However, B. subtilis Rec has a negligible ATPase activity, although ATP effectively inhibits dATP hydrolysis. In the presence of dATP, B. subtilis Rec catalyzes DNA strand transfer, assayed by the conversion of phi X174 linear duplex DNA and homologous circular single-stranded DNA to replicative form II (circular double-stranded DNA with a discontinuity in one strand). ATP does not support strand transfer by this protein. B. subtilis Rec catalyzes proteolytic cleavage of E. coli LexA repressor in a reaction that requires single-stranded DNA and nucleoside triphosphate. This result suggests that an SOS regulatory system like the E. coli system is present in B. subtilis. The B. subtilis enzyme does not promote any detectable cleavage of the E. coli bacteriophage lambda repressor.  相似文献   

6.
We quantitated the induction of the Bacillus subtilis Rec protein (the analog of Escherichia coli RecA protein) and the B. subtilis din-22 operon (representative of a set of DNA damage-inducible operons in B. subtilis) following DNA damage in Rec+ and DNA repair-deficient strains. After exposure to mitomycin C or UV irradiation, each of four distinct rec (recA1, recB2, recE4, and recM13) mutations reduced to the same extent the rates of both Rec protein induction (determined by densitometric scanning of immunoblot transfers) and din-22 operon induction (determined by assaying beta-galactosidase activity in din-22::Tn917-lacZ fusion strains). The induction deficiencies in recA1 and recE4 strains were partially complemented by the E. coli RecA protein, which was expressed on a plasmid in B. subtilis; the E. coli RecA protein had no effect on either induction event in Rec+, recB2, or recM13 strains. These results suggest that (i) the expression of both the B. subtilis Rec protein and the din-22 operon share a common regulatory component, (ii) the recA1 and recE4 mutations affect the regulation and/or activity of the B. subtilis Rec protein, and (iii) an SOS regulatory system like the E. coli system is highly conserved in B. subtilis. We also showed that the basal level of B. subtilis Rec protein is about 4,500 molecules per cell and that maximum induction by DNA damage causes an approximately fivefold increase in the rate of Rec protein accumulation.  相似文献   

7.
Deinococcus radiodurans strain rec30, which is a DNA damage repair-deficient mutant, has been estimated to be defective in the deinococcal recA gene. To identify the mutation site of strain rec30 and obtain information about the region flanking the gene, a 4.4-kb fragment carrying the wild-type recA gene was sequenced. It was revealed that the recA locus forms a polycistronic operon with the preceding cistrons (orf105a and orf105b). Predicted amino acid sequences of orf105a and orf105b showed substantial similarity to the competence-damage inducible protein (cinA gene product) from Streptococcus pneumoniae and the 2'-5' RNA ligase from Escherichia coli, respectively. By analyzing polymerase chain reaction (PCR) fragments derived from the genomic DNA of strain rec30, the mutation site in the strain was identified as a single G:C to A:T transition which causes an amino acid substitution at position 224 (Gly to Ser) of the deinococcal RecA protein. Furthermore, we succeeded in expressing both the wild-type and mutant recA genes of D. radiodurans in E. coli without any obvious toxicity or death. The gamma-ray resistance of an E. coli recA1 strain was fully restored by the expression of the wild-type recA gene of D. radiodurans that was cloned in an E. coli vector plasmid. This result is consistent with evidence that RecA proteins from many bacterial species can functionally complement E. coli recA mutants. In contrast with the wild-type gene, the mutant recA gene derived from strain rec30 did not complement E. coli recA1, suggesting that the mutant RecA protein lacks functional activity for recombinational repair.  相似文献   

8.
A recombinant plasmid carrying the recA gene of Leptospira biflexa serovar patoc was isolated from a cosmid library of genomic DNA by complementation of an Escherichia coli recA mutation. The cloned serovar patoc recA gene efficiently restored resistance to UV radiation and methyl methanesulfonate. Recombination proficiency was also restored, as measured by the formation of Lac+ recombinants from duplicated mutant lacZ genes. Additionally, the cloned recA gene increased the spontaneous and mitomycin C-induced production of lambda phage in lysogens of an E. coli recA mutant. The product of the cloned recA gene was identified in maxicells as a polypeptide with an Mr of 43,000. Antibodies prepared against the E. coli RecA protein cross-reacted with the serovar patoc RecA protein, indicating structural conservation. Southern hybridization data showed that the serovar patoc recA gene has diverged from the recA gene of L. interrogans, Leptonema illini, and E. coli. With the exception of the RecA protein of L. interrogans serovar hardjo, the RecA protein of the Leptospira serovars and L. illini were synthesized at elevated levels following treatment of cells with nalidixic acid. The level of detectable RecA correlated with previous studies demonstrating that free-living cells of L. biflexa serovars and L. illini were considerably more resistant to DNA-damaging agents than were those of parasitic L. interrogans serovars. RecA protein was not detected in cells of virulent Treponema pallidum or Borrelia burgdorferi.  相似文献   

9.
A recombinant plasmid carrying the recA gene of Leptospira biflexa serovar patoc was isolated from a cosmid library of genomic DNA by complementation of an Escherichia coli recA mutation. The cloned serovar patoc recA gene efficiently restored resistance to UV radiation and methyl methanesulfonate. Recombination proficiency was also restored, as measured by the formation of Lac+ recombinants from duplicated mutant lacZ genes. Additionally, the cloned recA gene increased the spontaneous and mitomycin C-induced production of lambda phage in lysogens of an E. coli recA mutant. The product of the cloned recA gene was identified in maxicells as a polypeptide with an Mr of 43,000. Antibodies prepared against the E. coli RecA protein cross-reacted with the serovar patoc RecA protein, indicating structural conservation. Southern hybridization data showed that the serovar patoc recA gene has diverged from the recA gene of L. interrogans, Leptonema illini, and E. coli. With the exception of the RecA protein of L. interrogans serovar hardjo, the RecA protein of the Leptospira serovars and L. illini were synthesized at elevated levels following treatment of cells with nalidixic acid. The level of detectable RecA correlated with previous studies demonstrating that free-living cells of L. biflexa serovars and L. illini were considerably more resistant to DNA-damaging agents than were those of parasitic L. interrogans serovars. RecA protein was not detected in cells of virulent Treponema pallidum or Borrelia burgdorferi.  相似文献   

10.
We cloned the gene (c1) which encodes the repressor of vegetative function of Pseudomonas aeruginosa bacteriophage D3. The cloned gene was shown to inhibit plating of D3 and the induction of D3 lysogens by UV irradiation. The efficiency of plating and prophage induction of the heteroimmune P. aeruginosa phage F116L were not affected by the presence of the cloned c1 gene of D3. When the D3 DNA fragment containing c1 was subcloned into pBR322 and introduced into Escherichia coli, it was shown to specifically inhibit the plating of phage lambda and the induction of the lambda prophage by mitomycin C. The plating of lambda imm434 phage was not affected. Analysis in minicells indicated that these effects correspond to the presence of a plasmid-encoded protein of 36,000 molecular weight. These data suggest the possibility that coliphage lambda and the P. aeruginosa phage D3 evolved from a common ancestor. The conservation of the functional similarities of their repressors may have occurred because of the advantage to these temperate phages of capitalizing on the potential of the evolutionarily conserved RecA protein to monitor the level of damage to the host genome.  相似文献   

11.
Functions of the Borrelia burgdorferi RecA protein were investigated in Escherichia coli recA null mutants. Complementation with B. burgdorferi recA increased survival of E. coli recA mutants by 3 orders of magnitude at a UV dose of 2,000 microJ/cm(2). The viability at this UV dose was about 10% that provided by the homologous recA gene. Expression of B. burgdorferi recA resulted in survival of E. coli at levels of mitomycin C that were lethal to noncomplemented hosts. B. burgdorferi RecA was as effective as E. coli RecA in mediating homologous recombination in E. coli. Furthermore, E. coli lambda phage lysogens complemented with B. burgdorferi recA produced phage even in the absence of UV irradiation. The level of phage induction was 55-fold higher than the level in cells complemented with the homologous recA gene, suggesting that B. burgdorferi RecA may possess an enhanced coprotease activity. This study indicates that B. burgdorferi RecA mediates the same functions in E. coli as the homologous E. coli protein mediates. However, the rapid loss of viability and the absence of induction in recA expression after UV irradiation in B. burgdorferi suggest that recA is not involved in the repair of UV-induced damage in B. burgdorferi. The primary role of RecA in B. burgdorferi is likely to be a role in some aspect of recombination.  相似文献   

12.
13.
To evaluate the importance of RecA in DNA double-strand break (DSB) repair, we examined the effect of low and high RecA concentrations such as 2500 and 100 000 molecules per cell expressed from the inducible Pspac promoter in Deinococcus radiodurans in absence or in presence of IPTG respectively. We showed that at low concentration, RecA has a negligible effect on cell survival after gamma-irradiation when bacteria were immediately plated on TGY agar whereas it significantly decreased the survival to gamma-irradiation of DeltaddrA cells while overexpression of RecA can partially compensate the loss of DdrA protein. In contrast, when cells expressing limited concentration of RecA were allowed to recover in TGY2X liquid medium, they showed a delay in mending DSB, failed to reinitiate DNA replication and were committed to die during incubation. A deletion of irrE resulted in sensitivity to gamma-irradiation and mitomycin C treatment. Interestingly, constitutive high expression of RecA compensates partially the DeltairrE sensitization to mitomycin C. The cells with low RecA content also failed to cleave LexA after DNA damage. However, neither a deletion of the lexA gene nor the expression of a non-cleavable LexA(Ind-) mutant protein had an effect on survival or kinetics of DNA DSB repair compared with their lexA+ counterparts in recA+ as well as in bacteria expressing limiting concentration of RecA, suggesting an absence of relationship between the absence of LexA cleavage and the loss of viability or the delay in the kinetics of DSB repair. Thus, LexA protein seems to play no major role in the recovery processes after gamma-irradiation in D. radiodurans.  相似文献   

14.
Abstract Degenerate PCR primers based on conserved RecA protein regions were used to amplify a portion of recE from Prevotella ruminicola strain 23, which was used as a probe to isolate the full-length recA gene from the P. ruminicola genomic library. The P. ruminicola recA gene encoded a protein of 340 amino acids with a molecular mass of 36.81 kDa. P. ruminicola RecA was highly similar to other RecA proteins and most closely resembled that of Bacteroides fragilis (75% identity). It alleviated the methyl methanesulfonate and mitomycin C sensitivities of Escherichia coli recA mutants, but did not restore the resistance to UV-light irradiation. Mitomycin C treatment of otherwise isogenic E. coli strains showed a higher level of prophage induction in a recA harboring lysogen.  相似文献   

15.
We have studied the levels of recA and umuC protein synthesis in Escherichia coli as a probe for regulatory and mechanistic events involved in mitomycin C mutagenesis. Both RecA and UmuC protein induction were greatly stimulated by mitomycin C in the wild-type strain, reached a peak at about 60 min for the recA gene, and at 90 min for the umuC gene, respectively, and maintained a plateau. The induction was blocked by recA and lexA(Ind-) mutations that conferred no mutagenesis on the cell. Mutation affecting uvrA protein markedly decreased induction of the recA gene as well as the umuC gene by mitomycin C. The results established that UvrA protein is involved in the induction of recA and umuC, and account, at least in part, for the mitomycin C nonmutability of uvrA mutants.  相似文献   

16.
RecA protein, which is essential for genetic recombination in Escherichia coli, was extensively purified from a strain of E. coli which contained the recA gene cloned in a plasmid (Sancar, A., and Rupp, W. D. (1979) Proc. Natl. Acad. Sci. U. S. A. 76, 3144-3148). Using the DNA-dependent ATPase activity of recA protein as an assay, we obtained about 60 mg of purified recA protein from 100 g of cells. Ten micrograms or 1 microgram of the purified protein exhibited only one detectable band with Mr approximately = 40,000 upon sodium dodecyl sulfate-acrylamide gel electrophoresis. More than 99% of the ATPase activity of purified recA protein was dependent on single-stranded DNA. Purified recA protein had no detectable DNase, topoisomerase, or ligase activities. The enzyme was stable for a least a year when stored at 0-4 degrees C. The half-life of the ATPase activity of 25 microM recA protein was 37 min at 51 degrees C. Purified recA protein binds to single-stranded and double-stranded DNA, unwinds duplex DNA by a mechanism that is stimulated by single-stranded DNA or oligonucleotides, and pairs homologous single strands with duplex DNA.  相似文献   

17.
Inhibition of Escherichia coli RecA coprotease activities by DinI.   总被引:2,自引:0,他引:2       下载免费PDF全文
T Yasuda  K Morimatsu  T Horii  T Nagata    H Ohmori 《The EMBO journal》1998,17(11):3207-3216
In Escherichia coli, the SOS response is induced upon DNA damage and results in the enhanced expression of a set of genes involved in DNA repair and other functions. The initial step, self-cleavage of the LexA repressor, is promoted by the RecA protein which is activated upon binding to single-stranded DNA. In this work, induction of the SOS response by the addition of mitomycin C was found to be prevented by overexpression of the dinI gene. dinI is an SOS gene which maps at 24.6 min of the E.coli chromosome and encodes a small protein of 81 amino acids. Immunoblotting analysis with anti-LexA antibodies revealed that LexA did not undergo cleavage in dinI-overexpressed cells after UV irradiation. In addition, the RecA-dependent conversion of UmuD to UmuD' (the active form for mutagenesis) was also inhibited in dinI-overexpressed cells. Conversely, a dinI-deficient mutant showed a slightly faster and more extensive processing of UmuD and hence higher mutability than the wild-type. Finally, we demonstrated, by using an in vitro reaction with purified proteins, that DinI directly inhibits the ability of RecA to mediate self-cleavage of UmuD.  相似文献   

18.
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   

19.
头状链轮丝菌(Streptoverticillium caespitosum)ATCC27422是抗肿瘤药物丝裂霉素C的主要产生菌。为了研究丝裂霉素C抗性的分子机制,实验通过鸟枪法克隆技术,从库中筛选得含有丝裂霉素C抗性基因(mcr)的6.6kb外源片段的克隆子,对此外源片段进行一系列亚克隆,将丝裂霉素C抗性基因定位在3.1kb的片段中。序列分析的结果表明,此3.1kb外源片段中存在一长度为134  相似文献   

20.
RecA protein is considered to be the most important participant in the radiation resistance of Deinococcus radiodurans. However, it is still unclear how RecA contributes to the resistance. In this study, we identified a new recA mutation (recA424) in the DNA-repair deficient mutant strain KI696, the phenotype of which is remarkably different from mutant strain rec30 carrying recA670. The properties of the gene products from the recA mutants were compared. recA424 could not complement the deficiency in Escherichia coli RecA, as found for recA670. In vitro, neither RecA424 nor RecA670 could promote DNA strand exchange under conditions in which wild-type RecA promoted the reaction, indicating that both RecA424 and Rec670 are defective in recombination activity. RecA424 promoted the autocleavage reaction of LexA in vitro, whereas RecA670 did not. The intracellular LexA level in KI696 was decreased following gamma-irradiation. However, the LexA level in strain rec30 was constant irrespective of irradiation. These results indicate that RecA424 retains co-protease activity, whereas RecA670 does not. While strain rec30 is extremely radiation sensitive, strain KI696 is only slightly sensitive. Together, these observations suggest that the co-protease activity rather than the recombination activity of RecA contributes to radiation resistance in D. radiodurans.  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