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.
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   

2.
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   

3.
The 55-kilobase plasmid, pLS20, of Bacillus subtilis (natto) 3335 promotes transfer of the tetracycline resistance plasmid pBC16 from B. subtilis (natto) to the Bacillus species B. anthracis, B. cereus, B. licheniformis, B. megaterium, B. pumilus, B. subtilis, and B. thuringiensis. Frequency of pBC16 transfer ranged from 2.3 x 10(-6) to 2.8 x 10(-3). Evidence for a plasmid-encoded conjugationlike mechanism of genetic exchange includes (i) pLS20+ strains, but not pLS20- strains, functioned as donors of pBC16; (ii) plasmid transfer was insensitive to the presence of DNase; and (iii) cell-free filtrates of donor cultures did not convert recipient cells to Tcr. Cotransfer of pLS20 and pBC16 in intraspecies matings and in matings with a restriction-deficient B. subtilis strain indicated that pLS20 was self-transmissible. In addition to mobilizing pBC16, pLS20 mediated transfer of the B. subtilis (natto) plasmid pLS19 and the Staphylococcus aureus plasmid pUB110. The fertility plasmid did not carry a selectable marker. To facilitate direct selection for pLS20 transfer, plasmid derivatives which carried the erythromycin resistance transposon Tn917 were generated. Development of this method of genetic exchange will facilitate the introduction of plasmid DNA into nontransformable species by use of transformable fertile B. subtilis or B. subtilis (natto) strains as intermediates.  相似文献   

4.
Plasmid transfer between strains of Bacillus thuringiensis subsp. israelensis was studied under a range of environmentally relevant laboratory conditions in vitro, in river water, and in mosquito larvae. Mobilization of pBC16 was detected in vitro at a range of temperatures, pH values, and available water conditions, and the maximum transfer ratio was 10(-3) transconjugant per recipient under optimal conditions. Transfer of conjugative plasmid pXO16::Tn5401 was also detected under this range of conditions. However, a maximum transfer ratio of 1.0 transconjugant per recipient was attained, and every recipient became a transconjugant. In river water, transfer of pBC16 was not detected, probably as a result of the low transfer frequency for this plasmid and the formation of spores by the introduced donor and recipient strains. In contrast, transfer of plasmid pXO16::Tn5401 was detected in water, but at a lower transfer ratio (ca. 10(-2) transconjugant per donor). The number of transconjugants increased over the first 7 days, probably as a result of new transfer events between cells, since growth of both donor and recipient cells in water was not detected. Mobilization of pBC16 was not detected in killed mosquito larvae, but transfer of plasmid pXO16::Tn5401 was evident, with a maximum rate of 10(-3) transconjugant per donor. The reduced transfer rate in insects compared to broth cultures may be accounted for by competition from the background bacterial population present in the mosquito gut and diet or by the maintenance of a large population of B. thuringiensis spores in the insects.  相似文献   

5.
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   

6.
The complete sequencing and annotation of the 181.7-kb Bacillus anthracis virulence plasmid pXO1 predicted 143 genes but could only assign putative functions to 45. Hybridization assays, PCR amplification, and DNA sequencing were used to determine whether pXO1 open reading frame (ORF) sequences were present in other bacilli and more distantly related bacterial genera. Eighteen Bacillus species isolates and four other bacterial species were tested for the presence of 106 pXO1 ORFs. Three ORFs were conserved in most of the bacteria tested. Many of the pXO1 ORFs were detected in closely related Bacillus species, and some were detected only in B. anthracis isolates. Three isolates, Bacillus cereus D-17, B. cereus 43881, and Bacillus thuringiensis 33679, contained sequences that were similar to more than one-half of the pXO1 ORF sequences examined. The majority of the DNA fragments that were amplified by PCR from these organisms had DNA sequences between 80 and 98% similar to that of pXO1. Pulsed-field gel electrophoresis revealed large potential plasmids present in both B. cereus 43881 (341 kb) and B. thuringiensis ATCC 33679 (327 kb) that hybridized with a DNA probe composed of six pXO1 ORFs.  相似文献   

7.
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   

8.
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   

9.
Abstract Two cryptic plasmids of 8.6 and 15 kb, originating from Bacillus thuringiensis , have been cloned in Escherichia coli . The determination of their physical map shows that the 8.6-kb plasmid harbors the transposon Tn 4430 and that the 15-kb plasmid carries Tn 4430 plus one copy of the IS 231 element. The replication regions were identified on the restriction maps and the segregational stability of derived plasmids containing these regions was analyzed in B. subtillis . The results indicate that the stability of these plasmids is negatively correlated to the temperature. After 30 generations, without selective pressure at 51°C, the two types of plasmids are lost.  相似文献   

10.
The strain H1.1 of Bacillus thuringiensis var. thuringiensis harbors three small cryptic plasmids: pGI1, pGI2, and pGI3 (8.2, 9.2, and 10.6 kb, respectively). Two of these plasmids (i.e., pGI2 and pGI3) were successfully cloned in their entirety into the vector pBR322, whereas only overlapping DNA fragments covering pGI1 were obtained in Escherichia coli. A curing-hybridization technique was used to obtain isolates of B. thuringiensis missing one or another small cryptic plasmid. These derivatives were examined for any change in a phenotypic trait, but no specific function could be assigned to one of these plasmids. Hybridization and restriction mapping data revealed that the transposon Tn4430 accounts for 45% of the pGI2 plasmid DNA.  相似文献   

11.
We present the microbiological and molecular characterization of bacteria isolated from four chimpanzees and one gorilla thought to have died of an anthrax-like disease in C?te d'Ivoire and Cameroon. These isolates differed significantly from classic Bacillus anthracis by the following criteria: motility, resistance to the gamma phage, and, for isolates from Cameroon, resistance to penicillin G. A capsule was expressed not only after induction by CO(2) and bicarbonate but also under normal growth conditions. Subcultivation resulted in beta-hemolytic activity and gamma phage susceptibility in some subclones, suggesting differences in gene regulation compared to classic B. anthracis. The isolates from C?te d'Ivoire and Cameroon showed slight differences in their biochemical characteristics and MICs of different antibiotics but were identical in all molecular features and sequences analyzed. PCR and Southern blot analyses confirmed the presence of both the toxin and the capsule plasmid, with sizes corresponding to the B. anthracis virulence plasmids pXO1 and pXO2. Protective antigen was expressed and secreted into the culture supernatant. The isolates possessed variants of the Ba813 marker and the SG-749 fragment differing from that of classic B. anthracis strains. Multilocus sequence typing revealed a close relationship of our atypical isolates with both classic B. anthracis strains and two uncommonly virulent Bacillus cereus and Bacillus thuringiensis isolates. We propose that the newly discovered atypical B. anthracis strains share a common ancestor with classic B. anthracis or that they emerged recently by transfer of the B. anthracis plasmids to a strain of the B. cereus group.  相似文献   

12.
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   

13.
Tn916-dependent mobilization of nonconjugative plasmids pUB110 and its derivative pUB110Deltam was compared. Deleting a 787-bp fragment from the pUB110 mob region created plasmid pUB110Deltam. Deletion of the mob region of pUB110 rendered the plasmid nontransferable by the conjugative plasmids of Bacillus thuringiensis subsp. israelensis. During matings between Bacillus subtilis (Tn916) and B. thuringiensis subsp. israelensis, however, Tn916-dependent mobilization of plasmids pUB110 and pUB110Deltam was observed at a frequency of approximately 2 x 10(-6) transconjugants per donor. The results show that Tn916-mediated conjugal transfer of plasmids is a mob-independent event. Jaworski and Clewell (J. Bacteriol 177; 6644-6651) recently demonstrated the presence of an IncP-like nicking site in the oriT of Tn916. These data suggest that a IncP-like nickling site is essential for Tn916-mediated plasmid transfer.  相似文献   

14.
J G Naglich  R E Andrews 《Plasmid》1988,20(2):113-126
The Staphylococcus aureus plasmids pC194 and pUB110 were introduced into Bacillus thuringiensis subsp. israelensis by using the Streptococcus faecalis transposon Tn916 as a mobilizing agent. Plasmid transfer occurred only when B. thuringiensis subsp. israelensis was mated with a B. subtilis donor that contained both pC194 and pUB110 and Tn916; plasmid transfer was not observed in the absence of the transposon. B. thuringiensis transconjugants resistant to chloramphenicol (Cmr) and tetracycline (Tetr) were detected at a frequency of 1.96 x 10(-6) per recipient cell, whereas the Tetr phenotype, but not the Cmr, was observed at a frequency of 1.09 x 10(-4). The converse, Cmr but not Tetr, was observed at a frequency of 2.94 X 10(-5). The transfer of pUB110 from B. subtilis to B. thuringiensis subsp. israelensis was observed at a frequency of 3.0 x 10(-6) per recipient cell but concomitant transfer of pUB110 and Tn916 was not observed. Mobilization of plasmid pE194 was not observed under these conditions. Transconjugants were detected in filter matings only, not in broth. The Tn916 phenotype was maintained during serial passage of B. thuringiensis without selection, whereas the pC194 phenotype was not. Unlike pC194, however, pUB110 remained stable in B. thuringiensis during several passages through nonselective medium. Southern hybridization analysis demonstrated that Tn916 had inserted into several different sites on the B. thuringiensis chromosome and that pC194 and pUB110 were maintained as an autonomous plasmid.  相似文献   

15.
The presence of one of the anthrax virulence plasmid pXO1 conserved fragments was analyzed in 24 Bacillus cereus and B. thuringiensis strains, including 6 B. thuringiensis subspecies, by polymerase chain reactions. Twelve out of 24 strains showed PCR-positive for an ORF101 homologous sequence. Two pXO1-ORF101-like fragments from a B. cereus B-4ac and a commercial B. thuringiensis kurstaki HD1 were cloned, sequenced and expressed in Escherichia coli. Toxicity assays revealed that the product encoded by the pXO1-ORF101-like fragment had no impact on either Vero cells or Chinese Hamster Ovary cells, suggesting that this fragment probably not contribute to enterotoxic activity. Sequence alignment of the pXO1-ORF101 from three Bacillus anthracis and ORF101-like fragments from other 12 B. cereus group isolates indicated high identity (more than 90%) and the presence of subgroup- and strain-specific SNPs among these fragments.  相似文献   

16.
Liu X  Wang D  Wang H  Feng E  Zhu L  Wang H 《PloS one》2012,7(1):e29875
The large plasmid pXO1 encoding the anthrax toxin is important for the virulence of Bacillus anthracis. It is essential to cure pXO1 from B. anthracis to evaluate its role in the pathogenesis of anthrax infection. Because conventional methods for curing plasmids (e.g., curing agents or growth at elevated temperatures) can induce mutations in the host chromosomal DNA, we developed a specific and reliable method to eliminate pXO1 from B. anthracis using plasmid incompatibility. Three putative replication origins of pXO1 were inserted into a temperature-sensitive plasmid to generate three incompatible plasmids. One of the three plasmids successfully eliminated the large plasmid pXO1 from B. anthracis vaccine strain A16R and wild type strain A16. These findings provided additional information about the replication/partitioning of pXO1 and demonstrated that introducing a small incompatible plasmid can generate plasmid-cured strains of B. anthracis without inducing spontaneous mutations in the host chromosome.  相似文献   

17.
A complete physical map of a Bacillus thuringiensis chromosome.   总被引:16,自引:4,他引:12       下载免费PDF全文
Bacillus thuringiensis is the source of the most widely used biological pesticide, through its production of insecticidal toxins. The toxin genes are often localized on plasmids. We have constructed a physical map of a Bacillus thuringiensis chromosome by aligning 16 fragments obtained by digestion with the restriction enzyme NotI. The fragments ranged from 15 to 1,350 kb. The size of the chromosome was 5.4 Mb. The NotI DNA fingerprint patterns of 12 different B. thuringiensis strains showed marked variation. The cryIA-type toxin gene was present on the chromosome in four strains, was extrachromosomal in four strains, and was both chromosomal and extrachromosomal in two strains. A Tn4430 transposon probe hybridized to 5 of the 10 cryIA-positive chromosomal fragments, while cryIA and the transposon often hybridized to different extrachromosomal bands. Ten of the strains were hemolytic when grown on agar plates containing human erythrocytes. Nine of the strains were positive when assayed for the presence of Bacillus cereus enterotoxin. We conclude that B. thuringiensis is very closely related to B. cereus and that the distinction between B. cereus and B. thuringiensis should be reconsidered.  相似文献   

18.
A 3.7-kilobase (kb) XbaI fragment harboring the cryIVB gene (L. Thorne, F. Garduno, T. Thompson, D. Decker, M. A. Zounes, M. Wild, A. M. Walfield, and T. J. Pollock, J. Bacteriol. 166:801-811, 1986) which encoded a 130-kilodalton (kDa) mosquitocidal toxin from a 110-kb plasmid of Bacillus thuringiensis subsp. israelensis 4Q2-72 was cloned into pUC12 and transformed into Escherichia coli. The clone with a recombinant plasmid (designated pBT8) was toxic to Aedes aegypti larvae. The fragment (3.7 kb) was ligated into pBC16 (tetracycline resistant [Tcr]) and transformed by the method of protoplast transformation into Bacillus sphaericus 1593 and 2362, which were highly toxic to Anopheles and Culex mosquito larvae but less toxic to Aedes larvae. After cell regeneration on regeneration medium, the Tcr plasmids from transformants (pBTC1) of both strains of B. sphaericus were prepared and analyzed. The 3.7-kb XbaI fragment from the B. thuringiensis subsp. israelensis plasmid was shown to be present by agarose gel electrophoresis and Southern blot hybridization. In addition, B. sphaericus transformants produced a 130-kDa mosquitocidal toxin which was detected by Western (immuno-) blot analysis with antibody prepared against B. thuringiensis subsp. israelensis 130-kDa mosquitocidal toxin. The 50% lethal concentrations of the transformants of strains 1593 and 2362 against A. aegypti larvae were 2.7 X 10(2) and 5.7 X 10(2) cells per ml, respectively. This level of toxicity was comparable to the 50% lethal concentration of B. thuringiensis subsp. israelensis but much higher than that of B. sphaericus 1593 and 2362 (4.7 X 10(4) cells per ml) against A. aegypti larvae.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

19.
A 3.7-kilobase (kb) XbaI fragment harboring the cryIVB gene (L. Thorne, F. Garduno, T. Thompson, D. Decker, M. A. Zounes, M. Wild, A. M. Walfield, and T. J. Pollock, J. Bacteriol. 166:801-811, 1986) which encoded a 130-kilodalton (kDa) mosquitocidal toxin from a 110-kb plasmid of Bacillus thuringiensis subsp. israelensis 4Q2-72 was cloned into pUC12 and transformed into Escherichia coli. The clone with a recombinant plasmid (designated pBT8) was toxic to Aedes aegypti larvae. The fragment (3.7 kb) was ligated into pBC16 (tetracycline resistant [Tcr]) and transformed by the method of protoplast transformation into Bacillus sphaericus 1593 and 2362, which were highly toxic to Anopheles and Culex mosquito larvae but less toxic to Aedes larvae. After cell regeneration on regeneration medium, the Tcr plasmids from transformants (pBTC1) of both strains of B. sphaericus were prepared and analyzed. The 3.7-kb XbaI fragment from the B. thuringiensis subsp. israelensis plasmid was shown to be present by agarose gel electrophoresis and Southern blot hybridization. In addition, B. sphaericus transformants produced a 130-kDa mosquitocidal toxin which was detected by Western (immuno-) blot analysis with antibody prepared against B. thuringiensis subsp. israelensis 130-kDa mosquitocidal toxin. The 50% lethal concentrations of the transformants of strains 1593 and 2362 against A. aegypti larvae were 2.7 X 10(2) and 5.7 X 10(2) cells per ml, respectively. This level of toxicity was comparable to the 50% lethal concentration of B. thuringiensis subsp. israelensis but much higher than that of B. sphaericus 1593 and 2362 (4.7 X 10(4) cells per ml) against A. aegypti larvae.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

20.
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