Skip to main content
  • Research article
  • Open access
  • Published:

A draft genome sequence and functional screen reveals the repertoire of type III secreted proteins of Pseudomonas syringae pathovar tabaci 11528

Abstract

Background

Pseudomonas syringae is a widespread bacterial pathogen that causes disease on a broad range of economically important plant species. Pathogenicity of P. syringae strains is dependent on the type III secretion system, which secretes a suite of up to about thirty virulence 'effector' proteins into the host cytoplasm where they subvert the eukaryotic cell physiology and disrupt host defences. P. syringae pathovar tabaci naturally causes disease on wild tobacco, the model member of the Solanaceae, a family that includes many crop species as well as on soybean.

Results

We used the 'next-generation' Illumina sequencing platform and the Velvet short-read assembly program to generate a 145X deep 6,077,921 nucleotide draft genome sequence for P. syringae pathovar tabaci strain 11528. From our draft assembly, we predicted 5,300 potential genes encoding proteins of at least 100 amino acids long, of which 303 (5.72%) had no significant sequence similarity to those encoded by the three previously fully sequenced P. syringae genomes. Of the core set of Hrp Outer Proteins that are conserved in three previously fully sequenced P. syringae strains, most were also conserved in strain 11528, including AvrE1, HopAH2, HopAJ2, HopAK1, HopAN1, HopI, HopJ1, HopX1, HrpK1 and HrpW1. However, the hrpZ1 gene is partially deleted and hopAF1 is completely absent in 11528. The draft genome of strain 11528 also encodes close homologues of HopO1, HopT1, HopAH1, HopR1, HopV1, HopAG1, HopAS1, HopAE1, HopAR1, HopF1, and HopW1 and a degenerate HopM1'. Using a functional screen, we confirmed that hopO1, hopT1, hopAH1, hopM1', hopAE1, hopAR1, and hopAI1' are part of the virulence-associated HrpL regulon, though the hopAI1' and hopM1' sequences were degenerate with premature stop codons. We also discovered two additional HrpL-regulated effector candidates and an HrpL-regulated distant homologue of avrPto1.

Conclusion

The draft genome sequence facilitates the continued development of P. syringae pathovar tabaci on wild tobacco as an attractive model system for studying bacterial disease on plants. The catalogue of effectors sheds further light on the evolution of pathogenicity and host-specificity as well as providing a set of molecular tools for the study of plant defence mechanisms. We also discovered several large genomic regions in Pta 11528 that do not share detectable nucleotide sequence similarity with previously sequenced Pseudomonas genomes. These regions may include horizontally acquired islands that possibly contribute to pathogenicity or epiphytic fitness of Pta 11528.

Background

Pseudomonas syringae is a widespread bacterial pathogen that causes disease on a broad range of economically important plant species. The species P. syringae is sub-divided into about 50 pathovars, each exhibiting characteristic disease symptoms and distinct host-specificities. P. syringae pathovar tabaci (Pta) causes wild-fire disease in soybean and tobacco plants [1, 2], characterised by chlorotic halos surrounding necrotic spots on the leaves of infected plants. Formation of halos is dependent on the beta-lactam tabtoxin, which causes ammonia accumulation in the host cell by inhibition of glutamine synthetase [3]. However, whether tabtoxin is an essential component of the disease process is unclear [4, 5].

Pathogenicity of P. syringae strains is dependent on the type III secretion system (T3SS). The T3SS secretes a suite of virulence 'effector' proteins into the host cytoplasm where they subvert the eukaryotic cell physiology and disrupt host defences [6–14]. Mutants lacking the T3SS do not secrete effectors, and as a consequence do not infect plants or induce disease symptoms. Thus, understanding effector action is central to understanding bacterial pathogenesis. A single P. syringae strain typically encodes about 30 different effectors [14]. However, different P. syringae strains have different complements of effector genes. The emerging view is that of a core of common effectors encoded by most strains, augmented by a variable set. Individual effectors appear to act redundantly with each other and are individually dispensable with a small or no loss to pathogen virulence [10]. Effectors are also thought to play an important role in determining host range. This is most clearly true when infections are restricted by host defences. Some plants have evolved specific mechanisms to recognise certain effectors; such recognition induces strong host defences which curtail infection. For example, expression of the T3SS effector HopQ1-1 from P. syringae pathovar tomato (Pto) DC3000 was sufficient to render Pta 11528 avirulent on Nicotiana benthamiana [15]. The opposite situation, in which acquisition of a novel effector gene confers the ability to infect new host plants, has not been demonstrated and remains speculative. However, heterologous expression of the effector gene avrPtoB conferred a plasmid-cured strain of P. syringae pathovar phaseolicola (Pph) with increased virulence [16]. We hope that further identification and characterisation of effector repertoires of particular strains will shine new light on their roles in determining host range. Finally, bacterial virulence is also likely to be influenced by other non-T3SS-dependent virulence factors such as toxins which are often co-regulated with the T3SS [17].

Complete genome sequences are available for strains representing three P. syringae pathovars: Pto, pathovar phaseolicola (Pph) and pathovar syringae (Psy) [18–20]. Comparisons of these have led to the identification of core effector gene sets and to explain some of the differences in host-specificity between pathovars. However, these three sequenced strains are representatives of three distinct phylogroups within the species P. syringae, and as such are phylogenetically quite distant [21, 22]. According to DNA-DNA hybridisation studies and ribotyping [21], P. syringae can be divided into 9 discrete genomospecies. Representative strains of Psy, Pph and Pto fell into genomospecies one, two and three respectively [21]. Recently, a strain of pathovar oryzae (genomospecies four) was sequenced [23]. A draft genome sequence was also published for Pto T1 [24], a strain closely related to Pto DC3000 but restricted to tomato hosts, whereas Pto DC3000 is able to cause disease on Arabidopsis. In the current study, we explore genetic differences at an intermediate phylogenetic resolution; that is, we compared the genome sequences of Pta 11528 to that of P. phaseolicola (Pph) 1448A, which resides within the same phylogroup but possesses a distinct host range and causes different disease symptoms.

Pto DC3000 was the first plant-pathogenic pseudomonad to have its genome sequenced, helping to establish the Arabidopsis-Pto system as the primary model for plant-microbe interactions. However, Arabidopsis is not a natural host of Pto, and it is important to develop alternative systems given the genetic variability of P. syringae strains, particularly in regard to effectors. We work on the interaction between Pta and the wild tobacco plant N. benthamiana, which offers certain advantages over Arabidopsis. Firstly, N. benthamiana is an important model for the Solanaceae, which includes many important crop species. The Pta-N. benthamiana interaction is a natural pathosystem. Lastly, N. benthamiana is an important model plant that is more amenable to biochemistry-based approaches and facile manipulation of gene expression such as virus-induced gene silencing (VIGS). Thus N. benthamiana provides experimental options for understanding plant-bacterial interactions. Strains of Pta can cause disease on N. benthamiana, but relatively few genetic sequence data are available for this pathovar.

In this study we generated a draft complete genome sequence of Pta 11528 and used a functional screen for HrpL-dependent genes to infer its repertoire of T3SS effectors and associated Hrp Outer Proteins (Hops), which differs significantly from that of its closest relative whose complete genome has previously been published (Pph 1448A). Pta 11528 does not encode functional homologues of HopAF1 or HrpZ1. This was surprising since HopAF1 was conserved in the three previously sequenced pathovars [18–20]. HrpZ1 is conserved in most strains of P. syringae that have been investigated, albeit with differences in amino acid sequence [25]. However, Pta strain 6605 and several other isolates from Japan, were previously shown to carry a major deletion leading to truncated HrpZ protein product [26]. Pta 11528 encodes several novel potential T3SS effectors for which no close orthologues have been reported. We also discovered several large genomic regions in Pta 11528 that do not share detectable nucleotide sequence similarity with previously sequenced Pseudomonas genomes. These regions may be horizontally acquired islands that possibly contribute to pathogenicity or epiphytic fitness of Pta 11528.

Results and discussion

Sequencing and assembly of the Pta 11528 genome

The Illumina sequencing platform provides a cost-effective and rapid means to generate nucleotide sequence data [27–29]. Although this method generates very short sequence reads, several recent studies have demonstrated that it is possible to assemble these short reads into good quality draft genome sequences [30–41].

We generated 12,096,631 pairs of 36-nucleotide reads for a total of 870,957,432 nucleotides. This represents approximately 145X depth of coverage assuming a genome size of six megabases. We used Velvet 0.7.18 [41] to assemble the reads de novo. Our resulting assembly had 71 supercontigs of mean length 85,604 nucleotides, an N50 number of eight, and N50 length of 317,167 nucleotides; that is, the eight longest supercontigs were all at least 317,167 nucleotides long and together covered more than 50% of the predicted genome size of six megabases. The largest supercontig was 606,547 nucleotides long. The total length of the 71 assembled supercontigs was 6,077,921 nucleotides. The G+C content of the assembly was 57.96%, similar to that of the previously sequenced P. syringae genomes (Table 1). The sequence data from this project have been deposited at DDBJ/EMBL/GenBank under the accession ACHU00000000. The version described in this paper is the first version, ACHU01000000. The data can also be accessed from the authors' website http://tinyurl.com/Pta11528-data and as Additional files submitted with this manuscript. In addition, an interactive genome browser is available from the authors' website http://tinyurl.com/Pta11528-browser.

Table 1 Comparison of Pta 11528 genome properties with those of previously sequenced P. syringae genomes [18–20, 83–85], [86-93].

We aligned the 71 Pta supercontigs against published complete Pseudomonas genome sequences using MUMMER [42]. The Pta 11528 genome was most similar to that of Pph 1448A, with 97.02% nucleotide sequence identity over the alignable portions. The next most similar genome was that of Pto DC3000, with less than 90% identity (Table 1). This pattern of sequence similarity is consistent with phylogenetic studies that placed strains of Pta in the same phylogroup as Pph and revealed a relatively distant relationship to Pto [21, 22].

Comparison of the protein complement of Pta 11528 versus Pph 1448A and other pseudomonads

Using the FgenesB annotation pipeline http://www.softberry.com, we identified 6,057 potential protein-coding genes, of which 5,300 were predicted to encode proteins of at least 100 amino acids long. Of 5,300 predicted Pta 11528 proteins, 575 (10.8%) had no detectable homology with Pph 1448A proteins (based on our criterion of an E-value less than 1e-10 using BLASTP). Of these 575 sequences, 303 had no detectable homologues in Psy B728a nor Pto DC3000. These 303 Pta-specific sequences had a median length of 198 amino acids whereas the median length of the 5,300 sequences was 216 amino acids. Automated gene prediction is not infallible and inevitably a subset of the predictions will be incorrect. The reliability of gene predictions is poorer for short sequences than for longer ones. This slight enrichment for very short sequences among the Pta-specific gene predictions might be explained by the inclusion of some open reading frames that are not functional genes among those 303. However, many of the predicted proteins showed significant similarity to other proteins in the NCBI NR databases (See Additional file 1: Table S1), confirming that these are likely to be genuine conserved genes.

Conservation of the T3SS apparatus and T3SS-dependent effectors

The Hop Database (HopDB, http://www.pseudomonas-syringae.org) provides a catalogue of confirmed and predicted hop genes [43]. Figure 1 lists the hop genes in HopDB for the three previously fully sequenced P. syringae genomes. A 'core' set of hop genes are conserved in all three previously sequenced pathovars: avrE1, hopAF1, hopAH2, hopAJ2, hopAK1, hopAN1, hopI1, hopJ1, hopX1, hrpK1, hrpW1 and hrpZ1. In addition to this core set, each genome contains additional hop genes that are found in only a subset of the sequenced strains. The Pta 11528 homologues of hop genes are listed in Table 2. Figure 1 also indicates those hop genes for which a close homologue was found to be encoded in Pta 11528.

Table 2 Homologues of known hop genes in Pta 11528. Homologues were detected by searching the Pta 11528 FgenesB-predicted protein sequences against HopDB http://www.pseudomonas-syringae.org using BLASTP
Figure 1
figure 1

Comparison of the hop gene complements of the three previously fully sequenced P. syringae genomes. Those hop genes that are conserved in Pta 11528 are shown in boldface and underlined. Pta 11528 also contains three hop genes that do not have orthologues in the sequenced genomes: hopAR1, hopF1 and hopW1. * No close homologue of avrPto1 was found in Pta 11528; however, there is a gene encoding a protein that shares 43% amino acid identity with Avr Pto 1 from Pto DC3000. ** In the Pta 11528 genome hrpZ1 appears to be a pseudogene.

In sequenced strains of P. syringae, the gene cluster encoding the T3SS apparatus is flanked by collections of effector genes termed the exchangeable effector locus (EEL) and the conserved effector locus (CEL). Together, these three genetic components comprise the Hrp pathogenicity island [44]. A core set of hop genes is located in the Hrp pathogenicity island [44], which is highly conserved between Pta 11528 and Pph 1448A (Figure 2), except that in Pta 11528 there is a deletion in hrpZ1 and an insertion in the hrpV-hrcU intergenic region. The core hop genes avrE1, hopAH2, hopAJ2, hopAK1, hopAN1, hopI1, hopJ1, hopX1 and hrpK1 are conserved in Pta 11528 and encode intact full-length proteins. Pta 11528 encodes a full-length HrpW1 protein, albeit with insertions of 69 and 12 nucleotides relative to the Pph 1448A sequence. However, there is a large deletion in hrpZ1 that likely renders it non-functional and hopAF1 is completely absent.

Figure 2
figure 2

Conservation of the Hrp pathogenicity island between Pph 1448A and Pta 11528. Panel A shows an alignment of the Pph 1448A Hrp pathogenicity island (lower track) against the homologous region in Pta 11528 (upper track), prepared using GenomeMatcher, which indicates similarity values by colour with dark blue, green, yellow and red representing increasing degrees of similarity [78]. Panel B shows the MAQ [79] alignment of the Pta 11528 Illumina reads (in black) and the BLASTN [80] alignment of the Pta 11528 de novo assembly (in green) against the Hrp region of the Pph 1448A genome.

Besides the core conserved hop genes, the Pta 11528 genome assembly contains full-length orthologues of hopR1, hopAS1, hopAE1 and hopV1, which are also found in Pph 1448A but are absent from Psy B728a and/or Pto DC3000.

The hrpZ1 gene encodes a harpin, which is not classified as a type III effector because it is not injected directly into host cells. Harpins are characteristically acidic, heat-stable and enriched for glycine, lack cysteine residues [8] and can induce defences in both host and non-host plants [45, 46]. HrpZ1 forms pores in the host membrane [47] suggesting a role in translocation of effectors across the host membrane. It also shows sequence-specific protein binding activity [48]. HrpZ1 can induce defences in both host and non-host plants and tobacco has been extensively used as the non-host plant species [45, 46]. The inactivation of hrpZ1 in Pta 11528 and other strains of Pta [26] may be an adaptive strategy and have been an important process in the stepwise progression towards compatibility, allowing Pta 11528 to avoid detection by the tobacco host plant. This is reminiscent of the "black holes" and other processes that inactivate genes whose expressed products are detrimental to a pathogenic lifestyle [49, 50]. One excellent example is the inactivation of cadA in genomes of Shigella species as compared to the genome of their closely related but non-pathogenic Escherichia coli strain [51, 52].

Pta 11528 contains highly conserved homologues of hopAB2, hopW, hopO1-1, hopT1-1, hopAG1, hopAH1, hopF1 and hopAR1, which are absent in Pph 1448A. Although absent from the Pph 1448A genome, hopAR1 and hopF1 have been identified in other strains of Pph [53–57]. In Pph 1302A, hopAR1 is located on the pathogenicity island PPH GI-1, though its genomic location varies between strains [56, 57]. PPH GI-1 is absent from the Pph 1448A genome [57]. The Pta 11528 genome (supercontig 1087) possesses a region of similarity to PTPH GI-1, but which contains a substantial number of insertions and deletions (Additional file 2: Figure S1). The Pta 11528 hopAR1 homologue (C1E_2036) is not located in the PPH GI-1 region; it falls on supercontig 672 about two kilobases upstream of a gene encoding a protein (C1E_2039) sharing 43% amino acid identity with Pto DC3000 avrPto1. In contrast to AvrPto1 from Pto DC3000, the AvrPto1 homologue (C1E_2039) from Pta 11528 is not recognised by the plant Pto/Prf system (S. Gimenez Ibanez and J. Rathjen, manuscript in preparation).

The homologues of hopAG1, hopAH1 and the degenerate hopAI1' are found within a region of the Pta 11528 genome that shares synteny with the chromosome of Psy B728a. This region is also conserved in Pto DC3000A, albeit with several deletions and insertions, suggesting that these effector genes are ancestral to the divergence of the pathovars and have been lost in Pph 1448A rather than having been laterally transferred laterally between Pta 11528 and Psy B728a. In Pto DC3000, hopAG1 (PSPTO_0901) has been disrupted by an insertion sequence (IS) element. This is consistent with a model of lineage-specific loss of certain ancestral effectors.

In Pto DC3000, hopO1-1 and hopT1-1 are located on the large plasmid pDC3000A; homologues of these effector-encoding genes are not found in Pph 1448A. The Pta 11528 genome contains a three kilobase region of homology to pDC3000 comprising homologues of these two effector genes and a homologue of the ShcO1 chaperone-encoding gene. These three genes are situated in a large (at least 50 kilobase) region of the Pta 11528 genome that has only limited sequence similarity with Pph 1448A. Two tRNA genes (tRNA-Pro and tRNA-Lys) are located at the boundary of this region (Figure 3), which would be consistent with this comprising a mobile island.

Figure 3
figure 3

A 90-kilobase region of the Pta 11528 genome containing homologues of hopT1-1 and hopO1-1. The G+C content is indicated by the plot near the top of the figure.

In plasmid pMA4326B from P. syringae pathovar maculicola (Pma), the hopW1 effector gene is immediately adjacent to a three-gene cassette comprising a resolvase, an integrase and exeA. This cassette is also found in plasmids and chromosomes of several human-pathogenic Gram-negative bacteria [58]. We found a homologue of this cassette along with a hopW1 homologue on supercontig 955 of the Pta 11528 genome assembly. Stavrinides and Guttman [58] proposed that the boundaries of the cassette lay upstream of the resolvase and upstream of hopW1. The presence of this four-gene unit in a completely different location in Pta 11528 is indeed consistent with the hypothesis that it represents a discrete mobile unit.

Several hop genes are located on the large plasmid of Pph 1448A. We found no homologues of these genes in Pta 11528, suggesting that the plasmid is not present in Pta 11528. Consistent with this, only a small proportion of the plasmid was alignable to our 36-nucleotide Illumina sequence reads (Figure 4). This reveals that a large component of the pathogen's effector arsenal is determined by its complement of plasmids. However, simple loss or gain of a plasmid does not explain all of the differences in effector complement since Pta 11528 lacks homologues of several Pph 1448A chromosomally-located effector-encoding hop genes hopG1, hopAF1, avrB4, hopF3 and hopAT1 as well as the non-effector hopAJ1. It also lacks homologues of the Pph 1448A degenerate effector gene hopAB3'.

Figure 4
figure 4

Limited conservation between the Pta 11528 genome sequence and the sequence of the Pph 1448A large plasmid. The MAQ [79] alignment of the Pta 11528 Illumina reads is shown in black. The thickness of the black track is proportional to the depth of coverage by Illumina reads. The BLASTN [80] alignment of the Pta 11528 de novo assembly against the plasmid sequence is shown in a green track, with the thickness of this single green track being proportional to sequence identity.

The regions of the Pph 1448A large plasmid that are apparently conserved in Pta 11528 include genes encoding the conjugal transfer system, suggesting the presence of one or more plasmids in this strain. We found an open reading frame (C1E_3950, located on supercontig 955 coordinates 59126-60394) encoding a protein with about 97% sequence identity to the RepA proteins characteristically encoded on pT23A-family plasmids (e.g. AAW01447; reviewed in [59]), suggesting that this 236 kilobase supercontig might represent a plasmid.

A functional screen for HrpL-regulated genes

We used a previously described functional screen [60] to complement our bioinformatics-based searches for type III effectors of Pta 11528. Our functional screen was based on two steps. The first step was employed to identify genes whose expression was regulated by the T3SS alternative sigma factor, HrpL. The second step was used to identify the subset of HrpL-regulated genes that encoded effectors. For Pta 11528, we employed only the first step to identify candidate effector genes based on induced expression by HrpL. A library was constructed from Pta 11528 into a broad-host range vector carrying a promoter-less GFP and mobilized into Pto lacking its endogenous hrpL but conditionally complemented with an arabinose-inducible hrpL. We used a fluorescence activated cell sorter (FACS) to select clones that carried HrpL-inducible promoters based on expression of GFP after growth in arabinose. Clones were sequenced and sequences were assembled. Clones representative of assembled supercontigs were verified again for HrpL regulation using FACS. Among the genes whose expression was confirmed to be HrpL-dependent were those encoding effectors hopAE1, hopI1, hopAR1, the avrPto1-like gene, hopF1, hopT1-1, hopO1-1, avrE1, hopX1, and the degenerate hopM1' and hopAI1' as well as known T3SS-associated genes hrpH (ORF1 of the CEL; [61]) and hrpW1. Interestingly, the screen also confirmed HrpL-dependent regulation of genes encoding a major facilitator superfamily (MFS) permease and a putative peptidase (Table 3).

Table 3 Pta 11528 genes confirmed by the functional screen to be under the transcriptional control of HrpL

Other differences in predicted proteomes of P. syringae strains

Host range and pathogenicity are likely to be further influenced by genes other than those associated with type III secretion. Virulence determinants in P. syringae include toxins as well as epiphytic fitness; that is, the ability to acquire nutrients and survive on the leaf surface [14]. Epiphytic fitness depends on quorum-sensing [62], chemotaxis [63], osmo-protection, extracellular polysaccharides, glycosylation of extracellular structures [64] iron uptake [65] and the ability to form biofilms. Cell-wall-degrading hydrolytic enzymes play a role in virulence in at least some plant-pathogenic pseudomonads [66]). Secretion systems (including type I, type II, type IV, type V, type VI and twin arginine transporter) may also contribute to both virulence and epiphytic fitness [67], whilst multidrug efflux pumps may confer resistance to plant-derived antimicrobials [68].

To identify differences between Pta 11528 and the previously sequenced Pph 1448A, Psy B728a and Pto DC3000 with respect to their repertoires of virulence factors, we performed BLASTP searches between the predicted proteomes. We found no significant differences in the repertoires of secretion systems between the proteomes. However, we found that Pta 11528 lacks homologues of several Pph 1448A polysaccharide modifying enzymes (glycosyl transferase PSPPH_0951, polysaccharide lyase PSPPH_1510, glycosyl transferase PSPPH_3642). Conversely, Pta 11528 encodes two glycosyl transferases (C1E_0355 and C1E_0361) and a thermostable glycosylase (C1E_4802) that do not have homologues in any of the three fully sequenced P. syringae genomes. This may imply differences in the extracellular polysaccharide profiles. In contrast to Pph 1448A, Pta 11528 lacks homologues of RhsA insecticidal toxins (PSPPH_4042 and PSPPH_4043). However, a tabtoxin biosynthesis gene cluster is found in the Pta 11528 genome and shows a high degree of conservation with the previously sequenced Pta BR2 tabtoxin biosynthesis cluster [69].

Pta 11528 encodes several enzymes that do not have homologues in any of the three fully sequenced P. syringae genomes (Table 4), including a predicted gluconolactonase (C1E_2553), a predicted dienelactone hydrolase (C1E_2589), a predicted nitroreductase (C1E_6026), and a sulphotransferase (C1E_6026). C1E_0903 shares 71.4% amino acid sequence identity with a predicted epoxide hydrolase (YP_745600.1) from Granulibacter bethesdensis CGDNIH1 [70] and has a significant match to the epoxide hydrolase N-terminal domain in the Pfam database (PF06441) [71, 72]. Epoxide hydrolases are found in P. aeruginosa and P. fluorescens PfO-1, but not in any other pseudomonads. It is possible that this gene product has a function in detoxification of host-derived secondary metabolites.

Table 4 Proteins encoded by the draft Pta 11528 genome that have no detectable homologues on three previously fully sequenced P. syringae genomes.

Pta protein C1E_6026 has a significant match to the sulphotransferase domain (Pfam:PF00685). Examples of this protein domain have not been found in other pseudomonads except for P. fluorescens PfO-1. Sulphotransferase proteins include flavonyl 3-sulphotransferase, aryl sulphotransferase, alcohol sulphotransferase, estrogen sulphotransferase and phenol-sulphating phenol sulphotransferase. These enzymes are responsible for the transfer of sulphate groups to specific compounds. The sulphotransferase gene (C1E_6026, 82% amino acid identity to P. fluorescens Pfl01_0157) overlaps a two kilobase Pta 11528-specific genomic island that also encodes a phage tail collar-protein encoding gene (C1E_5461, 61% amino acid identity to P. fluorescens Pfl01_0155) and an acetyltransferase (C1E_5459, 76% amino acid identity to P. fluorescens Pfl01_0148). We speculate that this region has been horizontally acquired in the Pta 11528 lineage via a bacteriophage.

An 80 kilobase region of Pta 11528 supercontig 684 contains two open reading frames (ORFs) (C1E_2584 and C1E_2585) whose respective predicted protein products show 48 and 55% amino acid identity to the C- and N-termini of a P. putida methyl-accepting chemotaxis protein (MCP) (PP_2643) and little similarity to any P. syringae protein. Since the N- and C-termini are divided into separate reading frames, this probably represents a degenerate pseudogene. Immediately downstream of these ORFs is a gene (C1E_2583) that specifies a MCP showing greatest sequence identity (70%) to PP_2643 from P. putida, whilst sharing only 65% identity to its closest homologue in P. syringae (PSPPH_4743). This region also encodes another MCP (C1E_2587) that shares only 50% amino acid identity with any previously sequenced P. syringae homologue. It remains to be tested whether these MCPs play a role in pathogenesis and/or epiphytic fitness.

Transcriptional regulators are not normally considered to be virulence factors. However, expression of virulence factors may be coordinated by and dependent on regulators. Moreover, heterologous expression of the RscS regulator was recently shown to be sufficient to transform a fish symbiont into a squid symbiont [73]. Pta 11528 encodes several predicted transcriptional regulators that are not found in Pto DC3000, Psy B728a and Pph 1448A. These include two predicted TetR-like proteins (C1E_0901 and C1E_6027), two predicted xenobiotic response element proteins (C1E_2056 and C1E_2563), a LacI-like protein (C1E_2286), a Cro/CI family protein (C1E_2570) and an IclR family protein (C1E_5715).

Pta 11528 encodes a novel pilin (C1E_2329) not found in previously sequenced P. syringae strains but sharing significant sequence similarity with a type IV pilin from P. aeruginosa [74]. Pilin is the major protein component of the type IV pili, which have functions in forming micro-colonies and biofilms, host-cell adhesion, signalling, phage-attachment, DNA uptake and surface motility, and have been implicated as virulence factors in animal-pathogenic bacteria [75]. The precise function of the C1E_2329 pilin is unknown but it may be involved in epiphytic fitness or plant-pathogenesis or could even be involved in an interaction with an insect vector.

Pta-specific genomic islands

We identified 102 genomic regions of at least one kilobase in length which gave no BLASTN matches against previously sequenced Pseudomonas genomes (Additional file 3: Table S2). Ten of the Pta 11528-specific regions are longer than 10 kilobases, the longest being 37.7, 21.8, 18.7, 17.9 and 16.6 kilobases. The 16.6 kilobase region corresponds to the tabtoxin biosynthesis gene cluster [69]. These regions will be good candidates for further study of the genetic basis for association of Pta with the tobacco host. For example, several of the islands encode MFS transporters and other efflux proteins that might be involved in protection from plant-derived antimicrobials (Additional file 3: Table S2).

Conclusion

We have generated a draft complete genome sequence for the Pta 11528 a pathogen that naturally causes disease in wild tobacco, an important model system for studying plant disease and immunity. From this sequence, combined with a functional screen, we were able to deduce the pathogen's repertoire of T3SS-associated Hop proteins. This has revealed some important differences between Pta and other pathovars with respect to the arsenal of T3SS effectors at their disposal for use against the host plant. We also revealed more than a hundred Pta-specific genomic regions that are not conserved in any other sequenced P. syringae, providing many potential leads for the further study of the Pta-tobacco disease system.

Methods

Sequence data

The previously published sequences of P. syringae pathovar phaseolicola 1448A [20], P. syringae pathovar syringae B728a [19], P. syringae pathovar tomato DC3000 [18] were downloaded from the NCBI FTP site ftp://ftp.ncbi.nih.gov/genomes/Bacteria/Pseudomonas_syringae_pv_B728a. The NCBI non-redundant (NR) Proteins database was downloaded from the NCBI FTP site ftp://ftp.ncbi.nih.gov/blast/db/ on 10th December 2008.

De novo sequence assembly and annotation

Solexa sequence data were assembled using Velvet 0.7.18 [41]. We used Softberry's FgenesB pipeline http://www.softberry.com to predict genes encoding rRNAs, tDNAs and proteins. Annotation of protein-coding genes by FgenesB was based on the NCBI NR Proteins database.

Prediction of HrpL-binding sites (Hrp boxes)

We built a profile hidden Markov model (HMM) based on a multiple sequence alignment of 26 known Hrp boxes from Pto DC3000 using hmmb from the HMMER 1.8.5 package http://hmmer.janelia.org. DNA sequence was scanned against this profile-HMM using hmmls from HMMER 1.8.5 with a bit-score cut-off of 12.0.

Functional screen for candidate type III effectors

Library preparation and the Flow cytometric-based screen for HrpL-induced genes of Pta 11528 were done according to [60].

Visualisation of data

We generated graphical views of genome alignments using CGView [76]. To visualise the annotation draft genome assembly of Pta 11528, we used the 'gbrowse' Generic Genome Browser [77].

Library preparation for Illumina sequencing

DNA was prepared from bacteria grown in L-medium using the Puregene Genomic DNA Purification Kit (Gentra Systems, Inc., Minneapolis, USA) according to manufacturer's instructions. A library for Illumina Paired-End sequencing was prepared from 5 mg DNA using a Paired-End DNA Sample Prep Kit (Pe-102-1001, Illumina, Inc., Cambridge, UK). DNA was fragmented by nebulisation for 6 min at a pressure of 32 psi. For end-repair and phosphorylation, sheared DNA was purified using QIAquick Nucleotide Removal Kit (Quiagen, Crawley, UK). The end repaired DNA was A-tailed and ada Pto rs were ligated according to manufacturer's instructions.

Size fractionation and purification of ligation products was performed using a 5% polyacrylamide gel run in TBE at 180V for 120 min. Gel slices were cut containing DNA in the 500 to 10 bp range. DNA was than extracted using 0.3 M sodium acetate and 2 mM EDTA [pH 8.0] followed by ethanol precipitation. Using 18 PCR cycles with primer PE1.0 and PE2.0 supplied by Illumina, 5' ada Pto r extension and enrichment of the library was performed. The library was finally purified using a QIAquick PCR Purification Kit and adjusted to a concentration of 10 nM in 0.1% Tween. The stock was kept at -20°C until used.

Sequencing

The flow cell was prepared according to manufacturer's instructions using a Paired-End Cluster Generation Kit (Pe-103-1001) and a Cluster Station. Sequencing reactions were performed on a 1G Genome Analyzer equipped with a Paired-End Module (Illumina, Inc., Cambridge, UK). 5 pM of the library were used to achieve ~20,000 to 25,000 clusters per tile. Capillary sequencing of avrE, HrpW1 and other individual genes was done on an ABI 3730. PCR products were directly sequenced after treatment with ExoI and SAP. Primer sequences are available upon request from JHC.

Verification of Illumina sequence data

Three of the core hop genes in Pta 11528 appeared to be degenerate, based on the de novo assembly of short Illumina sequence reads. The avrE1 gene appeared to have a 20-nucleotide deletion, hrpZ1 a 325-nucleotide deletion, whilst hrpW1 appeared to have three insertions of 22, 6 and 12 nucleotides. Currently, the reliability of de novo sequence assembly from short Illumina reads has not been fully characterised. In particular, repetitive and low-complexity sequence might generate artefacts in assembled supercontigs. Therefore, we checked these putative insertions and deletions by aligning the Illumina sequence reads against the relevant regions of both the Pph 1448A reference genome sequence and our Pta 11528 assembly. As an additional control, we also performed Velvet assembies on previously published Illumina short-read data from Psy B728a [35]. We found that the B728a avreE1, hrpZ1, hrpW1 and hopAF1 were assembled intact [Additional file 4: Figure S2], indicating that there is nothing inherently 'un-assemble-able' about these gene sequences. Sequence alignment is much more robust than de novo assembly and is not subject to assembly artefacts. The alignments supported the presence of a large deletion in hrpZ1. However, the alignments were not consistent with the assembly for avrE1 and hrpW1. Therefore, we amplified the Pta 11528 avrE1 and hrpW1 genes by PCR and verified their sequences by capillary sequencing [Additional file 5: Table S3]. This confirmed that the apparent deletion in avrE1 was an artefact of the de novo assembly and that the avrE1 sequence encodes a full-length protein product. Furthermore, transient expression of avrE1 in N. benthamiana induces cell death (S. Gimenez Ibanez and J. Rathjen, unpublished). Capillary sequencing also confirmed that the de novo assembly of hrpW1 was incorrect and that Pta 11528 encodes a full-length HrpW1 protein, albeit with repetitive sequence insertions of 69 and 12 nucleotides relative to the Pph 1448A sequence.

The absence of hopAF1 from Pta 11528 is supported not only by the de novo assembly, but also by the absence of aligned (unassembled) reads. As an additional control for the degeneracy of hopAF1 and hrpZ1, we performed the same bioinformatics and sequencing protocols to Psy B728a [35] and recovered hopAF1 and hrpZ1 intact in the de novo assembly assembly (Additional file 4: Figure S1).

Sequence data

In addition to the data available from Genbank accession ACHU00000000, the Velvet assembly and predicted protein sequences are provided in FastA format in Additional files 6 and Additional file 7.

Bioinformatics tools

We used GenomeMatcher [78] for generating and visualising whole-genome alignments. For aligning short Illumina sequence reads against a reference genome, we used MAQ [79] and for other sequence alignments and searches we used BLAST [80]. We used previously published complete genomes as reference sequences for comparative analyses [81–85].

Abbreviations

CEL:

conserved effector locus

EEL:

exchangeable effector locus

HMM:

hidden Markov model

HopDB:

Hop database

MCP:

methyl-accepting chemotaxis protein

PCR:

polymerase chain reaction

Pma Pseudomonas syringae :

pathovar maculicola

Pph Pseudomonas syringae :

pathovar phaseolicola

PPH GI-1 Pph:

genomic island 1

Pta Pseudomonas syringae :

pathovar tabaci

Psy Pseudomonas syringae :

pathovar syringae

Pto Pseudomonas syringae :

pathovar tomato

VIGS:

virus-induced gene silencing

IS:

insertion sequence.

References

  1. Gasson MJ: Indicator Technique for Antimetabolic Toxin Production by Phytopathogenic Species of Pseudomonas. Appl Environ Microbiol. 1980, 39: 25-29.

    PubMed Central  CAS  PubMed  Google Scholar 

  2. Ribeiro R de LD, Hagedorn DJ, Durbin RD, Uchytil TF: Characterization of the Bacterium Inciting Bean Wildfire in Brazil. Phytopathology. 1979, 69: 208-212. 10.1094/Phyto-69-208.

    Article  Google Scholar 

  3. Thomas MD, Langston-Unkefer PJ, Uchytil TF, Durbin RD: Inhibition of Glutamine Synthetase from Pea by Tabtoxinine-beta-lactam. Plant Physiol. 1983, 71: 912-915. 10.1104/pp.71.4.912.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  4. Durbin RD: Bacterial phytotoxins: Mechanisms of action. Cell Mol Life Sci. 1991, 47: 776-783. 10.1007/BF01922457.

    Article  CAS  Google Scholar 

  5. Turner JG, Taha RR: Contribution of tabtoxin to the pathogenicity of Pseudomonas syringae pv. tabaci. Physiol plant pathol. 1984, 25: 55-69. 10.1016/0048-4059(84)90017-1.

    Article  CAS  Google Scholar 

  6. Chang JH, Goel AK, Grant SR, Dangl JL: Wake of the flood: ascribing functions to the wave of type III effector proteins of phytopathogenic bacteria. Curr Opin Microbiol. 2004, 7: 11-18. 10.1016/j.mib.2003.12.006.

    Article  CAS  PubMed  Google Scholar 

  7. Collmer A, Lindeberg M, Petnicki-Ocwieja T, Schneider DJ, Alfano JR: Genomic mining type III secretion system effectors in Pseudomonas syringae yields new picks for all TTSS prospectors. Trends Microbiol. 2002, 10: 462-469. 10.1016/S0966-842X(02)02451-4.

    Article  CAS  PubMed  Google Scholar 

  8. Cornelis GR, Van Gijsegem F: Assembly and function of type III secretory systems. Annu Rev Microbiol. 2000, 54: 735-774. 10.1146/annurev.micro.54.1.735.

    Article  CAS  PubMed  Google Scholar 

  9. Cunnac S, Lindeberg M, Collmer A: Pseudomonas syringae type III secretion system effectors: repertoires in search of functions. Curr Opin Microbiol. 2009, 12: 53-60. 10.1016/j.mib.2008.12.003.

    Article  CAS  PubMed  Google Scholar 

  10. Grant SR, Fisher EJ, Chang JH, Mole BM, Dangl JL: Subterfuge and manipulation: type III effector proteins of phytopathogenic bacteria. Annu Rev Microbiol. 2006, 60: 425-449. 10.1146/annurev.micro.60.080805.142251.

    Article  CAS  PubMed  Google Scholar 

  11. Greenberg JT, Vinatzer BA: Identifying type III effectors of plant pathogens and analyzing their interaction with plant cells. Curr Opin Microbiol. 2003, 6: 20-28. 10.1016/S1369-5274(02)00004-8.

    Article  CAS  PubMed  Google Scholar 

  12. Jin Q, Thilmony R, Zwiesler-Vollick J, He SY: Type III protein secretion in Pseudomonas syringae. Microbes Infect. 2003, 5: 301-310. 10.1016/S1286-4579(03)00032-7.

    Article  CAS  PubMed  Google Scholar 

  13. Lindeberg M, Cartinhour S, Myers CR, Schechter LM, Schneider DJ, Collmer A: Closing the circle on the discovery of genes encoding Hrp regulon members and type III secretion system effectors in the genomes of three model Pseudomonas syringae strains. Mol Plant Microbe Interact. 2006, 19: 1151-1158. 10.1094/MPMI-19-1151.

    Article  CAS  PubMed  Google Scholar 

  14. Lindeberg M, Myers CR, Collmer A, Schneider DJ: Roadmap to new virulence determinants in Pseudomonas syringae: insights from comparative genomics and genome organization. Mol Plant Microbe Interact. 2008, 21: 685-700. 10.1094/MPMI-21-6-0685.

    Article  CAS  PubMed  Google Scholar 

  15. Wei CF, Kvitko BH, Shimizu R, Crabill E, Alfano JR, Lin NC, Martin GB, Huang HC, Collmer A: A Pseudomonas syringae pv. tomato DC3000 mutant lacking the type III effector HopQ1-1 is able to cause disease in the model plant Nicotiana benthamiana. Plant J. 2007, 51: 32-46. 10.1111/j.1365-313X.2007.03126.x.

    Article  CAS  PubMed  Google Scholar 

  16. de Torres M, Mansfield JW, Grabov N, Brown IR, Ammouneh H, Tsiamis G, Forsyth A, Robatzek S, Grant M, Boch J: Pseudomonas syringae effector AvrPtoB suppresses basal defence in Arabidopsis. Plant J. 2006, 47: 368-382. 10.1111/j.1365-313X.2006.02798.x.

    Article  CAS  PubMed  Google Scholar 

  17. Peñaloza-Vázquez A, Preston GM, Collmer A, Bender CL: Regulatory interactions between the Hrp type III protein secretion system and coronatine biosynthesis in Pseudomonas syringae pv. tomato DC3000. Microbiology. 2000, 146: 2447-2456.

    Article  PubMed  Google Scholar 

  18. Buell CR, Joardar V, Lindeberg M, Selengut J, Paulsen IT, Gwinn ML, Dodson RJ, Deboy RT, Durkin AS, Kolonay JF, Madupu R, Daugherty S, Brinkac L, Beanan MJ, Haft DH, Nelson WC, Davidsen T, Zafar N, Zhou L, Liu J, Yuan Q, Khouri H, Fedorova N, Tran B, Russell D, Berry K, Utterback T, Van Aken SE, Feldblyum TV, D'Ascenzo M, Deng WL, Ramos AR, Alfano JR, Cartinhour S, Chatterjee AK, Delaney TP, Lazarowitz SG, Martin GB, Schneider DJ, Tang X, Bender CL, White O, Fraser CM, Collmer A: The complete genome sequence of the Arabidopsis and tomato pathogen Pseudomonas syringae pv. tomato DC3000. Proc Natl Acad Sci USA. 2003, 100: 10181-10186. 10.1073/pnas.1731982100.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  19. Feil H, Feil WS, Chain P, Larimer F, DiBartolo G, Copeland A, Lykidis A, Trong S, Nolan M, Goltsman E, Thiel J, Malfatti S, Loper JE, Lapidus A, Detter JC, Land M, Richardson PM, Kyrpides NC, Ivanova N, Lindow SE: Comparison of the complete genome sequences of Pseudomonas syringae pv. syringae B728a and pv. tomato DC3000. Proc Natl Acad Sci USA. 2005, 102: 11064-11069. 10.1073/pnas.0504930102.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  20. Joardar V, Lindeberg M, Jackson RW, Selengut J, Dodson R, Brinkac LM, Daugherty SC, Deboy R, Durkin AS, Giglio MG, Madupu R, Nelson WC, Rosovitz MJ, Sullivan S, Crabtree J, Creasy T, Davidsen T, Haft DH, Zafar N, Zhou L, Halpin R, Holley T, Khouri H, Feldblyum T, White O, Fraser CM, Chatterjee AK, Cartinhour S, Schneider DJ, Mansfield J, Collmer A, Buell CR: Whole-genome sequence analysis of Pseudomonas syringae pv. phaseolicola 1448A reveals divergence among pathovars in genes involved in virulence and transposition. J Bacteriol. 2005, 187: 6488-6498. 10.1128/JB.187.18.6488-6498.2005.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  21. Gardan L, Shafik H, Belouin S, Broch R, Grimont F, Grimont PA: DNA relatedness among the pathovars of Pseudomonas syringae and description of Pseudomonas tremae sp. nov. and Pseudomonas cannabina sp. nov. (ex Sutic and Dowson 1959). Int J Syst Bacteriol. 1999, 49: 469-478.

    Article  CAS  PubMed  Google Scholar 

  22. Sarkar SF, Guttman DS: Evolution of the core genome of Pseudomonas syringae, a highly clonal, endemic plant pathogen. Appl Environ Microbiol. 2004, 70: 1999-2012. 10.1128/AEM.70.4.1999-2012.2004.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  23. Reinhardt JA, Baltrus DA, Nishimura MT, Jeck WR, Jones CD, Dangl JL: De novo assembly using low-coverage short read sequence data from the rice pathogen Pseudomonas syringae pv. oryzae. Genome Res. 2009, 19: 294-305. 10.1101/gr.083311.108.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  24. Almeida NF, Yan S, Lindeberg M, Studholme DJ, Schneider DJ, Condon B, Liu H, Viana CJ, Warren A, Evans C, Kemen E, Maclean D, Angot A, Martin GB, Jones JD, Collmer A, Setubal JC, Vinatzer BA: A draft genome sequence of Pseudomonas syringae pv. tomato T1 reveals a type III effector repertoire significantly divergent from that of Pseudomonas syringae pv. tomato DC3000. Mol Plant Microbe Interact. 2009, 22: 52-62. 10.1094/MPMI-22-1-0052.

    Article  CAS  PubMed  Google Scholar 

  25. Inoue Y, Takikawa Y: The hrpZ and hrpA genes are variable, and useful for grouping Pseudomonas syringae bacteria. J Gen Plant Pathol. 2006, 72: 26-33. 10.1007/s10327-005-0240-1.

    Article  CAS  Google Scholar 

  26. Taguchi F, Tanaka R, Kinoshita S, Ichinose Y, Imura Y, Andi S, Toyoda K, Tomonori Shiraishi T, Yamada T: Harpinpsta from Pseudomonas syringae pv. tabaci Is Defective and Deficient in Its Expression and HR-inducing Activity. Pseudomonas syringae. 2001, 67: 116-123.

    CAS  Google Scholar 

  27. Holt RA, Jones SJ: The new paradigm of flow cell sequencing. Genome Res. 2008, 18: 839-846. 10.1101/gr.073262.107.

    Article  CAS  PubMed  Google Scholar 

  28. Mardis ER: Next-generation DNA sequencing methods. Annu Rev Genomics. 2008, 9: 387-402. 10.1146/annurev.genom.9.081307.164359.

    Article  CAS  Google Scholar 

  29. MacLean D, Jones JDG, Studholme DJ: Application of 'next-generation' sequencing technologies to microbial genetics. Nature Rev Microbiol. 2009, 7: 287-296. 10.1038/nrmicro2088.

    Google Scholar 

  30. Aury JM, Cruaud C, Barbe V, Rogier O, Mangenot S, Samson G, Poulain J, Anthouard V, Scarpelli C, Artiguenave F, Wincker P: High quality draft sequences for prokaryotic genomes using a mix of new sequencing technologies. BMC Genomics. 2008, 9: 603-10.1186/1471-2164-9-603.

    Article  PubMed Central  PubMed  Google Scholar 

  31. Butler J, MacCallum I, Kleber M, Shlyakhter IA, Belmonte MK, Lander ES, Nusbaum C, Jaffe DB: ALLPATHS: de novo assembly of whole-genome shotgun microreads. Genome Res. 2008, 18: 810-820. 10.1101/gr.7337908.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  32. Chaisson M, Pevzner P, Tang H: Fragment assembly with short reads. Bioinformatics. 2004, 20: 2067-2074. 10.1093/bioinformatics/bth205.

    Article  CAS  PubMed  Google Scholar 

  33. Chaisson MJ, Pevzner PA: Short read fragment assembly of bacterial genomes. Genome Res. 2008, 18: 324-330. 10.1101/gr.7088808.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  34. Dohm JC, Lottaz C, Borodina T, Himmelbauer H: SHARCGS, a fast and highly accurate short-read assembly algorithm for de novo genomic sequencing. Genome Res. 2007, 17: 1697-706. 10.1101/gr.6435207.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  35. Farrer RA, Kemen E, Jones JDG, Studholme DJ: De novo assembly of the Pseudomonas syringae pv. syringae B728a genome using Illumina/Solexa short sequence reads. FEMS Microbiol Lett. 2009, 291: 103-111. 10.1111/j.1574-6968.2008.01441.x.

    Article  CAS  PubMed  Google Scholar 

  36. Hernandez D, Franvßois P, Farinelli L, Ostervs M, Schrenzel J: De novo bacterial genome sequencing: millions of very short reads assembled on a desktop computer. Genome Res. 2008, 18: 802-809. 10.1101/gr.072033.107.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  37. Jeck WR, Reinhardt JA, Baltrus DA, Hickenbotham MT, Magrini V, Mardis ER, Dangl JL, Jones CD: Extending assembly of short DNA sequences to handle error. Bioinformatics. 2007, 23: 2942-2944. 10.1093/bioinformatics/btm451.

    Article  CAS  PubMed  Google Scholar 

  38. Salzberg SL, Sommer DD, Puiu D, Lee VT: Gene-boosted assembly of a novel bacterial genome from very short reads. PLoS Comput Biol. 2008, 4: e1000186-10.1371/journal.pcbi.1000186.

    Article  PubMed Central  PubMed  Google Scholar 

  39. Sundquist A, Ronaghi M, Tang H, Pevzner P, Batzoglou S: Whole-genome sequencing and assembly with high-throughput, short-read technologies. PLoS ONE. 2007, 2: e484-10.1371/journal.pone.0000484.

    Article  PubMed Central  PubMed  Google Scholar 

  40. Warren RL, Sutton GG, Jones SJ, Holt R: Assembling millions of short DNA sequences using SSAKE. Bioinformatics. 2007, 23: 500-501. 10.1093/bioinformatics/btl629.

    Article  CAS  PubMed  Google Scholar 

  41. Zerbino DR, Birney E: Velvet: algorithms for de novo short read assembly using de Bruijn graphs. Genome Res. 2008, 18: 821-829. 10.1101/gr.074492.107.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  42. Kurtz S, Phillippy A, Delcher AL, Smoot M, Shumway M, Antonescu C, Salzberg SL: Versatile and open software for comparing large genomes. Genome Biol. 2004, 5: R12-10.1186/gb-2004-5-2-r12.

    Article  PubMed Central  PubMed  Google Scholar 

  43. Lindeberg M, Stavrinides J, Chang JH, Alfano JR, Collmer A, Dangl JL, Greenberg JT, Mansfield JW, Guttman DS: Proposed guidelines for a unified nomenclature and phylogenetic analysis of type III Hop effector proteins in the plant pathogen Pseudomonas syringae. Mol Plant Microbe Interact. 2005, 18: 275-282. 10.1094/MPMI-18-0275.

    Article  CAS  PubMed  Google Scholar 

  44. Alfano JR, Charkowski AO, Deng WL, Badel JL, Petnicki-Ocwieja T, van Dijk K, Collmer A: The Pseudomonas syringae Hrp pathogenicity island has a tripartite mosaic structure composed of a cluster of type III secretion genes bounded by exchangeable effector and conserved effector loci that contribute to parasitic fitness and pathogenicity in plants. Proc Natl Acad Sci USA. 2000, 97: 4856-4861. 10.1073/pnas.97.9.4856.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  45. Tampakaki AP, Panopoulos NJ: Elicitation of hypersensitive cell death by extracellularly targeted HrpZPsph produced in planta. Mol Plant Microbe Interact. 2000, 13: 1366-74. 10.1094/MPMI.2000.13.12.1366.

    Article  CAS  PubMed  Google Scholar 

  46. Preston G, Huang HC, He SY, Collmer A: The HrpZ proteins of Pseudomonas syringae pvs. syringae, glycinea, and tomato are encoded by an operon containing Yersinia ysc homologs and elicit the hypersensitive response in tomato but not soybean. Mol Plant Microbe Interact. 1995, 8: 717-732.

    Article  CAS  PubMed  Google Scholar 

  47. Lee J, Klessig DF, Nürnberger T: A harpin binding site in tobacco plasma membranes mediates activation of the pathogenesis-related gene HIN1 independent of extracellular calcium but dependent on mitogen-activated protein kinase activity. Plant Cell. 2001, 13: 1079-1093. 10.1105/tpc.13.5.1079.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  48. Li CM, Haapalainen M, Lee J, Nürnberger T, Romantschuk M, Taira S: Harpin of Pseudomonas syringae pv. phaseolicola harbors a protein binding site. Mol Plant Microbe Interact. 2005, 18: 60-66. 10.1094/MPMI-18-0060.

    Article  CAS  PubMed  Google Scholar 

  49. Maurelli AT: Black holes, antivirulence genes, and gene inactivation in the evolution of bacterial pathogens. FEMS Microbiol Lett. 2007, 267: 1-8. 10.1111/j.1574-6968.2006.00526.x.

    Article  CAS  PubMed  Google Scholar 

  50. Maurelli AT, Fernández RE, Bloch CA, Rode CK, Fasano A: "Black holes" and bacterial pathogenicity: a large genomic deletion that enhances the virulence of Shigella spp. and enteroinvasive Escherichia coli. Proc Natl Acad Sci USA. 1998, 95: 3943-3948. 10.1073/pnas.95.7.3943.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  51. Casalino M, Latella MC, Prosseda G, Ceccarini P, Grimont F, Colonna B: Molecular evolution of the lysine decarboxylase-defective phenotype in Shigella sonnei. Int J Med Microbiol. 2005, 294: 503-512. 10.1016/j.ijmm.2004.11.001.

    Article  CAS  PubMed  Google Scholar 

  52. Day WA, Fernández RE, Maurelli AT: Pathoadaptive mutations that enhance virulence: genetic organization of the cadA regions of Shigella spp. Infect Immun. 2001, 69: 7471-7480. 10.1128/IAI.69.12.7471-7480.2001.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  53. Tsiamis G, Mansfield JW, Hockenhull R, Jackson RW, Sesma A, Athanassopoulos E, Bennett MA, Stevens C, Vivian A, Taylor JD, Murillo J: Cultivar-specific avirulence and virulence functions assigned to avrPphF in Pseudomonas syringae pv. phaseolicola, the cause of bean halo-blight disease. EMBO J. 2000, 19: 3204-3214. 10.1093/emboj/19.13.3204.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  54. Jackson RW, Athanassopoulos E, Tsiamis G, Mansfield JW, Sesma A, Arnold DL, Gibbon MJ, Murillo J, Taylor JD, Vivian A: Identification of a pathogenicity island, which contains genes for virulence and avirulence, on a large native plasmid in the bean pathogen Pseudomonas syringae pathovar phaseolicola. Proc Natl Acad Sci USA. 1999, 96: 10875-10880. 10.1073/pnas.96.19.10875.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  55. Jackson RW, Mansfield JW, Arnold DL, Sesma A, Paynter CD, Murillo J, Taylor JD, Vivian A: Excision from tRNA genes of a large chromosomal region, carrying avr Pph B, associated with race change in the bean pathogen, Pseudomonas syringae pv. phaseolicola. Mol Microbiol. 2000, 38: 186-197. 10.1046/j.1365-2958.2000.02133.x.

    Article  CAS  PubMed  Google Scholar 

  56. Jenner C, Hitchin E, Mansfield J, Walters K, Betteridge P, Teverson D, Taylor J: Gene-for-gene interactions between Pseudomonas syringae pv. phaseolicola and Phaseolus. Mol Plant Microbe Interact. 1991, 4: 553-562.

    Article  CAS  PubMed  Google Scholar 

  57. Pitman AR, Jackson RW, Mansfield JW, Kaitell V, Thwaites R, Arnold DL: Exposure to host resistance mechanisms drives evolution of bacterial virulence in plants. Curr Biol. 2005, 15: 2230-2235. 10.1016/j.cub.2005.10.074.

    Article  CAS  PubMed  Google Scholar 

  58. Stavrinides J, Guttman DS: Nucleotide sequence and evolution of the five-plasmid complement of the phytopathogen Pseudomonas syringae pv. maculicola ES4326. J Bacteriol. 2004, 186: 5101-5115. 10.1128/JB.186.15.5101-5115.2004.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  59. Sundin GW: Genomic insights into the contribution of phytopathogenic bacterial plasmids to the evolutionary history of their hosts. Annu Rev Phytopathol. 2007, 45: 129-151. 10.1146/annurev.phyto.45.062806.094317.

    Article  CAS  PubMed  Google Scholar 

  60. Chang JH, Urbach JM, Law TF, Arnold LW, Hu A, Gombar S, Grant SR, Ausubel FM, Dangl JL: A high-throughput, near-saturating screen for type III effector genes from Pseudomonas syringae. Proc Natl Acad Sci USA. 2005, 102: 2549-2554. 10.1073/pnas.0409660102.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  61. Oh HS, Kvitko BH, Morello JE, Collmer A: Pseudomonas syringae lytic transglycosylases coregulated with the type III secretion system contribute to the translocation of effector proteins into plant cells. J Bacteriol. 2007, 189: 8277-8289. 10.1128/JB.00998-07.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  62. Von Bodman SB, Bauer WD, Coplin DL: Quorum sensing in plant-pathogenic bacteria. Annu Rev Phytopathol. 2003, 41: 455-482. 10.1146/annurev.phyto.41.052002.095652.

    Article  CAS  PubMed  Google Scholar 

  63. Brencic A, Winans SC: Detection of and response to signals involved in host-microbe interactions by plant-associated bacteria. Microbiol Mol Biol Rev. 2005, 69: 155-294. 10.1128/MMBR.69.1.155-194.2005.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  64. Taguchi F, Shibata S, Suzuki T, Ogawa Y, Aizawa S, Takeuchi K, Ichinose Y: Effects of glycosylation on swimming ability and flagellar polymorphic transformation in Pseudomonas syringae pv. tabaci 6605. J Bacteriol. 2008, 190: 764-768. 10.1128/JB.01282-07.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  65. Cha JY, Lee JS, Oh JI, Choi JW, Baik HS: Functional analysis of the role of Fur in the virulence of Pseudomonas syringae pv. tabaci 11528: Fur controls expression of genes involved in quorum-sensing. Biochem Biophys Res Commun. 2008, 366: 281-287. 10.1016/j.bbrc.2007.11.021.

    Article  CAS  PubMed  Google Scholar 

  66. Jakob K, Kniskern JM, Bergelson J: The role of pectate lyase and the jasmonic acid defense response in Pseudomonas viridiflava virulence. Mol Plant Microbe Interact. 2007, 20: 146-158. 10.1094/MPMI-20-2-0146.

    Article  CAS  PubMed  Google Scholar 

  67. Preston GM, Studholme DJ, Caldelari I: Profiling the secretomes of plant pathogenic Proteobacteria. FEMS Microbiol Rev. 2005, 29: 331-360. 10.1016/j.femsre.2004.12.004.

    Article  CAS  PubMed  Google Scholar 

  68. Barabote RD, Johnson OL, Zetina E, San Francisco SK, Fralick JA, San Francisco MJ: Erwinia chrysanthemi tolC is involved in resistance to antimicrobial plant chemicals and is essential for phytopathogenesis. J Bacteriol. 2003, 185: 5772-5778. 10.1128/JB.185.19.5772-5778.2003.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  69. Kinscherf TG, Willis DK: The biosynthetic gene cluster for the beta-lactam antibiotic tabtoxin in Pseudomonas syringae. J Antibiot. 2005, 58: 817-821. 10.1038/ja.2005.109.

    Article  CAS  PubMed  Google Scholar 

  70. Greenberg DE, Porcella SF, Zelazny AM, Virtaneva K, Sturdevant DE, Kupko JJ, Barbian KD, Babar A, Dorward DW, Holland SM: Genome sequence analysis of the emerging human pathogenic acetic acid bacterium Granulibacter bethesdensis. J Bacteriol. 2007, 189: 8727-8736. 10.1128/JB.00793-07.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  71. Arand M, Hemmer H, Dürk H, Baratti J, Archelas A, Furstoss R, Oesch F: Cloning and molecular characterization of a soluble epoxide hydrolase from Aspergillus niger that is related to mammalian microsomal epoxide hydrolase. Biochem J. 1999, 344: 273-280. 10.1042/0264-6021:3440273.

    PubMed Central  CAS  PubMed  Google Scholar 

  72. Finn RD, Tate J, Mistry J, Coggill PC, Sammut SJ, Hotz HR, Ceric G, Forslund K, Eddy SR, Sonnhammer EL, Bateman A: The Pfam protein families database. Nucleic Acids Res. 2008, 36: D281-8. 10.1093/nar/gkm960.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  73. Mandel MJ, Wollenberg MS, Stabb EV, Visick KL, Ruby EG: A single regulatory gene is sufficient to alter bacterial host range. Nature. 2009, 12: 215-8. 10.1038/nature07660.

    Article  Google Scholar 

  74. Pasloske BL, Sastry PA, Finlay BB, Paranchych W: Two unusual pilin sequences from different isolates of Pseudomonas aeruginosa. J Bacteriol. 1988, 170: 3738-3741.

    PubMed Central  CAS  PubMed  Google Scholar 

  75. Pizarro-Cerdá J, Cossart P: Bacterial adhesion and entry into host cells. Cell. 2006, 124: 715-727. 10.1016/j.cell.2006.02.012.

    Article  PubMed  Google Scholar 

  76. Stothard P, Wishart DS: Circular genome visualization and exploration using CGView. Bioinformatics. 2005, 21: 537-539. 10.1093/bioinformatics/bti054.

    Article  CAS  PubMed  Google Scholar 

  77. Stein LD, Mungall C, Shu S, Caudy M, Mangone M, Day A, Nickerson E, Stajich JE, Harris TW, Arva A, Lewis S: The generic genome browser: a building block for a model organism system database. Genome Res. 2002, 12: 1599-1610. 10.1101/gr.403602.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  78. Ohtsubo Y, Ikeda-Ohtsubo W, Nagata Y, Tsuda M: GenomeMatcher: a graphical user interface for DNA sequence comparison. BMC Bioinformatics. 2008, 9: 376-10.1186/1471-2105-9-376.

    Article  PubMed Central  PubMed  Google Scholar 

  79. Li H, Ruan J, Durbin R: Mapping short DNA sequencing reads and calling variants using mapping quality scores. Genome Res. 2008, 18: 1851-1858. 10.1101/gr.078212.108.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  80. Altschul SF, Gish W, Miller W, Myers EW, Lipman DJ: Basic local alignment search tool. J Mol Biol. 1990, 215: 403-410.

    Article  CAS  PubMed  Google Scholar 

  81. Camargo LE, da Silva AC, Moon DH, Takita MA, Lemos EG, Machado MA, Ferro MI, da Silva FR, Goldman MH, Goldman GH, Lemos MV, El-Dorry H, Tsai SM, Carrer H, Carraro DM, de Oliveira RC, Nunes LR, Siqueira WJ, Coutinho LL, Kimura ET, Ferro ES, Harakava R, Kuramae EE, Marino CL, Giglioti E, Abreu IL, Alves LM, do Amaral AM, Baia GS, Blanco SR, Brito MS, Cannavan FS, Celestino AV, da Cunha AF, Fenille RC, Ferro JA, Formighieri EF, Kishi LT, Leoni SG, Oliveira AR, Rosa VE, Sassaki FT, Sena JA, de Souza AA, Truffi D, Tsukumo F, Yanai GM, Zaros LG, Civerolo EL, Simpson AJ, Almeida NF, Setubal JC, Kitajima JP: Comparative analyses of the complete genome sequences of Pierce's disease and citrus variegated chlorosis strains of Xylella fastidiosa. J Bacteriol. Edited by: Young JP, Crossman LC, Johnston AW, Thomson NR, Ghazoui ZF, Hull KH, Wexler M. 2003, 185: 1018-26. 10.1128/JB.185.3.1018-1026.2003.

    Google Scholar 

  82. Curson AR, Todd JD, Poole PS, Mauchline TH, East AK, Quail MA, Churcher C, Arrowsmith C, Cherevach I, Chillingworth T, Clarke K, Cronin A, Davis P, Fraser A, Hance Z, Hauser H, Jagels K, Moule S, Mungall K, Norbertczak H, Rabbinowitsch E, Sanders M, Simmonds M, Whitehead S, Parkhill J: The genome of Rhizobium leguminosarum has recognizable core and accessory components. Genome Biol. 2006, 7: R34-10.1186/gb-2006-7-4-r34.

    Article  PubMed Central  PubMed  Google Scholar 

  83. Hayden HS, Gillett W, Saenphimmachak C, Lim R, Zhou Y, Jacobs MA, Chang J, Rohmer L, D'Argenio DA, Palmieri A, Levy R, Haugen E, Wong GK, Brittnacher MJ, Burns JL, Miller SI, Olson MV, Kaul R: Large-insert genome analysis technology detects structural variation in Pseudomonas aeruginosa clinical strains from cystic fibrosis patients. Genomics. 2008, 91: 530-537. 10.1016/j.ygeno.2008.02.005.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  84. Iacono M, Villa L, Fortini D, Bordoni R, Imperi F, Bonnal RJ, Sicheritz-Ponten T, De Bellis G, Visca P, Cassone A, Carattoli A: Whole-genome pyrosequencing of an epidemic multidrug-resistant Acinetobacter baumannii strain belonging to the European clone II group. Antimicrob Agents Chemother. 2008, 52: 2616-2625. 10.1128/AAC.01643-07.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  85. Salanoubat M, Genin S, Artiguenave F, Gouzy J, Mangenot S, Arlat M, Billault A, Brottier P, Camus JC, Cattolico L, Chandler M, Choisne N, Claudel-Renard C, Cunnac S, Demange N, Gaspin C, Lavie M, Moisan A, Robert C, Saurin W, Schiex T, Siguier P, Thébault P, Whalen M, Wincker P, Levy M, Weissenbach J, Boucher CA: Genome sequence of the plant pathogen Ralstonia solanacearum. Nature. 2002, 415: 497-502. 10.1038/415497a.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We would like to thank Dr. Larry Arnold and Theresa Law for their assistance in the functional screen for candidate type III effector genes of Pta 11528, Jodie Pike for performing the Illumina sequencing, Eric Kemen for advice and help with Illumina sequencing and Kee Hoon Sohn for handling and maintaining the bacterial strains. We thank Ashley Chu and Caitlin Thireault for assistance with capillary sequencing and Dr. Jonathan Urbach for guidance on searching for hrp-boxes. We thank Robert Jackson, Gail Preston and two anonymous reviewers for very helpful comments on the manuscript. Financial support by the Gatsby Charitable Foundation is gratefully acknowledged.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to David J Studholme.

Additional information

Authors' contributions

DJS and DM performed the sequence assembly and all subsequent bioinformatics analyses. SGI prepared the DNA libraries and performed phenotypic characterisation of Pta 11528. JHC performed the functional screen for HrpL-dependent genes and analysed the resulting data. JR, DJS and JHC conceived of the study, participated in its design. DJS, JR and JHC wrote the manuscript. All authors read, approved and made contributions to the manuscript.

Electronic supplementary material

12864_2009_2279_MOESM1_ESM.html

Additional file 1: Table S1. Proteins encoded in the Pta 11528 draft with no detectable homologue in previously sequenced P. syringae genomes (Pto DC3000, Psy B728a and Pph 1448A). Proteins implicated in mobile genetic elements are shaded in cyan. Other proteins for which a function could be predicted by homology are shaded in yellow. (HTML 104 KB)

12864_2009_2279_MOESM2_ESM.pdf

Additional file 2: Figure S1. Alignment of the Pph 1302A PPHGI-1 pathogenicity island against the Pta 11528 genome assembly. The Pta 11528 genome sequence is in the upper track, aligned against the Pph 1302A PPHGI-1 pathogenicity island sequence. (Genbank: AJ870974). (PDF 34 KB)

12864_2009_2279_MOESM3_ESM.html

Additional file 3: Table S2. Regions of the Pta 11528 genome with no nucleotide sequence similarity to the genomes of Pto DC3000, Pss B728a and Pph 1448A. (HTML 67 KB)

12864_2009_2279_MOESM4_ESM.pdf

Additional file 4: Figure S2. The avrE1 , hrpZ1 , hrpW1 and hopAF1 genes are recovered intact in a de novo sequence assembly of Illumina short sequence reads from Psy B728a. We assembled a 40 × deep dataset (reference 35) of paired 36-nucleotide reads from Psy B728a genomic DNA using Velvet 0.7.18, using the same protocol as for the Pta 11528 data. Panel A shows the MAQ alignment of the B728a Illumina reads (in black) and the blastn alignment of the B728a de novo assembly (in green) against the avrE1 gene in the B728a genome. Panel B shows the alignments against hrpZ1. Panel C shows the alignments against hrpW1. Panel D shows the alignments against hopAF1. (PDF 1 MB)

12864_2009_2279_MOESM5_ESM.html

Additional file 5: Table S3. Verification of predicted genes by capillary sequencing. We verified a selection of genes predicted from the Illumina-based Pta11528 genome sequence assembly by capillary sequencing of cloned PCR products. Sequence reads were trimmed to remove poor quality nucleotide calls and the trimmed sequences were aligned against predicted proteins using TBLASTN. (HTML 82 KB)

Additional file 6:The Pta 11528 draft genome assembly, in FastA format, generated using Velvet(FNA 6 MB)

12864_2009_2279_MOESM7_ESM.faa

Additional file 7:Protein sequences, in FastA format, predicted in the Pta 11528 draft genome assembly using FgenesB(FAA 2 MB)

Authors’ original submitted files for images

Rights and permissions

This article is published under license to BioMed Central Ltd. This is an Open Access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/2.0), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

Reprints and permissions

About this article

Cite this article

Studholme, D.J., Ibanez, S.G., MacLean, D. et al. A draft genome sequence and functional screen reveals the repertoire of type III secreted proteins of Pseudomonas syringae pathovar tabaci 11528. BMC Genomics 10, 395 (2009). https://0-doi-org.brum.beds.ac.uk/10.1186/1471-2164-10-395

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://0-doi-org.brum.beds.ac.uk/10.1186/1471-2164-10-395

Keywords