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 15 毫秒
1.
The leaf colonization strategies of two bacterial strains were investigated. The foliar pathogen Pseudomonas syringae pv. syringae strain B728a and the nonpathogen Pantoea agglomerans strain BRT98 were marked with a green fluorescent protein, and surface (epiphytic) and subsurface (endophytic) sites of bean and maize leaves in the laboratory and the field were monitored to see if populations of these strains developed. The populations were monitored using both fluorescence microscopy and counts of culturable cells recovered from nonsterilized and surface-sterilized leaves. The P. agglomerans strain exclusively colonized epiphytic sites on the two plant species. Under favorable conditions, the P. agglomerans strain formed aggregates that often extended over multiple epidermal cells. The P. syringae pv. syringae strain established epiphytic and endophytic populations on asymptomatic leaves of the two plant species in the field, with most of the P. syringae pv. syringae B728a cells remaining in epiphytic sites of the maize leaves and an increasing number occupying endophytic sites of the bean leaves in the 15-day monitoring period. The epiphytic P. syringae pv. syringae B728a populations appeared to originate primarily from multiplication in surface sites rather than from the movement of cells from subsurface to surface sites. The endophytic P. syringae pv. syringae B728a populations appeared to originate primarily from inward movement through the stomata, with higher levels of multiplication occurring in bean than in maize. A rainstorm involving a high raindrop momentum was associated with rapid growth of the P. agglomerans strain on both plant species and with rapid growth of both the epiphytic and endophytic populations of the P. syringae pv. syringae strain on bean but not with growth of the P. syringae pv. syringae strain on maize. These results demonstrate that the two bacterial strains employed distinct colonization strategies and that the epiphytic and endophytic population dynamics of the pathogenic P. syringae pv. syringae strain were dependent on the plant species, whereas those of the nonpathogenic P. agglomerans strain were not.  相似文献   

2.
The genome sequence of more than 100 Pseudomonas syringae strains has been sequenced to date; however only few of them have been fully assembled, including P. syringae pv. syringae B728a. Different strains of pv. syringae cause different diseases and have different host specificities; so, UMAF0158 is a P. syringae pv. syringae strain related to B728a but instead of being a bean pathogen it causes apical necrosis of mango trees, and the two strains belong to different phylotypes of pv.syringae and clades of P. syringae. In this study we report the complete sequence and annotation of P. syringae pv. syringae UMAF0158 chromosome and plasmid pPSS158. A comparative analysis with the available sequenced genomes of other 25 P. syringae strains, both closed (the reference genomes DC3000, 1448A and B728a) and draft genomes was performed. The 5.8 Mb UMAF0158 chromosome has 59.3% GC content and comprises 5017 predicted protein-coding genes. Bioinformatics analysis revealed the presence of genes potentially implicated in the virulence and epiphytic fitness of this strain. We identified several genetic features, which are absent in B728a, that may explain the ability of UMAF0158 to colonize and infect mango trees: the mangotoxin biosynthetic operon mbo, a gene cluster for cellulose production, two different type III and two type VI secretion systems, and a particular T3SS effector repertoire. A mutant strain defective in the rhizobial-like T3SS Rhc showed no differences compared to wild-type during its interaction with host and non-host plants and worms. Here we report the first complete sequence of the chromosome of a pv. syringae strain pathogenic to a woody plant host. Our data also shed light on the genetic factors that possibly determine the pathogenic and epiphytic lifestyle of UMAF0158. This work provides the basis for further analysis on specific mechanisms that enable this strain to infect woody plants and for the functional analysis of host specificity in the P. syringae complex.  相似文献   

3.
In order to cope with pathogens, plants have evolved sophisticated mechanisms to sense pathogenic attacks and to induce defence responses. The N‐acyl‐homoserine lactone (AHL)‐mediated quorum sensing in bacteria regulates diverse physiological processes, including those involved in pathogenicity. In this work, we study the interactions between AHL‐producing transgenic tobacco plants and Pseudomonas syringae pv. tabaci 11528 (P. syringae 11528). Both a reduced incidence of disease and decrease in the growth of P. syringae 11528 were observed in AHL‐producing plants compared with wild‐type plants. The present data indicate that plant‐produced AHLs enhance disease resistance against this pathogen. Subsequent RNA‐sequencing analysis showed that the exogenous addition of AHLs up‐regulated the expression of P. syringae 11528 genes for flagella production. Expression levels of plant defence genes in AHL‐producing and wild‐type plants were determined by quantitative real‐time polymerase chain reaction. These data showed that plant‐produced AHLs activated a wide spectrum of defence responses in plants following inoculation, including the oxidative burst, hypersensitive response, cell wall strengthening, and the production of certain metabolites. These results demonstrate that exogenous AHLs alter the gene expression patterns of pathogens, and plant‐produced AHLs either directly or indirectly enhance plant local immunity during the early stage of plant infection.  相似文献   

4.
Mutations in the global regulatory genes gacS and gacA render Pseudomonas syringae pv. syringae strain B728a completely nonpathogenic in foliar infiltration assays on bean plants. It had been previously demonstrated that gac genes regulate alginate production in Pseudomonas species, while other published work indicated that alginate is involved in the pathogenic interaction of P. syringae on bean plants. Together, these results suggested that the effects of gacS and gacA mutations on virulence in B728a might stem directly from a role in regulating alginate. In this report, we confirm a role for gac genes in both algD expression and alginate production in B728a. However, B728a mutants completely devoid of detectable alginate were as virulent as the wild-type strain in our assay. Thus, factors other than, or in addition to, a deficiency of alginate must be involved in the lack of pathogenicity observed with gacS and gacA mutants.  相似文献   

5.
Foliar bacterial pathogens have to penetrate the plant tissue and access the interior of the apoplast in order to initiate the pathogenic phase. The entry process is driven by chemotaxis towards plant‐derived compounds in order to locate plant openings. However, information on plant signals recognized by bacterial chemoreceptors is scarce. Here, we show that the perception of GABA and l‐Pro, two abundant components of the tomato apoplast, through the PsPto‐PscC chemoreceptor drives the entry of Pseudomonas syringae pv. tomato into the tomato apoplast. The recognition of both compounds by PsPto‐PscC caused chemoattraction to both amino acids and participated in the regulation of GABA catabolism. Mutation of the PsPto‐PscC chemoreceptor caused a reduced chemotactic response towards these compounds which in turn impaired entry and reduced virulence in tomato plants. Interestingly, GABA and l‐Pro levels significantly increase in tomato plants upon pathogen infection and are involved in the regulation of the plant defence response. This is an example illustrating how bacteria respond to plant signals produced during the interaction as cues to access the plant apoplast and to ensure efficient infection.  相似文献   

6.
The epiphyte Pseudomonas syringae pv. syringae 22d/93 (Pss22d) produces the rare amino acid 3-methylarginine (MeArg), which is highly active against the closely related soybean pathogen Pseudomonas syringae pv. glycinea. Since these pathogens compete for the same habitat, Pss22d is a promising candidate for biocontrol of P. syringae pv. glycinea. The MeArg biosynthesis gene cluster codes for the S-adenosylmethionine (SAM)-dependent methyltransferase MrsA, the putative aminotransferase MrsB, and the amino acid exporter MrsC. Transfer of the whole gene cluster into Escherichia coli resulted in heterologous production of MeArg. The methyltransferase MrsA was overexpressed in E. coli as a His-tagged protein and functionally characterized (Km, 7 mM; kcat, 85 min−1). The highly selective methyltransferase MrsA transfers the methyl group from SAM into 5-guanidino-2-oxo-pentanoic acid to yield 5-guanidino-3-methyl-2-oxo-pentanoic acid, which then only needs to be transaminated to result in the antibiotic MeArg.Microbial plant pathogens cause severe losses in agriculture each year (1). For example, the plant pathogen Pseudomonas syringae pv. glycinea is responsible for bacterial blight of soybean, a leaf spot disease of great economic impact. Besides chemical treatment, biocontrol agents that antagonize microbial plant pathogens are gaining increasing importance in fighting plant diseases (6, 11, 27). In a screening for possible biocontrol strains, an epiphytic bacterium showing a strong and selective activity against the pathogen P. syringae pv. glycinea was isolated from soybean leaves (29). The strain was characterized as Pseudomonas syringae pv. syringae 22d/93 (Pss22d). The antagonism of Pss22d against P. syringae pv. glycinea has been demonstrated successfully in vitro and in planta under greenhouse and field conditions (19, 29). In order to identify the molecular basis of the antagonism of Pss22d against P. syringae pv. glycinea, we focused on its secondary metabolites. Besides the well-known lipodepsipeptides syringomycin and syringopeptin (3), Pss22d produces the rare amino acid 3-methylarginine (MeArg) (5). As little as 20 nmol of MeArg strongly and selectively inhibits P. syringae pv. glycinea but no other pseudomonads in vitro (29). Since the inhibition can be compensated for by l-arginine supplementation but not by any other essential amino acid, it is likely that the toxin acts as an inhibitor of the arginine biosynthesis pathway or an arginine-dependent pathway, such as nitric oxide formation (13, 16). Feeding experiments and Tn5 transposon mutagenesis suggested that MeArg is produced by an S-adenosyl methionine (SAM)-dependent methyltransferase (5) converting the enol of 5-guanidino-2-oxo-pentanoic acid to 5-guanidino-3-methyl-2-oxo-pentanoic acid. An analogous reaction is known to occur with the methyltransferases GlmT, DptI, and LptI, which form 3-methylglutamate from α-ketoglutarate (18). On the way to MeArg, only a transaminase catalyzing the formation of MeArg from 5-guanidino-3-methyl-2-oxo-pentanoic acid and an amino acid exporter to secrete the toxin would be needed.Here, we describe the identification and functional characterization of the MeArg biosynthesis gene cluster from the epiphyte Pss22d.  相似文献   

7.
Many defense mechanisms contribute to the plant immune system against pathogens, involving the regulation of different processes of the primary and secondary metabolism. At the same time, pathogens have evolved mechanisms to hijack the plant defense in order to establish the infection and proliferate. Localization and timing of the host response are essential to understand defense mechanisms and resistance to pathogens (Rico et al. 2011). Imaging techniques, such as fluorescence imaging and thermography, are a very valuable tool providing spatial and temporal information about a series of plant processes. In this study, bean plants challenged with two pathovars of Pseudomonas syringae have been investigated. Pseudomonas syringae pv. phaseolicola 1448A and P. syringae pv. tomato DC3000 elicit a compatible and incompatible interaction in bean, respectively. Both types of host–pathogen interaction triggered different changes in the activity of photosynthesis and the secondary metabolism. We conclude that the combined analysis of leaf temperature, chlorophyll fluorescence and green fluorescence emitted by phenolics allows to discriminate compatible from incompatible P. syringae–Phaseolus vulgaris interactions in very early times of the infection, prior to the development of symptoms. These can constitute disease signatures that would allow an early identification of emerging plagues in crops.  相似文献   

8.
The epiphyte Pseudomonas syringae pv. syringae 22d / 93 (Pss22d), isolated from soybean leaves, had been characterized as a promising and species‐specific biocontrol strain in vitro and in planta against the plant pathogen P. syringae pv. glycinea (Psg), which causes bacterial blight of soybean. Three toxins are known to be produced by Pss22d: syringomycin, syringopeptin and 3‐methylarginine (MeArg). In contrast to syringopeptin and syringomycin, MeArg inhibited the growth of Psg in vitro. To examine if the toxins produced by Pss22d are responsible for antagonistic effects in planta, the pathogen Psg was co‐inoculated with either Pss22d wild‐type, a syringopeptin/syringomycin‐negative double mutant (Pss22d.ΔsypA/syrE), or a MeArg‐negative mutant (Pss22d.1) into wounds of pin‐pricked leaves of greenhouse‐grown soybean plants, respectively. In all three cases, the wild‐type Pss22d and its toxin‐deficient mutants prevented development of disease symptoms normally caused by Psg. These results indicated that neither syringopeptin, nor syringomycin, nor MeArg was required for Pss22d’s antagonistic activity in planta. Consequently, factors other than the three toxins may contribute to the intra‐species antagonism in planta.  相似文献   

9.
Bacteria and plant derived volatile organic compounds have been reported as the chemical triggers that elicit induced resistance in plants. Previously, volatile organic compounds (VOCs), including acetoin and 2,3-butanediol, were found to be emitted from plant growth-promoting rhizobacteria (PGPR) Bacillus subtilis GB03, which had been shown to elicit ISR and plant growth promotion. More recently, we reported data that stronger induced resistance could be elicited against Pseudomonas syringae pv maculicola ES4326 in plants exposed to C13 VOC from another PGPR Paenibacillus polymyxa E681 compared with that of strain GB03. Here, we assessed whether another long hydrocarbon C16 hexadecane (HD) conferred protection to Arabidopsis from infection of a biotrophic pathogen, P. syringae pv maculicola and a necrotrophic pathogen, Pectobacterium carotovorum subsp carotovorum. Collectively, long-chain VOCs can be linked to a plant resistance activator for protecting plants against both biotrophic and necrotrophic pathogens at the same time.  相似文献   

10.
In a recent screen for novel virulence factors involved in the interaction between Pseudomonas savastanoi pv. savastanoi and the olive tree, a mutant was selected that contained a transposon insertion in a putative cyclic diguanylate (c‐di‐GMP) phosphodiesterase‐encoding gene. This gene displayed high similarity to bifA of Pseudomonas aeruginosa and Pseudomonas putida. Here, we examined the role of BifA in free‐living and virulence‐related phenotypes of two bacterial plant pathogens in the Pseudomonas syringae complex, the tumour‐inducing pathogen of woody hosts, P. savastanoi pv. savastanoi NCPPB 3335, and the pathogen of tomato and Arabidopsis, P. syringae pv. tomato DC3000. We showed that deletion of the bifA gene resulted in decreased swimming motility of both bacteria and inhibited swarming motility of DC3000. In contrast, overexpression of BifA in P. savastanoi pv. savastanoi had a positive impact on swimming motility and negatively affected biofilm formation. Deletion of bifA in NCPPB 3335 and DC3000 resulted in reduced fitness and virulence of the microbes in olive (NCPPB 3335) and tomato (DC3000) plants. In addition, real‐time monitoring of olive plants infected with green fluorescent protein (GFP)‐tagged P. savastanoi cells displayed an altered spatial distribution of mutant ΔbifA cells inside olive knots compared with the wild‐type strain. All free‐living phenotypes that were altered in both ΔbifA mutants, as well as the virulence of the NCPPB 3335 ΔbifA mutant in olive plants, were fully rescued by complementation with P. aeruginosa BifA, whose phosphodiesterase activity has been demonstrated. Thus, these results suggest that P. syringae and P. savastanoi BifA are also active phosphodiesterases. This first demonstration of the involvement of a putative phosphodiesterase in the virulence of the P. syringae complex provides confirmation of the role of c‐di‐GMP signalling in the virulence of this group of plant pathogens.  相似文献   

11.
To construct differentially-marked derivatives of our model wild-type strain, Pseudomonas syringae pv. syringae B728a (a causal agent of bacterial brown spot disease in snap bean plants), for field experiments, we selected a site in the gacS-cysM intergenic region for site-directed insertion of antibiotic resistance marker cassettes. In each of three field experiments, population sizes of the site-directed chromosomally marked B728a derivatives in association with snap bean plants were not significantly different from that of the wild-type strain. Inserts of up to 7 kb of DNA in the intergenic region did not measurably affect fitness of B728a in the field. The site is useful for site-directed genomic insertions of single copies of genes of interest.  相似文献   

12.
Pseudomonas syringae pv. tomato, the causal agent of bacterial speck of tomato, and the plant growth-promoting bacterium Azospirillum brasilense were inoculated onto tomato plants, either alone, as a mixed culture, or consecutively. The population dynamics in the rhizosphere and foliage, the development of bacterial speck disease, and their effects on plant growth were monitored. When inoculated onto separate plants, the A. brasilense population in the rhizosphere of tomato plants was 2 orders of magnitude greater than the population of P. syringae pv. tomato (107 versus 105 CFU/g [dry weight] of root). Under mist chamber conditions, the leaf population of P. syringae pv. tomato was 1 order of magnitude greater than that of A. brasilense (107 versus 106 CFU/g [dry weight] of leaf). Inoculation of seeds with a mixed culture of the two bacterial strains resulted in a reduction of the pathogen population in the rhizosphere, an increase in the A. brasilense population, the prevention of bacterial speck disease development, and improved plant growth. Inoculation of leaves with the mixed bacterial culture under mist conditions significantly reduced the P. syringae pv. tomato population and significantly decreased disease severity. Challenge with P. syringae pv. tomato after A. brasilense was established in the leaves further reduced both the population of P. syringae pv. tomato and disease severity and significantly enhanced plant development. Both bacteria maintained a large population in the rhizosphere for 45 days when each was inoculated separately onto tomato seeds (105 to 106 CFU/g [dry weight] of root). However, P. syringae pv. tomato did not survive in the rhizosphere in the presence of A. brasilense. Foliar inoculation of A. brasilense after P. syringae pv. tomato was established on the leaves did not alleviate bacterial speck disease, and A. brasilense did not survive well in the phyllosphere under these conditions, even in a mist chamber. Several applications of a low concentration of buffered malic acid significantly enhanced the leaf population of A. brasilense (>108 CFU/g [dry weight] of leaf), decreased the population of P. syringae pv. tomato to almost undetectable levels, almost eliminated disease development, and improved plant growth to the level of uninoculated healthy control plants. Based on our results, we propose that A. brasilense be used in prevention programs to combat the foliar bacterial speck disease caused by P. syringae pv. tomato.  相似文献   

13.
Nitrilase enzymes catalyse the hydrolysis of nitrile compounds to the corresponding carboxylic acid and ammonia, and have been identified in plants, bacteria and fungi. There is mounting evidence to support a role for nitrilases in plant–microbe interactions, but the activity of these enzymes in plant pathogenic bacteria remains unexplored. The genomes of the plant pathogenic bacteria Pseudomonas syringae pv. syringae B728a and Pseudomonas syringae pv. tomato DC3000 contain nitrilase genes with high similarity to characterized bacterial arylacetonitrilases. In this study, we show that the nitrilase of P. syringae pv. syringae B728a is an arylacetonitrilase, which is capable of hydrolysing indole-3-acetonitrile to the plant hormone indole-3-acetic acid, and allows P. syringae pv. syringae B728a to use indole-3-acetonitrile as a nitrogen source. This enzyme may represent an additional mechanism for indole-3-acetic acid biosynthesis by P. syringae pv. syringae B728a, or may be used to degrade and assimilate aldoximes and nitriles produced during plant secondary metabolism. Nitrilase activity was not detected in P. syringae pv. tomato DC3000, despite the presence of a homologous nitrilase gene. This raises the interesting question of why nitrilase activity has been retained in P. syringae pv. syringae B728a and not in P. syringae pv. tomato DC3000.  相似文献   

14.
We studied the response of glutathione‐ and ascorbate‐related antioxidant systems of the two tomato cultivars to Pseudomonas syringae pv. tomato infection. In the inoculated susceptible A 100 cultivar a substantial decrease in reduced glutathione (GSH) content, oxidised glutathione accumulation and GSH redox ratio decline as well as glutathione peroxidase activity increase were found. The enhanced glutathione reductase activity was insufficient to keep the glutathione pool reduced. A transiently increased dehydroascorbic acid (DHA) content and ascorbic acid (AA) redox ratio decrease together with ascorbate peroxidase activity suppression were observed. Adversely to the progressive reduction in GSH pool size, AA content tended to increase but the changes were more modest than those of GSH. By contrast, in interaction with the resistant Ontario cultivar the glutathione pool homeostasis was maintained throughout P. syringae attack and no significant effect on the ascorbate pool was observed. Moreover, in the resistant interaction there was a significantly higher constitutive and pathogen‐induced glutathione‐S‐transferase (GST) activity. The relationship between GST activity and DHA content found in this study indicates that this enzyme could also act as dehydroascorbate reductase. These results reflect the differential involvement of GSH and AA in tomato‐P. syringae interaction and, in favour of the former, they clearly indicate the role of GSH and GSH‐utilizing enzymes in resistance to P. syringae. The maintenance of glutathione pool homeostasis and GST induction appear to contribute to tissue inaccessibility to bacterial attack.  相似文献   

15.
16.
Osmoregulated periplasmic glucans (OPGs) are intrinsic components of the Gram-negative bacterial envelope and are important for bacterial-host interactions. The OPGs of Pseudomonas syringae pv. syringae have been known to be highly branched linear glucans ranging from 6 to 13 glucose residues devoid of any substituents, while having backbone structure similar to those of Escherichia coli and Erwinia chrysanthemi. Here, we report for the first time succinylated and large-sized OPGs from P. syringae pv. syringae. The glucans were isolated with trichloroacetic acid treatment and various chromatographic techniques. These were further characterized by thin-layer chromatography, matrix-assisted laser desorption/ionization time of flight mass spectrometer, and 1D 1H nuclear magnetic resonance spectroscopy. The results demonstrate that novel anionic glucans with one succinyl residue at the C-6 position of the glucose unit as well as neutral glucans including large-sized glucans with up to 28 degrees of polymerization are produced in P. syringae pv. syringae. Furthermore, the succinylated and large-sized OPGs of P. syringae pv. syringae are necessary for hypoosmotic adaptation.  相似文献   

17.
The host-specific plant pathogen Pseudomonas syringae elicits the hypersensitive response (HR) in nonhost plants and secretes the HrpZ harpin in culture via the Hrp (type III) secretion system. Previous genetic evidence suggested the existence of another harpin gene in the P. syringae genome. hrpW was found in a region adjacent to the hrp cluster in P. syringae pv. tomato DC3000. hrpW encodes a 42.9-kDa protein with domains resembling harpins and pectate lyases (Pels), respectively. HrpW has key properties of harpins. It is heat stable and glycine rich, lacks cysteine, is secreted by the Hrp system, and is able to elicit the HR when infiltrated into tobacco leaf tissue. The harpin domain (amino acids 1 to 186) has six glycine-rich repeats of a repeated sequence found in HrpZ, and a purified HrpW harpin domain fragment possessed HR elicitor activity. In contrast, the HrpW Pel domain (amino acids 187 to 425) is similar to Pels from Nectria haematococca, Erwinia carotovora, Erwinia chrysanthemi, and Bacillus subtilis, and a purified Pel domain fragment did not elicit the HR. Neither this fragment nor the full-length HrpW showed Pel activity in A230 assays under a variety of reaction conditions, but the Pel fragment bound to calcium pectate, a major constituent of the plant cell wall. The DNA sequence of the P. syringae pv. syringae B728a hrpW was also determined. The Pel domains of the two predicted HrpW proteins were 85% identical, whereas the harpin domains were only 53% identical. Sequences hybridizing at high stringency with the P. syringae pv. tomato hrpW were found in other P. syringae pathovars, Pseudomonas viridiflava, Ralstonia (Pseudomonas) solanacearum, and Xanthomonas campestris. ΔhrpZ::nptII or hrpW::ΩSpr P. syringae pv. tomato mutants were little reduced in HR elicitation activity in tobacco, whereas this activity was significantly reduced in a hrpZ hrpW double mutant. These features of hrpW and its product suggest that P. syringae produces multiple harpins and that the target of these proteins is in the plant cell wall.  相似文献   

18.
Pseudomonas syringae pv. tomato, the causal agent for bacterial speck of tomato, produces the phytotoxin coronatine. A 5.3-kilobase XhoI fragment from the chromosomal region controlling toxin production was cloned into the plasmid pGB2, and the resulting recombinant plasmid, pTPR1, was tested for its ability to serve as a diagnostic probe for P. syringae pv. tomato. In a survey of 75 plant-associated bacteria, pTPR1 hybridized exclusively to those strains that produced coronatine. The detection limit for this probe, which was labeled with the Chemiprobe nonradioactive reporter system, was approximately 4 × 103 CFU of lesion bacteria. During the 1989 growing season, a total of 258 leaf and fruit lesions from nine tomato fields were screened for P. syringae pv. tomato by using pTPR1 and the culture method of detection. The best agreement between the two methods, 90%, occurred early in the season with samples taken from relatively young (5-week-old) plants. Young plants also had a higher percentage of P. syringae pv. tomato-positive lesions. P. syringae pv. tomato was the only coronatine producer recovered from the nine tomato fields. All 244 P. syringae pv. tomato strains isolated during this study reacted strongly with the probe. The P. syringae pv. tomato population of healthy field tomato leaves was determined by a pTPR1 colony hybridization procedure. Every probe-positive colony that was isolated and characterized was identified as P. syringae pv. tomato. The pTPR1 probe should expedite disease diagnosis and facilitate epidemiological studies of this pathogen. It also should aid in screening transplant seedlings for bacterial speck infestation.  相似文献   

19.
The first outbreaks of bacterial canker of kiwifruit caused by Pseudomonas syringae pv. actinidiae biovar 3 were detected in France in 2010. P. syringae pv. actinidiae causes leaf spots, dieback, and canker that sometimes lead to the death of the vine. P. syringae pv. actinidifoliorum, which is pathogenic on kiwi as well, causes only leaf spots. In order to conduct an epidemiological study to track the spread of the epidemics of these two pathogens in France, we developed a multilocus variable-number tandem-repeat (VNTR) analysis (MLVA). MLVA was conducted on 340 strains of P. syringae pv. actinidiae biovar 3 isolated in Chile, China, France, Italy, and New Zealand and on 39 strains of P. syringae pv. actinidifoliorum isolated in Australia, France, and New Zealand. Eleven polymorphic VNTR loci were identified in the genomes of P. syringae pv. actinidiae biovar 3 ICMP 18744 and of P. syringae pv. actinidifoliorum ICMP 18807. MLVA enabled the structuring of P. syringae pv. actinidiae biovar 3 and P. syringae pv. actinidifoliorum strains in 55 and 16 haplotypes, respectively. MLVA and discriminant analysis of principal components revealed that strains isolated in Chile, China, and New Zealand are genetically distinct from P. syringae pv. actinidiae strains isolated in France and in Italy, which appear to be closely related at the genetic level. In contrast, no structuring was observed for P. syringae pv. actinidifoliorum. We developed an MLVA scheme to explore the diversity within P. syringae pv. actinidiae biovar 3 and to trace the dispersal routes of epidemic P. syringae pv. actinidiae biovar 3 in Europe. We suggest using this MLVA scheme to trace the dispersal routes of P. syringae pv. actinidiae at a global level.  相似文献   

20.
Strains of Pseudomonas syringae pv. syringae were isolated from healthy and diseased stone fruit tissues sampled from 43 orchard sites in California in 1995 and 1996. These strains, together with P. syringae strains from other hosts and pathovars, were tested for pathogenicity and the presence of the syrB and syrC genes and were genetically characterized by using enterobacterial repetitive intergenic consensus (ERIC) primers and PCR. All 89 strains of P. syringae pv. syringae tested were moderately to highly pathogenic on Lovell peach seedlings regardless of the host of origin, while strains of other pathovars exhibited low or no pathogenicity. The 19 strains of P. syringae pv. syringae examined by restriction fragment length polymorphism analysis contained the syrB and syrC genes, whereas no hybridization occurred with 4 strains of other P. syringae pathovars. The P. syringae pv. syringae strains from stone fruit, except for a strain from New Zealand, generated ERIC genomic fingerprints which shared four fragments of similar mobility. Of the P. syringae pv. syringae strains tested from other hosts, only strains from rose, kiwi, and pear generated genomic fingerprints that had the same four fragments as the stone fruit strains. Analysis of the ERIC fingerprints from P. syringae pv. syringae strains showed that the strains isolated from stone fruits formed a distinct cluster separate from most of the strains isolated from other hosts. These results provide evidence of host specialization within the diverse pathovar P. syringae pv. syringae.  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