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 31 毫秒
1.
G R Banks  S G Sedgwick 《Biochemistry》1986,25(20):5882-5889
When the Escherichia coli RecA protein is UV irradiated in the presence of [alpha-32P]ATP, a labeled protein--ATP adduct is formed. All the experimental evidence indicates that, in forming such an adduct, the ATP becomes specifically immobilized in the catalytically relevant ATP binding site. The adduct can also be identified after irradiation of E. coli cell lysates in a similar manner. This direct ATP photolabeling of RecA proteins has been used to identify regions of the polypeptide chain involved in the binding of ATP. The photolabeling of a RecA protein that lacks wild-type carboxy-terminal amino acids is not detectable. A RecA protein in which the amino-terminal sequence NH2-Ala-Ile-Asp-Glu-Asn- is replaced by NH2-Thr-Met-Ile-Thr-Asn-Ser-Ser-Ser- is only about 5% as efficiently photolabeled as the wild-type protein. Both of these RecA protein constructions, however, contain all the elements previously implicated, directly or indirectly, in the binding of ATP. ATP-photolabeled RecA protein has also been chemically cleaved at specific amino acids in order to identify regions of the polypeptide chain to which the nucleotide becomes covalently photolinked. The evidence is consistent with a region comprising amino acids 116-170. Thus, this work and that of others suggest that several disparate regions of the unfolded polypeptide chain may combine to form the ATP binding site upon protein folding or may influence binding through long-range effects.  相似文献   

2.
ATP hydrolysis during SOS induction in Escherichia coli.   总被引:2,自引:1,他引:1  
Changes in cellular ATP concentration during SOS induction in strains of Escherichia coli with different levels of RecA and LexA proteins were studied. UV irradiation of RecA+ strains induced a twofold increase in the ATP concentration around the first 20 min, followed by a decrease to the values of nonirradiated cells. On the other hand, mutants defective in RecA protein or with either deficient RecA protease activity or cleavage-resistant LexA repressor did not show any decrease, suggesting that ATP consumption is related to LexA repressor hydrolysis. Furthermore, strains presenting a constitutive synthesis of RecA protein showed the same changes in ATP concentration as the wild-type strain. Likewise, the presence in a RecA+ strain of a LexA(Def) protein, which is defective in its capacity for binding specifically to SOS operators, did not disturb the changes in ATP when compared with the LexA+ RecA+ strain. Moreover, after UV irradiation, a LexA(Def) RecA- double mutant showed an important increase in ATP concentration, which remained elevated for at least 120 min after UV treatment.  相似文献   

3.
The deficiency in UV mutagenesis in uvrD3 recB21 strains of E. coli is almost completely overcome by constitutive activation of RecA protein and expression of the SOS system (by recA730 or 43 degrees C treated recA441 lexA71). When SOS was expressed but RecA protein not self-activated (recA441 lexA71 at 30 degrees C), uvrD3 recB21 still reduced UV mutagenesis at low doses. The uvrD3 recB21 combination is therefore inhibiting activation of RecA protein. It is suggested that the DNA unwinding activity of the products of the uvrD and recB genes may be involved in generating single-stranded DNA needed to activate RecA protein both for the cleavage of LexA repressor and for a further role in UV mutagenesis.  相似文献   

4.
F Masek  M Skorvaga  M Sedliaková 《Gene》1989,78(1):195-199
Dimer excision was followed in Escherichia coli K-12 AB1157 DM49 lexA3 mutant (whose repressor is not cleavable with RecA protease), and in E. coli K-12 AB2497[pGC3] carrying the cloned lexA gene. In either case din genes could not be efficiently derepressed. In such cells ultraviolet (UV) irradiation caused an extensive DNA degradation, which was not observed in cells with derepressed din genes. Even after a high UV dose (70 J/m2) dimers were being excised efficiently. However, progressive DNA degradation interfered with the precise detection of unexcised dimers. We conclude that induction of din genes is required for filling some of the gaps and for prevention of DNA degradation, but not for excision itself.  相似文献   

5.
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   

6.
Salmonella typhimurium has a SOS regulon which resembles that of Escherichia coli. recA mutants of S. typhimurium have already been isolated, but no mutations in lexA have been described yet. In this work, two different lexA mutants of S. typhimurium LT2 have been constructed on a sulA background to prevent cell death and further characterized. The lexA552 and lexA11 alleles contain an insertion of the kanamycin resistance fragment into the carboxy- and amino-terminal regions of the lexA gene, respectively. SOS induction assays indicated that both lexA mutants exhibited a LexA(Def) phenotype, although SOS genes were apparently more derepressed in the lexA11 mutant than in the lexA552 mutant. Like lexA(Def) of E. coli, both lexA mutations only moderately increased the UV survival of S. typhimurium, and the lexA552 strain was as mutable as the lexA+ strain by UV in the presence of plasmids encoding MucAB or E. coli UmuDC (UmuDCEc). In contrast, a lexA11 strain carrying any of these plasmids was nonmutable by UV. This unexpected behavior was abolished when the lexA11 mutation was complemented in trans by the lexA gene of S. typhimurium. The results of UV mutagenesis correlated well with those of survival to UV irradiation, indicating that MucAB and UmuDCEc proteins participate in the error-prone repair of UV damage in lexA552 but not in lexA11. These intriguing differences between the mutagenic responses of lexA552 and lexA11 mutants to UV irradiation are discussed, taking into account the different degrees to which the SOS response is derepressed in these mutants.  相似文献   

7.
DNA repair mechanisms affecting cytotoxicity by streptozotocin in E. coli   总被引:2,自引:0,他引:2  
Mechanisms underlying cytotoxicity by the monofunctional nitrosourea streptozotocin (STZ) were evaluated in DNA repair-deficient E. coli mutants. Strains not proficient in recombinational repair which lack either RecA protein or RecBC gene products were highly sensitive to STZ. In contrast, cells that constitutively synthesize RecA protein and cannot initiate SOS repair mechanisms because of uncleavable LexA repressor (recAo98 lexA3) were resistant to this drug compared to a lexA3 strain. Further, E. coli cells lacking both 3-methyladenine DNA glycosylases I (tag) and II (alkA) also were highly sensitive to STZ. DNA synthesis was most inhibited by STZ in recA and alkA tag E. coli mutants, but was suppressed less markedly in wild-type and recBC cells. DNA degradation was most extensive in recA E. coli after STZ treatment, while comparable in recBC, alkA tag, and wild-type cells. Although increased single-stranded DNA breaks were present after STZ treatment in recA and recBC mutants compared to the wild type, no significant increase in DNA single-stranded breaks was noted in alkA tag E. coli. Further, DNA breaks in recBC cells were repaired, while those present in recA cells were not. These findings establish the critical importance of both recombinational repair and 3-methyladenine DNA glycosylase in ameliorating cytotoxic effects and DNA damage caused by STZ in E. coli.  相似文献   

8.
Bunny K  Liu J  Roth J 《Journal of bacteriology》2002,184(22):6235-6249
The LexA protein of Escherichia coli represses the damage-inducible SOS regulon, which includes genes for repair of DNA. Surprisingly, lexA null mutations in Salmonella enterica are lethal even with a sulA mutation, which corrects lexA lethality in E. coli. Nine suppressors of lethality isolated in a sulA mutant of S. enterica had lost the Fels-2 prophage, and seven of these (which grew better) had also lost the Gifsy-1 and Gifsy-2 prophages. All three phage genomes included a homologue of the tum gene of coliphage 186, which encodes a LexA-repressed cI antirepressor. The tum homologue of Fels-2 was responsible for lexA lethality and had a LexA-repressed promoter. This basis of lexA lethality was unexpected because the four prophages of S. enterica LT2 are not strongly UV inducible and do not sensitize strains to UV killing. In S. enterica, lexA(Ind(-)) mutants have the same phenotypes as their E. coli counterparts. Although lexA null mutants express their error-prone DinB polymerase constitutively, they are not mutators in either S. enterica or E. coli.  相似文献   

9.
The Escherichia coli SOS response to DNA damage is modulated by the RecA protein, a recombinase that forms an extended filament on single-stranded DNA and hydrolyzes ATP. The RecA K72R ( recA2201 ) mutation eliminates the ATPase activity of RecA protein. The mutation also limits the capacity of RecA to form long filaments in the presence of ATP. Strains with this mutation do not undergo SOS induction in vivo . We have combined the K72R variant of RecA with another mutation, RecA E38K ( recA730 ). In vitro , the double mutant RecA E38K/K72R ( recA730,2201 ) mimics the K72R mutant protein in that it has no ATPase activity. The double mutant protein will form long extended filaments on ssDNA and facilitate LexA cleavage almost as well as wild-type, and do so in the presence of ATP. Unlike recA K72R, the recA E38K/K72R double mutant promotes SOS induction in vivo after UV treatment. Thus, SOS induction does not require ATP hydrolysis by the RecA protein, but does require formation of extended RecA filaments. The RecA E38K/K72R protein represents an improved reagent for studies of the function of ATP hydrolysis by RecA in vivo and in vitro .  相似文献   

10.
11.
C Lesca  C Petit  M Defais 《Biochimie》1991,73(4):407-409
The SOS response is induced in E coli following treatments that interfere with DNA replication. The response is under the control of the recA and the lexA genes. Strains defective in LexA repressor constitutively express SOS proteins. However, SOS repair does not reach its maximum level in these strains. Instead, an activation of RecA protein and de novo protein synthesis are required for full repair. We have analyzed by 2-dimensional gel electrophoresis the induction of proteins after UV irradiation of lexA(Def) bacteria. Proteins which might participate in SOS repair are induced under these conditions.  相似文献   

12.
13.
Lee AM  Singleton SF 《Biochemistry》2006,45(14):4514-4529
The Escherichia coli RecA protein is the prototypical member of a family of molecular motors that transduces ATP binding and hydrolysis for mechanical function. While many general mechanistic features of RecA action are known, specific structural and functional insights into the molecular basis of RecA activation remain elusive. Toward a more complete understanding of the interdependence between ATP and DNA binding by RecA, we report the characterization of a mutant RecA protein wherein the aspartate residue at position 100 within the ATP binding site is replaced by arginine. Physiologically, D100R RecA was characterized by an inducible, albeit reduced, activation of the SOS response and a diminished ability to promote cellular survival after UV irradiation. Biochemically, the D100R substitution caused a surprisingly modest perturbation of RecA-ATP interactions and an unexpected and significant decrease in the affinity of RecA for ssDNA. Moreover, in vitro assays of RecA activities requiring the coordinated processing of ATP and DNA revealed (1) a 2-5-fold decrease in steady-state turnover of ATP; (2) no formation of mixed nucleoprotein filaments when wild-type and D100R RecA compete for limiting ssDNA; and (3) no formation of strand exchange reaction products. Taken together, these results suggest that the D100R mutational effects on isolated RecA activities combine synergistically to perturb its higher-order functions. We conclude that the replacement of Asp100 resulted in a change in the electrostatic complementarity between RecA monomers during active filament assembly that prevents the protein from fully accessing the active multimeric state.  相似文献   

14.
The deduced amino acid sequence of Gluconobacter oxydans RecA protein shows 75.2, 69.4, and 66.2% homology with those from Aquaspirillum magnetotacticum, Escherichia coli, and Pseudomonas aeruginosa, respectively. The amino acid residues essential for function of the recombinase, protease, and ATPase in E. coli recA protein are conserved in G. oxydans. Of 24 amino acid residues believed to be the ATP binding domain of E. coli RecA, 17 are found to be identical in G. oxydans RecA. Interestingly, nucleotide sequence alignment between the SOS box of G. orphans recA gene and those from different microorganisms revealed that all the DNA sequences examined have dyad symmetry that can form a stem-loop structure. A G. oxydans recA-deficient mutant (LCC96) was created by allelic exchange using the cloned recA gene that had been insertionally inactivated by a kanamycin-resistance cassette. Such replacement of the wild-type recA with a kanamycin resistance gene in the chromosome was further verified by Southern hybridization. Phenotypically, the recA-deficient mutant is significantly more sensitive to UV irradiation than the wild-type strain, suggesting that the recA gene of G. oxydans ATCC9324 plays a role in repairing DNA damage caused by UV irradiation. Moreover, the mutant strain is much more plasmid transformable than its parent strain, illustrating that G. oxydans LCC96 could be used as a host to take up the recombinant plasmid for gene manipulation.  相似文献   

15.
Site-directed mutagenesis in the Escherichia coli recA gene   总被引:1,自引:0,他引:1  
C Cazaux  F Larminat  M Defais 《Biochimie》1991,73(2-3):281-284
Escherichia coli RecA protein plays a fundamental role in genetic recombination and in regulation and expression of the SOS response. We have constructed 6 mutants in the recA gene by site-directed mutagenesis, 5 of which were located in the vicinity of the recA430 mutation responsible for a coprotease deficient phenotype and one which was at the Tyr 264 site. We have analysed the capacity of these mutants to accomplish recombination and to express SOS functions. Our results suggest that the region including amino acid 204 and at least 7 amino acids downstream interacts not only with LexA protein but also with ATP. In addition, the mutation at Tyr 264 shows that this amino acid is essential for RecA activities in vivo, probably because of its involvement in an ATP binding site, as previously shown in vitro.  相似文献   

16.
The recA gene of Synechococcus sp. strain PCC 7002 was detected and cloned from a lambda gtwes genomic library by heterologous hybridization by using a gene-internal fragment of the Escherichia coli recA gene as the probe. The gene encodes a 38-kilodalton polypeptide which is antigenically related to the RecA protein of E. coli. The nucleotide sequence of a portion of the gene was determined. The translation of this region was 55% homologous to the E. coli protein; allowances for conservative amino acid replacements yield a homology value of about 74%. The cyanobacterial recA gene product was proficient in restoring homologous recombination and partial resistance to UV irradiation to recA mutants of E. coli. Heterologous hybridization experiments, in which the Synechococcus sp. strain PCC 7002 recA gene was used as the probe, indicate that a homologous gene is probably present in all cyanobacterial strains.  相似文献   

17.
The survival of Escherichia coli following treatment with a low dose (1-3 mM) of hydrogen peroxide (H(2)O(2)) that causes extensive mode-one killing of DNA repair mutants is stimulated by the induction of the SOS regulon. Results for various mutants indicate that induction of recA and RecA protein-mediated recombination are critical factors contributing to the repair of H(2)O(2)-induced oxidative DNA damage. However, because DNA damage activates RecA protein's coprotease activity essential to cleavage of LexA repressor protein and derepression of all SOS genes, it is unclear to what extent induction of RecA protein stimulates this repair. To make this determination, we examined mode-one killing of DeltarecA cells carrying plasmid-borne recA (P(tac)-recA(+)) and constitutively expressing a fully induced level of wild-type RecA protein when SOS genes other than recA are non-inducible in a lexA3 (Ind(-)) genetic background or inducible in a lexA(+) background. At a H(2)O(2) dose resulting in maximal killing, DeltarecA lexA3 (Ind(-)) cells with P(tac)-recA(+) show 40-fold greater survival than lexA3 (Ind(-)) cells with chromosomal recA having a low, non-induced level of RecA protein. However, they still show 10- to 15-fold lower survival than wild-type cells and DeltarecA lexA(+) cells with P(tac)-recA(+). To determine if the inducible RuvA protein stimulates survival, we examined a ruvA60 mutant that is defective for the repair of UV-induced DNA damage. This mutant also shows 10- to 15-fold lower survival than wild-type cells. We conclude that while induction of RecA protein has a pronounced stimulatory effect on the recombinational repair of H(2)O(2)-induced oxidative DNA damage, the induction of other SOS proteins such as RuvA is essential for wild-type repair.  相似文献   

18.
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   

19.
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   

20.
The presence of the plasmid colicinogenic factor Ib-P9 in Escherichia coli wild type cells is shown to increase bacterial survival after UV irradiation and the action of N-methyl-N'-nitro-N-nitrosoguanidine. The ability of the plasmid to cause the UV protection is observed in uvrA, uvrB, uvrC, polA, recB, recF E. coli strains, but the plasmid does not restore the UV resistance of the mutant cells to the wild type level. The protective effect of the plasmid CoI Ib-P9 depends on the recA+lexA+ genotype of the cells. The inhibition of protein synthesis (amino acid starvation) before and after UV irradiation does not prevent the UV protection by ColIb-P9. The nature of the plasmid-associated repair functions is discussed.  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