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.
A recombinant plasmid, pSM2513, containing an 8.5 kb DNA insert was isolated from a genomic library of Serratia marcescens by using interspecific complementation. This plasmid conferred resistance to methyl methanesulphonate and UV irradiation upon recA mutants of Escherichia coli and enhanced recombination proficiency, as measured by Hfr-mediated conjugation, in recA mutants of E. coli. Furthermore, when recA mutants of E. coli harbouring pSM2513 were subjected to UV irradiation, filamentation of the cells was observed. This did not occur upon UV irradiation of the same mutants harbouring the cloning vector alone. These results imply that the S. marcescens recA gene on pSM2513 is functionally similar to the E. coli recA gene in several respects. Restriction enzyme analysis and subcloning studies revealed that the S. marcescens recA gene was located on a 2.7 kb Bg/II-KpnI fragment of pSM2513, and its gene product of approximately 39 kDa resembled the E. coli RecA protein in molecular mass. Using transformation-mediated marker rescue, a recA mutant of S. marcescens was successfully constructed; its proficiency both in homologous recombination and in DNA repair was abolished compared with its parent.  相似文献   

2.
Although previously reported attempts to construct recA null mutants in Streptomyces spp. have been unsuccessful, we have used the suicide plasmid pErmdeltaRecA to inactivate the recA gene in Streptomyces rimosus by gene disruption. pErmdeltaRecA carries the erythromycin resistance gene ermE and a 451-bp fragment of the S. rimosus recA gene (encoding amino acids 2-151). An erythromycin-resistant clone with single plasmid integration into the recA gene on the chromosome was analyzed in detail. This clone possesses one inactive copy of recA which lacks the entire promoter region and the ATG start codon, and a second, truncated gene that encodes only first 151 amino acids of the RecA protein. This S. rimiosus rec A mutant can therefore be considered a completely RecA-deficient strain. The mutant strain is highly sensitive to UV light. Introduction of a plasmid carrying the wild type S. rimosus recA gene completely restored the UV resistance of the recA mutant to wild-type levels. recA genes encoding RecA proteins with short deletions at the C-terminus (21 and 51 amino acids) could not fully rescue the UV sensitivity of the S. rimosus recA strain, when introduced in the same way.  相似文献   

3.
Temperature-sensitive integration plasmids carrying internal fragments of the Streptomyces lividans TK24 recA gene were constructed and used to inactivate the chromosomal recA gene of S. lividans by gene disruption and gene replacement. Integration of these plasmids resulted in recA mutants expressing C-terminally truncated RecA proteins, as deduced from Southern hybridization experiments. Mutants FRECD2 in which the last 42 amino acids, comprising the variable part of bacterial RecA proteins, had been deleted retained the wild-type phenotype. The S. lividans recA mutant FRECD3 produced a RecA protein lacking 87 amino acids probably including the interfilament contact site. FRECD3 was more sensitive to UV and MMS than the wild-type. Its ability to undergo homologous recombination was impaired, but not completely abolished. Integration of the disruption plasmid pFRECD3 in S. coelicolor“Müller” caused the same mutant phenotype as S. lividans FRECD3. In spite of many attempts no S. lividans recA mutants with deletions of 165 C-terminal amino acids or more were isolated. Furthermore, the recA gene could not be replaced by a kanamycin resistance cassette. These experiments indicate a crucial role of the recA gene in ensuring viability of Streptomyces.  相似文献   

4.
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   

5.
6.
7.
The recombinase RecA plays a crucial role in homologous recombination and the SOS response in bacteria. Although recA mutants usually are defective in homologous recombination and grow poorly, they nevertheless can be isolated in almost all bacteria. Previously, considerable difficulties were experienced by several laboratories in generating recA null mutations in Streptomyces, and the only recA null mutants isolated (from Streptomyces lividans) appeared to be accompanied by a suppressing mutation. Using gene replacement mediated by Escherichia coli-Streptomyces conjugation, we generated recA null mutations in a series of Streptomyces coelicolor A3(2) strains. These recA mutants were very sensitive to mitomycin C but only moderately sensitive to UV irradiation, and the UV survival curves showed wide shoulders, reflecting the presence of a recA-independent repair pathway. The mutants segregated minute colonies with low viability during growth and produced more anucleate spores than the wild type. Some crosses between pairs of recA null mutants generated no detectable recombinants, showing for the first time that conjugal recombination in S. coelicolor is recA mediated, but other mutants retained the ability to undergo recombination. The nature of this novel recombination activity is unknown.  相似文献   

8.
Temperature-sensitive integration plasmids carrying internal fragments of the Streptomyces lividans TK24 recA gene were constructed and used to inactivate the chromosomal recA gene of S. lividans by gene disruption and gene replacement. Integration of these plasmids resulted in recA mutants expressing C-terminally truncated RecA proteins, as deduced from Southern hybridization experiments. Mutants FRECD2 in which the last 42 amino acids, comprising the variable part of bacterial RecA proteins, had been deleted retained the wild-type phenotype. The S. lividans recA mutant FRECD3 produced a RecA protein lacking 87 amino acids probably including the interfilament contact site. FRECD3 was more sensitive to UV and MMS than the wild-type. Its ability to undergo homologous recombination was impaired, but not completely abolished. Integration of the disruption plasmid pFRECD3 in S. coelicolor“Müller” caused the same mutant phenotype as S. lividans FRECD3. In spite of many attempts no S. lividans recA mutants with deletions of 165 C-terminal amino acids or more were isolated. Furthermore, the recA gene could not be replaced by a kanamycin resistance cassette. These experiments indicate a crucial role of the recA gene in ensuring viability of Streptomyces. Received: 20 December 1996 / Accepted: 25 March 1997  相似文献   

9.
In Escherichia coli, constituents of the main recombination pathway are provided by the genes recA (RecA protein) and recBCD (RecBCD enzyme). Recombination in conjugation experiments and repair of UV damage of E. coli mutants deleted for recA, for recBCD or for recA plus recBCD were restored, although to different degrees, by the cloned recA and recBCD genes from Serratia marcescens or Proteus mirabilis. When both recombination enzymes were from the same species, repair and recombination efficiencies had the order E. coli greater than S. marcescens greater than P. mirabilis. However, the P. mirabilis recA plus recBCD genes resulted in higher levels of repair and recombination than those obtained with one component from P. mirabilis (recA or recBCD) and the other from E. coli or S. marcescens. The data provide evidence for the similarity of RecABCD pathways of recombination among enteric bacteria and suggest an in vivo advantage of an intraspecies combination of P. mirabilis RecA protein and RecBCD enzyme over interspecies combinations. This could point to a cooperation between these basic recombination enzymes. The molecular processes which could be involved are discussed.  相似文献   

10.
Abstract The recA gene of Chlamydia trachomatis was isolated by complementation of an Escherichia coli recA mutant. The cloned gene restored resistance to methyl methanesulfonate in E. coli recA mutants. The DNA sequence of the chlamydial gene was determined and the deduced protein sequence compared with other RecA proteins. In E. coli recA deletion mutants, the cloned gene conferred moderate recombinational activity as assayed by Hfr matings. The chlamydial recA gene was efficient in repairing alkylated DNA but less so in repairing of UV damage when compared with the E. coli homologue. As detected by an SOS gene fusion, a small but measurable amount of LexA co-cleavage was indicated.  相似文献   

11.
A recombinant plasmid carrying the recA gene of Aeromonas caviae was isolated from an A. caviae genomic library by complementation of an Escherichia coli recA mutant. The plasmid restored resistance to both UV irradiation and to the DNA-damaging agent methyl methanesulfonate in the E. coli recA mutant strain. The cloned gene also restored recombination proficiency as measured by the formation of lac+ recombinants from duplicated mutant lacZ genes and by the ability to propagate a strain of phage lambda (red gam) that requires host recombination functions for growth. The approximate location of the recA gene on the cloned DNA fragment was determined by constructing deletions and by the insertion of Tn5, both of which abolished the ability of the recombinant plasmid to complement the E. coli recA strains. A. caviae recA::Tn5 was introduced into A. caviae by P1 transduction. The resulting A. caviae recA mutant strain was considerably more sensitive to UV light than was its parent. Southern hybridization analysis indicated that the A. caviae recA gene has diverged from the recA genes from a variety of gram-negative bacteria, including A. hydrophila and A. sobria. Maxicell labeling experiments revealed that the RecA protein of A. caviae had an Mr of about 39,400.  相似文献   

12.
Streptomyces lividans TK23 gives rise to chloramphenicol-sensitive (Cml(s)) mutants at a frequency of about 0.5%. This is due to the frequent occurrence of very large chromosomal deletions removing the corresponding chloramphenicol resistance gene. A mutant in which the recA gene has been disrupted (S. lividans FrecD3 [G. Muth, D. Frese, A. Kleber, and W. Wohlleben, personal communication]) segregated about 70 times more chloramphenicol-sensitive mutants than the parental strain. An enhancement of the deletion frequency was responsible for this mutator phenotype. The amplifiable locus AUD1 has a duplicated structure in some S. lividans strains and is frequently highly amplified in some mutants generated by genetic instability. The chromosomal AUD1 is not amplified in strain TK23 because of the lack of one duplication. Nevertheless, AUD1-derived amplifiable units presenting the typical duplicated organization amplified very well in TK23 when carried on a plasmid. No amplification of these units was observed in the recA mutant. The ability to amplify was restored when the wild-type recA gene was introduced into the plasmid carrying the amplifiable unit. These results suggest that the RecA protein plays a role in reducing the level of genetic instability and chromosomal deletions and show that the recA gene is necessary to achieve high-copy-number amplification of AUD1.  相似文献   

13.
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   

14.
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   

15.
A recombinant plasmid carrying the recA gene of Vibrio cholerae was isolated from a V. cholerae genomic library, using complementation in Escherichia coli. The plasmid complements a recA mutation in E. coli for both resistance to the DNA-damaging agent methyl methanesulfonate and recombinational activity in bacteriophage P1 transductions. After determining the approximate location of the recA gene on the cloned DNA fragment, we constructed a defined recA mutation by filling in an XbaI site located within the gene. The 4-base pair insertion resulted in a truncated RecA protein as determined by minicell analysis. The mutation was spontaneously recombined onto the chromosome of a derivative of V. cholerae strain P27459 by screening for methyl methanesulfonate-sensitive variants. Southern blot analysis confirmed the presence of the inactivated XbaI site in the chromosome of DNA isolated from one of these methyl methanesulfonate-sensitive colonies. The recA V. cholerae strain was considerably more sensitive to UV light than its parent, was impaired in homologous recombination, and was deficient in induction of a temperate vibriophage upon exposure to UV light. We conclude that the V. cholerae RecA protein has activities which are analogous to those described for the RecA protein of E. coli.  相似文献   

16.
The rec-1 gene of Haemophilus influenzae was cloned into a shuttle vector that replicates in Escherichia coli as well as in H. influenzae. The plasmid, called pRec1, complemented the defects of a rec-1 mutant in repair of UV damage, transformation, and ability of prophage to be induced by UV radiation. Although UV resistance and recombination were caused by pRec1 in E. coli recA mutants, UV induction of lambda and UV mutagenesis were not. We suggest that the ability of the H. influenzae Rec-1 protein to cause cleavage of repressors but not the recombinase function differs from that of the E. coli RecA protein.  相似文献   

17.
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   

18.
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   

19.
The argC gene of Streptomyces clavuligerus encoding N-acetylglutamyl-phosphate reductase (AGPR) has been cloned by complementation of argC mutants Streptomyces lividans 1674 and Escherichia coli XC33. The gene is contained in an open reading frame of 1,023 nucleotides which encodes a protein of 340 amino acids with a deduced molecular mass of 35,224 Da. The argC gene is linked to argE, as shown by complementation of argE mutants of E. coli. Expression of argC from cloned DNA fragments carrying the gene leads to high levels of AGPR in wild-type S. lividans and in the argC mutant S. lividans 1674. Formation of AGPR is repressed by addition of arginine to the culture medium. The protein encoded by the argC gene is very similar to the AGPRs of Streptomyces coelicolor, Bacillus subtilis, and E. coli and, to a lesser degree, to the homologous enzymes of Saccharomyces cerevisiae and Anabaena spp. A conserved PGCYPT domain present in all the AGPR sequences suggests that this may be the active center of the protein. Transformation of S. clavuligerus 328, an argC auxotroph deficient in clavulanic acid biosynthesis, with plasmid pULML30, carrying the cloned argC gene, restored both prototrophy and antibiotic production.  相似文献   

20.
Isolation and characterization of the Vibrio cholerae recA gene   总被引:4,自引:1,他引:3       下载免费PDF全文
A 3.6-kilobase PstI fragment was isolated from a Vibrio cholerae chromosomal DNA library and shown to encode RecA-like activity in complementation studies with Escherichia coli recA mutants. Although DNA hybridization experiments failed to detect any homology between the E. coli and V. cholerae recA genes, hyperimmune antiserum produced against purified E. coli RecA protein recognized epitopes shared by the V. cholerae protein. The V. cholerae chromosomal fragments, when cloned and transferred to E. coli, provided the missing recA functions, including resistance to the alkylating agent methyl methanesulfonate, resistance to UV irradiation, and promotion of homologous recombination in Hfr mating experiments.  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