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 593 毫秒
1.
Organization of the adenyl cyclase (cya) locus of Rhizobium meliloti   总被引:2,自引:0,他引:2  
  相似文献   

2.
3.
4.
S Fanning  F O'Gara 《Gene》1988,71(1):57-64
The Rhizobium meliloti (Rm) lacZ gene provides a convenient model to investigate patterns of gene regulation in these agronomically important bacteria. A gene encoding beta-galactosidase (beta Gal) activity was cloned from R. meliloti by complementing a lactose-negative Escherichia coli mutant. A series of Sau3A subclones was generated in pBR322, and the coding region for the beta Gal-coding gene was localized to a 2.4-kb core fragment. In E. coli 'maxicells', these lacZ subclones produced a 79-kDa polypeptide, irrespective of the fragment size demonstrating that the translation initiation signal(s) are located on the 2.4-kb fragment. Transposon Tn5 mutagenesis and BAL 31 deletion analysis showed that the expression of the Rm lacZ gene in E. coli was dependent on the tetracycline-resistance promoter of pBR322. The cloned sequence was required for beta Gal synthesis in Rhizobium since mutants generated by reverse genetics lack this enzyme and were specifically defective in lactose catabolism.  相似文献   

5.
运用生色基因标记黄瓜根围促生菌(PGPR)筛选菌株   总被引:10,自引:1,他引:9  
采用三亲交配方法 ,通过Tn7转座系统将lacZY标记基因导入黄瓜根围促生菌 (PG PR)筛选菌株PseudomonasfluorescensCN1 1 6和PseudomonascorrugataCN31的利福平抗性突变株中 ;标记假单胞菌菌株则被赋予了利用乳糖作为唯一碳源的能力 ,在只有乳糖的M9培养基上生长能分解X Gal,菌落显出特有的蓝色 ;经Southern杂交分析 ,证明标记基因lacZY存在于转化菌株的染色体上 ;经验证标记菌株标记性状稳定 ,与对应的野生菌株比较其它性状如培养性状、形态特征、生防效果等基本不变 ;PGPR菌株利福平抗性和生色基因标记的结合 (双标记 )能最大限度地将土壤中引入的PGPR菌株与土著细菌分开 ,检测下限可达 1 0CFU mL ,为PGPR在根围的分子生态学研究提供了一个较好的工具。  相似文献   

6.
Serratia entomophila UC9 (A1MO2), which causes amber disease in the New Zealand grass grub Costelytra zealandica, was subjected to transposon (TnphoA)-induced mutagenesis. A mutant (UC21) was found to be nonpathogenic (Path-) to grass grub larvae in bioassays and was shown, by Southern hybridization, to contain a single TnphoA insertion. This mutant failed to adhere to the gut wall (Adn-) of the larvae and also failed to produce pili (Pil-). A comparative study of the total protein profiles of wild-type S. entomophila UC9 and mutant UC21 revealed that the mutant lacked an approximately 44-kDa protein and overexpressed an approximately 20-kDa protein. Transfer of cosmids containing homologous wild-type sequences into mutant strain UC21 restored wild-type phenotypes (Path+, Pil+, and Adn+). One of the complementing cosmids (pSER107) conferred piliation on Pil- Escherichia coli HB101. The TnphoA insertion in UC21 was mapped within an 8.6-kb BamHI fragment common to the complementing cosmids, and we designated this gene locus amb-1. Six gene products with molecular masses of 44, 36, 34, 33, 20, and 18 kDa were detected in E. coli minicells exclusive to the cloned 8.6-kb fragment (pSER201A). The 44-kDa gene product was not detected in E. coli minicells containing the cloned mutant fragment. Saturation mutagenesis of this fragment produced four unlinked insertional mutations with active fusions to TnphoA. These active fusions disrupted the expression of one or more gene products encoded by amb-1. The 8.6-kb fragment cloned in the opposite orientation (pSER201B) expressed only a 20-kDa protein. We propose that these are the products of structural and/or regulatory genes involved in adhesion and/or piliation which are prerequisites in the S. entomophila-grass grub interaction leading to amber disease.  相似文献   

7.
8.
Mutations in several dna genes of Escherichia coli, when introduced into a strain with a lac fusion in the SOS gene sulA, resulted in formation of blue colonies on plates containing 5-bromo-4-chloro-3-indolyl-beta-D-galactoside (X-Gal). Unexpectedly, several lines of evidence indicated that the blue colony color was not primarily due to induction of the SOS system but rather was due to a membrane defect, along with the replication defect, making the cell X-Gal extrasensitive (phenotypically Xgx), possibly because of enhanced permeability to X-Gal or leakage of beta-galactosidase. (i) In most cases, beta-galactosidase specific activity increased only two- to threefold. (ii) Mutations conferring tolerance to colicin E1 resulted in blue colony color with no increase in beta-galactosidase specific activity. (iii) Mutations in either the dnaA, dnaB, dnaC, dnaE, dnaG, or ssb gene, when introduced into a strain containing a bioA::lac fusion, produced a blue colony color without an increase in beta-galactosidase synthesis. These lac fusion strains can serve as signal strains to detect dna mutations as well as membrane mutations. By localized mutagenesis of the 92-min region of the chromosome of the sulA::lac signal strain and picking blue colonies, we isolated a novel ssb allele that confers the same extreme UV sensitivity as a delta recA allele, which is a considerably greater sensitivity than that conferred by the two well-studied ssb alleles, ssb-1 and ssb-113. The technique also yielded dnaB mutants; fortuitously, uvrA mutants were also found.  相似文献   

9.
S Mito  Q M Zhang    S Yonei 《Journal of bacteriology》1993,175(9):2645-2651
Gene fusions in Escherichia coli that showed increased beta-galactosidase expression in response to treatment with a superoxide radical (O2-) generator, methyl viologen (MV), were obtained. These fusions were constructed by using a Mud(Ap lac) phage to insert the lactose structural genes randomly into the E. coli chromosome. Ampicillin-resistant colonies were screened for increased expression of beta-galactosidase on X-Gal (5-bromo-4-chloro-3-indolyl-beta-D-galactopyranoside) plates containing MV at 1.25 micrograms/ml. Other O2- generators, menadione and plumbagin, also induced beta-galactosidase activity in these fusion strains. The induction by these drugs occurred only under aerobic conditions. Hyperoxygenation also elicited an induction of the fusions. On the other hand, no significant induction was observed with hydrogen peroxide and cumene hydroperoxide. The induction of these fusions by MV was not dependent on the peroxide stress control mediated by the oxyR gene or on the recA-dependent SOS system. These fusions were named soi (superoxide inducible)::lacZ. The induction of beta-galactosidase was significantly reduced by introducing a soxS::Tn10 locus into the fusion strains, indicating that the soi genes are members of the soxRS regulon. Five of the fusions were located in 6 to 26 min of the E. coli genetic map, while three fusions were located in 26 to 36 min, indicating that these fusions are not related to genes already known to be inducible by O2- under the control of soxRS. At least five mutants containing the soi::lacZ fusion were more sensitive to MV and menadione than the wild-type strain, suggesting that the products of these soi genes play an important role in protection against oxidative stress.  相似文献   

10.
A 16-kb BamHI fragment of the lactose plasmid pNZ63 from Leuconostoc lactis NZ6009 was cloned in Escherichia coli MC1061 by using pACYC184 and was found to express a functional beta-galactosidase. Deletion and complementation analysis showed that the coding region for beta-galactosidase was located on a 5.8-kb SalI-BamHI fragment. Nucleotide sequence analysis demonstrated that this fragment contained two partially overlapping genes, lacL (1,878 bp) and lacM (963 bp), that could encode proteins with calculated sizes of 72,113 and 35,389 Da, respectively. The L. lactis beta-galactosidase was overproduced in E. coli by using a lambda pL expression system. Two new proteins with M(r)s of 75,000 and 36,000 appeared upon induction of PL. The N-terminal sequences of these proteins corresponded to those deduced from the lacL and lacM gene sequences. Mutation and deletion analysis showed that lacL expression is essential for LacM production and that both the lacL and lacM genes are required for the production of a functional beta-galactosidase in E. coli. The deduced amino acid sequences of the LacL and LacM proteins showed considerable identity with the sequences of the N- and C-terminal parts, respectively, of beta-galactosidases from other lactic acid bacteria or E. coli. DNA and protein sequence alignments suggest that the L. lactis lacL and lacM genes have been generated by an internal deletion in an ancestral beta-galactosidase gene.  相似文献   

11.
12.
A digoxigenin-labeled DNA probe that was complementary to the gene ptsH and the beginning of the gene ptsI was used to clone a 3.2-kb HincII-BamHI restriction fragment containing the complete ptsI gene of Staphylococcus carnosus. The restriction fragment was cloned in the antisense orientation to the lac promoter in the low-copy-number vector pSU18. The nucleotide sequences of the ptsI gene, which encodes enzyme I (EC 2.7.3.9), and the corresponding flanking regions were determined. The primary translation product, derived from the nucleotide sequence, consists of 574 amino acids and has a calculated molecular weight of 63,369. Amino acid sequence comparison showed 47% similarity to enzyme I of Escherichia coli and 37% similarity to the enzyme I domain of the multiphosphoryl transfer protein of Rhodobacter capsulatus. The histidinyl residue at position 191 could be identified as the probable phosphoenolpyruvate-dependent phosphorylation site of enzyme I of S. carnosus because of sequence homologies with the peptide sequences of enzyme I-active sites of Enterococcus faecalis and Lactococcus lactis. Several in vivo and in vitro complementation studies with the enzyme I ptsI genes of S. carnosus and the E. coli ptsI mutant JLT2 were carried out. The generation times and interaction between enzyme I with histidine-containing protein from gram-positive and gram-negative bacteria were measured in a phosphoryl group transfer test.  相似文献   

13.
Five acetic acid-sensitive mutants of Acetobacter aceti subsp. aceti no. 1023 were isolated by mutagenesis with N-methyl-N'-nitro-N-nitrosoguanidine. Three recombinant plasmids that complemented the mutations were isolated from a gene bank of the chromosome DNA of the parental strain constructed in Escherichia coli by using cosmid vector pMVC1. One of these plasmids (pAR1611), carrying about a 30-kilobase-pair (kb) fragment that conferred acetic acid resistance to all five mutants, was further analyzed. Subcloning experiments indicated that a 8.3-kb fragment was sufficient to complement all five mutations. To identify the mutation loci and genes involved in acetic acid resistance, insertional inactivation was performed by insertion of the kanamycin resistance gene derived from E. coli plasmid pACYC177 into the cloned 8.3-kb fragment and successive integration into the chromosome of the parental strain. The results suggested that three genes, designated aarA, aarB, and aarC, were responsible for expression of acetic acid resistance. Gene products of these genes were detected by means of overproduction in E. coli by use of the lac promoter. The amino acid sequence of the aarA gene product deduced from the nucleotide sequence was significantly similar to those of the citrate synthases (CSs) of E. coli and other bacteria. The A. aceti mutants defective in the aarA gene were found to lack CS activity, which was restored by introduction of a plasmid containing the aarA gene. A mutation in the CS gene of E. coli was also complemented by the aarA gene. These results indicate that aarA is the CS gene.  相似文献   

14.
In Shigella flexneri, the ompB locus (containing the ompR and envZ genes) was found to modulate expression of the vir genes, which are responsible for invasion of epithelial cells. vir gene expression was markedly enhanced under conditions of high osmolarity (300 mosM), similar to that encountered in tissues both extra- and intracellularly. Two ompB mutants were constructed and tested for virulence and for osmotic regulation of vir genes. An envZ::Tn10 mutant remained invasive, although its virulence was significantly decreased as a result of its inability to survive intracellularly. By using a vir::lac operon fusion, this mutation was shown to decrease beta-galactosidase expression both in low- and high-osmolarity conditions but did not affect vir expression in response to changes in osmolarity. A delta ompB deletion mutant was also constructed via allelic exchange with an in vitro-mutagenized ompB locus of Escherichia coli. This mutation severely impaired virulence and abolished expression of the vir::lac fusion in both low- and high-osmolarity conditions. Therefore, a two-component regulatory system modulates virulence according to environmental conditions. In addition, the mutation affecting a spontaneous avirulent variant of S. flexneri serotype 5, M90T, has been mapped at the ompB locus and was complemented by the cloned E. coli ompB locus. Introduction of the vir::lac fusion into this mutant did not result in the expression of beta-galactosidase (Lac-).  相似文献   

15.
The amino acid sequence of the first 30 residues of fragment C of tetanus toxin was determined, and a mixture of 32 complementary oligonucleotides, each 17 bases long, was synthesized. A 2-kilobase (kb) EcoI fragment of Clostridium tetani DNA was identified by Southern blotting and was cloned into the Escherichia coli plasmid vector pAT153 with the 32P-labeled oligonucleotide mixture as a probe. A second 3.2-kb Bg/II fragment was identified and cloned with the 2-kb EcoRI fragment as a probe. The nucleotide sequence of 1.8 kb of this DNA was determined and was shown to encode the entire fragment C and a portion of fragment B of tetanus toxin. The tetanus DNA was expressed in E. coli with pWRL507, a plasmid vector containing the trp promoter and a portion of the trpE gene. The trpE-tetanus fusion proteins were visualized by sodium dodecyl sulfate-polyacrylamide gel electrophoresis and were shown to react with anti-fragment C antibody.  相似文献   

16.
17.
Identification and cloning of a fur regulatory gene in Yersinia pestis.   总被引:37,自引:15,他引:22       下载免费PDF全文
Yersinia pestis is one of many microorganisms responding to environmental iron concentrations by regulating the synthesis of proteins and an iron transport system(s). In a number of bacteria, expression of iron uptake systems and other virulence determinants is controlled by the Fur regulatory protein. DNA hybridization analysis revealed that both pigmented and nonpigmented cells of Y. pestis possess a DNA locus homologous to the Escherichia coli fur gene. Introduction of a Fur-regulated beta-galactosidase reporter gene into Y. pestis KIM resulted in iron-responsive beta-galactosidase activity, indicating that Y. pestis KIM expresses a functional Fur regulatory protein. A cloned 1.9-kb ClaI fragment of Y. pestis chromosomal DNA hybridized specifically to the fur gene of E. coli. The coding region of the E. coli fur gene hybridized to a 1.1-kb region at one end of the cloned Y. pestis fragment. The failure of this clone to complement an E. coli fur mutant suggests that the 1.9-kb clone does not contain a functional promoter. Subcloning of this fragment into an inducible expression vector restored Fur regulation in an E. coli fur mutant. In addition, a larger 4.8-kb Y. pestis clone containing the putative promoter region complemented the Fur- phenotype. These results suggest that Y. pestis possesses a functional Fur regulatory protein capable of interacting with the E. coli Fur system. In Y. pestis Fur may regulate the expression of iron transport systems and other virulence factors in response to iron limitation in the environment. Possible candidates for Fur regulation in Y. pestis include genes involved in ferric iron transport as well as hemin, heme/hemopexin, heme/albumin, ferritin, hemoglobin, and hemoglobin/haptoglobin utilization.  相似文献   

18.
Cloning chromosomal lac genes of Klebsiella pneumoniae   总被引:4,自引:0,他引:4  
C MacDonald  M Riley 《Gene》1983,24(2-3):341-345
The chromosomal gene for beta-galactosidase from Klebsiella pneumoniae strain T17R1 and associated regulatory genes have been cloned as a 5-kb HindIII fragment in the pBR322 plasmid vector. The beta-galactoside permease gene is not present in a functional form in the 5-kb fragment. The K. pneumoniae genes are expressed in an Escherichia coli host. The synthesis of beta-galactosidase is inducible by isopropyl-beta-D-galactosidase (IPTG) and is sensitive to catabolite repression. There appears to be greater homology between the K. pneumoniae and E. coli structural genes for beta-galactosidase than there is between the respective repressor genes.  相似文献   

19.
An 11.6-kilobase (kb) region of a 34-kb fragment of Escherichia coli DNA that encodes the K1 capsular polysaccharide genes is necessary for translocation of the K1 polysaccharide to the bacterial cell surface. This 11.6-kb region contains a gene, kpsD, encoding a 60-kilodalton protein. The kpsD gene was localized to a 2.4-kb PstI-BamHI fragment. Cells harboring a Tn1000 insertion in kpsD did not synthesize the 60-kilodalton protein and did not express polysaccharide on the cell surface. Immunodiffusion and rocket immunoelectrophoresis of cell extracts, however, demonstrated that K1 polysaccharide was synthesized by these cells. We present evidence that the kpsD gene product is synthesized as a precursor and that the processed form is located in the periplasmic space. Analysis of alkaline phosphatase activity of a kpsD-phoA fusion demonstrated that kpsD expression was under positive regulation. A 260-base-pair AluI fragment located within the kpsD coding sequence was used as a probe and was found to hybridize to chromosomal DNA from E. coli that synthesizes the K2, K5, K7, K12, and K13 capsular polysaccharides but not K3 and K100. These results suggest that the kpsD gene product may be required for export not only of K1 but for other K antigens as well.  相似文献   

20.
Merodiploid derivatives bearing an F-linked lac operon (i(+), o(+), z(+), y(+), a(+)) from Escherichia coli were prepared from a Proteus mirabilis strain unable to utilize lactose and from a lac deletion strain of E. coli. A suitable growth medium was found in which the episomal element in the P. mirabilis derivative was sufficiently stable to allow induction of the episome-borne lac operon and thus to permit a comparison of the activities and properties of E. coli lac products in the intracellular environments of P. mirabilis and E. coli. In both derivatives the episomal lac operon was shown to be repressed in the absence of inducer. Kinetics of induction with gratuitous inducer (isopropyl-1-thio-beta-d-galactoside) were similar for both beta-galactosidase activity (beta-d-galactoside galactohydrolase, EC 3.4.1.23) and beta-galactoside transport activity in both derivatives, although the ratio of galactoside transport to beta-galactosidase activity was approximately 1.6-fold higher in the E. coli derivative. Comparison of beta-galactosidase and M-protein (lac y gene product)-specific activities indicated coordinate expression of the induced lac operon in both derivatives. Quantitatively, the maximal beta-galactosidase specific activity was two or three times higher for the E. coli derivative. A significant sodium azide inhibition (65% inhibition by 10 mM sodium azide) of lactose permease-mediated transport of o-nitrophenyl-beta-galactoside from an outside region of high concentration to an inside region of very low concentration ("downhill transport") was observed for the P. mirabilis derivative. Identical conditions for the E. coli derivative yielded only about 15% inhibition. Active transport of thiomethyl-beta-galactoside was similar for both derivatives, the major difference being that active transport was more sensitive to azide poisoning in the P. mirabilis derivative. Preliminary examination of the thiomethyl-beta-galactoside derivatives following active transport did not demonstrate the accumulation of a phosphorylated product in either strain but did reveal an unidentified derivative present in the P. mirabilis merodiploid extract which was not detectable in the E. coli merodiploid.  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