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.
The recA genes of Proteus vulgaris, Erwinia carotovora, Shigella flexneri and Escherichia coli B/r have been isolated and introduced into Escherichia coli K-12. All the heterologous genes restore resistance to killing by UV irradiation and the mutagen 4-nitroquinoline-1-oxide in RecA- E. coli K-12 hosts. Recombination proficiency is also restored as measured by formation of Lac+ recombinants from duplicated mutant lacZ genes and the ability to propagate phage lambda derivatives requiring host recombination functions for growth (Fec-). The cloned heterologous genes increase the spontaneous induction of lambda prophage in lysogens of a recA strain. Addition of mitomycin C stimulates phage production in cells carrying the E. coli B/r and S. flexneri recA genes, but little or no stimulation is seen in cells carrying the E. carotovora and P. vulgaris recA genes. After treatment with nalidixic acid, the heterologous RecA proteins are synthesized at elevated levels, a result consistent with their regulation by the E. coli K-12 LexA repressor. Southern hybridization and preliminary restriction analysis indicate divergence among the coding sequences, but antibodies prepared against the E. coli K-12 RecA protein cross-react with the heterologous enzymes, indicating structural conservation among these proteins.  相似文献   

2.
Summary The complete nucleotide sequences of therecA genes fromEscherichia coli B/r,Shigella flexneri, Erwinia carotovora andProteus vulgaris were determined. The DNA sequence of the coding region of theE. coli B/r gene contained a single nucleotide change compared with theE. coli K12 gene sequence whereas theS. flexneri gene differed at 7 residues. In both cases, the predicted proteins were identical in primary structure to theE. coli K12 RecA protein. The DNA sequences of the recA genes fromE. carotovora andP. vulgaris were 80% and 74% homologous, respectively, to theE. coli K12 gene. The predicted amino acid sequences of theE. carotovora andP. vulgaris RecA proteins were 91% and 85% identical respectively, to that ofE. coli K12. The RecA proteins from bothP. vulgaris andE. carotovora diverged significantly in sequence in the last 50 residues whereas they showed striking conservation throughout the first 300 amino acids which include an ATP-binding region and a subunit interaction domain. A putative LexA repressor binding site was localized upstream of each of the heterologous genes.  相似文献   

3.
The RecA protein of Deinococcus radiodurans (RecA(Dr)) is essential for the extreme radiation resistance of this organism. The RecA(Dr) protein has been cloned and expressed in Escherichia coli and purified from this host. In some respects, the RecA(Dr) protein and the E. coli RecA (RecA(Ec)) proteins are close functional homologues. RecA(Dr) forms filaments on single-stranded DNA (ssDNA) that are similar to those formed by the RecA(Ec). The RecA(Dr) protein hydrolyzes ATP and dATP and promotes DNA strand exchange reactions. DNA strand exchange is greatly facilitated by the E. coli SSB protein. As is the case with the E. coli RecA protein, the use of dATP as a cofactor permits more facile displacement of bound SSB protein from ssDNA. However, there are important differences as well. The RecA(Dr) protein promotes ATP- and dATP-dependent reactions with distinctly different pH profiles. Although dATP is hydrolyzed at approximately the same rate at pHs 7.5 and 8.1, dATP supports an efficient DNA strand exchange only at pH 8.1. At both pHs, ATP supports efficient DNA strand exchange through heterologous insertions but dATP does not. Thus, dATP enhances the binding of RecA(Dr) protein to ssDNA and the displacement of ssDNA binding protein, but the hydrolysis of dATP is poorly coupled to DNA strand exchange. The RecA(Dr) protein thus may offer new insights into the role of ATP hydrolysis in the DNA strand exchange reactions promoted by the bacterial RecA proteins. In addition, the RecA(Dr) protein binds much better to duplex DNA than the RecA(Ec) protein, binding preferentially to double-stranded DNA (dsDNA) even when ssDNA is present in the solutions. This may be of significance in the pathways for dsDNA break repair in Deinococcus.  相似文献   

4.
5.
We have analyzed the nature of RecA protein-RecA protein interactions using an affinity column prepared by coupling RecA protein to an agarose support. When radiolabeled soluble proteins from Escherichia coli are applied to this column, only the labeled RecA protein from the extract was selectively retained and bound tightly to the affinity column. Efficient binding of purified 35S-labeled RecA protein required Mg2+, and high salt did not interfere with the binding of RecA protein to the column. Complete removal of the bound enzyme from the affinity column required treatment with guanidine HCl (5 M) or urea (8 M). These and other properties suggest that hydrophobic interactions contribute significantly to RecA protein subunit recognition in solution. Using a series of truncated RecA proteins synthesized in vitro, we have obtained evidence that at least some of the sequences involved in protein recognition are localized within the first 90 amino-terminal residues of the protein. Based on the observation that RecA proteins from three heterologous bacteria are specifically retained on the E. coli RecA affinity column, it is likely that this binding domain is highly conserved and is required for interaction and association of RecA protein monomers. Stable ternary complexes of RecA protein and single-stranded DNA were formed in the presence of the nonhydrolyzable ATP analog adenosine 5'-O-(thiotriphosphate) and applied to the affinity columns. Most of the complexes formed with M13 DNA could be eluted in high salt, whereas a substantial fraction of those formed with the oligonucleotide (dT)25-30 remained bound in high salt and were quantitatively eluted with guanidine HCl (5 M). The different binding properties of these RecA protein-DNA complexes likely reflect differences in the availability of a hydrophobic surface on RecA protein when it is bound to long polynucleotides compared to short oligonucleotides.  相似文献   

6.
Introduction of the rol genes of Shigella dysenteriae 1 and Escherichia coli K-12 into Shigella flexneri carrier strains expressing the heterologous S. dysenteriae type 1 lipopolysaccharide resulted in the formation of longer chains of S. dysenteriae 1 O antigen. In bacteria producing both homologous and heterologous O antigen, this resulted in a reduction of the masking of heterologous O antigen by homologous lipopolysaccharide and an increased immune response induced by intraperitoneal immunization of mice by recombinant bacteria. The rol genes of S. dysenteriae 1 and E. coli K-12 were sequenced, and their gene products were compared with the S. flexneri Rol protein. The primary sequence of S. flexneri Rol differs from both E. coli K-12 and S. dysenteriae 1 Rol proteins only at positions 267 and 270, which suggests that this region may be responsible for the difference in biological activities.  相似文献   

7.
The outer membranes of many gram-negative bacteria contain a major heat-modifiable protein which shows serological cross-reactivity with the OmpA protein of Escherichia coli K-12. Using the cloned gene for the E. coli K12 protein as a DNA-DNA hybridization probe, we were able to identify the corresponding genes from Shigella dysenteriae. Enterobacter aerogenes, and Serratia marcescens. These were cloned in a phage lambda vector, and their expression in E. coli K-12 was studied. All three OmpA proteins were fully produced and correctly exported to the outer membrane. In several cases, complete or partial restoration of known function of the E. coli K-12 protein was observed.  相似文献   

8.
Radioactive elongation factor Tu coded by either the tufA or the tufB gene of Escherichia coli K-12 was isolated from cells incubated with a mixture of radioactive amino acids after infection with the defective lambda phage particles that carry either of these genes. Two-dimensional chromatographic analyses of tryptic digests of the tufB gene product revealed about 50 radioactive spots. These same spots plus an additional one were also found in tryptic digests of the tufA gene product. Furthermore, these peptide maps are qualitatively the same as those of the elongation factor Tu obtained from two separate isolates of uninfected E. coli K-12 or from rel+ and relA strains of E. coli B. Because the number of spots recovered is consistent with the number of trypsin-sensitive sites, these analyses indicate that the tufA and tufB genes have not significantly diverged from each other.  相似文献   

9.
Survival and induction of the SOS system by 5-azacytidine, an analog of cytidine, were studied in Escherichia coli K-12. This compound did not produce any effect on the viability of dcm and dam dcm mutants. Furthermore, recA430 and lexA1 strains (both mutations interfere with LexA repressor cleavage but not recombination proficiency) were more resistant than the wild-type strain of E. coli K-12. In contrast, recBC and recA13 mutants were more sensitive to 5-azacytidine than the wild type. Transient exposure of E. coli to 5-azacytidine for 60 min induced both recA-dependent inhibition of cell division and induction of lambda prophage in Dcm+ strains but not in Dcm- mutants. Expression of both functions was dependent on recBC exonuclease. On the other hand, 5-azacytidine was unable to trigger the induction of umuCD and mucB genes and no amplification of RecA protein synthesis in either Dcm+ or Dcm- strains was observed. These last results are in agreement with previously reported data suggesting that there is a discrimination in the expression of the several SOS functions and that some SOS genes may be induced without amplification of RecA protein synthesis.  相似文献   

10.
The strict human pathogen Neisseria gonorrhoeae is the only causative agent of the sexually transmitted infection gonorrhea. The recA gene from N. gonorrhoeae is essential for DNA repair, natural DNA transformation, and pilin antigenic variation, all processes that are important for the pathogenesis and persistence of N. gonorrhoeae in the human population. To understand the biochemical features of N. gonorrhoeae RecA (RecA(Ng)), we overexpressed and purified the RecA(Ng) and SSB(Ng) proteins and compared their activities to those of the well-characterized E. coli RecA and SSB proteins in vitro. We observed that RecA(Ng) promoted more strand exchange at early time points than RecA(Ec) through DNA homologous substrates, and exhibited the highest ATPase activity of any RecA protein characterized to date. Further analysis of this robust ATPase activity revealed that RecA(Ng) is more efficient at displacing SSB from ssDNA and that RecA(Ng) shows higher ATPase activity during strand exchange than RecA(Ec). Using substrates created to mimic the cellular processes of DNA transformation and pilin antigenic variation we observed that RecA(Ec) catalyzed more strand exchange through a 100 bp heterologous insert, but that RecA(Ng) catalyzed more strand exchange through regions of microheterology. Together, these data suggest that the processes of ATP hydrolysis and DNA strand exchange may be coupled differently in RecA(Ng) than in RecA(Ec). This difference may explain the unusually high ATPase activity observed for RecA(Ng) with the strand exchange activity between RecA(Ng) and RecA(Ec) being more similar.  相似文献   

11.
Outer membrane proteins of various strains of Escherichia coli were compared using three different systems of sodium dodecyl sulfate (SDS)-polyacrylamide gel electrophoresis. The outer membranes of E. coli K-12, E. coli B, and E. coli J-5 had distinctive protein compositions. As regards proteins which interact with peptidoglycan, E. coli K-12 contained O-8 and O-9, while E. coli B possessed one protein which migrated to the position of O-9. Although E. coli J-5 possessed two such proteins, O-8' and O-9', their positions on polyacrylamide gel were different from those of O-8 and O-9. Protein O-7, which migrates slightly more slowly than O-8, was found specifically in E. coli K-12. Proteins O-10 and O-11 were found in all strains tested, although the relative amounts were different depending on the strain. Strains of E. coli K-12 and E. coli J-5 gave three major bands, O-2a, O-2b, and O-3, in the region of high molecular weight. These proteins were repressed by iron in the cultivation media. Strains of E. coli B, on the other hand, gave only O-2b and O-3. E. coli J-5 gave two other major bands in this region, but the amounts were not controlled by iron in the cultivation media.  相似文献   

12.
A plasmid containing the human HSP70 gene was used to transfect and express the protein in Escherichia coli. The bacterial product was a fusion protein containing 640 amino acids of HSP70, plus 33 additional NH2 terminal amino acids; 12 from the bacterial expression vector and 21 from a 5' human sequence that is not normally translated. It was partially purified by ion-exchange and ATP-Sepharose affinity column chromatography. The bacterially produced human HSP70 protein was then compared with HSP70 obtained from cultured 293 cells. Both shared the same staphylococcal V8 protease peptide fragment pattern, ATP binding, and a weak ATPase activity (about 10-15 nmol ATP hydrolyzed per milligram protein per minute at 30 degrees C). The bacterially produced human HSP70 protein differed in its V8 protease pattern with an E. coli ATP-binding protein that corresponded in molecular mass to the E. coli dnaK gene product. Mutants in the human HSP70 gene were constructed which significantly reduced a predicted major alpha-helical domain in the HSP70 molecule that has partial homology to an ATP-binding site of several protein kinases. One HSP70 mutant clone contained a deletion of 20% at the NH2 terminus, and expressed a 57-kDa product, while the other was missing the middle 50% of the gene (40-kDa product). Neither protein fragment bound to an ATP affinity column, suggesting that ATP binding to HSP70 may be conformationally affected by a region about 20% internal to the NH2 terminal end of the molecule. Recently, a similar location of the ATP-binding site has been reported by Milarski and Morimoto (27).  相似文献   

13.
The structural genes of ADPglucose pyrophosphorylase (glgC) and glycogen synthase (glgA) from Salmonella typhimurium LT2 were cloned on a 5.8-kilobase-pair insert in the SalI site of pBR322. A single strand specific radioactive probe containing the N terminus of the Escherichia coli K-12 glgC gene in M13mp8 was used to hybridize against a S. typhimurium genomic library in lambda 1059. DNA from a plaque showing a positive hybridization signal was isolated, subcloned into pBR322, and transformed into E. coli K-12 RR1 and E. coli G6MD3 (a mutant with a deletion of the glg genes). Transformants were stained with iodine for the presence of glycogen. E. coli K-12 RR1 transformants stained dark brown, whereas G6MD3 transformants stained greenish yellow, and they both were shown to contain a 5.8-kilobase-pair insert in the SalI site of pBR322, designated pPL301. Enzyme assays of E. coli K-12 G6MD3 harboring pPL301 restored ADPglucose pyrophosphorylase and glycogen synthase activities. The specific activities of ADPglucose pyrophosphorylase and glycogen synthase in E. coli K-12 RR1(pPL301) were increased 6- to 7-fold and 13- to 15-fold, respectively. Immunological and kinetic studies showed that the expressed ADPglucose pyrophosphorylase activity in transformed E. coli K-12 G6MD3 cells was very similar to that of the wild-type enzyme.  相似文献   

14.
15.
The periplasmic histidine transport system (permease) of Escherichia coli and Salmonella typhimurium is composed of a soluble, histidine-binding receptor located in the periplasm and a complex of three membrane-bound proteins of which one, HisP, was shown previously to bind ATP. These permeases are energized by ATP. HisP is a member of a family of membrane transport proteins which is conserved in all periplasmic permeases and is presumed to be involved in coupling the energy of ATP to periplasmic transport. In this paper the nature of the ATP-binding site of HisP has been explored by identification of some of the residues that come into contact with ATP. HisP was derivatized with 8-azido-ATP (N3ATP). Both the underivatized and the derivatized forms of HisP were solubilized, purified, and digested with trypsin. The resulting tryptic peptides were resolved by high pressure liquid chromatography, and peptides modified by N3ATP were isolated and sequenced. Two peptides, X and Z, spanning amino acid residues 16-23 and 31-45, were found to contain sites of N3ATP attachment at His19 and Ser41, respectively. Both peptides are close to the amino-terminal end of HisP; peptide Z is located in one of the well conserved regions comprising the nucleotide-binding consensus motifs of the energy-coupling components of these permeases. These consensus motifs are found in many purine nucleotide-binding proteins. The relationship between the location of these residues and the overall structure of the ATP-binding site is discussed.  相似文献   

16.
The photoaffinity label 8-azidoadenosine 5'-triphosphate (N3-ATP) was used to covalently modify the recA protein from Escherichia coli within its ATP-binding site. We have previously demonstrated that N3-ATP modification of recA protein is specific for the ATP-binding site and have isolated a unique tryptic peptide (T31), spanning residues 257-280, that contains the exclusive site of attachment of this ATP analog (Knight, K. L., and McEntee, K. (1985) J. Biol. Chem. 260, 867-872). We performed a secondary proteolytic digestion of the [alpha-32P]N3-ATP-labeled T31 peptide using Staphylococcus aureus V8 protease and purified the resulting peptide fragments by high-pressure liquid chromatography (HPLC). Based on a comparison of the amino acid compositions of all purified fragments and sequence analysis of one labeled fragment we determined that Tyr-264 is the exclusive site of N3-ATP attachment in recA protein. Photoaffinity labeling of recA protein was also performed in the presence of single-stranded DNA. Following trypsin treatment and separation of peptides by HPLC we showed that tryptic peptide T31 contained the exclusive site of N3-ATP attachment. A secondary proteolytic digestion was performed on both [alpha-32P]N3ATP-modified T31 and unmodified T31 using alpha-chymotrypsin. Comparison of the HPLC profiles and amino acid compositions of the resulting fragments was consistent with Tyr-264 as the exclusive site of N3-ATP attachment to recA protein.  相似文献   

17.
The time course of the intracellular ATP concentration in several UV-irradiated RecA protease constitutive (Cptc) mutants of E. coli has been studied. All Cptc mutants harboring a mutation in region 3 of the RecA protein (including amino acid residues 298-301) increased ATP after UV damage but without any subsequent decrease. Nevertheless, these mutants induced the SOS response after UV irradiation. Likewise, truncated RecA proteins lacking region 3 are also unable to carry out massive ATP hydrolysis in UV-irradiated cells. On the other hand, mutants in region 1 (including amino acids 25-39) or 2 (amino acids 157-184) of the RecA protein showed an increase in ATP concentration during the first 20 min following UV irradiation, which dropped afterwards to the basal level. All these data indicate that region 3 of the RecA protein must be involved in the ATP hydrolysis process. Furthermore, a relationship between the quantity of the UV-mediated ATP produced and the strength of the different RecA Cptc mutants has also been found. Accordingly, both lexA71::Tn5 and null lexA mutants of E. coli only show a cellular ATP increase after UV irradiation when containing a multicopy plasmid carrying either a wild-type lexA or a lexA (Ind-) gene.  相似文献   

18.
The fec region of the Escherichia coli chromosome determines a citrate-dependent iron(III) transport system. The nucleotide sequence of fec revealed five genes, fecABCDE, which are transcribed from fecA to fecE. The fecA gene encodes a previously described outer membrane receptor protein. The fecB gene product is formed as a precursor protein with a signal peptide of 21 amino acids; the mature form, with a molecular weight of 30,815, was previously found in the periplasm. The fecB genes of E. coli B and E. coli K-12 differed in 3 nucleotides, of which 2 gave rise to conservative amino acid exchanges. The fecC and fecD genes were found to encode very hydrophobic polypeptides with molecular weights of 35,367 and 34,148, respectively, both of which are localized in the cytoplasmic membrane. The fecE product was a rather hydrophilic but cytoplasmic membrane-bound protein of Mr 28,189 and contained regions of extensive homology to ATP-binding proteins. The number, structural characteristics, and locations of the FecBCDE proteins were typical for a periplasmic-binding-protein-dependent transport system. It is proposed that after FecA- and TonB-dependent transport of iron(III) dicitrate across the outer membrane, uptake through the cytoplasmic membrane follows the binding-protein-dependent transport mechanism. FecC and FecD exhibited homologies to each other, to the N- and C-terminal halves of FhuB of the iron(III) hydroxamate transport system, and to BtuC of the vitamin B12 transport system. FecB showed some homology to FhuD, suggesting that the latter may function in the same manner as a binding protein in iron(III) hydroxamate transport. The close homology between the proteins of the two iron transport systems and of the vitamin B12 transport system indicates a common evolution for all three systems.  相似文献   

19.
The RecA residues Lys248 and Glu96 are closely opposed across the RecA subunit-subunit interface in some recent models of the RecA nucleoprotein filament. The K248R and E96D single mutant proteins of the Escherichia coli RecA protein each bind to DNA and form nucleoprotein filaments but do not hydrolyze ATP or dATP. A mixture of K248R and E96D single mutant proteins restores dATP hydrolysis to 25% of the wild type rate, with maximum restoration seen when the proteins are present in a 1:1 ratio. The K248R/E96D double mutant RecA protein also hydrolyzes ATP and dATP at rates up to 10-fold higher than either single mutant, although at a reduced rate compared with the wild type protein. Thus, the K248R mutation partially complements the inactive E96D mutation and vice versa. The complementation is not sufficient to allow DNA strand exchange. The K248R and E96D mutations originate from opposite sides of the subunit-subunit interface. The functional complementation suggests that Lys248 plays a significant role in ATP hydrolysis in trans across the subunit-subunit interface in the RecA nucleoprotein filament. This could be part of a mechanism for the long range coordination of hydrolytic cycles between subunits within the RecA filament.  相似文献   

20.
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   

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