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 harboring cry4A, cry4B and cry11A from Bacillus thuringiensis subsp. israelensis and binary toxin genes from Bacillus sphaericus has been constructed. The three cry genes were placed under the control of the cry4B promoter whereas the binary toxin gene was controlled by its native promoter. The expression of toxins in Escherichia coli harboring the resulting plasmid, p4BDA-5142, was investigated. Cry4B expression was highest compared to other toxins. Although the level of toxin expression was low compared with E. coli expressing single toxins, the recombinant E. coli strain harboring p4BDA-5142 exhibited broad range mosquito-larvicidal activity against all Aedes, Culex and Anopheles larvae. This work has shown that the development of the recombinant plasmid can be used to broaden the host range spectrum of the appropriate bacterial host for mosquito control.  相似文献   

2.
To facilitate the analysis of genetic determinants carried by large resident plasmids of Bacillus anthracis, a mating system was developed which promotes plasmid transfer among strains of B. anthracis, B. cereus, and B. thuringiensis. Transfer of the selectable tetracycline resistance plasmid pBC16 and other plasmids from B. thuringiensis to B. anthracis and B. cereus recipients occurred during mixed incubation in broth. Two plasmids, pXO11 and pXO12, found in B. thuringiensis were responsible for plasmid mobilization. B. anthracis and B. cereus transcipients inheriting either pXO11 or pXO12 were, in turn, effective donors. Transcipients harboring pXO12 were more efficient donors than those harboring pXO11; transfer frequencies ranged from 10(-4) to 10(-1) and from 10(-8) to 10(-5), respectively. Cell-to-cell contact was necessary for plasmid transfer, and the addition of DNase had no effect. The high frequencies of transfer, along with the fact that cell-free filtrates of donor cultures were ineffective, suggested that transfer was not phage mediated. B. anthracis and B. cereus transcipients which inherited pXO12 also acquired the ability to produce parasporal crystals (Cry+) resembling those produced by B. thuringiensis, indicating that pXO12 carries a gene(s) involved in crystal formation. Transcipients which inherited pXO11 were Cry-. This mating system provides an efficient method for interspecies transfer of a large range of Bacillus plasmids by a conjugation-like process.  相似文献   

3.
We used site-directed mutagenesis to introduce both a NdeI restriction endonuclease site and an initiator codon at the junction of the leader and structural gene sequences of the metallo-beta-lactamase of Bacillus cereus 5/B/6. This construct allowed us to clone just the beta-lactamase structural gene sequence into an Escherichia coli expression vector. E. coli cells were transformed with the recombinant plasmid, the B. cereus beta-lactamase was expressed, and these E. coli cells were disrupted by sonic oscillation. When the resultant suspensions were clarified by ultracentrifugation, the B. cereus beta-lactamase represented 15% of the total protein in the supernatant. Subsequent gel filtration and ion-exchange chromatography allowed the first reported purification to homogeneity of the B. cereus beta-lactamase from E. coli with an 87% recovery and an overall yield of 17 mg of enzyme per liter of cell culture. The electrophoretic mobilities of the enzyme expressed in and purified from E. coli and the enzyme purified directly from B. cereus were identical in both native and sodium dodecyl sulfate gel electrophoreses. As with the B. cereus enzyme, Km and Vmax (using cephalosporin C as substrate) for the enzyme purified from E. coli were 0.39 mM and 1333 units/mg protein, respectively. Likewise, the Co(II)-reconstituted enzyme purified from E. coli, which retained 29% of the activity of the Zn(II) enzyme, had electronic absorption spectra with maxima at 347, 551, 617, and 646 nm with extinction coefficients of 900, 250, 173, and 150 M-1 cm-1, respectively.  相似文献   

4.
The transfer of plasmids by mating from four Bacillus thuringiensis subspecies to Bacillus anthracis and Bacillus cereus recipients was monitored by selecting transcipients which acquired plasmid pBC16 (Tcr). Transcipients also inherited a specific large plasmid from each B. thuringiensis donor at a high frequency along with a random array of smaller plasmids. The large plasmids (ca. 50 to 120 megadaltons), pXO13, pXO14, pXO15, and pXO16, originating from B. thuringiensis subsp. morrisoni, B. thuringiensis subsp. toumanoffi, B. thuringiensis subsp. alesti, and B. thuringiensis subsp. israelensis, respectively, were demonstrated to be responsible for plasmid mobilization. Transcipients containing any of the above plasmids had donor capability, while B. thuringiensis strains cured of each of them were not fertile, indicating that the plasmids confer conjugation functions. Confirmation that pXO13, pXO14, and pXO16 were self-transmissible was obtained by the isolation of fertile B. anthracis and B. cereus transcipients that contained only pBC16 and one of these plasmids. pXO14 was efficient in mobilizing the toxin and capsule plasmids, pXO1 and pXO2, respectively, from B. anthracis transcipients to plasmid-cured B. anthracis or B. cereus recipients. DNA-DNA hybridization experiments suggested that DNA homology exists among pXO13, pXO14, and the B. thuringiensis subsp. thuringiensis conjugative plasmids pXO11 and pXO12. Matings performed between strains which each contained the same conjugative plasmid demonstrated reduced efficiency of pBC16 transfer. However, in many instances when donor and recipient strains contained different conjugative plasmids, the efficiency of pBC16 transfer appeared to be enhanced.  相似文献   

5.
A synthetic oligodeoxynucleotide probe was used to clone the gene encoding the phosphatidylcholine-preferring phospholipase C of Bacillus cereus. The sequence of a 2050-bp restriction fragment containing the gene was determined. Analysis of the gene-derived amino acid (aa) sequence showed that this exoenzyme is probably synthesized as a 283-aa precursor with a 24-aa signal peptide and a 14-aa propeptide. The mature, secreted enzyme comprises 245 aa residues. Sonicates of Escherichia coli HB101 carrying the gene on a multicopy plasmid showed phospholipase C activity. This activity was inhibited by Tris, a known inhibitor of the B. cereus enzyme and also by antiserum raised against pure B. cereus phospholipase C. We conclude therefore that the gene is expressed in E. coli. The cloning and sequencing described here complete the first step toward using in vitro mutagenesis for investigations of the structure-function relationships of B. cereus phospholipase C.  相似文献   

6.
Recent characterisations of plasmids related to the anthrax virulence plasmids pXO1 and pXO2 in clinical isolates of Bacillus cereus and Bacillus thuringiensis have contributed to the emerging picture of a virulence-associated plasmid pool in the B. cereus sensu lato group. The family of pXO2-like plasmids includes the conjugative plasmid pAW63 from the biopesticide strain B. thuringiensis subsp. kurstaki HD73 and the heretofore cryptic plasmid pBT9727 from the clinical strain B. thuringiensis subsp. konkukian 97-27. Comparative sequence analysis of these three plasmids suggested that they were derived from an ancestral conjugative plasmid, with pAW63 retaining its self-transfer capabilities, and pXO2 having lost them through genetic drift. Such properties had not been investigated in pBT9727, but sequence homologies led us to predict that it may possess self-transfer capabilities. Here, we report that pBT9727 is indeed conjugative, and is able to promote its own transfer as well as that of small mobilisable plasmids.  相似文献   

7.
The substrate stereospecificity of phosphatidylinositol-specific phospholipase C from Bacillus cereus is examined using the resolved optical isomers of synthetic myo-inositol 1-(4-nitrophenyl phosphate), a chromogenic substrate for the phospholipase. The synthetic route employs mild acid-labile protecting groups and separation of the substituted myo-inositol enantiomers as the (-)-camphanyl ester diastereomers. Measurements of the initial rates of cleavage of the D and L enantiomers of the nitrophenyl substrate by phosphatidylinositol-specific phospholipase C from B. cereus show that this enzyme is essentially stereospecific for the D enantiomer. Under identical conditions, the rate of cleavage of the L isomer is less than 0.2% of that observed for the D isomer. The same is observed for the highly homologous enzyme from Bacillus thuringiensis. There is no measurable inhibition by the L enantiomer of the B. cereus enzyme acting on the D enantiomer, even when the molar ratio of L:D is 5, indicating that binding of the L enantiomer to the phospholipase is negligible. Thus, the enzyme active site is exquisitely sensitive to the stereochemistry of the myo-inositol group of the substrate.  相似文献   

8.
9.
Transfer of chromosomal genes and plasmids in Bacillus thuringiensis   总被引:1,自引:0,他引:1  
A low frequency of chromosomal gene transfer from Bacillus thuringiensis to Bacillus cereus was detected by cell mating, with a tryptophan marker being the most frequently transferred gene among four that were tested. The process was resistant to DNase and was not mediated by cell filtrates. Among several B. thuringiensis subspecies tested, transfer was best with a derivative of B. thuringiensis subsp. kurstaki HD1, which lost several plasmids. All of the B. cereus recombinants contained at least one plasmid from the donor B. thuringiensis; frequently, it was a plasmid that encoded a protoxin gene. In matings with B. thuringiensis subsp. kurstaki HD1, a 29-megadalton plasmid that contained a ca. 2.5-kilobase region of homology with the chromosome was always transferred. No detectable transfer of chromosomal genes was found in B. thuringiensis subsp. kurstaki HD1 strains lacking this plasmid, suggesting that there may be chromosome mobilization.  相似文献   

10.
A low frequency of chromosomal gene transfer from Bacillus thuringiensis to Bacillus cereus was detected by cell mating, with a tryptophan marker being the most frequently transferred gene among four that were tested. The process was resistant to DNase and was not mediated by cell filtrates. Among several B. thuringiensis subspecies tested, transfer was best with a derivative of B. thuringiensis subsp. kurstaki HD1, which lost several plasmids. All of the B. cereus recombinants contained at least one plasmid from the donor B. thuringiensis; frequently, it was a plasmid that encoded a protoxin gene. In matings with B. thuringiensis subsp. kurstaki HD1, a 29-megadalton plasmid that contained a ca. 2.5-kilobase region of homology with the chromosome was always transferred. No detectable transfer of chromosomal genes was found in B. thuringiensis subsp. kurstaki HD1 strains lacking this plasmid, suggesting that there may be chromosome mobilization.  相似文献   

11.
The possibility of homologous and heterologous transception of Cry+ plasmids in Bacillus thuringiensis is demonstrated. Cry+ plasmids from crystal bearing strain of Bacillus thuringiensis were transferred into acrystalline strain belonging to H5 serotype by mutual incubation. The donor strain was previously marked by the transmissive plasmid pAM beta 1 coding for erythromycin and lincomycin resistance. The transcipients having acquired the ability to synthesize delta-endotoxin were referred to H5 serotype due to their phenotype. By analogous method Cry+ plasmid was transferred from Bacillus thuringiensis to Bacillus cereus. Bacillus cereus strain GP7 was used as a recipient strain resistant to tetracycline. The presence of delta-endotoxin in transcipients was confirmed by bioprobes and immunoenzyme assay. To prove the transfer of Cry+ plasmid the plasmid profiles of the parent strains and transcipients have been analyzed. The formation of cellular contacts during mutual incubation of Bacillus thuringiensis and Bacillus cereus strains was demonstrated by electron microscopic study of ultrafine cuts.  相似文献   

12.
Bacteriophage CP-51, a generalized transducing phage for Bacillus anthracis, B. cereus, and B. thuringiensis, mediates transduction of plasmid DNA. B. cereus GP7 harbors the 2.8-megadalton multicopy tetracycline resistance plasmid, pBC16. B. thuringiensis 4D11A carries pC194, the 1.8-megadalton multicopy chloramphenicol resistance plasmid. When phage CP-51 was propagated on these strains, it transferred the plasmid-encoded antibiotic resistances to the nonvirulent Weybridge (Sterne) strain of B. anthracis, to B. cereus 569, and to strains of several B. thuringiensis subspecies. The frequency of transfer was as high as 10(-5) transductants per PFU. Tetracycline-resistant and chloramphenicol-resistant transductants contained newly acquired plasmid DNA having the same molecular weight as that contained in the donor strain. Antibiotic-resistant transductants derived from any of the three species were effective donors of plasmids to recipients from all three species.  相似文献   

13.
The plasmids of the members of the Bacillus cereus sensu lato group of organisms are essential in defining the phenotypic traits associated with pathogenesis and ecology. For example, Bacillus anthracis contains two plasmids, pXO1 and pXO2, encoding toxin production and encapsulation, respectively, that define this species pathogenic potential, whereas the presence of a Bt toxin-encoding plasmid defines Bacillus thuringiensis isolates. In this study the plasmids from B. cereus isolates that produce emetic toxin or are linked to periodontal disease were sequenced and analyzed. Two periodontal isolates examined contained almost identical approximately 272-kb plasmids, named pPER272. The emetic toxin-producing isolate contained one approximately 270-kb plasmid, named pCER270, encoding the cereulide biosynthesis gene cluster. Comparative sequence analyses of these B. cereus plasmids revealed a high degree of sequence similarity to the B. anthracis pXO1 plasmid, especially in a putative replication region. These plasmids form a newly defined group of pXO1-like plasmids. However, these novel plasmids do not contain the pXO1 pathogenicity island, which in each instance is replaced by plasmid specific DNA. Plasmids pCER270 and pPER272 share regions that are not found in any other pXO1-like plasmids. Evolutionary studies suggest that these plasmids are more closely related to each other than to other identified B. cereus plasmids. Screening of a population of B. cereus group isolates revealed that pXO1-like plasmids are more often found in association with clinical isolates. This study demonstrates that the pXO1-like plasmids may define pathogenic B. cereus isolates in the same way that pXO1 and pXO2 define the B. anthracis species.  相似文献   

14.
A Z Ge  R M Pfister  D H Dean 《Gene》1990,93(1):49-54
Conditions for hyperexpression, in Escherichia coli, of the Bacillus thuringiensis var, kurstaki gene, cryIA9(c)73, encoding an insecticidal crystal protein, CryIA(c)73, were investigated by varying the promoter type, host cell, plasmid copy number, the second codon and number of terminators. The cryIA(c)73 gene was cloned into three E. coli expression vectors, pKK223-3 (Ptac promoter), pET-3a (P phi 10 promoter), and pUC19 (Ptac promoter). The level of cryIA(c)73 expression was measured by ELISA and compared to total cellular protein over growth periods of 24 and 48 h. Maximum expression levels of 284 microgram CryIA(C)73/ml (48% of cellular protein) were obtained in shake flasks with the Ptac promoter in E. coli JM103. Optimal conditions were found to be low-copy-number plasmid (pBR322 ori), 48 h of growth, in lon+ cells. A change of the gene's second codon to AAA can improve expression by two to three fold but is undetectable in the presence of a strong E. coli promoter. The cryIA(c)73 gene product, in E. coli, formed crystals with the same lattice structure as the native crystals formed in B. thuringiensis (as visualized by electron microscopy). Bioassay results (insect toxicity and specificity) of the crystal produced in E. coli were similar to that produced in B. thuringiensis.  相似文献   

15.
rpoB and gyr genes (and their fragments) of chromosomal DNA of bacteria from Bacillus cereus group - B. anthracis, B. cereus, and B. thuringiensis - which are the potential markers for their genotyping were sequenced and phylogenetic trees were constructed. Sets of primers for species-specific detection of B. anthracis, B. cereus, and B. thuringiensis by multiplex polymerase chain reaction were designed. Also primers sets, which allow to differentiate strains of B. anthracis with various plasmid profiles (containing both plasmids (pXO1+, pXO2+), and without one (pXO1+, pXO2- or pXO1-, pXO2+) or both plasmids (pXO1-, pXO2-), determining pathogenic characteristics of the strains, were developed. For multiplex PCR primer sets were optimized on the annealing temperature of primers and amplicon length. Itwas shown that phylogenetic tree can be applied as an indicator of reliability and accuracy of taxonomical classification of microorganisms' species and subspecies. Comparison of pXO1 and pXO2 plasmid sequences of B. anthracis showed that these plasmids contain 18 and 4 palindrome sequences respectively which can potentially form thermodynamically stable hairpin-loop structures.  相似文献   

16.
Spores and vegetative cells of Bacillus thuringiensis were more sensitive to UV light than were spores or cells of plasmid-cured B. thuringiensis strains or of the closely related Bacillus cereus. Introduction of B. thuringiensis plasmids into B. cereus by cell mating increased the UV sensitivity of the cells and spores. Protoxins encoded by one or more B. thuringiensis plasmids were not involved in spore sensitivity, since a B. thuringiensis strain conditional for protoxin accumulation was equally sensitive at the permissive and nonpermissive temperatures. In addition, introduction of either a cloned protoxin gene, the cloning vector, or another plasmid not containing a protoxin gene into a plasmid-cured strain of B. thuringiensis all increased the UV sensitivity of the spores. Although the variety of small, acid-soluble proteins was the same in the spores of all strains examined, the quantity of dipicolinic acid was about twice as high in the plasmid-containing strains, and this may account for the differences in UV sensitivity of the spores. The cells of some strains harboring only B. thuringiensis plasmids were much more sensitive than cells of any of the other strains, and the differences were much greater than observed with spores.  相似文献   

17.
The replication origins of three large Bacillus thuringiensis plasmids, derived from B. thuringiensis HD263 subsp. kurstaki, have been cloned in Escherichia coli and sequenced. The replication origins, designated ori 43, ori 44, and ori 60, were isolated from plasmids of 43, 44, and 60 MDa, respectively. Each cloned replication origin exhibits incompatibility with the resident B. thuringiensis plasmid from which it was derived. Recombinant plasmids containing the three replication origins varied in their ability to transform strains of B. thuringiensis, Bacillus megaterium, and Bacillus subtilis. Analysis of the derived nucleotide and amino acid sequences indicates that the replication origins are nonhomologous, implying independent derivations. No significant homology was found to published sequences of replication origins derived from the single-stranded DNA plasmids of gram-positive bacteria, and shuttle vectors containing the three replication origins do not appear to generate single-stranded DNA intermediates in B. thuringiensis. The replication origin regions of the large plasmids are each characterized by a single open reading frame whose product is essential for replication in B. thuringiensis. The putative replication protein of ori 60 exhibits partial homology to the RepA protein of the Bacillus stearothermophilus plasmid pTB19. The putative replication protein of ori 43 exhibits weak but extensive homology to the replication proteins of several streptococcal plasmids, including the open reading frame E replication protein of the conjugative plasmid pAM beta 1. The nucleotide sequence of ori 44 and the amino acid sequence of its putative replication protein appear to be nonhomologous to other published replication origin sequences.  相似文献   

18.
Spores and vegetative cells of Bacillus thuringiensis were more sensitive to UV light than were spores or cells of plasmid-cured B. thuringiensis strains or of the closely related Bacillus cereus. Introduction of B. thuringiensis plasmids into B. cereus by cell mating increased the UV sensitivity of the cells and spores. Protoxins encoded by one or more B. thuringiensis plasmids were not involved in spore sensitivity, since a B. thuringiensis strain conditional for protoxin accumulation was equally sensitive at the permissive and nonpermissive temperatures. In addition, introduction of either a cloned protoxin gene, the cloning vector, or another plasmid not containing a protoxin gene into a plasmid-cured strain of B. thuringiensis all increased the UV sensitivity of the spores. Although the variety of small, acid-soluble proteins was the same in the spores of all strains examined, the quantity of dipicolinic acid was about twice as high in the plasmid-containing strains, and this may account for the differences in UV sensitivity of the spores. The cells of some strains harboring only B. thuringiensis plasmids were much more sensitive than cells of any of the other strains, and the differences were much greater than observed with spores.  相似文献   

19.
Bacillus thuringiensis subsp. kurstaki total DNA was digested with BglII and cloned into the BamHI site of plasmid pUC9 in Escherichia coli. A recombinant plasmid, pHBHE, expressed a protein of 135,000 daltons that was toxic to caterpillars. A HincII-SmaI double digest of pHBHE was then ligated to BglII-cut plasmid pBD64 and introduced into Bacillus subtilis by transformation. The transformants were identified by colony hybridization and confirmed by Southern blot hybridization. A 135,000-dalton protein which bound to an antibody specific for the crystal protein of B. thuringiensis was detected from the B. subtilis clones containing the toxin gene insert in either orientation. A toxin gene insert cloned into a PvuII site distal from the two drug resistance genes of the pBD64 vector also expressed a 135,000-dalton protein. These results suggest that the toxin gene is transcribed from its own promoter. Western blotting of proteins expressed at various stages of growth revealed that the crystal protein expression in B. subtilis begins early in the vegetative phase, while in B. thuringiensis it is concomitant with the onset of sporulation. The cloned genes when transferred to a nonsporulating strain of B. subtilis also expressed a 135,000-dalton protein. These results suggest that toxin gene expression in B. subtilis is independent of sporulation. Another toxin gene encoding a 130,000- to 135,000-dalton protein was cloned in E. coli from a library of B. thuringiensis genes established in lambda 1059. This gene was then subcloned in B. subtilis. The cell extracts from both clones were toxic to caterpillars. Electron microscope studies revealed the presence of an irregular crystal inclusion in E. coli and a well-formed bipyramidal crystal in B. subtilis clones similar to the crystals found in B. thuringiensis.  相似文献   

20.
A total of 75 strains (including 5 reference strains) of Bacillus amyloliquefaciens, B. cereus, B. circulans, B. licheniformis, B. megaterium, B. pumilus, B. sphaericus, B. subtilis, and B. thuringiensis and 36 species-unidentified Bacillus strains were surveyed for plasmids by cesium chloride-ethidium bromide equilibrium centrifugation of cell lysates in a study of antibiotic resistance in host cells. Of the 111 strains, 13 (including 3 reference strains) were found to harbor plasmids, and 5 of the 13 showed antibiotic resistance. This antibiotic resistance appeared not to be due to the plasmids, however, because the trait was not cured by cultivation of cells in nutrient medium containing ethidium bromide (1 mug/ml), sodium dodecyl sulfate (0.2 mug/ml), or novobiocin (1 mug/ml), except in one strain, in which kanamycin and streptomycin resistances were cured by novobiocin. One strain of B. amyloliquefaciens, S294, was found to harbor a plasmid, pFTB14, which differed from the plasmid species of classes 1 to 6 in B. subtilis and B. amyloliquefaciens, as determined by restriction analysis and DNA contour length determination. However, in DNA-DNA hybridization on a filter after Southern blotting from an agarose gel, the pFTB14 DNA hybridized with plasmids of classes 1 to 5. Three strains of B. thuringiensis each carried at least 4 to 11 plasmid species, whereas no plasmids were detected in four strains of B. cereus, which, in relation to B. thuringiensis, is closely related taxonomically and has highly homologous DNA sequences. The plasmid DNAs prepared from species other than B. subtilis and B. amyloliquefaciens did not hybridize with that of pFTB14.  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