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 46 毫秒
1.
A combined physical and genetic map of theCorynebacterium glutamicum ATCC 13032 chromosome was constructed using pulsed-field gel electrophoresis (PFGE) and hybridizations with cloned gene probes. Total genomic DNA was digested with the meganucleasesSwaI (5′-ATTTAAAT-3′),PacI (5′-TTAATTAA-3′), andPmeI (5′-GTTTAAAC-3′) yielding 26, 27, and 23 fragments, respectively. The chromosomal restriction fragments were then separated by PFGE. By summing up the lengths of the fragments generated with each of the three enzymes, a genome size of 3082 +/- 20 kb was determined. To identify adjacentSwaI fragments, a genomic cosmid library ofC. glutamicum was screened for chromosomal inserts containingSwaI sites. Southern blots of the PFGE gels were hybridized with these linking clones to connect theSwaI fragments in their natural order. By this method, about 90% of the genome could be ordered into three contigs. Two of the remaining gaps were closed by cross-hybridization of blottedSwaI digests using as probesPacI andPmeI fragments isolated from PFGE gels. The last gap in the chromosomal map was closed by hybridization experiments using partialSwaI digestions, thereby proving the circularity of the chromosome. By hybridization of gene probes toSwaI fragments separated by PFGE about 30 genes, including rRNA operons, IS element and transposon insertions were localized on the physical map.  相似文献   

2.
A Tn5-derived mobile element has been constructed to identify genes and promoters related to pathogenesis and virulence inPseudomonas syringaepv.phaseolicola.To enhance the rate of mutation this Tn5derivative was constructed carrying a mutant transposase which was placed incisto the transposable element, but just outside the inverted repeats, therefore eliminating secondary transposition and increasing the stability of the insertion. The new element also contains a promoterlesscat(chloramphenicol acetyltransferase) gene as reporter to allow for positive selection of promoters being expressed under specific conditions. To facilitate cloning and manipulations inEscherichia coli,a ColE1 origin of replication has been included within the transposable element as well as the Mob region from the broad-host-range plasmid RP4, which allows this element to be efficiently mobilized by a triparental mating or by using anE. colistrain such as S17-1 to provide thetrafunctions. Sites for the rare cuttersPacI andPmeI have also been included to facilitate locating the insertions on aPacI and/orPmeI physical map. This construction combines the properties of both a mobilizable plasmid and a transposon and therefore has been termed pTn5cat.It is almost the same size as the wild-type Tn5, 5877 bp, and has successfully been tested inP.s. phaseolicolaandXanthomonas campestrispv.campestris.  相似文献   

3.
A combined physical and genetic map of theCorynebacterium glutamicum ATCC 13032 chromosome was constructed using pulsed-field gel electrophoresis (PFGE) and hybridizations with cloned gene probes. Total genomic DNA was digested with the meganucleasesSwaI (5-ATTTAAAT-3),PacI (5-TTAATTAA-3), andPmeI (5-GTTTAAAC-3) yielding 26, 27, and 23 fragments, respectively. The chromosomal restriction fragments were then separated by PFGE. By summing up the lengths of the fragments generated with each of the three enzymes, a genome size of 3082 +/- 20 kb was determined. To identify adjacentSwaI fragments, a genomic cosmid library ofC. glutamicum was screened for chromosomal inserts containingSwaI sites. Southern blots of the PFGE gels were hybridized with these linking clones to connect theSwaI fragments in their natural order. By this method, about 90% of the genome could be ordered into three contigs. Two of the remaining gaps were closed by cross-hybridization of blottedSwaI digests using as probesPacI andPmeI fragments isolated from PFGE gels. The last gap in the chromosomal map was closed by hybridization experiments using partialSwaI digestions, thereby proving the circularity of the chromosome. By hybridization of gene probes toSwaI fragments separated by PFGE about 30 genes, including rRNA operons, IS element and transposon insertions were localized on the physical map.  相似文献   

4.
A 13-kb DNA fragment containing oriC and the flanking genes thdF, orf900, yidC, rnpA, rpmH, oriC, dnaA, dnaN, recF, and gyrB was cloned from the gram-negative plant pathogen Xanthomonas campestris pv. campestris 17. These genes are conserved in order with other eubacterial oriC genes and code for proteins that share high degrees of identity with their homologues, except for orf900, which has a homologue only in Xylella fastidiosa. The dnaA/dnaN intergenic region (273 bp) identified to be the minimal oriC region responsible for autonomous replication has 10 pure AT clusters of four to seven bases and only three consensus DnaA boxes. These findings are in disagreement with the notion that typical oriCs contain four or more DnaA boxes located upstream of the dnaA gene. The X. campestris pv. campestris 17 attB site required for site-specific integration of cloned fragments from filamentous phage Lf replicative form DNA was identified to be a dif site on the basis of similarities in nucleotide sequence and function with the Escherichia coli dif site required for chromosome dimer resolution and whose deletion causes filamentation of the cells. The oriC and dif sites were located at 12:00 and 6:00, respectively, on the circular X. campestris pv. campestris 17 chromosome map, similar to the locations found for E. coli sites. Computer searches revealed the presence of both the dif site and XerC/XerD recombinase homologues in 16 of the 42 fully sequenced eubacterial genomes, but eight of the dif sites are located far away from the 6:00 point instead of being placed opposite the cognate oriC. The differences in the relative position suggest that mechanisms different from that of E. coli may participate in the control of chromosome replication.  相似文献   

5.
Plant pathogen Xanthomonas campestris pv. campestris produces cis-11-methyl-2-dodecenoic acid (diffusible signal factor [DSF]) as a cell-cell communication signal to regulate biofilm dispersal and virulence factor production. Previous studies have demonstrated that DSF biosynthesis is dependent on the presence of RpfF, an enoyl-coenzyme A (CoA) hydratase, but the DSF synthetic mechanism and the influence of the host plant on DSF biosynthesis are still not clear. We show here that exogenous addition of host plant juice or ethanol extract to the growth medium of X. campestris pv. campestris could significantly boost DSF family signal production. It was subsequently revealed that X. campestris pv. campestris produces not only DSF but also BDSF (cis-2-dodecenoic acid) and another novel DSF family signal, which was designated DSF-II. BDSF was originally identified in Burkholderia cenocepacia to be involved in regulation of motility, biofilm formation, and virulence in B. cenocepacia. Functional analysis suggested that DSF-II plays a role equal to that of DSF in regulation of biofilm dispersion and virulence factor production in X. campestris pv. campestris. Furthermore, chromatographic separation led to identification of glucose as a specific molecule stimulating DSF family signal biosynthesis in X. campestris pv. campestris. 13C-labeling experiments demonstrated that glucose acts as a substrate to provide a carbon element for DSF biosynthesis. The results of this study indicate that X. campestris pv. campestris could utilize a common metabolite of the host plant to enhance DSF family signal synthesis and therefore promote virulence.  相似文献   

6.
Variations in the outer membrane proteins (OMPs) and lipopolysaccharides (LPSs) of 54 isolates belonging to 16 different pathovars of Xanthomonas campestris were characterized. OMP samples prepared by sarcosyl extraction of cell walls and LPS samples prepared by proteinase K treatment of sonicated cells were analyzed by sodium dodecyl sulfate-polyacrylamide gel electrophoresis in the presence of 4 M urea. In general, the OMP and LPS profiles within each pathovar were very similar but different from the profiles of other pathovars. Heterogeneity in OMP and LPS profiles was observed within X. campestris pv. campestris, X. campestris pv. translucens, and X. campestris pv. vesicatoria. LPSs were isolated from six X. campestris pathovars, which fell into two major groups on the basis of O antigenicity. The O antigens of X. campestris pv. begoniae, X. campestris pv. graminis, and X. campestris pv. translucens cross-reacted with each other; the other group consisted of X. campestris pv. campestris, X. campestris pv. pelargonii, and X. campestris pv. vesicatoria. A chemical analysis revealed a significant difference between the compositions of the neutral sugars of the LPSs of those two groups; the LPSs of the first group contained xylose and a 6-deoxy-3-O-methyl hexose, whereas the LPSs of the other group lacked both sugars.  相似文献   

7.
Xanthomonas campestris strains that cause disease in citrus were compared by restriction endonuclease analysis of DNA fragments separated by pulsed-field gel electrophoresis and by DNA reassociation. Strains of X. campestris pv. citrumelo, which cause citrus bacterial spot, were, on average, 88% related to each other by DNA reassociation, although these strains exhibited diverse restriction digest patterns. In contrast, strains of X. campestris pv. citri groups A and B, which cause canker A and canker B, respectively, had relatively homogeneous restriction digest patterns. The groups of strains causing these three different citrus diseases were examined by DNA reassociation and were found to be from 55 to 63% related to one another. Several pathovars of X. campestris, previously shown to cause weakly aggressive symptoms on citrus, ranged from 83 to 90% similar to X. campestris pv. citrumelo by DNA reassociation. The type strain of X. campestris pv. campestris ranged from 30 to 40% similar in DNA reassociation experiments to strains of X. campestris pv. citrumelo and X. campestris pv. citri groups A and B. Whereas DNA reassociation quantified the difference between relatively unrelated groups of bacterial strains, restriction endonuclease analysis distinguished between closely related strains.  相似文献   

8.
The chemotaxis towards sucrose and yeast extract of nine strains of Xanthomonas campestris representing pathovars campestris, armoraciae, translucens, vesicatoria, and pelargonii was analyzed by using swarm plates. Unexpectedly, each of these strains formed small or reduced swarms typical of nonmotile or nonchemotactic bacteria. With time, however, chemotactic cells appeared on the swarm plates as blebs of bacteria. These cells were strongly chemotactic and were concomitantly deficient in exopolysaccharide production. The switch from the wild type (exopolysaccharide producing and nonchemotactic) to the swarmer type (exopolysaccharide deficient and chemotactic) appeared irreversible ex planta in bacteriological medium. However, in radish leaves swarmer-type strains of X. campestris pv. campestris were able to revert to the wild type. Swarmer-type derivatives of two X. campestris pv. campestris wild-type isolates showed reduced virulence and growth in the host plants cauliflower and radish. However, exocellular complementation of X. campestris pv. campestris Hrp (nonpathogenic) mutant was achieved by coinoculation with a swarmer-type strain.  相似文献   

9.
Xanthomonas campestris pathovar campestris causes black rot, a vascular disease on cruciferous plants, including Arabidopsis thaliana. The gene XC1553 from X. campestris pv. campestris strain 8004 encodes a protein containing leucine-rich repeats (LRRs) and appears to be restricted to strains of X. campestris pv. campestris. LRRs are found in a number of type III-secreted effectors in plant and animal pathogens. These prompted us to investigate the role of the XC1553 gene in the interaction between X. campestris pv. campestris and A. thaliana. Translocation assays using the hypersensitive-reaction-inducing domain of X. campestris pv. campestris AvrBs1 as a reporter revealed that XC1553 is a type III effector. Infiltration of Arabidopsis leaf mesophyll with bacterial suspensions showed no differences between the wild-type strain and an XC1553 gene mutant; both strains induced disease symptoms on Kashmir and Col-0 ecotypes. However, a clear difference was observed when bacteria were introduced into the vascular system by piercing the central vein of leaves. In this case, the wild-type strain 8004 caused disease on the Kashmir ecotype, but not on ecotype Col-0; the XC1553 gene mutant became virulent on the Col-0 ecotype and still induced disease on the Kashmir ecotype. Altogether, these data show that the XC1553 gene, which was renamed avrACXcc8004, functions as an avirulence gene whose product seems to be recognized in vascular tissues.  相似文献   

10.
A combined physical and genetic map of the Pseudomonas putida KT2440 genome was constructed from data obtained by pulsed-field gel electrophoresis techniques (PFGE) and Southern hybridization. Circular genome size was estimated at 6.0 Mb by adding the sizes of 19 SwaI, 9 PmeI, 6 PacI, and 6 I-CeuI fragments. A complete physical map was achieved by combining the results of (i) analysis of PFGE of the DNA fragments resulting from digestion of the whole genome with PmeI, SwaI, I-CeuI, and PacI as well as double digestion with combinations of these enzymes and (ii) Southern hybridization analysis of the whole wild-type genome digested with different enzymes and hybridized against a series of probes obtained as cloned genes from different pseudomonads of rRNA group I and Escherichia coli, as P. putida DNA obtained by PCR amplification based on sequences deposited at the GenBank database, and by labeling of macrorestriction fragments of the P. putida genome eluted from agarose gels. As an alternative, 10 random mini-Tn5-Km mutants of P. putida KT2440 were used as a source of DNA, and the band carrying the mini-Tn5 in each mutant was identified after PFGE of a series of complete chromosomal digestions and hybridization with the kanamycin resistance gene of the mini-Tn5 as a probe. We established a circular genome map with an average resolution of 160 kb. Among the 63 genes located on the genetic map were key markers such as oriC, 6 rrn loci (rnnA to -F), recA, ftsZ, rpoS, rpoD, rpoN, and gyrB; auxotrophic markers; and catabolic genes for the metabolism of aromatic compounds. The genetic map of P. putida KT2440 was compared to those of Pseudomonas aeruginosa PAO1 and Pseudomonas fluorescens SBW25. The chromosomal backbone revealed some similarity in gene clustering among the three pseudomonads but differences in physical organization, probably as a result of intraspecific rearrangements.  相似文献   

11.
Plasmid-borne copper resistance genes from copper-resistant strains of Xanthomonas campestris pv. vesicatoria from California, Florida, and Oklahoma shared structural similarities. A strain of X. campestris pv. campestris also contained plasmid-borne copper resistance genes similar to the resistance genes from X. campestris pv. vesicatoria. Furthermore, a region of the copper resistance genes from X. campestris pv. vesicatoria 07882 hybridized with copA, the first gene of the copper resistance operon (cop) of Pseudomonas syringae pv. tomato. A copper-inducible protein of similar size to CopA was detected by Western blot (immunoblot) analysis from the wild-type strain 07882 and from the cloned copper resistance genes of 07882 introduced into a copper-sensitive strain of X. campestris pv. vesicatoria. A low level of hybridization was observed with chromosomal DNA from other xanthomonads when the copper resistance genes from strain 07882 were used as probes.  相似文献   

12.
The efficacy of copper bactericides for control of Xanthomonas campestris pv. vesicatoria in eastern Oklahoma tomato fields was evaluated. Copper bactericides did not provide adequate control, and copper-resistant (Cur) strains of the pathogen were isolated. The Cur genes in these strains were located on a large indigenous plasmid designated pXV10A. The host range of pXV10A was investigated; this plasmid was efficiently transferred into 8 of 11 X. campestris pathovars. However, the transfer of pXV10A to other phytopathogenic genera was not detected. DNA hybridization experiments were performed to characterize the Cur genes on pXV10A. A probe containing subcloned Cur genes from X. campestris pv. vesicatoria E3C5 hybridized to pXV10A; however, a subclone containing Cur genes from P. syringae pv. tomato PT23 failed to hybridize to pXV10A. Further DNA hybridization experiments were performed to compare pXV10A with pXvCu plasmids, a heterogenous group of Cur plasmids present in strains of X. campestris pv. vesicatoria from Florida. These studies indicated that the Cur genes on pXV10A and pXvCu plasmids share nucleotide sequence homology and may have a common origin. Further experiments showed that these plasmids are distinctly different because pXV10A did not contain sequences homologous to IS476, an insertion sequence present on pXvCu plasmids.  相似文献   

13.
《Gene》1998,206(2):223-228
A physical map of the Zymomonas mobilis ZM4 genome has been constructed from the results of reciprocal Southern hybridization with PmeI, PacI, and NotI-digested genomic DNA fragments and linking cosmid clones. Restriction enzyme-digested Z. mobilis ZM4 genome was electrophoresed with phage lambda DNA concatemers as a size standard in a Bio-Rad CHEF-DRII pulsed-field gel electrophoresis (PFGE) system. The restriction enzyme PmeI generated 15 fragments (3–625 kb), and PacI produced 19 fragments (7–525 kb). Each size of restriction fragment was calculated by comparison to the size of phage lambda DNA concatemers, and the genome size of Z. mobilis ZM4 was estimated to be 2085.5 kb. The 19 known genes and three rrn operons were localized on the map.  相似文献   

14.
A system allowing transposon mutagenesis of cloned DNA fragments in Escherichia coli with Tn4431, which carries the promotorless luciferase (lux) operon of Vibrio fischeri, has been developed. The transposon delivery plasmid, pDS1, based on an IncF replicon, is thermosensitive in replication and mobilizable to many Gram-negative bacteria. We used pDS1 for Tn4431-saturation mutagenesis of a 10-kb DNA fragment of Xanthomonas campestris pv. campestris (X.c.c.) in E. coli and showed that the expression of the lux operon was dependent on orientation and location of the transposon. Transfer of a specific Tn4431 insertion to X.c.c. allowed the determination of the bioluminescence phenotype in planta. Correspondence to: U. B. Priefer  相似文献   

15.
Thirty-two plasmid insertion mutants were independently isolated from two strains of Xanthomonas campestris pv. campestris in Taiwan. Of the 32 mutants, 14 (44%), 8 (25%), and 4 (12%) mutants resulted from separate insertions of an IS3 family member, IS476, and two new insertion sequences (IS), IS1478 and IS1479. While IS1478 does not have significant sequence homology with any IS elements in the EMBL/GenBank/DDBJ database, IS1479 demonstrated 73% sequence homology with IS1051 in X. campestris pv. dieffenbachiae, 62% homology with IS52 in Pseudomonas syringae pv. glycinea, and 60% homology with IS5 in Escherichia coli. Based on the predicted transposase sequences as well as the terminal nucleotide sequences, IS1478 by itself constitutes a new subfamily of the widespread IS5 family, whereas IS1479, along with IS1051, IS52, and IS5, belongs to the IS5 subfamily of the IS5 family. All but one of the IS476 insertions had duplications of 4 bp at the target sites without sequence preference and were randomly distributed. An IS476 insertion carried a duplication of 952 bp at the target site. A model for generating these long direct repeats is proposed. Insertions of IS1478 and IS1479, on the other hand, were not random, and IS1478 and IS1479 each showed conservation of PyPuNTTA and PyTAPu sequences (Py is a pyrimidine, Pu is a purine, and N is any nucleotide) for duplications at the target sites. The results of Southern blot hybridization analysis indicated that multiple copies of IS476, IS1478, and IS1479 are present in the genomes of all seven X. campestris pv. campestris strains tested and several X. campestris pathovars.  相似文献   

16.
Inoculation of mature leaves of turnip (Brassica campestris) with the incompatible Xanthomonas campestris pv vitians resulted in the induction of β-1,3-glucanase and chitinase/lysozyme (CHL) activity. No increase in the basal activity of β-1,3-glucanase was observed after inoculation of leaves with heat- or rifampicin-killed X. c. vitians, Escherichia coli, or sterile water. Inoculation with the compatible X. campestris pv campestris resulted in a slower induction of glucanase than that seen with X. c. vitians. In contrast, all bacteria caused an induction of CHL activity. One major β-1,3-glucanase (molecular mass 36.5 kilodaltons, isoelectric point [pl] ~8.5) was purified from both inoculated and untreated leaves by ion-exchange chromatography. The enzyme degraded laminarin by an endo-glycolytic mechanism. Two major CHL isozymes (CHL 1 and CHL 2, molecular mass 30 kilodaltons and pl 9.4 and 10.2, respectively) were purified from X. c. vitians inoculated leaves by affinity chromatography on a chitin column followed by ion-exchange chromatography. Both enzymes degraded chitin by an endo-glycolytic mechanism although the ratio of lysozyme to chitinase specific activities for CHL 1 and CHL2 were different. The induction of CHL 1 was associated with the hypersensitive reaction caused by X. c. vitians whereas all other treatments induced largely CHL 2.  相似文献   

17.
A 13-kb DNA fragment containing oriC and the flanking genes thdF, orf900, yidC, rnpA, rpmH, oriC, dnaA, dnaN, recF, and gyrB was cloned from the gram-negative plant pathogen Xanthomonas campestris pv. campestris 17. These genes are conserved in order with other eubacterial oriC genes and code for proteins that share high degrees of identity with their homologues, except for orf900, which has a homologue only in Xylella fastidiosa. The dnaA/dnaN intergenic region (273 bp) identified to be the minimal oriC region responsible for autonomous replication has 10 pure AT clusters of four to seven bases and only three consensus DnaA boxes. These findings are in disagreement with the notion that typical oriCs contain four or more DnaA boxes located upstream of the dnaA gene. The X. campestris pv. campestris 17 attB site required for site-specific integration of cloned fragments from filamentous phage phiLf replicative form DNA was identified to be a dif site on the basis of similarities in nucleotide sequence and function with the Escherichia coli dif site required for chromosome dimer resolution and whose deletion causes filamentation of the cells. The oriC and dif sites were located at 12:00 and 6:00, respectively, on the circular X. campestris pv. campestris 17 chromosome map, similar to the locations found for E. coli sites. Computer searches revealed the presence of both the dif site and XerC/XerD recombinase homologues in 16 of the 42 fully sequenced eubacterial genomes, but eight of the dif sites are located far away from the 6:00 point instead of being placed opposite the cognate oriC. The differences in the relative position suggest that mechanisms different from that of E. coli may participate in the control of chromosome replication.  相似文献   

18.
Previous studies have indicated that the yellow pigments (xanthomonadins) produced by phytopathogenic Xanthomonas bacteria are unimportant during pathogenesis but may be important for protection against photobiological damage. We used a Xanthomonas campestris pv. campestris parent strain, single-site transposon insertion mutant strains, and chromosomally restored mutant strains to define the biological role of xanthomonadins. Although xanthomonadin mutant strains were comparable to the parent strain for survival when exposed to UV light; after their exposure to the photosensitizer toluidine blue and visible light, survival was greatly reduced. Chromosomally restored mutant strains were completely restored for survival in these conditions. Likewise, epiphytic survival of a xanthomonadin mutant strain was greatly reduced in conditions of high light intensity, whereas a chromosomally restored mutant strain was comparable to the parent strain for epiphytic survival. These results are discussed with respect to previous results, and a model for epiphytic survival of X. campestris pv. campestris is presented.  相似文献   

19.
Polymerase chain reaction (PCR) amplification of total DNA was evaluated as a method to distinguish Xanthomonas campestris pv. pelargonii from other pathovars within this species. Two sets of highly conserved enterobacterial consensus sequences were used as targets for PCR amplification: (a) enterobacterial repetitive intergenic consensus [ERIC] and (b) repetitive extragenic palindromic [REP] sequences. Nucleic acid was extracted from a total of 37 isolates of bacteria: 19 isolates ofX campestris pv. pelargonii and 18 isolates representing 10 other pathovars of X. campestris. After PCR amplification using the ERIC primer pair the DNA fingerprints of X. campestris pv, pelargonii contained two major DNA products (estimated size 500 and 740 pp) that were conserved among all 19 isolates. With the REP primer pair, the fingerprints were more complex and major DNA products ranging from -690 to 1650 bp were detected. Using information from both ERIC- and REP-primed Imgerprints, the X. campestris pv. pelargonii fingerprints were distinguishable from the fingerprints of the other pathovars examined: pvs. citrumelo. citri, beganiae, vittans B and C. phaseoli. campestris, manihotis, juglandis, carotae and pruni.  相似文献   

20.
Host factors that are important for infection of Xanthomonas campestris pv. citri by the filamentous bacteriophage cf were investigated by transposon mutagenesis with Tn5tac1. A mutant, XT501, that was resistant to cf infection was recovered, showing that the gene inactivated by the transposon is required for infection by the phage but not for cf replication or assembly. A 1.7-kb SacI-ApaI DNA fragment from XT501 containing the bacterial DNA flanking one end of the transposon was cloned and shown to be required for cf infection. Nucleotide sequence analysis of the 1.7-kb fragment reveals the presence of an ORF that encodes a protein of 146 amino acids. This protein shows 42% identity to the type 4 prepilin encoded by the pilA genes of other bacteria. The pilA gene of X. campestris pv. citri is thus essential for infection by the bacteriophage cf.  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