You are on page 1of 7

International Journal of Systematic and Evolutionary Microbiology (2007), 57, 2881–2887 DOI 10.1099/ijs.0.

65220-0

Phylogenetic analysis of Xanthomonas species by


comparison of partial gyrase B gene sequences
Neil Parkinson,1 Valentine Aritua,2 John Heeney,1 Claire Cowie,1
Janice Bew1 and David Stead1
Correspondence 1
Central Science Laboratory, Sand Hutton, York YO41 1LZ, UK
Neil Parkinson 2
National Agricultural Biotechnology Centre, Kawanda Agricultural Research Institute,
n.parkinson@csl.gov.uk
PO Box 7065, Kampala, Uganda

The genus Xanthomonas currently comprises 27 species with validly published names that are
important crop and horticultural pathogens. We have constructed a phylogram from alignment of
gyrase B (gyrB) sequences for all xanthomonad species, both to indicate inter-species
relatedness and as an aid for rapid and accurate species-level identification. The phylogeny
indicated a monophyletic group, with X. albilineans and X. sacchari as the most ancestral species.
Three species, X. hyacinthi, X. translucens and X. theicola, formed an early-branching group.
Three clades were supported by high bootstrap values: group 1 comprised X. cucurbitae, X.
cassavae and X. codiaei; group 2 comprised X. arboricola, X. campestris, X. populi, X. hortorum, X.
gardneri and X. cynarae; group 3 contained the remaining species, within which two further
clades, supported by a 100 % bootstrap value, were identified. Group 3A comprised X.
axonopodis, X. euvesicatoria, X. perforans and X. melonis, together with X. alfalfae, X. citri and X.
fuscans, whose names were recently validly published. Group 3B contained the monocot
pathogens X. vasicola and X. oryzae. Two recently identified species, X. cynarae and X. gardneri,
were poorly discriminated and were related closely to X. hortorum. Three species, X. perforans, X.
euvesicatoria and X. alfalfae, had identical gyrB sequences. Partial sequencing of a further five
genes from these species found only minor sequence differences that confirmed their close
relatedness. Although branch lengths between species varied, indicating different degrees of
genetic distinctiveness, the majority (n521) were well-differentiated, indicating the utility of the
method as an identification tool, and we now use this method for routine diagnosis of
xanthomonad species.

INTRODUCTION The complexity of the genus and resource requirements for


comprehensive pairwise analyses required for DNA–DNA
Classification of species within the genus Xanthomonas,
hybridization analysis make this technique unsuitable for
which was hitherto based on biochemical characteristics
routine identification in most diagnostic laboratories. A
(Van den Mooter & Swings, 1990), underwent major revision
rapid characterization method based on analysing cell-wall
after a comprehensive spectrophotometric DNA–DNA
fatty acids using GC was developed and applied to the
hybridization study of the genus (Vauterin et al., 1995),
identification of plant pathogens (Stead, 1989, 1992).
which resulted in the recognition of 20 constituent species.
Whilst this method can be very useful for typing of some
Following this study, names of seven further species have
xanthomonads, it does not provide the necessary resolu-
been validly published: X. cynarae (Trébaol et al., 2000), X.
tion or reproducibility to be useful for all species. More
euvesicatoria, X. perforans, X. gardneri (Jones et al., 2004;
recently, a major and comprehensive analysis of the genus
Euzéby, 2006), X. citri, X. fuscans and X. alfalfae (Gabriel
was completed (Rademaker et al., 2005) that compared
et al., 1989; Euzéby, 2007; Schaad et al., 2005, 2006, 2007).
amplicon profiles produced by using PCR primers
designed to repeated sequences dispersed in bacterial
The GenBank/EMBL/DDBJ accession numbers for the gyrB sequences
of xanthomonad species determined in this study are EU007516–
genomes (Martin et al., 1992; Louws et al., 1994;
EU007542 and DQ676938. Thwaites et al., 1999). This study utilized all three
commonly used repetitive PCR methods (REP, BOX and
Supplementary tables showing type strains of all xanthomonad species
used in the study, PCR and sequencing primers and a similarity matrix ERIC) to produce a high-resolution identification scheme,
indicating all inter-species percentage nucleotide identities are available and encompassed many of the pathovars of X. campestris
with the online version of this paper. that could not be resolved in the earlier study by Vauterin

65220 G 2007 Crown copyright Printed in Great Britain 2881


N. Parkinson and others

et al. (1995) using DNA–DNA hybridization. Comparison setting (100 uC) in a dry block heater. After cooling, the prepared
of gel profiles, however, still presents problems outside DNA suspensions were frozen at 220 uC prior to use, when 1 ml
reference laboratories, and inter-laboratory standardization preparation was added to 24 ml PCR reagent mix. Centrifuging of the
boiled cells or other purification treatment was not required.
can be problematic. DNA sequencing studies of the 16S
rRNA gene and the 16S–23S intergenic region of As xanthomonads can have very high G+C contents, a blend of Taq
Xanthomonas species (Hauben et al., 1997; Gonçalves & polymerase and a proof-reading polymerase was used for all
Rosato, 2002) have indicated the phylogenetic position of amplifications (Long PCR Enzyme mix; Fermentas). The PCR buffer
included magnesium and was as supplied by the enzyme supplier. The
the genus within the Gammaproteobacteria. Whilst the reaction components comprised: forward primer (10 pmol ml21),
16S–23S phylogeny had greater resolution than that 0.75 ml; reverse primer (10 pmol ml21), 0.75 ml; 106 PCR buffer
produced from the 16S study, most of the species within (including 15 mM MgCl2), 2.5 ml; dNTP mix (2.5 mM each), 2.0 ml;
the genus fell into a single group, and neither method Long PCR Enzyme mix, 0.125 ml; prepared DNA, 1.0 ml; water,
could be used reliably to differentiate the majority of 17.9 ml. The final concentrations of reagents were: MgCl2, 1.5 mM;
species. dNTPs, 0.2 mM; primers, 0.3 mM each; enzyme mix, 0.75 units (all
per 25 ml reaction volume). PCR and sequencing primers are shown
Recently, DNA sequencing of genes encoding conserved in Supplementary Table S2, available in IJSEM Online. We routinely
proteins involved in essential cell processes that constitute used standard primers that produced large yields of PCR product for
the ‘core genome’ has been used to produce phylogenies all species except X. albilineans and related xanthomonads; for these
species, the alternative conserved primer set was used. The conserved
that are more resolving than 16S and 16S–23S studies.
primers were designed from accrued gyrB alignments from this study
Increasingly, this approach is being adopted for bacterial and can be used for gyrB amplification from all species. The PCR
identification, as the method provides a convenient and cycling conditions were 2.5 min at 94 uC, then 30 s at 94 uC, 45 s at
rapid means of providing information on the relatedness of 50 uC and 1 min at 68 uC for 34 cycles, then 7 min at 68 uC to end.
species and strains. This study aimed to produce a
Purity and yield of PCR product for sequencing were checked by
xanthomonad phylogeny based on partial sequence align- running 5 ml reaction mixture on a 1.5 % agarose gel and post-
ments of the gyrB gene, both to indicate xanthomonad staining using ethidium bromide, according to standard protocols
relatedness and as a simple and reliable means of species- (Thwaites et al., 1999). The remaining mixture was purified by using a
level identification. commercial kit to remove residual primers and other reaction
components (Wizard PCR Clean-up kit; Promega). After elution of
As published DNA–DNA hybridization values are available bound DNA in 50 ml water, the equivalent of approximately 12 mg
for species comparisons (Vauterin et al., 1995), it was DNA (usually 5–10 ml purified product) was tranferred to a fresh
possible to relate percentage gyrB sequence identities to microfuge tube and lyophilized prior to sequencing from a
these values for several species comparisons, and to commercial service (MWG Ltd), using a Sanger sequencing-based
produce a scatter graph to clarify the relationship between fluorescent-label PCR method. The sequences, provided in FASTA
format files, were cropped further to provide a 530 nt sequence,
the two methods for quantifying DNA sequence similarity.
corresponding to nt 1807–2336 in the gyrB sequence from X.
During the study, it was found that X. euvesicatoria, X. euvesicatoria strain 85-10. The cropped sequences all start with
perforans and X. alfalfae had identical gyrB sequences. To ATCACCGG; the end of the sequence finishes (with some variants)
address this anomaly and to clarify relatedness between with AGCTGTG.
these species, partial sequences from an additional five
genes were analysed. Concatenated sequences from these Sequencing of X. euvesicatoria, X. perforans and X. alfalfae
loci were produced and compared in a phylogeny along using additional gene loci. PCR amplification and sequencing
with similar sequences derived from genome sequences for primers for a further five loci [sigma factor 70 (rpoD), aconitate
hydratase B (acnB), phosphoglucoisomerase (pgi), glyceraldehyde-3-
X. axonopodis pv. citri, X. oryzae and X. campestris. phosphate dehydrogenase (gap) and phosphofructokinase (pfkA)]
were designed from the X. euvesicatoria genome sequence (strain 85-
10) available from GenBank (accession no. CAJ23349) (Thieme et al.,
METHODS 2005). PCR conditions, cycling parameters and sequencing protocol
were the same as used above for the gyrB analysis. PCR and
Bacterial strains. Type strains of all xanthomonad species used in sequencing primers for multi-locus sequence typing of X. euvesica-
the study were obtained from international culture collections toria, X. perforans and X. alfalfae are shown in Supplementary Table
(mostly the National Collection of Plant Pathogenic Bacteria; S3, available in IJSEM Online.
NCPPB) and their reference numbers, together with alternative strain
references from other collections, are indicated in Supplementary Concatenated sequences were compared with the same loci derived
Table S1, available in IJSEM Online. Species are as indicated in the from GenBank for X. campestris and X. axonopodis pv. citri (da Silva
phylogram (Fig. 1). et al., 2002), as well as for X. oryzae (Lee et al., 2005).

gyrB locus and PCR amplification. Primers were designed from Sequence alignment and phylogenetic analysis. All loci were
xanthomonad alignments of gyrB sequences available from GenBank sequenced using both forward and reverse primers. Unprocessed gyrB
and corresponded to the C-terminal region of the protein. In relation sequences were cropped to a common start and finish sequence to
to the gyrB gene from X. euvesicatoria (strain 85-10; GenBank produce 530 nt sequences that were aligned by using the CLUSTAL W
accession no. CAJ23349), the region amplified corresponded to program (Thompson et al., 1994), available at http://www.ebi.ac.uk/
aa 603–778, of a total of 832 residues for this protein. DNA, prepared clustalw. Phylogenies were rooted by using the gyrB sequence from a
from suspensions of freshly grown xanthomonad cultures, in water laboratory isolate that, by 16S rRNA gene sequencing, was related
was adjusted to an A650 of 0.1 using a spectrophotometer, and then most closely to the genus Stenotrophomonas. The TREECON program
heated in a microfuge tube for 6 min at the highest temperature (Van de Peer & De Wachter, 1993) was used for phylogenetic analysis.

2882 International Journal of Systematic and Evolutionary Microbiology 57


Xanthomonas phylogeny

Fig. 1. gyrB phylogeny of the genus Xanthomonas produced by using the neighbourhood-joining method. Sequences were
analysed by using the method of Jukes & Cantor (1969). Percentage bootstrap values .50 % from 500 samplings are
indicated. All species are type strains, except for X. hortorum pv. pelargonii. The phylogeny was rooted by using a laboratory
isolate that, by 16S rRNA gene sequencing, was related most closely to the genus Stenotrophomonas. Bar, 0.1 substitutions
per site.

Distance estimation was done by using two methods available in the nucleotide identities is provided in Supplementary Table
program (Jukes & Cantor, 1969; Tajima & Nei, 1984). Tree S4, available in IJSEM Online.
construction used the neighbour-joining method (Saitou & Nei,
1987). The two most ancestral species within the phylogeny were
X. albilineans and X. sacchari. The relative phylogenetic
positions of these species, and of the clade containing X.
RESULTS theicola, X. translucens and X. hyacinthi, were supported by
low bootstrap values. To account for this ambiguity in the
Xanthomonas gyrB phylogeny relationship between these basal species, which occurs
often in the most ancestral regions of phylogenies (Sarker
The phylogeny indicates a monophyletic classification, in & Guttman, 2004), they are designated ‘early-branching
agreement with previous studies comparing 16S rRNA species’ in the phylogram. In contrast, three major clades
gene and 16S–23S intergenic sequences (Fig. 1). Sequence representing the remaining species were supported by high
alignments revealed the absence of any insertions or bootstrap values (.77 %). Group 1 comprised X. cassavae,
deletions, consistent with a single origin for the genus. X. codiaei and X. cucurbitae. Group 2 comprised X.
Of 530 nucleotide positions, 55.3 % were invariant. Inter- arboricola, X. campestris, X. populi, X. hortorum, X. gardneri
species sequence similarities ranged between 75 and 100 %; and X. cynarae, which are all species indigenous to Europe.
a similarity matrix indicating all inter-species percentage Group 3 comprises a large group of the remaining species,

http://ijs.sgmjournals.org 2883
N. Parkinson and others

and contained two further clades (3A and 3B) that were
supported by a high bootstrap value (100 %). Clade 3A
comprised X. perforans, X. euvesicatoria, X. alfalfae, X.
melonis, X. axonopodis, X. fuscans, X. citri and X. bromi,
whilst clade 3B comprised the monocot pathogens X.
vasicola and X. oryzae.
Most of the reference type species (n521) were distin-
guished clearly according to branch length. Exceptions
were X. perforans, X. euvesicatoria and X. alfalfae, which
had identical sequences. The relatedness of these species
was investigated further by partial sequencing of an
additional five loci and these results are described below.
The type strains of X. hortorum (pv. hederae), X. gardneri
and X. cynarae were distinguished only by very short
branch lengths compared with other species. The pelargonii
pathovar of X. hortorum, which was used in DNA–DNA Fig. 2. Relationship between DNA–DNA hybridization and gyrB
hybridization studies to support X. cynarae and X. gardneri sequence. Mean DNA–DNA hybridization values for species
as distinct species, was separated from the type strain of X. comparisons (Vauterin et al., 1995) are plotted against corres-
hortorum with a branch length similar to those distinguish- ponding comparisons using percentage gyrB identities. The paired
ing other species of Xanthomonas, and was supported by a species comparisons were: X. populi vs X. hortorum; X. populi vs
high bootstrap value (97 %). X. cassavae; X. populi vs X. codiaei; X. populi vs X. hyacinthi; X.
populi vs X. sacchari; X. codiaei vs X. cassavae; X. codiaei vs X.
bromi; X. codiaei vs X. vesicatoria; X. codiaei vs X. hyacinthi; X.
Relationship between gyrB sequences and codiaei vs X. sacchari; X. vesicatoria vs X. hyacinthi; X. vesicatoria
reported DNA–DNA hybridization values for vs X. sacchari; X. hyacinthi vs X. sacchari; X. bromi vs X. cassavae;
paired species comparisons X. oryzae vs X. cassavae; X. vasicola vs X. cassavae; X. populi vs X.
fragariae; X. codiaei vs X. fragariae; X. populi vs X. translucens; X.
The percentage gyrB sequence similarity and published
hortorum vs X. cucurbitae; X. codiaei vs X. cucurbitae; X. oryzae vs
DNA–DNA hybridization values (Vauterin et al., 1995) for
X. vasicola.
24 species comparisons (Fig. 2) were selected for plotting
on a scatter graph. As the reported hybridization values
were means from groups of strains, those species
comparison values with high standard errors (SEM) were DISCUSSION
avoided. The paired sequence identity and DNA–DNA
Until recently, phylogenetic analysis of plant-pathogenic
hybridization values were plotted as a scatter plot by using
bacteria based on protein-encoding genes has been used to
the SigmaPlot software package. The scatter plot (Fig. 2)
analyse diversity within a species. Endoglucanase and hrpB
indicates a clear correlation between the published DNA–
genes have provided a detailed description of sequevar
DNA hybridization data and gyrB percentage sequence
diversity within Ralstonia solanacearum, which identified
identity for the same-paired species comparisons, sup-
five major phylotypes associated with distinct geographical
ported by a Pearson correlation coefficient of 0.755.
regions (Poussier et al., 2000; Fegan & Prior, 2005). Seven
loci were analysed in a large multi-locus sequence study of
Relatedness between X. euvesicatoria, Pseudomonas syringae (Sarkar & Guttman, 2004; Sarkar
X. perforans and X. alfalfae, determined by using et al., 2006) that provided high-resolution analyses of
additional gene loci lineages within five major clades identified within this
species. Sequence analysis revealed that recombination
Concatenated rpoD, acnB, pgi, gap and pfkA sequences were
accounted for a very small proportion of nucleotide
cropped to a total length of 2720 nt. The CLUSTAL W
divergence compared with other bacteria. This was
sequence alignment indicated scores of 99 % sequence
suggested to be attributable to adaptation to specific host
identity for all three comparisons of X. euvesicatoria, X.
plants, which restricted contact with other lineages and
perforans and X. alfalfae. The unrooted phylogeny
minimized genetic exchange. This feature may have
produced from the alignment (Fig. 3) indicated that these
contributed to a phylogeny that reflects very accurately
three species were most similar to X. citri, in agreement
the evolutionary history of this species. As multiple loci
with the gyrB phylogeny. Branch lengths between the three
were analysed, it was possible to compare phylogenies that
species were very short, indicating that none of the
revealed close agreement between six of the seven loci
additional five loci sequenced produced a species-level
examined.
differentiation. The close sequence similarity found for all
five additional loci confirms a close level of core-genome Most recently, phylogenies have been developed that
relatedness for X. euvesicatoria, X. perforans and X. alfalfae. encompass species within higher-level taxa, including the

2884 International Journal of Systematic and Evolutionary Microbiology 57


Xanthomonas phylogeny

Fig. 3. Core-genome relatedness of X. perfor-


ans, X. euvesicatoria and X. alfalfae. The
unrooted neighbour-joining phylogram was
produced from concatenated sequences
derived from the rpoD, acnB, pgi, gap and
pfkA loci. Sequences from the same loci for X.
axonopodis pv. citri (strain 306), X. oryzae
(KACC 10331) and X. campestris (ATCC
33913T) were derived from genome
sequences (da Silva et al., 2002; Lee et al.,
2005; Thieme et al., 2005). Bar, 0.05 sub-
stitutions per site.

genus Pseudomonas and the family Enterobacteriaceae (Ait value. Clade 3A comprises both monocot and eudicot
Tayeb et al., 2005; Paradis et al., 2005). Both of these pathogens, including economically important pathogens of
studies revealed good correlation between established bean, citrus, tomato and pepper. Clade 3B comprises the
reference species and their phylogenetic identification. sorghum and rice pathogens X. vasicola and X. oryzae.
The two genes used to study the family Enterobacteriaceae
DNA–DNA hybridization data are considered to be the
(Paradis et al., 2005), elongation factor Tu (tuf) and F-
‘gold standard’ to measure relatedness between species. The
ATPase b-subunit (atpD), produced phylogenies of similar
scatter graph clearly indicates a correlation between DNA
resolution. However, comparison of the two loci used for
hybridization and the gyrB phylogeny, which is supported
analysis of the genus Pseudomonas (Ait Tayeb et al., 2005)
by the high Pearson correlation coefficient (0.755), and this
revealed that the rpoB sequences produced a phylogeny of
forms the basis for the general agreement between the
approximately three times higher resolution than that
two methods of distinguishing xanthomonad species. The
produced by using the alternative rrs gene, indicating that
deviation from the calculated trend line for many
choice of locus can affect the degree of strain discrimina-
comparisons may be attributable to experimental error,
tion afforded by a phylogeny.
inherent in DNA–DNA hybridization studies. The presence
The gyrB sequences used in this study to produce the of genetic elements associated with the flexible genome,
phylogeny (Fig. 1) were sufficiently distinct to discriminate especially those of large size or that occur as multiple
the majority of established species in the genus copies, including phage, transposable elements, insertion
Xanthomonas. The phylogram indicates that the genus is sequences and plasmids, could all potentially affect DNA
a monophyletic group derived from a common ancestor reassociation kinetics and hybridization values. These
that acquired the ability to exploit plant tissue as an elements are a further potential source of deviation
environmental niche. Four of the five species that branch at between DNA–DNA hybridization- and nucleotide
the base of the phylogeny, and which probably represent sequence-based measures of relatedness.
early-evolving species, are pathogens of monocot hosts, i.e.
The gyrB phylogeny clearly offers greater resolution than
X. albilineans, X. sacchari, X. hyacinthi and X. translucens,
those produced previously from the 16S rRNA gene or
and it is possible that the genus first arose as a monocot
16S–23S intergenic loci (Hauben et al., 1997; Gonçalves &
pathogen. Earlier studies using the 16S or 16S–23S loci
Rosato, 2002), and has resolved X. citri and X. fuscans,
(Hauben et al., 1997; Gonçalves & Rosato, 2002) also found
which were previously classified as pathovars of X.
X. albilineans, X. hyacinthi, X. theicola, X. translucens and
axonopodis. These two species, whose names were recently
X. sacchari to be the most ancestral species. The gyrB
validly published, were identified by using a DNA–DNA
phylogeny identifies three major inter-species groups that
hybridization technique (Schaad et al., 2005) that may be
are well supported by bootstrap analysis (groups 1–3),
more discriminating than the spectrophotometric assay
whilst a fourth comprises early-branching species of
used previously (Vauterin et al., 1995). The gyrB phylogeny
uncertain relatedness. Extensive species diversification has
supports the distinctiveness of these two species.
occurred within groups 2 and 3, which constitute the
majority of the species within the genus. A bootstrap-value The inability of the gyrB phylogeny to discriminate X.
threshold of 70 % is often used as a guideline to indicate a euvesicatoria, X. perforans and X. alfalfae was investigated
high level of confidence in phylogram branches, and the by partial sequencing of a further five genes. None of these
separation of groups 2 and 3 is well supported with a five loci was able to distinguish any of the three species
bootstrap value of 78 %. Group 3 is the largest species-level with branch lengths consistent with species-level differ-
group, and the most basal members of this clade comprised entiation, confirming their close core-genome relatedness
X. fragariae, X. pisi and the tomato pathogen X. vesicatoria. (Fig. 3). Similarity between X. euvesicatoria and X. alfalfae
Clades 3A and 3B were supported by a 100 % bootstrap has been found recently in the REP-PCR study reported by

http://ijs.sgmjournals.org 2885
N. Parkinson and others

Rademaker et al. (2005), which placed both species in the delineate a species. The phylogeny could distinguish 24 of
same group, designated 9.2. Relatively high DNA–DNA the currently accepted species and clearly there is excellent
hybridization values (58 %) between X. perforans and X. potential for gyrB sequencing as a convenient and accurate
euvesicatoria have been reported (Jones et al., 2004). DNA– means of indicating xanthomonad relatedness at the
DNA hybridization data differ from sequence analysis in species and, possibly, strain or pathovar level, and we
that the former are a measure of whole-genome DNA now use the method for routine xanthomonad identifica-
similarity, which may include elements of the flexible tion. We are currently sequencing the gyrB locus for all of
genome, e.g. phage, transposable elements and plasmids, the xanthomonad pathovar type strains to provide a
that are often variable within strains of a species. These comprehensive phylogeny of the genus Xanthomonas.
components could affect DNA reassociation kinetics and
lead to differential DNA–DNA hybridization values. A
large 190 or 155 kb plasmid has been reported in 75 % of ACKNOWLEDGEMENTS
X. euvesicatoria strains (Wichmann et al., 2005). The This work was supported by the Plant Health Division of DEFRA. The
genome-sequenced strain of X. euvesicatoria, 85-10, con- licence number for working with the quarantine bacteria is PHL
tains a large plasmid and three others, totalling 241 686 bp 251C/ 5574 (02/2007).
and representing 4.66 % of the genome, which, taking into
account that most plasmids are present in more than a
single copy, constitutes a significant proportion of the REFERENCES
genome. Whilst the plasmid contents of X. perforans and X.
Ait Tayeb, L., Ageron, E., Grimont, F. & Grimont, P. (2005). Molecular
alfalfae are unknown, it is possible that the presence of phylogeny of the genus Pseudomonas based on rpoB sequences and
plasmids in X. euvesicatoria may have contributed to application for identification of isolates. Res Microbiol 156, 763–773.
DNA–DNA hybridization values and this may explain the da Silva, A. C. R., Ferro, J. A., Reinach, F. C., Farah, C. S., Furlan, L.,
disparity between the DNA hybridization and gyrB Quaggio, R., Monteiro-Vitorello, C., Van Sluys, M. M., Almeida, N. &
sequence data. The close core-genome relatedness of these other authors (2002). Comparison of the genomes of two
three species that infect pepper, tomato and alfalfa suggests Xanthomonas pathogens with differing host specificities. Nature
that evolution of the pathogens to enable utilization of 417, 459–463.
different hosts may be proceeding relatively rapidly, before Euzéby, J. (2006). In Validation of the Publication of New Names and
substantial substitutions in the core genome have had time New Combinations Previously Effectively Published Outside the IJSB,
to accumulate. List 109. Int J Syst Evol Microbiol 56, 925–927.
Euzéby, J. (2007). In List of New Names and New Combinations
Two other species, X. cynarae and X. gardneri, could not be Previously Effectively, but not Validly, Published, List 115. Int J Syst
distinguished from each other or from the type strain of X. Evol Microbiol 57, 893–897.
hortorum by using the gyrB phylogeny. The reference strain Fegan, M. & Prior, P. (2005). How complex is the ‘‘Ralstonia
of X. hortorum used for comparison in the DNA–DNA solanacearum species complex’’? In Bacterial Wilt Disease and the
hybridization study to discriminate X. cynarae from X. Ralstonia solanacearum Complex, pp. 449–461. Edited by C. Allen,
hortorum (Trébaol et al., 2000) was the pathotype reference P. Prior & A. Hayward. Saint Paul, MN: American Phytopathological
strain of X. hortorum pv. pelargonii, rather than the species Society.
type strain, which belongs to X. hortorum pv. hederae. The Gabriel, D. W., Kingsley, M. T., Hunter, J. E. & Gottwald, T. (1989).
study by Vauterin et al. (1995) recognized heterogeneity Reinstatement of Xanthomonas citri (ex Hasse) and X. phaseoli (ex
Smith) to species and reclassification of all X. campestris pv. citri
within X. hortorum and reported DNA–DNA hybridization
strains. Int J Syst Bacteriol 39, 14–22.
values of 77 % between these two pathovars. The gyrB
Gonçalves, E. R. & Rosato, Y. B. (2002). Phylogenetic analysis of
phylogeny placed X. hortorum pv. pelargonii as a distinct
Xanthomonas species based upon 16S–23S rDNA intergenic spacer
lineage with a branch length close to that differentiating sequences. Int J Syst Evol Microbiol 52, 355–361.
other species, indicating that this pathotype was a poor
Hauben, L., Vauterin, L., Swings, J. & Moore, E. (1997). Comparison
choice to represent the species in the DNA–DNA of 16S ribosomal DNA sequences of all Xanthomonas species. Int J
hybridization studies. Consequently, X. cynarae may not Syst Bacteriol 47, 328–335.
have been differentiated adequately from X. hortorum by Jones, J. B., Lacy, G. H., Bouzar, H., Stall, R. E. & Schaad, N. W.
using DNA–DNA hybridization. The same X. hortorum pv. (2004). Reclassification of the xanthomonads associated with bacterial
pelargonii reference strain was also used to distinguish X. spot disease of tomato and pepper. Syst Appl Microbiol 27, 755–762.
gardneri as a separate species from X. hederae (Jones et al., Jukes, T. & Cantor, C. (1969). Evolution of protein molecules. In
2004). We could not find reported DNA–DNA hybridiza- Mammalian Protein Metabolism, pp. 21–132. Edited by H.N. Munro.
tion data comparing X. cynarae and X. gardneri. The New York: Academic Press.
phylogeny illustrates how the greater resolution of the gyrB Lee, B. M., Park, Y. J., Park, D. S., Kang, H. W., Kim, J. G., Song, E. S.,
phylogeny compared with DNA–DNA hybridization can Park, I. C., Yoon, U. H., Hahn, J. H. & other authors (2005). The
clarify xanthomonad relatedness. genome sequence of Xanthomonas oryzae pathovar oryzae
KACC10331, the bacterial blight pathogen of rice. Nucleic Acids Res
One issue that may need to be addressed in the future, if a 33, 577–586.
phylogenetic approach is used for classification, is to decide Louws, F. J., Fulbright, D. W., Stephens, C. T. & De Bruijn, F. J. (1994).
what level of sequence divergence should be used to Specific genomic fingerprints of phytopathogenic Xanthomonas and

2886 International Journal of Systematic and Evolutionary Microbiology 57


Xanthomonas phylogeny

Pseudomonas pathovars and strains, generated with repetitive Schaad, N. W., Postnikova, E., Lacy, G., Sechler, A., Agarkova, I.,
sequences and PCR. Appl Environ Microbiol 60, 2286–2295. Stromberg, P. E., Stromberg, V. K. & Vidaver, A. K. (2007).
Martin, B., Humbert, O., Camara, M., Guenzi, E., Walker, J., Mitchell, T., Xanthomonas alfalfae sp. nov., Xanthomonas citri sp. nov. and
Andrew, P., Prudhomme, M., Alloing, G. & other authors (1992). A Xanthomonas fuscans sp. nov. In List of New Names and New
highly conserved repeated DNA element located in the chromosome of Combinations Previously Effectively, but not Validly, Published,
Streptococcus pneumoniae. Nucleic Acids Res 20, 3479–3483. Validation List no. 115. Int J Syst Evol Microbiol 57, 893–897.
Paradis, S., Boissnot, M., Paquette, N., Bélanger, D., Martel, E., Stead, D. E. (1989). Grouping of Xanthomonas campestris pathovars
Boudreau, D., Picard, J., Ouellette, M., Roy, P. & Bergeron, M. (2005). of cereals and grasses by cellular fatty acid profiling. EPPO Bull 19,
Phylogeny of the Enterobacteriaceae based on encoding elongation 51–68.
factor Tu and F-ATPase b-subunit. Int J Syst Evol Microbiol 55, Stead, D. E. (1992). Grouping of plant-pathogenic and some other
2013–2025. Pseudomonas spp. by using cellular fatty acid profiles. Int J Syst
Poussier, S., Prior, P., Luisetti, J., Hayword, C. & Fegan, M. (2000). Bacteriol 42, 281–295.
Partial sequencing of the hrpB and endoglucanase genes confirms and Tajima, F. & Nei, M. (1984). Estimation of evolutionary distance
expands the known diversity within the Ralstonia solanacearum between nucleotide sequences. Mol Biol Evol 1, 269–285.
species complex. Syst Appl Microbiol 23, 479–486.
Thieme, F., Koebnik, R., Bekel, T., Berger, C., Boch, J., Büttner, D.,
Rademaker, J. L. W., Louws, F. J., Schultz, M. H., Rossbach, U., Caldana, C., Gaigalat, L., Goesmann, A. & other authors (2005).
Vauterin, L., Swings, J. & de Bruijn, F. J. (2005). A comprehensive Insights into genome plasticity and pathogenicity of the plant
species to strain taxonomic framework for Xanthomonas. pathogenic bacterium Xanthomonas campestris pv. vesicatoria revealed
Phytopathology 95, 1098–1111. by the complete genome sequence. J Bacteriol 187, 7254–7266.
Saitou, N. & Nei, M. (1987). The neighbour-joining method: a new Thompson, J. D., Higgins, D. G. & Gibson, T. J. (1994). CLUSTAL W:
method of constructing phylogenetic trees. Mol Biol Evol 4, 406–425. improving the sensitivity of progressive multiple sequence alignment
Sarkar, S. F. & Guttman, D. S. (2004). Evolution of the core genome through sequence weighting, position-specific gap penalties and
of Pseudomonas syringae, a highly clonal endemic plant pathogen. weight matrix choice. Nucleic Acids Res 22, 4673–4680.
Appl Environ Microbiol 70, 1999–2012. Thwaites, R., Mansfield, J., Eden-Green, S. & Seal, S. (1999). RAPD
Sarkar, S. F., Gordon, J. S., Martin, G. B. & Guttman, D. S. (2006). and rep PCR-based fingerprinting of vascular bacterial pathogens of
Comparative genomics of host-specific virulence in Pseudomonas Musa spp. Plant Pathol 48, 121–128.
syringae. Genetics 174, 1041–1056. Trébaol, G., Gardan, C., Manceau, J., Tanguy, Y., Trilly, Y. & Boury, S.
Schaad, N. W., Postnikova, E., Lacy, G. H., Sechler, A., Agarkova, I., (2000). Genomic and phenotypic characterisation of Xanthomonas
Stromberg, P. E., Stromberg, V. K. & Vidaver, A. K. (2005). cynarae; a new species causing bacterial bract spot of artichoke
Reclassification of Xanthomonas campestris pv. citri (ex Hasse 1915) (Cynara scolymus L.). Int J Syst Evol Microbiol 50, 1471–1478.
Dye 1978 forms A, B/C/D, and E as X. smithii subsp. citri (ex Hasse) Van de Peer, Y. & De Wachter, R. (1993). TREECON: a software package
sp. nov. nom. rev. comb. nov., X. fuscans subsp. aurantifolii (ex for the construction and drawing of RT evolutionary trees. Comput
Gabriel 1989) sp. nov. nom. rev. comb. nov., and X. alfalfae subsp. Appl Biosci 9, 177–182.
citrumelo (ex Riker and Jones) Gabriel et al., 1989 sp. nov. nom. rev.
comb. nov.; X. campestris pv. malvacearum (ex Smith 1901) Dye 1978 Van den Mooter, M. & Swings, J. (1990). Numerical analysis of 295
as X. smithii subsp. smithii nov. comb. nov. nom. nov.; X. campestris phenotypic features of 266 Xanthomonas strains and related strains
pv. alfalfae (ex Riker and Jones, 1935) Dye 1978 as X. alfalfae subsp. and an improved taxonomy of the genus. Int J Syst Bacteriol 40,
alfalfae (ex Riker et al., 1935) sp. nov. nom. rev.; and ‘‘var. fuscans’’ of 348–369.
X. campestris pv. phaseoli (ex Smith, 1987) Dye 1978 as X. fuscans Vauterin, L., Hoste, B., Kersters, K. & Swings, J. (1995).
subsp. fuscans sp. nov. Syst Appl Microbiol 28, 494–518. Reclassification of Xanthomonas. Int J Syst Bacteriol 45, 472–489.
Schaad, N. W., Postnikova, E., Lacy, G., Sechler, A., Agarkova, I., Wichmann, G., Ritchie, D., Kousik, C. & Bergelson, J. (2005).
Stromberg, P. E., Stromberg, V. K. & Vidaver, A. K. (2006). (Erratum) Reduced genetic variation occurs among genes of the highly clonal
Emended classification of xanthomonad pathogens on citrus. Syst plant pathogen Xanthomonas axonopodis pv. vesicatoria, including the
Appl Microbiol 29, 690–695. effector gene avrBs2. Appl Environ Microbiol 71, 2418–2432.

http://ijs.sgmjournals.org 2887

You might also like