ArticlePDF Available

Comparative ICE genomics: Insights into the evolution of the SXT/R391 family of ICEs

PLOS
PLOS Genetics
Authors:

Abstract and Figures

Integrating and conjugative elements (ICEs) are one of the three principal types of self-transmissible mobile genetic elements in bacteria. ICEs, like plasmids, transfer via conjugation; but unlike plasmids and similar to many phages, these elements integrate into and replicate along with the host chromosome. Members of the SXT/R391 family of ICEs have been isolated from several species of gram-negative bacteria, including Vibrio cholerae, the cause of cholera, where they have been important vectors for disseminating genes conferring resistance to antibiotics. Here we developed a plasmid-based system to capture and isolate SXT/R391 ICEs for sequencing. Comparative analyses of the genomes of 13 SXT/R391 ICEs derived from diverse hosts and locations revealed that they contain 52 perfectly syntenic and nearly identical core genes that serve as a scaffold capable of mobilizing an array of variable DNA. Furthermore, selection pressure to maintain ICE mobility appears to have restricted insertions of variable DNA into intergenic sites that do not interrupt core functions. The variable genes confer diverse element-specific phenotypes, such as resistance to antibiotics. Functional analysis of a set of deletion mutants revealed that less than half of the conserved core genes are required for ICE mobility; the functions of most of the dispensable core genes are unknown. Several lines of evidence suggest that there has been extensive recombination between SXT/R391 ICEs, resulting in re-assortment of their respective variable gene content. Furthermore, our analyses suggest that there may be a network of phylogenetic relationships among sequences found in all types of mobile genetic elements.
The boundaries of the 5 hotspots. The boundaries between conserved and hotspot variable regions are shown. Black typeface indicates conserved sequence, while color indicates variable sequence. Numbers in parentheses indicate the number of intervening nucleotides. The thin dotted lines indicate continuations of variable DNA. Bold letters indicate a non-conserved base. (A) Hotspot 1, which is present between traJ and traL. Line 1: SXT, ICEVchInd4, ICEPalBan1; Line 2: R391, ICEPdaSpa1, ICEVchBan5, ICEVchInd5, ICEPmiUSA, ICESpuPO1, ICEVflInd, ICEVchMoz10, ICEVchBan9; Line 3: ICEVchMex1. (B) Hotspot 2, which is present between traA and s054. Line 1: SXT, ICEVchInd4, ICEPmiUSA, ICEVflInd, ICEVchInd5, ICEPalBan1, ICEVchBan5; Line 2: ICEPdaSpa1, ICEVchMoz10, ICEVchBan9; Line 3: R391; Line 4: ICEVchMex1; Line 5: ICESpuPO1. (C) Hotspot 3, which is present between s073 and traF. Line 1: SXT, ICEVchInd4; Line 2: ICEVchMex1, ICEVflInd; Line 3: ICEVchMoz10, ICEVchBan9, ICEVchInd5, ICEPalBan1, ICEVchBan5. Line 4: R391; Line 5: ICEPmiUSA; Line 6: ICESpuPO1; Line 7: ICEPdaSpa1. (D) Hotspot 4, which is present between traN and s063. Line 1: SXT, ICEVchInd4. Line 2: ICEVchInd5, ICEVchBan5; Line 3: ICESpuPO1, ICEPmiUSA; Line 4: R391, ICEVchMoz10, ICEVchBan9, ICEVflInd. Line 5: ICEPdaSpa1; Line 6: ICEPalBan1; Line 7: ICEVchMex1. (E) Hotspot 5, which is present between s026 and traI. Line 1: SXT, ICEVchInd4, ICEPdaSpa1, R391, ICEVchMoz10, ICEVchBan9; Line 2: ICEPmiUSA; Line 3: ICESpuPO1, ICEPalBan1, ICEVflInd1; Line 4: ICEVchInd5, ICEVchBan5; Line 5: ICEVchMex1. doi:10.1371/journal.pgen.1000786.g003
… 
Comparison of the SXT/R391 core genome with the genome of pIP1202 and defining the minimal functional SXT/R391 gene set. (A) Alignment of the conserved core genes of SXT/R391 ICEs with the genome of the IncA/C conjugative plasmid pIP1202 from Yersinia pestis. The top line shows the same core ICE genes shown in Figure 2A. ORFs are color coded as follows: DNA processing, yellow; mating pair formation, orange; DNA recombination and repair, green; integration/excision, red; replication, purple; regulation, gray; entry exclusion, blue; homologous genes of unknown function, black; genes without corresponding counterparts in ICEs and pIP1202, white. Numbers shown in the middle represent % identity between the orthologous proteins encoded by SXT and pIP1202 [GenBank:NC_009141]. The positions of the hotspots in SXT/R391 ICEs are marked by downward pointing arrowheads. For pIP1202, the size of the sequences (which include IncA/C backbone DNA as well as variable DNA) found at these locations as well as resistance markers are indicated by upward pointing arrowheads. aphA, aadA and strAB confer resistance to aminoglycosides. sul1 and sul2 confer resistance to sulfonamides. cat, bla SHV-1 , tetAR, qacED1 and merRTPCADE confer resistance to chloramphenicol, b-lactams, tetracyclines, quaternary ammonium compounds and mercury ions, respectively. Detailed descriptions of the conserved backbone of the IncA/C conjugative plasmids have been published elsewhere [48,50]. Regions that were deleted from SXT to investigate the function of genes of unknown function (see panel B) are indicated with straight lines. Dotted lines indicate that the deletion included DNA in the adjacent hotspots. (B) Influence of deletion of genes of unknown function on the frequency of SXT transfer. The mean values and standard deviations from three independent experiments are shown. * indicates that the frequency of transfer was below the detection level (,10 28 ). Deletion mutants SXTDa, SXTDk and SXTDl, transferred at frequencies that were not significantly different from that of wild-type SXT (data not shown). (C) Proposed minimal set of genes necessary for a functional SXT/R391 ICE. int, integration/excision module; mob, DNA processing module; mpf, mating pair formation modules; reg, regulation module. doi:10.1371/journal.pgen.1000786.g004
… 
Content may be subject to copyright.
Comparative ICE Genomics: Insights into the Evolution of
the SXT/R391 Family of ICEs
Rachel A. F. Wozniak
1,2,3
, Derrick E. Fouts
3
, Matteo Spagnoletti
4
, Mauro M. Colombo
4
, Daniela
Ceccarelli
5
, Genevie
`ve Garriss
5
, Christine De
´ry
5
, Vincent Burrus
5
*, Matthew K. Waldor
1,2,6
*
1Channing Laboratory, Brigham and Women’s Hospital, Harvard Medical School, Boston, Massachusetts, United States of America, 2Department of Genetics, Tufts
Medical School, Boston, Massachusetts, United States of America, 3J. Craig Venter Institute, Rockville, Maryland, United States of America, 4Dipartimento di Biologia
Cellulare e dello Sviluppo, Universita
´di Roma La Sapienza, Rome, Italy, 5Centre d’E
´tude et de Valorisation de la Diversite
´Microbienne, De
´partement de Biologie,
Universite
´de Sherbrooke, Sherbrooke, Que
´bec, Canada, 6Howard Hughes Medical Institute, Chevy Chase, Maryland, United States of America
Abstract
Integrating and conjugative elements (ICEs) are one of the three principal types of self-transmissible mobile genetic
elements in bacteria. ICEs, like plasmids, transfer via conjugation; but unlike plasmids and similar to many phages, these
elements integrate into and replicate along with the host chromosome. Members of the SXT/R391 family of ICEs have been
isolated from several species of gram-negative bacteria, including Vibrio cholerae, the cause of cholera, where they have
been important vectors for disseminating genes conferring resistance to antibiotics. Here we developed a plasmid-based
system to capture and isolate SXT/R391 ICEs for sequencing. Comparative analyses of the genomes of 13 SXT/R391 ICEs
derived from diverse hosts and locations revealed that they contain 52 perfectly syntenic and nearly identical core genes
that serve as a scaffold capable of mobilizing an array of variable DNA. Furthermore, selection pressure to maintain ICE
mobility appears to have restricted insertions of variable DNA into intergenic sites that do not interrupt core functions. The
variable genes confer diverse element-specific phenotypes, such as resistance to antibiotics. Functional analysis of a set of
deletion mutants revealed that less than half of the conserved core genes are required for ICE mobility; the functions of
most of the dispensable core genes are unknown. Several lines of evidence suggest that there has been extensive
recombination between SXT/R391 ICEs, resulting in re-assortment of their respective variable gene content. Furthermore,
our analyses suggest that there may be a network of phylogenetic relationships among sequences found in all types of
mobile genetic elements.
Citation: Wozniak RAF, Fouts DE, Spagnoletti M, Colombo MM, Ceccarelli D, et al. (2009) Comparative ICE Genomics: Insights into the Evolution of the SXT/R391
Family of ICEs. PLoS Genet 5(12): e1000786. doi:10.1371/journal.pgen.1000786
Editor: Ivan Matic, Universite
´Paris Descartes, INSERM U571, France
Received October 7, 2009; Accepted November 24, 2009; Published December 24, 2009
Copyright: ß2009 Wozniak et al. This is an open-access article distributed under the terms of the Creative Commons Attribution License, which permits
unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited.
Funding: MKW acknowledges support from NIH (AI R37-42347) and HHMI. GG is the recipient of a PhD research scholarship from FQRNT. VB holds a Canada
Research Chair in molecular biology, impact and evolution of bacterial mobile elements and is grateful for support from the Natural Sciences and Engineering
Research Council of Canada (Discovery Grant Program). MS was supported by a grant from PRIN 2007 - Italy, and DC was supported by a fellowship from Cenci
Bolognetti-Institut Pasteur Foundation, Italy. Sequencing at the JCVI was funded by the National Institute of Allergy and Infectious Diseases, National Institutes of
Health, Department of Health and Human Services under contract number N01-AI-30071. The funders had no role in study design, data collection and analysis,
decision to publish, or preparation of the manuscript.
Competing Interests: The authors have declared that no competing interests exist.
* E-mail: Vincent.Burrus@USherbrooke.ca (VB); mwaldor@rics.bwh.harvard.edu (MKW)
Introduction
There are three types of self-transmissible mobile genetic
elements: plasmids, bacteriophages and integrative conjugative
elements (ICEs). All three classes of elements enable horizontal
transmission of genetic information and all have had major
impacts on bacterial evolution [1–4]. ICEs, (aka conjugation
transposons), like plasmids, are transmitted via conjugation;
however, unlike plasmids, ICEs integrate into and replicate along
with the chromosome. Following integration, ICEs can excise
from the chromosome and form circular molecules that are
intermediates in ICE transfer. Plasmids and phages have been the
subject of more extensive study than ICEs and while there is
growing understanding of the molecular aspects of several ICEs
[5–10], to date there have been few reports of comparative ICE
genomics [11,12] and consequently understanding of ICE
evolution is only beginning to be unraveled.
Diverse ICEs have been identified in a variety of gram-positive
and gram–negative organisms [13]. These elements utilize a
variety of genes to mediate the core ICE functions of chromosome
integration, excision and conjugation. In addition to a core gene
set, ICEs routinely contain genes that confer specific phenotypes
upon their hosts, such as resistance to antibiotics and heavy metals
[14–18], aromatic compound degradation [19] or nitrogen
fixation [20].
SXT is an ,100 Kb ICE that was originally discovered in Vibrio
cholerae O139 [16], the first non-O1 serogroup to cause epidemic
cholera [21]. SXT encodes resistances to several antibiotics,
including sulfamethoxazole and trimethoprim (which together are
often abbreviated as SXT) that had previously been useful in the
treatment of cholera. Since the emergence of V. cholerae O139 on
the Indian subcontinent in 1992, SXT or a similar ICE has been
found in most clinical isolates of V. cholerae, including V. cholerae
serogroup O1, from both Asia and Africa. Other vibrio species
besides V. cholerae have also been found to harbor SXT-related
ICEs [22]. Furthermore, SXT-like ICEs are not restricted to
vibrio species, as such ICEs have been detected in Photobacterium
damselae,Shewanella putrefaciens and Providencia alcalifaciens [23–25].
PLoS Genetics | www.plosgenetics.org 1 December 2009 | Volume 5 | Issue 12 | e1000786
Moreover, Hochhut et al [26] found that SXT is genetically and
functionally related to the so-called ‘Inc J’ element R391, which
was derived from a South African Providencia rettgeri strain isolated
in 1967 [27]. It is now clear that Inc J elements are SXT-related
ICEs that were originally misclassified as plasmids. In the
laboratory, SXT has a fairly broad host range and can be
transmitted between a variety of gram-negative organisms [16].
The SXT/R391 family of ICEs is now known to include more
than 30 elements that have been detected in clinical and
environmental isolates of several species of c- proteobacteria from
disparate locations around the globe [28]. SXT/R391 ICEs are
grouped together as an ICE family because they all encode a
nearly identical integrase, Int. Int, a tyrosine recombinase, is
considered a defining feature of these elements because it enables
their site-specific integration into the 59end of prfC, a conserved
chromosomal gene that encodes peptide chain release factor 3
[29]. Int mediates recombination between nearly identical element
and chromosome sequences, attP and attB respectively [29]. When
an SXT/R391 ICE excises from the chromosome, Int, aided by
Xis, a recombination directionality factor, mediates the reverse
reaction - recombination between the extreme right and left ends
(attR and attL) of the integrated element - thereby reconstituting
attP and attB [6,29]. The excised circular SXT form is thought to
be the principal substrate for its conjugative transfer. The genes
that encode activities required for SXT transfer (tra genes) were
originally found to be distantly related to certain plasmid tra genes
[30–32]. The tra genes encode proteins important for processing
DNA for transfer, mating pair formation and generating the
conjugation machinery. Regulation of SXT excision and transfer
is at least in part governed by a pathway that resembles the
pathway governing the lytic development of the phage lambda.
Agents that damage DNA and induce the bacterial SOS response
are thought to stimulate the cleavage and inactivation of SetR, an
SXT encoded lcI-related repressor, which represses expression of
setD and setC, transcription activators that promote expression of
int and tra genes [5].
The complete nucleotide sequences of SXT (99.5kb) and R391
(89kb) were the first SXT/R391 ICE family genomes to be
reported [14,32]. Comparative [33] and functional genomic
analyses [5,32] revealed that these 2 ICEs share a set of conserved
core genes that mediate their integration/excision (int and xis),
conjugative transfer (various tra genes), and regulation (setR,setCD).
In addition to the conserved genes, these 2 ICEs contain element
specific genes that confer element specific properties such as
resistance to antibiotics or heavy metals. Interestingly, many of
these genes were found in identical locations in SXT and R391,
leading Beaber et al [33] to propose that there are ‘hotspots’ where
SXT/R391 ICEs can acquire new DNA. The genomes of two
additional SXT/R391 ICEs, ICEPdaSpa1, isolated from Photo-
bacterium damselae [23], and ICESpuPO1, derived from an
environmental isolate of Shewanella putrefaciens [24] are now also
known. These two genomes also share most of the conserved set of
core genes present in SXT and R391 and contain element specific
DNA.
Determination of the sequences of SXT/R391 family ICE
genomes was a fairly arduous task due to their size and
predominantly chromosomal localization. Here, we developed a
method to capture and then sequence complete SXT/R391 ICE
genomes. In addition, we identified 3 as yet unannotated SXT/
R391 ICE genomes in the database of completed bacterial
genomes. Comparative analyses of the 13 SXT/R391 genomes
now available allowed us to greatly refine our understanding of the
organization and conservation of the core genes that are present in
all members of this ICE family. Comparative and functional
analyses also facilitated our proposal of the minimal functional
SXT/R391 ICE genome. Furthermore, this work provides new
knowledge of the considerable diversity of genes and potential
accessory functions encoded by the variable DNA found in these
mobile elements. Finally, this comparative genomics approach has
allowed us to garner clues regarding the evolution of this class of
mobile elements.
Results/Discussion
An ICE capture system
To date, ICE sequencing has been cumbersome because it has
typically required construction of chromosome-derived cosmid
libraries and screening for sequences that hybridize to ICE probes
[23,32]. We constructed a vector (pIceCap) that enables capture of
complete SXT/R391 ICE genomes on a low-copy plasmid to
simplify the protocol for ICE sequencing. This plasmid is a
derivative of the single-copy modified F plasmid pXX704 [34,35],
which contains a minimal set of genes for F replication and
segregation but lacks genes enabling conjugation. We modified
pXX704 to include an ,400bp fragment that encompasses the
SXT/R391 attachment site (attB) and thereby enabled Int-catalyzed
site-specific recombination between attB on pIceCap and attP on an
excised and transferred ICE to drive ICE capture (Figure 1).
Conjugations between an SXT/R391 ICE-bearing donor strain
and an E. coli recipient deleted for prfC (and thus chromosomal attB)
and harboring pIceCap yielded exconjugants containing the
transferred ICE integrated into pIceCap (Figure 1). We used the
DprfC recipient to bias integration of the transferred ICE into
pIceCap rather than the chromosome. In these experiments, we
selected for exconjugants containing the transferred ICE integrated
into pIceCap, using an antibiotic marker present on the ICE as well
as a marker present in pIceCap. The low copy IceCap::ICE plasmid
was then isolated and used as a substrate for shotgun sequencing.
We also found that the IceCap::ICE plasmids were transmissible.
Thus, in principle this technique should facilitate capture of ICEs
that do not harbor genes conferring resistance to antibiotics, by
mating out the IceCap::ICE plasmid into a new recipient and
selecting for the marker on pIceCap.
SXT/R391 ICEs included in this analysis
A list of the 13 SXT/R391 ICEs whose genomes were analyzed
and compared in this study is shown in Table 1. All of the ICEs
Author Summary
Integrative and conjugative elements (ICEs) are a class of
mobile genetic elements that are key mediators of
horizontal gene flow in bacteria. These elements integrate
into the host chromosome, yet are able to excise and
transfer via conjugation. Our understanding of ICE
evolution is rudimentary. Here, we developed a method
to capture ICEs on plasmids, thus facilitating their
sequencing. Comparative analyses of the DNA sequences
of ICEs from the same family revealed that they have an
identical genetic structure consisting of syntenous, highly
conserved core genes that are interrupted by clusters of
diverse variable genes. Unexpectedly, many genes in the
core backbone proved non-essential for ICE transfer.
Comparisons of the variable gene content in the ICEs
analyzed revealed that these elements are mosaics whose
genomes have been shaped by inter–ICE recombination.
Finally, our work suggests that ICEs contribute to a larger
gene pool that connects all types of mobile elements.
Comparative ICE Genomics
PLoS Genetics | www.plosgenetics.org 2 December 2009 | Volume 5 | Issue 12 | e1000786
included in our analyses contain an int gene that was amplifiable
using PCR primers for int
sxt
[29]. They were isolated on 4
continents and from the Pacific Ocean during a span of more
than 4 decades. They are derived from 7 different genera of
c-proteobacteria and the ICEs derived from V. cholerae strains are
from both clinical and environmental isolates of 3 different
V. cholerae serogroups.
Five of these ICE genome sequences were determined at the J.
Craig Venter Institute (JCVI) using the ICE capture system
described above (Table 1, rows 1–5). In addition, we sequenced
ICEVflInd1, also at the JCVI, by isolating cosmids that
encompassed this V. fluvialis derived ICE prior to developing the
ICE capture technique (Table 1, row 6). Table 1 (rows 7–10) also
includes 4 previously unannotated ICE genomes that we found in
BLAST searches of the NCBI database of completed but as yet
unannotated genomes; 3 of these ICEs are clearly members of
SXT/R391 ICE family since they are integrated into their
respective host’s prfC locus and contain int genes that are predicted
to encode Int proteins that are 99% identical to Int
sxt
. The fourth
element, ICEVchBan8 does not encode an Int
sxt
orthologue;
however, this element contains nearly identical homologues of
most of the known conserved core SXT/R391 ICE family genes.
ICEVchBan8 will be discussed in more detail below but since it
does not contain an Int
sxt
orthologue it is not considered a member
of the SXT/R391 family of ICEs and thus not included in our
comparative study. Finally, Table 1 also includes the 4 SXT/R391
ICEs that were previously sequenced (Table 1, rows 11–14).
Despite the diversity of our sources for SXT/R391 ICEs, the
genomes of two pairs of ICEs that we analyzed proved to be very
similar. SXT
MO10
and ICEVchInd4 only differed by 13 SNPs in 7
genes and by the absence from ICEVchInd4 of dfr18, a gene
conferring trimethoprim resistance. These ICEs were derived from
V. cholerae O139 strains isolated in India from different cities at
different times: SXT
MO10
from Chennai in 1992 and ICEVchInd4
from Kolkata in 1997. The high degree of similarity of these two
ICE genomes suggests that ICEs can be fairly stable over time.
ICEVchBan9 and ICEVchMoz10 were also extremely similar
although ICEVchMoz10 lacks dfrA1, another allele for trimethoprim
resistance. These two ICEs were derived from V. cholerae O1 strains
from Bangladesh (1994) and Mozambique (2004) respectively. The
great similarity of these ICEs suggests that there has been spread of
SXT-related ICEs between Asia and Africa in recent times. Studies
of CTX prophage genomes have also suggested the spread of V.
cholerae strains between these continents [36].
General structure and sizes of SXT/R391 genomes
The ICEs listed in Table 1 were initially compared using
MAUVE [37] and LAGAN [38], programs that enable visuali-
zation of conserved and variable regions on a global scale. All of
the SXT/R391 ICEs we analyzed share a common structure and
have sizes ranging from 79,733 bp to 108,623 bp (Table 1 and
Figure 2). They contain syntenous sets of 52 conserved core genes
(Figure 2A) that total approximately 47kb and encode proteins
with an average of 97% identity to those encoded by SXT. All of
the individual ICEs also contain DNA that is relatively specific for
individual elements (Figure 2B); the differences in the sizes of the
variable regions accounts for the range in ICE sizes.
Five sites within the conserved SXT/R391 ICE structure have
variable DNA present in all of the ICEs in Figure 2. Four of these
sites were previously termed ‘hotspots’ for ICE acquisition of new
DNA [33]. Due to similarities between SXT and R391, the fifth
hotspot only became apparent through our comparison of the 13
ICEs examined here. Each of these hotspots (HS1 to HS5 in
Figure 2B) is found in an intergenic region (see below), suggesting
that the acquisition of these variable DNA regions has not
interrupted core ICE gene functions. In addition, some of the
ICEs have variable DNA inserted in additional intergenic
locations or in rumB (labeled I–IV in Figure 2B). Previous analyses
[32] indicated that the insertion in rumB, did not impair SXT
transmissibility. Overall, comparison of these 13 SXT/R391 ICE
genomes suggests that: 1) these elements consist of the same
perfectly syntenous and nearly identical 52 core genes that serve as
a scaffold (see below) capable of mobilizing a large range of
variable DNA; and 2) selection pressure to maintain ICE mobility
has restricted insertions of variable DNA into sites that do not
interrupt core functions.
Figure 1. Schematic of the ICE capture system. Conjugation between a donor strain bearing a chromosomal ICE and a DprfC recipient strain
harboring pIceCap, which contains attB, yields exconjugants that contain the transferred ICE integrated into pIceCap. Exconjugants were selected for
using a marker on pIceCap and on the ICE. attR and attL represent the right and left ICE-chromosome junctions.
doi:10.1371/journal.pgen.1000786.g001
Comparative ICE Genomics
PLoS Genetics | www.plosgenetics.org 3 December 2009 | Volume 5 | Issue 12 | e1000786
Table 1. SXT/R391 ICE family members analyzed in this study.
ICE Host strain Site and year of isolation
Size
(bp)
% Identity to
Int
SXT
Resistance profile Notable Variable Genes
Genbank Accession
Number
Strain or ICE
References
ICEVchMex1 Vibrio cholerae non
O1-O139
San Luis Potosi, Mexico 2001 82839 99% (410/413) -Fic family protein, diguanylate cyclase,
restriction modification system
GQ463143 [66]
ICEVchInd4 Vibrio cholerae O139 Kolkata, India 1997 95491 100% (413/413) floR, strBA, sul2 Toxin-antitoxin system GQ463141 [54]
ICEVchInd5 Vibrio cholerae O1 Sevagram, India 1994 97847 99% (409/413) floR, strBA, sul2, dfrA1 AraC family transcription regulator,
glyoxoylase abx resistance
GQ463142 This study
ICEVchBan5 Vibrio cholerae O1 Bangladesh, 1998 102131 99% (409/413) floR, strBA, sul2, dfrA1 AraC family transcription regulator,
glyoxoylase abx resistance
GQ463140 [29]
ICEPalBan1 Providencia alcalifaciens Bangladesh, 1999 96586 99% (409/413) floR, strBA, sul2, dfrA1 Toxin-antitoxin system, phenazine
biosynthesis protein, lysine exporter
GQ463139 [25]
ICEVflInd1 Vibrio fluvialis Kolkata, India 2002 91369
(a)
99% (409/413) dfr18, floR, strBA, sul2 Toxin-antitoxin system GQ463144 [22]
ICEVchMoz10 Vibrio cholerae O1 Beira, Mozambique 2004 104495 99% (409/413) floR, strBA, sul2, tetA’ AraC family transcription regulator,
glyoxoylase abx resistance,
ATP-dependent Lon protease
ACHZ00000000 [67]
ICEPmiUsa1 Proteus mirabilis Maryland, United States 1986 79733 99% (409/413) -ATP-dependent helicase AM942759 [68]
ICEVchBan9 Vibrio cholerae O1 Matlab, Bangladesh 1994 106124 99% (409/413) floR, strBA, sul2, dfrA1,
tetA’
AraC family transcription regulator,
glyoxoylase abx resistance,
ATP-dependent Lon protease
CP001485 [69]
ICEVchBan8 Vibrio cholerae non O1-O139 Bangladesh, 2001 105790
(a)
25% (76/301) -Toxin-antitoxin system NZ_AAUU00000000 This study
SXT
MO10
Vibrio cholerae O139 Chennai, India 2002 99452 100% dfr18, floR, strBA, sul2 Toxin-antitoxin system AY055428 [16]
R391 Providencia rettgeri Pretoria, South Africa 1967 88532 99% (410/413) kanR, merRTPCA Sulfate transporter, universal
stress protein
AY090559 [27]
ICEPdaSpa1 Photobacterium damselae Galicia, Spain 2003 102985 99% (412/413) tetAR ATP-dependent Lon protease,
heat-shock protein
AJ870986 [23]
ICESpuPO1 Shewanella putrefaciens 630m, Pacific Ocean 2000 108623 99% (409/413) - Zn/Co/Cd efflux system, restriction
modification system
CP000503 [24]
(a)
The sequence is not complete and therefore the true size is not known.
doi:10.1371/journal.pgen.1000786.t001
Comparative ICE Genomics
PLoS Genetics | www.plosgenetics.org 4 December 2009 | Volume 5 | Issue 12 | e1000786
The SXT/R391 ICE core genes
The 52 core genes present in all the SXT/R391 ICEs analyzed
include sets of genes that are known to be required for the key ICE
functions of integration/excision, conjugative transfer and regu-
lation [32] as well as many genes of unknown function. Most genes
of known or putative (based on homology) function (coded by gray
shading or hatch marks in Figure 2A) are clustered with genes that
have related functions. For example, int and xis, genes required for
integration and excision, are adjacent and setR, and setC/D, the key
SXT regulators are near each other at the extreme 39end of the
elements, although separated by 4 conserved genes of unknown
function. Each ICE also has four gene clusters implicated in
conjugative DNA processing and transfer (shown in light gray in
Figure 2A). Finally, each of the ICEs has a nearly identical origin
of transfer (oriT), a cis-acting DNA site that is thought to be nicked
to initiate DNA processing events during conjugative transfer [39],
in the same relative location.
The conserved core genes include approximately as many genes
of unknown function as genes of known function. Some of the
genes of unknown function are found either interspersed amongst
gene clusters that likely comprise functional modules (e.g s091
between traD and s043) while others are grouped together (e.g.
most genes between traN and traF). In several cases, the
interspersed genes appear to be part of operons with genes of
known function (e.g. s086-s082 maybe in an operon with setDC).
Variable ICE DNA
In addition to sharing 52 core genes, all of the ICE genomes
analyzed contain variable DNA regions, ranging in size from 676
to 29,210 bp. Most of the variable DNA sequences are found in 5
intergenic hotspots (Figure 2B). However, some ICEs contain
additional variable DNA inserts outside the 5 hotspots. For
example, SXT and five other ICEs in Figure 2 have variable DNA
segments, corresponding to related ISCR2 elements, disrupting
rumB (Figure 2B, site III). ISCR2 elements are IS91-like
transposable elements that tend to accumulate antibiotic resistance
genes [40]. Interestingly, it is unusual for the contents of the
hotspots and other variable regions to be found in only one ICE.
Figure 2. Structure of the genomes of 13 SXT/R391 ICEs. (A) The upper line represents the set of core genes (thick arrows) and sequences
common to all 13 SXT/R391 genomes analyzed. Hatched ORFs indicate genes involved in site-specific excision and integration (xis and int), error-
prone DNA repair (rumAB), DNA recombination (bet and exo) or entry exclusion (eex). Dark gray ORFs correspond to genes involved in regulation
(setCDR). Light gray ORFs represent genes encoding the conjugative transfer machinery, and white ORFs represent genes of unknown function. (B)
Variable ICE regions are shown with colors according to the elements in which they were originally described SXT (blue), R391 (red), ICEPdaSpa1
(green), ICESpuPO1 (purple), ICEVchMex1 (yellow), ICEPalBan1 (orange), ICEVchInd5 (turquoise), ICEPmiUSA1 (olive), ICEVchBan9 (pink), ICEVflInd1
(light green). Thin arrows indicate the sites of insertion for each variable region and HS1–HS5 represent hotspots 1–5. Roman numerals indicate
variable regions not considered true hotspots. Cm, chloramphenicol; Hg, mercury; Kn, kanamycin; Sm, streptomycin; Su, sulfamethoxazole; Tc,
tetracycline; Tm, trimethoprim. * indicates that s073 is absent from ICEPdaSpa1.
a
ICEVchMoz10, which lacks dfrA1 in the integron structure, does not
confer resistance to Tm.
b
The purple gene content of ICEVflInd1 was deduced from partial sequencing, PCR analysis and comparison with ICESpuPO1.
doi:10.1371/journal.pgen.1000786.g002
Comparative ICE Genomics
PLoS Genetics | www.plosgenetics.org 5 December 2009 | Volume 5 | Issue 12 | e1000786
Instead, the variable gene content of most of the ICEs shown in
Figure 2B is found in more than one ICE. For example,
ICESpuPO1, ICEPalBan1, and ICEVflInd1, all have identical
contents in hotspot 5 (lavender genes in hotspot 5 in Figure 2B);
however, the contents of the other hotspots in these 3 elements are
almost entirely different. Thus, the variable gene content of the
SXT/R391 ICEs reveals that these elements are mosaics. The
overlapping distribution of variable DNA segments seen in the
ICEs in Figure 2B suggests that recombination among this family
of mobile elements may be extensive. In addition, in some
instances, the variable regions appear subject to additional genetic
modifications. For example, ICEPdaSpa1 and ICEVchBan9
contain ICE-specific DNA nested within the shared sequences
inserted at hotspot 5 DNA (the green and pink genes in hotspot 5
in these elements, Figure 2B).
The variable genes encode a large array of functions and only a
few will be discussed here. A complete list of the diverse genes
found in the hotspots is found in Table S1. Although we cannot
predict functions for many genes found in the hotspots, since they
lack homology to genes of known function, at least a subset of the
known genes seem likely to confer an adaptive advantage upon
their hosts. Most of the ICE antibiotic resistance genes are found
within transposon-like structures (e.g., the ISCR2 elements noted
above) but four ICEs contain a dfrA1 cassette, which confers
resistance to trimethoprim [25], in a class IV integron located in
hotspot 3. A disproportionate number of variable genes are likely
involved in DNA modification, recombination or repair, as they
are predicted to encode diverse putative restriction-modification
systems, helicases and endonucleases. Such genes may provide the
host with barriers to invasion by foreign DNA including phage
infection and/or promote the integrity of the ICE genome during
its transfer between hosts. Three ICEs contain genes that encode
diguanylate cyclases [41] in hotspot 3. These enzymes catalyze the
formation of cyclic-diguanosine monophosphate (c-di-GMP), a
second messenger molecule that regulates biofilm formation,
motility and virulence in several organisms including V. cholerae
[42,43]. Most SXT/R391 ICEs contain mosA and mosT in hotspot
2. These two genes encode a novel toxin-antitoxin pair that
promotes SXT maintenance by killing or severely inhibiting the
growth of cells that have lost this element [44]. Not all ICEs in the
SXT/R391 family contain mosAT; however, those lacking these
genes may encode similar systems to prevent ICE loss. For
instance, R391 and ICEVchMex1 contain two genes (orf2 and orf3)
encoding a predicted HipA-like toxin and a predicted transcrip-
tional repressor distantly related to the antitoxin HipB.
Locations of the ICE variable genes
The variable regions found in the 5 hotspots are found
exclusively in intergenic regions, punctuating the conserved ICE
backbone (Figure 2). The boundaries between the conserved and
variable sequences were mapped on the nucleotide level and
compared (Figure 3A–3E). Each hotspot had a distinct boundary.
Remarkably, even though the contents of the variable regions
markedly differ, with few exceptions the left and right boundaries
between conserved and variable DNA for each hotspot was
identical among all the ICEs (Figure 3). For example, the left
junctions of the inserts in hotspot 2 immediately follow the stop
codon of traA and the right junctions are exactly 79 bp upstream of
the start of s054 (Figure 3B), despite the fact that the DNA
contents within these borders greatly differ. In hotspot 2, the right
junction appears to begin with a 15 bp sequence that has two
variants (Figure 3B, brown & light brown sequence). These
sequences may reflect the presence of earlier insertions that have
since been partially replaced. A similar pattern was found adjacent
to the left boundary of hotspot 4 in several ICEs (Figure 3D, lines
3–6). Once an insertion is acquired, the number of permissive sites
for the addition of new variable DNA likely increases.
There are two exceptions to the precise boundaries between
variable and conserved DNA. Hotspot 1 and hotspot 3 in
ICEVchMex1 and ICEPdaSpa1, respectively, contain variable
DNA that extends beyond the boundary exhibited by all the other
ICEs in these locations (Figure 3A, line 3, and Figure 3C, line 7).
The only boundary that could not be identified was the left border
of hotspot 5, the region containing genes between s026 and traI.As
discussed below, s026 is the least conserved core gene and its
variability obscured any consensus sequence abutting the variable
DNA. Perhaps this border has eroded because s026 is not required
for ICE mobility [32].
The relative precision of most boundaries between conserved
and variable DNA sequences in all the ICEs analyzed suggests that
a particular recombination mechanism, such as bet/exo-mediated
recombination, may explain the acquisition of the variable regions.
However, at this point, we cannot exclude the possibility that the
precise locations for variable DNA insertions simply reflects
selection for optimal ICE fitness; i.e., ICEs can optimally
accommodate variable DNA in these locations while preserving
their essential functions.
Similarity of SXT/R391 ICE and IncA/C plasmid core genes
Unexpectedly, BLAST analyses revealed that most of the
conserved core SXT/R391 genes are also present in IncA/C
conjugative plasmids. These multidrug resistance plasmids are
widely distributed among Salmonella and other enterobacterial
isolates from agricultural sources [45,46]. Recently, members of
this family of plasmids have also been identified in Yersinia pestis,
including from a patient with bubonic plague [47], and in aquatic
c-proteobacteria [48], including Vibrio cholerae [49,50]. To date, the
closest known relatives of the SXT/R391 transfer proteins are
found in the IncA/C plasmids. Every predicted SXT transfer
protein is encoded by the IncA/C plasmid pIP1202 isolated from
Y. pestis [50] and the identities of these predicted protein sequences
vary from 34 to 78% (Figure 4A). Furthermore, there is perfect
synteny between the four gene clusters encoding the respective
conjugative machineries of these two mobile elements (yellow and
orange genes in Figure 4A). Despite the extensive similarity of the
SXT and IncA/C conjugative transfer systems, these plasmids lack
homologues of setR and setD/C as well as int/xis, suggesting that
regulation of conjugative transfer differs between these elements.
The similarity of IncA/C plasmids and SXT/R391 ICEs is not
limited to genes important for conjugal DNA transfer. Ten genes
of unknown function (shown in black in Figure 4A), some of which
are interspersed within likely tra gene operons and some of which
are clustered together between traN and traF, are similar in the two
elements. Furthermore, most of these ten genes are in identical
locations in the two elements. Both elements also contain
homologs of bet and exo (shown in green in Figure 4A); these are
the only known homologs of the lRed recombination genes found
outside of bacteriophages. Together, the similarity of DNA
sequences and organization of SXT/R391 ICEs and IncA/C
plasmids suggests that these elements have a common ancestor.
The fact that the contents of the hotspots in the two classes of
elements are entirely distinct suggests that their evolutionary paths
diverged prior to acquisition of these variable DNA segments.
The minimal functional SXT/R391 ICE gene set
The conservation of the 52 core genes in all 13 SXT/R391
ICEs analyzed suggested that many or even all of these genes
would be required for key ICE functions of excision/integration,
Comparative ICE Genomics
PLoS Genetics | www.plosgenetics.org 6 December 2009 | Volume 5 | Issue 12 | e1000786
conjugative transfer and regulation. The presence of ten ICE core
genes of unknown function in IncA/C plasmids (black genes in
Figure 4A) is also consistent with the hypothesis that these genes
might be required for ICE transfer. However, our previous work
demonstrated that not all genes recognized here as part of the
conserved core gene set are required for SXT transfer. Beaber et al
showed that deletion of rumB – s026 (which includes 5 cores genes)
from SXT had no detectable influence on SXT excision or
transfer [32]. Therefore, we systematically deleted all of the core
ICE genes whose contributions had not previously been assessed,
in order to explore the hypothesis that these genes (especially those
also present in IncA/C plasmids) would be essential for ICE
transfer and to define the minimum functional SXT/R391 gene
set.
Surprisingly, deletion of most of the ICE core genes of unknown
function, including genes with homologues in IncA/C plasmids,
did not alter SXT transfer efficiency. Deletion of s002 or s003,
which are located downstream of int in all SXT/R391 ICEs, did
not alter the frequency of SXT transfer; similarly, deletion of
s082,s083, and s084, core genes of unknown function that are
found near the opposite end of SXT/R391 ICEs but not in
IncA/C plasmids, also did not influence SXT transfer frequency
(Figure 4B). Furthermore, deletion of s091, which is found
between traD and s043 in ICEs and IncA/C plasmids, did not
reduce SXT transfer (Figure 4B). In contrast, deletion of s043,
which has weak homology to traJ in the F plasmid (a gene
important in DNA processing) and is located in a transfer cluster
containing traI and traD, abolished transfer (Figure 4B, Dd),
suggesting that s043, here re-named traJ is required for SXT
transfer. It is unlikely that the transfer defect of SXTDtraJ can be
explained by polar effects of the deletion on downstream genes,
since traJ appears to be the last gene of an operon found
immediately upstream of hotspot 1. Similarly, deletion of s054,
which is found immediately 59of traC and is homologous to a
disulfide-bond isomerase dsbC, also abolished transfer (Figure 4B,
De). Interestingly, disulfide bond-isomerases are present in several
other conjugative systems [51]. However, it is not clear at this
point if the deletion of s054 from SXT accounts for the transfer
defect of SXTDs054, since we could not restore transfer by
complementation.
Additionally, Beaber et al found that deletion of s060 through
s073 in SXT, which includes 7 genes that are also found in IncA/
C plasmids reduced SXT transfer more than 100-fold [32]. We
constructed several smaller deletions in this region and found that
Figure 3. The boundaries of the 5 hotspots. The boundaries between conserved and hotspot variable regions are shown. Black typeface
indicates conserved sequence, while color indicates variable sequence. Numbers in parentheses indicate the number of intervening nucleotides. The
thin dotted lines indicate continuations of variable DNA. Bold letters indicate a non-conserved base. (A) Hotspot 1, which is present between traJ and
traL. Line 1: SXT, ICEVchInd4, ICEPalBan1; Line 2: R391, ICEPdaSpa1, ICEVchBan5, ICEVchInd5, ICEPmiUSA, ICESpuPO1, ICEVflInd, ICEVchMoz10,
ICEVchBan9; Line 3: ICEVchMex1. (B) Hotspot 2, which is present between traA and s054. Line 1: SXT, ICEVchInd4, ICEPmiUSA, ICEVflInd, ICEVchInd5,
ICEPalBan1, ICEVchBan5; Line 2: ICEPdaSpa1, ICEVchMoz10, ICEVchBan9; Line 3: R391; Line 4: ICEVchMex1; Line 5: ICESpuPO1. (C) Hotspot 3, which is
present between s073 and traF. Line 1: SXT, ICEVchInd4; Line 2: ICEVchMex1, ICEVflInd; Line 3: ICEVchMoz10, ICEVchBan9, ICEVchInd5, ICEPalBan1,
ICEVchBan5. Line 4: R391; Line 5: ICEPmiUSA; Line 6: ICESpuPO1; Line 7: ICEPdaSpa1. (D) Hotspot 4, which is present between traN and s063. Line
1: SXT, ICEVchInd4. Line 2: ICEVchInd5, ICEVchBan5; Line 3: ICESpuPO1, ICEPmiUSA; Line 4: R391, ICEVchMoz10, ICEVchBan9, ICEVflInd. Line
5: ICEPdaSpa1; Line 6: ICEPalBan1; Line 7: ICEVchMex1. (E) Hotspot 5, which is present between s026 and traI. Line 1: SXT, ICEVchInd4, ICEPdaSpa1,
R391, ICEVchMoz10, ICEVchBan9; Line 2: ICEPmiUSA; Line 3: ICESpuPO1, ICEPalBan1, ICEVflInd1; Line 4: ICEVchInd5, ICEVchBan5; Line 5: ICEVchMex1.
doi:10.1371/journal.pgen.1000786.g003
Comparative ICE Genomics
PLoS Genetics | www.plosgenetics.org 7 December 2009 | Volume 5 | Issue 12 | e1000786
deletion of s063, which is also found in pIP1202, reduced the
transfer frequency of SXT by ,100-fold, nearly the same amount
as deleting the entire region (Figure 4B). Complementation
analyses revealed that the absence of s063 accounted for the
transfer defect of SXTDs063 (data not shown). Even though
SXTDs063 was still capable of transfer, in our view, the drastic
reduction in the transfer frequency of this mutant warrants
inclusion of s063 into the minimum functional SXT ICE genome
(shown in Figure 4C). Other deletions in this region, including
deletions of bet,exo,s067,s068 and s070, which have orthologues in
IncA/C plasmids, resulted in #10-fold reductions in transfer
frequency. We therefore did not include these genes in the
minimal functional core SXT/R391 genome (Figure 4C).
The findings from our experiments testing the transfer
frequencies of SXT derivatives harboring core gene deletions
(shown in Figure 4B), coupled with our previous work demon-
strating the requirements for the predicted SXT tra genes in the
element’s transfer [32], suggest a minimal functional SXT/R391
ICE structure as shown in Figure 4C. This minimum element is
,29.7 kb and consists of 25 genes. Genes with related functions,
which in some cases encode proteins that likely form large
functional complexes (such as the conjugation apparatus), are
grouped together in the minimal genome. At the left end of the
minimum ICE genomes are xis and int, the integration/excision
module of SXT/R391 ICEs. In the minimal ICE genome, the
ICE oriT and mobI, which encodes a protein required for SXT
transfer [39], are no longer separated from the other genes (traIDJ)
that are also thought to play roles in the DNA processing events
required for conjugative DNA transfer. The genes required for
formation of the conjugation machinery, including the pilus, and
mating pair formation and stabilization [32,39] are divided
between three clusters (denoted mpf1-3 in Figure 4C). Finally, at
the right end of the minimal functional genome are the genes that
regulate ICE transfer (setC/D and setR). Thus, the minimal
Figure 4. Comparison of the SXT/R391 core genome with the genome of pIP1202 and defining the minimal functional SXT/R391
gene set. (A) Alignment of the conserved core genes of SXT/R391 ICEs with the genome of the IncA/C conjugative plasmid pIP1202 from Yersinia
pestis. The top line shows the same core ICE genes shown in Figure 2A. ORFs are color coded as follows: DNA processing, yellow; mating pair
formation, orange; DNA recombination and repair, green; integration/excision, red; replication, purple; regulation, gray; entry exclusion, blue;
homologous genes of unknown function, black; genes without corresponding counterparts in ICEs and pIP1202, white. Numbers shown in the
middle represent % identity between the orthologous proteins encoded by SXT and pIP1202 [GenBank:NC_009141]. The positions of the hotspots in
SXT/R391 ICEs are marked by downward pointing arrowheads. For pIP1202, the size of the sequences (which include IncA/C backbone DNA as well as
variable DNA) found at these locations as well as resistance markers are indicated by upward pointing arrowheads. aphA,aadA and strAB confer
resistance to aminoglycosides. sul1 and sul2 confer resistance to sulfonamides. cat,bla
SHV-1
,tetAR,qacED1 and merRTPCADE confer resistance to
chloramphenicol, b-lactams, tetracyclines, quaternary ammonium compounds and mercury ions, respectively. Detailed descriptions of the conserved
backbone of the IncA/C conjugative plasmids have been published elsewhere [48,50]. Regions that were deleted from SXT to investigate the function
of genes of unknown function (see panel B) are indicated with straight lines. Dotted lines indicate that the deletion included DNA in the adjacent
hotspots. (B) Influence of deletion of genes of unknown function on the frequency of SXT transfer. The mean values and standard deviations from
three independent experiments are shown. * indicates that the frequency of transfer was below the detection level (,10
28
). Deletion mutants SXTDa,
SXTDk and SXTDl, transferred at frequencies that were not significantly different from that of wild-type SXT (data not shown). (C) Proposed minimal
set of genes necessary for a functional SXT/R391 ICE. int, integration/excision module; mob, DNA processing module; mpf, mating pair formation
modules; reg, regulation module.
doi:10.1371/journal.pgen.1000786.g004
Comparative ICE Genomics
PLoS Genetics | www.plosgenetics.org 8 December 2009 | Volume 5 | Issue 12 | e1000786
functional SXT/R391 ICE is relatively small and organized into 3
discrete functional modules that mediate excision/integration,
conjugation, and regulation.
Even though deletion of 27 out of 52 SXT/R391 ICE core
genes proved to have little or no effect on SXT transfer frequency,
and hence these genes were not included in Figure 4C, it is
reasonable to presume that these genes encode functions that
enhance ICE fitness given their conservation. For example, the
presence of highly conserved bet and exo genes in all SXT/R391
ICEs suggests that there has been selection pressure to maintain
this ICE-encoded recombination system that promotes ICE
diversity by facilitating inter ICE recombination (G Garriss, MK
Waldor, V Burrus, in press). A key challenge for future studies will
be to determine how core genes of unknown function promote
ICE fitness.
Variations in the similarity of core genes
To identify genes in the SXT/R391 core genome that may be
subject to different selection pressures, we compared the percent
identity of each ICE’s core genes to the corresponding SXT gene
(Figure 5). Most of the ICEs’ core genes exhibited 94% to 98%
identity on the nucleotide level to SXT’s core genes. There was no
discernable difference in the degree of conservation of most core
genes that were or were not part of the minimal ICE, suggesting
that there are equal selective pressures on essential and non-
essential genes. However, we identified 8 genes (s026,traI,orfZ,
s073,traF,eex,s086, and setR) that exhibit significantly different
degrees of conservation (Figure 5 and Figure S1). Three of these
showed unusually high conservation, while the other 5 had below
average conservation. Two of the highly conserved genes, setR and
s086, are found at the extreme 39end of the elements. The
conservation of setR may reflect the key role of this gene in
controlling SXT gene expression. S086 may also play a role in
regulating SXT transfer [52]. The other highly conserved gene,
orfZ, is found between bet and exo and has no known function.
s026 and s073 are the most divergent of all the genes in the
backbone. s026 encodes a hypothetical protein with homologues in
many gram negative organisms. Although S026 is predicted to
contain a conserved domain, COG2378, which has a putative role
in transcription regulation, this protein is not required for SXT
transfer [32]. The significant divergence of s026 along with its lack
of essentiality suggests that this gene could become a pseudogene.
A similar argument could be made for s073, which encodes a
hypothetical protein that is also not required for ICE transfer.
However, this argument does not hold for traI or traF, two genes
which are essential for ICE transfer. Although the reasons which
account for the different degrees of conservation of these 8 core
genes are hard to ascertain at this point, the data in Figure 5
suggests that individual core genes are subject to different
evolutionary pressures.
Figure 5. Variations in the nucleotide conservation of core ICE genes. The nucleotide sequence of each core gene from each ICE was
compared to the corresponding sequence in SXT using pairwise BLASTn analyses to determine the percent identities. The average values for all of the
ICEs, excluding SXT and ICEVchInd4, are shown in the inset.
doi:10.1371/journal.pgen.1000786.g005
Comparative ICE Genomics
PLoS Genetics | www.plosgenetics.org 9 December 2009 | Volume 5 | Issue 12 | e1000786
Comparisons of core gene phylogenies
We created phylogenetic trees for each core gene based on their
respective nucleotide sequences to further explore the evolution of
the conserved backbone of SXT/R391 ICEs. Since we found such
a high degree of conservation for most of the core genes, the
bootstrap values for most of these trees were relatively low. Thus,
we concentrated on the most polymorphic genes found in Figure 5,
s026,s073,traI, and eex, for phylogenetic analyses. As shown in
Figure 6A, the trees for s026,traI and s073 exhibit 3 distinct
branching patterns. The lack of similarity in these phylogenetic
trees suggests that either individual core genes have evolved
independently or that high degrees of recombination mask their
common evolutionary history. The latter hypothesis seems more
likely since experimental findings have revealed that SXT/R391
ICEs can co-exist in a host chromosome in tandem [26] and
recombination between tandem elements can yield novel hybrid
ICEs with considerable frequency [53] (G Garriss, MK Waldor, V
Burrus, in press). Also, as noted above, the distributions of variable
genes among the ICEs shown in Figure 2 also supports the idea
that inter-ICE recombination is commonplace.
Unlike most core genes, the trees for traG and eex were similar.
In these two trees, the ICEs segregate into two evolutionarily
distinct groups (Figure 6B), confirming and extending previous
observations that revealed that there are two groups of eex and traG
sequences in SXT/R391 ICEs [54]. These two groups correspond
to the two functional SXT/R391 ICE exclusion groups.
Interactions between traG and eex of the same group mediate
ICE exclusion [55]. Thus, the identical 2 clusters of traG and eex
Figure 6. Phylogenetic analysis of several core ICE genes. Nucleotide sequences of the indicated core genes were used to generate the
phylogenetic trees shown. Bootstrap values are indicated at branch points. The individual scale bars represent genetic distances and reflect the
number of substitutions per residue.
doi:10.1371/journal.pgen.1000786.g006
Comparative ICE Genomics
PLoS Genetics | www.plosgenetics.org 10 December 2009 | Volume 5 | Issue 12 | e1000786
sequences observed in their respective trees reveals the co-
evolution of the traG/eex functional unit. The two groups of eex
sequences can also be observed in Figure 5 where the bifurcating
pattern reveals the 2 exclusion groups. This pattern is difficult to
discern for traG, perhaps because of the large size of this multi-
functional gene.
ICEVchBan8, an SXT-like ICE that lacks Int
sxt
The sequence of ICEVchBan8, which was derived from a non-
O1, non-O139 V. cholerae strain, is incomplete but it appears to
contain 49 out of 52 SXT/R391 core genes. However, since this
strain lacks Int
sxt
it was not included in our comparative analyses
above. It is not known if ICEVchBan8 is capable of excision or
transmission; however, it contains a P4-like integrase and a
putative xis. It is tempting to speculate that the genome of
ICEVchBan8 provides an illustration of how acquisition (presum-
ably via recombination) of a new integration/excision module may
generate a novel ICE family.
Perspectives
Comparative analysis of the genomes of the 13 SXT/R391
ICEs studied here has greatly refined our understanding of this
group of mobile genetic elements. These elements, which have
been isolated from 4 continents and the depths of the Pacific
Ocean, all have an identical genetic structure, consisting of the
same syntenous set of 52 conserved core genes that are interrupted
by clusters of diverse variable genes. All the elements have
insertions of variable DNA segments in the same five intergenic
hotspots that interrupt the conserved backbone. Furthermore,
some of the elements have additional insertions outside the
hotspots; however, in all cases the acquisition of variable DNA has
not compromised the integrity of the core genes required for ICE
mobility. Functional analyses revealed that less than half of the
conserved genes are necessary for ICE transmissibility and the
contributions of the 27 core genes of unknown function to ICE
fitness remains an open question. Finally, several observations
presented here suggest that recombination between SXT/R391
ICEs has been a major force in shaping the genomes of this
widespread family of mobile elements.
Although comparisons of the 13 ICE genomes analyzed here
strongly suggest that these mobile elements have undergone
extensive recombination during their evolutionary histories, there
is a remarkable degree of similarity among the SXT/R391 ICEs.
All of these ICEs consist of the same syntenous and nearly identical
52 genes. In contrast, other families of closely related mobile
elements, such as lambdoid or T4-like phages for example, exhibit
greater diversity [56,57]. Since the elements that we sequenced
were isolated from several different host species and from diverse
locations, the great degree of similarity of the SXT/R391 ICE
family does not likely reflect bias in the elements that we
sequenced. It is possible that this family of mobile elements is a
relatively recent creation of evolution and has yet to undergo
significant diversification.
To date, relatively few formal comparative genomic analyses of
other ICE families have been reported. Mohd-Zain et al [11]
identified several diverse ICEs and genomic islands that shared a
largely syntenous set of core genes with ICEHin1056, an ICE
originally identified in Haemophilus influenzae. However, even
though these elements share a similar genomic organization, they
exhibit far greater variability in the sites of insertion of variable
DNA and in the degree of conservation in their core genes
compared to SXT/R391 ICEs. Thus, although this group of
elements appears to share a common ancestor, they seem to have
diverged earlier in evolutionary history than the SXT/R391 ICEs.
However, when comparative genomic analyses were restricted to
ICEHin1056-related ICEs found in only two Haemophilus sp., Juhas
et al found that, like the SXT/R391 family of ICEs, these 7
ICEHin1056-related ICEs share greater than 90% similarity at the
DNA level in their nearly syntenous set of core genes [12]. It will
be interesting to learn the extent of conservation of genetic
structure and DNA sequence in additional ICE families to obtain a
wider perspective on ICE evolution.
Comparative genomic studies of bacteriophages have led to the
idea that the full range of phage sequences are part of common but
extremely diverse gene pool [58,59]. The SXT/R391 ICE
genomes suggest that there may be an even larger network of
phylogenetic relationships linking sequences found in all types of
mobile genetic elements including phages, plasmids, ICEs and
transposons. The genomes of SXT/R391 ICEs appear to be
amalgams of genes commonly associated with other types of
mobile elements. Many of the ICE core genes are usually
associated with phages, such as int,bet,exo and setR, or with
plasmids, such as the tra genes. Additionally, the SXT/R391 ICEs
and IncA/C plasmids clearly have a common ancestor, as we
found that the entire set of SXT/R391 tra genes are also present in
IncA/C plasmids. Thus, the genes present in all types of mobile
genetic elements appear to contribute to a common gene pool
from which novel variants of particular elements (such as
ICEVchBan8) or perhaps even novel types of mobile genetic
elements can arise.
Materials and Methods
ICE Sequencing
ICEPalBan1, ICEVchMex1, ICEVchInd4, ICEVchInd5 and
ICEVchBan5 were isolated using the plasmid capture system
described in Figure 1. The SXT chromosomal attachment
sequence, attB, was introduced into the modified F plasmid
pXX704 [34] to create pIceCap. This plasmid was then
introduced into a DprfC derivative of the Tc
R
E. coli strain
CAG18439. Exconjugants derived from matings between this
strain and those harboring the 5 ICEs listed above resulted in
strains carrying a pIceCap::ICE plasmid. Once captured, the
plasmids were isolated using the Qiagen plasmid midi kit for low-
copy plasmids (Qiagen). Isolated pIceCap::ICE plasmids were
then sequenced.
ICEVflInd genome was determined by sequencing several
overlapping cosmids that encompassed this ICE’s genome. Briefly,
genomic DNA from a Vibrio fluvialis strain carrying ICEVflInd was
prepared using the GNome DNA kit (QBIOgene). Sau3A1
restricted genomic DNA was used to create a SuperCos1
(Stratagene)-based cosmid library according the manufacture’s
instructions. The library was subsequently screened for cosmids
containing ICE-specific sequences using PCR with primers to
conserved core ICE sequences. Four cosmids containing overlap-
ping ICEVflInd sequences were identified and sequenced.
The genomes of 6 ICEs were sequenced by the Sanger random
shotgun method [60]. Briefly, small insert plasmid libraries (2–
3 kb) were constructed by random nebulization and cloning of
pIceCap::ICE DNA or of cosmid DNA for ICEVflInd. In the
initial random sequencing phase, 8–12 fold sequence coverage was
achieved. The sequences of either pIceCap or pSuperCos were
subtracted and the remaining sequences were assembled using the
Celera Assembler [61]. An initial set of open reading frames
(ORFs) that likely encode proteins was identified using GLIM-
MER [62], and those shorter than 90 base pairs (bp) as well as
some of those with overlaps eliminated.
Comparative ICE Genomics
PLoS Genetics | www.plosgenetics.org 11 December 2009 | Volume 5 | Issue 12 | e1000786
Bioinformatics
Nucleotide and amino acid conservation were assessed with the
appropriate BLAST algorithms. ICEs were aligned using clustalW
with default settings [63]. MAUVE [37] and LAGAN [38] were
used to identify core genes in Figure 2. To map the boundaries of
the hotspots, sequence comparisons were made using MAUVE
and then manually compared to find boundaries between
conserved and variable DNA as shown in Figure 3.
Phylogenetic trees were generated from alignments of nucleotide
sequences using the neighbor-joining method as implemented by
ClustalX software, version 2.011 [64]. The reliability of each tree
was subjected to a bootstrap test with 1000 replications. Trees
were edited using FigTree 1.22 (http://tree.bio.ed.ac.uk/software/
figtree/).
Generation and testing of SXT deletion mutants
CAG81439 harboring SXT was used as the host strain to create
the SXT deletion mutants shown in Figure 3; the deletions were
constructed using one-step gene inactivation as previously
described [44,65]. The primers used to create the deletion mutants
are available upon request. Matings were conducted as previously
described [16,44] using deletion mutants and a Kn
R
E. coli
recipient, CAG18420. Exconjugants were selected on LB agar
plates containing chloramphenicol, 20mg/ml (for SXT selection)
and kanamycin, 50 mg/ml. The frequency of exconjugant
formation was calculated by dividing the number of exconjugants
by the number of donors.
Supporting Information
Figure S1 Variations in the conservation of individual core ICE
genes. The percent identity of the nucleotide sequence of each
core gene and oriT versus the corresponding sequence in SXT was
calculated for all ICEs studied. The average values for each gene
(as shown in the inset of Figure 5) were then used in one-way
ANOVA comparisons to determine genes that exhibit significantly
more or less conservation compared to other core genes. p-values
of one-way ANOVA comparisons of each core ICE gene are
shown. The grid represents all pair-wise comparisons, and the
color indicates the level of significance as follows: red: p,.001,
orange: p,.01, and yellow: p,.05. Genes that exhibited a p-
value,.05 when compared with at least 50% of all other core
genes are discussed in the text.
Found at: doi:10.1371/journal.pgen.1000786.s001 (0.56 MB TIF)
Table S1 Contents of the hotspots.
Found at: doi:10.1371/journal.pgen.1000786.s002 (0.14 MB
DOC)
Acknowledgments
We thank Brigid Davis and Fre´de´rique Le Roux for helpful comments on
the manuscript. We thank Robert Hall for facilitating the work described
here and Yoshiharu Yamaichi for suggesting the design of pIceCap.
Author Contributions
Conceived and designed the experiments: RAFW VB MKW. Performed
the experiments: RAFW DEF MS DC GG CD VB. Analyzed the data:
RAFW MS DC VB MKW. Contributed reagents/materials/analysis tools:
MMC. Wrote the paper: RAFW VB MKW.
References
1. Frost LS, Leplae R, Summers AO, Toussaint A (2005) Mobile genetic elements:
the agents of open source evolution. Nat Rev Microbiol 3: 722–732.
2. Gogarten JP, Townsend JP (2005) Horizonta l gene transfer, genome innovation
and evolution. Nat Rev Microbiol 3: 679–687.
3. Lawrence JG, Hendrickson H (2005) Genome evolution in bacteria: order
beneath chaos. Curr Opin Microbiol 8: 572–578.
4. Ochman H, Lerat E, Daubin V (2005) Examining bacterial species under the specter
of gene transfer and exchange. Proc Natl Acad Sci U S A 102 Suppl 1: 6595–6599.
5. Beaber JW, Hochhut B, Waldor MK (2004) SOS response promotes horizontal
dissemination of antibiotic resistance genes. Nature 427: 72–74.
6. Burrus V, Waldor MK (2003) Control of SXT integration and excision.
J Bacteriol 185: 5045–5054.
7. Auchtung JM, Lee CA, Mons on RE, Lehman AP, Grossman AD (2005)
Regulation of a Bacillus subtilis mobile genetic element by intercellular signaling
and the global DNA damage response. Proc Natl Acad Sci U S A 102:
12554–12559.
8. Caparon MG, Scott JR (1989) Excision and insertio n of the conjugative
transposon Tn916 involves a novel recombination mechanism. Cell 59:
1027–1034.
9. Cheng Q, Paszkiet BJ, Shoemaker NB, Gardner JF, Salyers AA (2000)
Integration and excision of a Bacteroides conjugative transposon, CTnDOT.
J Bacteriol 182: 4035–4043.
10. Hagege J, Pernodet JL, Friedmann A, Guerineau M (1993) Mode and origin of
replication of pSAM2, a conjugative integrating element of Streptomyces
ambofaciens. Mol Microbiol 10: 799–812.
11. Mohd-Zain Z, Turner SL, Cerdeno-Tarraga AM, Lilley AK, Inzana TJ, et al.
(2004) Transferable antibiotic resistance elements in Haemophilus influenzae
share a common evolutionary origin with a diverse family of syntenic genomic
islands. J Bacteriol 186: 8114–8122.
12. Juhas M, Power PM, Harding RM, Ferguson DJ, Dimopoulou ID, et al. (2007)
Sequence and functional analyses of Haemophilus spp. genomic islands.
Genome Biol 8: R237.
13. Burrus V, Waldor MK (2004) Shaping bacterial genomes with integrative and
conjugative elements. Res Microbiol 155: 376–386.
14. Boltner D, MacMahon C, Pembroke JT, Strike P, Osborn AM (2002) R391: a
conjugative integrating mosaic comprised of phage, plasmid, and transposon
elements. J Bacteriol 184: 5158–5169.
15. Whittle G, Shoemake r NB, Salyers AA (2002) The role of Bacteroi des
conjugative transposons in the dissemination of antibiotic resistance genes. Cell
Mol Life Sci 59: 2044–2054.
16. Waldor MK, Tschape H, Mekalanos JJ (1996) A new type of conjugative
transposon encodes resistance to sulfamethoxazole, trimethoprim, and strepto-
mycin in Vibrio cholerae O139. J Bacteriol 178: 4157–4165.
17. Shoemaker NB, Barber RD, Salyers AA (1989) Cloning and characterization of
a Bacteroides conjugal tetracycline-erythromycin resistance element by using a
shuttle cosmid vector. J Bacteriol 171: 1294–1302.
18. Franke AE, Clewell DB (1981) Evidence for a chromosome-borne resistance
transposon (Tn916) in Streptococcus faecalis that is capable of ‘‘conjugal’’
transfer in the absence of a conjugative plasmid. J Bacteriol 145: 494–502.
19. Ravatn R, Studer S, Springael D, Zehnder AJ, van der Meer JR (1998)
Chromosomal integration, tandem amplification, and deamplification in Pseudo-
monas putida F1 of a 105-kilobase genetic element containing the chlorocatechol
degradative genes from Pseudomonas sp. Strain B13. J Bacteriol 180: 4360–4369.
20. Sullivan JT, Ronson CW (1998) Evolution of rhizobia by acquisition of a 500-kb
symbiosis island that integrates into a phe-tRNA gene. Proc Natl Acad Sci U S A
95: 5145–5149.
21. Cholera Working Group (1993) Large epidemic of cholera-like disease in
Bangladesh caused by Vibrio cholerae O139 synonym Bengal. Lancet 342: 387–390.
22. Ahmed AM, Shinoda S, Shimamoto T (2005) A variant type of Vibrio cholerae
SXT element in a multidrug-resistant strain of Vibrio fluvialis. FEMS Microbiol
Lett 242: 241–247.
23. Osorio CR, Marrero J, Wozniak RA, Lemos ML, Burrus V, et al. (2008)
Genomic and functional analysis of ICEPdaSpa1, a fish-pathogen-derived SXT-
related integrating conjugative element that can mobilize a virulence plasmid.
J Bacteriol 190: 3353–3361.
24. Pembroke JT, Piterina AV (2006) A novel ICE in the genome of Shewanella
putrefaciens W3-18-1: comparison with the SXT/R391 ICE-like elements.
FEMS Microbiol Lett 264: 80–88.
25. Hochhut B, Lotfi Y, Mazel D, Faruqu e SM, Woodgate R, et al. (2001)
Molecular analysis of antibiotic resistance gene clusters in vibrio cholerae O139
and O1 SXT constins. Antimicrob Agents Chemother 45: 2991–3000.
26. Hochhut B, Beaber JW, Woodgate R, Waldor MK (2001) Formation of
chromosomal tandem arrays of the SXT element and R391, two conjugative
chromosomally integrating elements that share an attachment site. J Bacteriol
183: 1124–1132.
27. Coetzee JN, Datta N, Hedges RW (1972) R factors from Proteus rettgeri. J Gen
Microbiol 72: 543–552.
28. Burrus V, Marrero J, Waldor MK (2006) The current ICE age: biology and
evolution of SXT-related integrating conjugative elements. Plasmid 55:
173–183.
Comparative ICE Genomics
PLoS Genetics | www.plosgenetics.org 12 December 2009 | Volume 5 | Issue 12 | e1000786
29. Hochhut B, Waldor MK (1999) Site-specific integration of the conjugal Vibrio
cholerae SXT element into prfC. Mol Microbiol 32: 99–110.
30. Maeda K, Nojiri H, Shintani M, Yoshida T, Habe H, et al. (2003) Complete
nucleotide sequence of carbazole/dioxin-degrading plasmid pCAR1 in Pseudo-
monas resinovorans strain CA10 indicates its mosaicity and the presence of large
catabolic transposon Tn4676. J Mol Biol 326: 21–33.
31. Murata T, Ohnishi M, Ara T, Kaneko J, Han CG, et al. (2002) Complete
nucleotide sequence of plasmid Rts1: implications for evolution of large plasmid
genomes. J Bacteriol 184: 3194–3202.
32. Beaber JW, Hochhut B, Waldor MK (2002) Gen omic and functional analyses of
SXT, an integrating antibiotic resistance gene transfer element derived from
Vibrio cholerae. J Bacteriol 184: 4259–4269.
33. Beaber JW, Burrus V, Hochhut B, Waldor MK (2002) Comparison of SXT and
R391, two conjugative integrating elements: definition of a genetic backbone for
the mobilization of resistance determinants. Cell Mol Life Sci 59: 2065–2070.
34. Niki H, Hiraga S (1997) Subcellular distribution of actively partitioning F
plasmid during the cell division cycle in E. coli. Cell 90: 951–957.
35. Yamaichi Y, Niki H (2004) migS, a cis-acting site that affects bipolar positioning
of oriC on the Escherichia coli chromosome. Embo J 23: 221–233.
36. Faruque SM, Tam VC, Chowdhury N, Diraphat P, Dziejman M, et al. (2007)
Genomic analysis of the Mozambique strain of Vibrio cholerae O1 reveals the
origin of El Tor strains carrying classical CTX prophage. Proc Natl Acad
Sci U S A 104: 5151–5156.
37. Darling AC, Mau B, Blattner FR, Perna NT (2004) Mauve: multiple alignment
of conserved genomic sequence with rearrangements. Genome Res 14:
1394–1403.
38. Brudno M, Do CB, Cooper GM, Kim MF, Davydov E, et al. (2003) LAGAN
and Multi-LAGAN: efficient tools for large-scale multiple alignment of genomic
DNA. Genome Res 13: 721–731.
39. Ceccarelli D, Daccord A, Rene M, Burrus V (2008) Identification of the orig in of
transfer (oriT) and a new gene required for mobilization of the SXT/R391
family of integrating conjugative elements. J Bacteriol 190: 5328–5338.
40. Toleman MA, Bennett PM, Walsh TR (2006) ISCR elements: novel gene-
capturing systems of the 21st century? Microbiol Mol Biol Rev 70: 296–316.
41. Bordeleau E, Brouillette E, Robichaud N, Burrus V (2009) Beyond antibiotic
resistance: integrating conjugative elements of the SXT/R391 family that
encode novel diguanylate cyclases participate to c-di-GMP signalling in Vibrio
cholerae. Environ Microbiol.
42. Jenal U, Malone J (2006) Mechanisms of cyclic-di-GMP signaling in bacteria.
Annu Rev Genet 40: 385–407.
43. Romling U, Gomelsky M, Galperin MY (2005) C-di-GMP: the dawning of a
novel bacterial signalling system. Mol Microbiol 57: 629–639.
44. Wozniak RA, Waldor MK (2009) A toxin-antitoxin system promotes the
maintenance of an integrative conjugative element. PLoS Genet 5: e1000439.
doi:10.1371/journal.pgen.1000439.
45. Lindsey RL, Fedorka-Cray PJ, Frye JG, Meinersmann RJ (2009) Inc A/C
plasmids are prevalent in multidrug-resistant Salmonella enterica isolates. Appl
Environ Microbiol 75: 1908–1915.
46. Llanes C, Gabant P, Couturier M, Bayer L, Plesiat P (1996) Molecular analysis
of the replication elements of the broad-host-range RepA/C replicon. Plasmid
36: 26–35.
47. Galimand M, Guiyoule A, Gerbaud G, Rasoamanana B, Chanteau S, et al.
(1997) Multidrug resistance in Yersinia pestis mediated by a transferable
plasmid. N Engl J Med 337: 677–680.
48. Welch TJ, Fricke WF, McDermott PF, White DG, Rosso ML, et al. (2007)
Multiple antimicrobial resistance in plague: an emerging public health risk.
PLoS One 2: e309. doi:10.1371/journal.pone.0000309.
49. Pan JC, Ye R, Wang HQ, Xiang HQ, Zhang W, et al. (2008) Vibrio cholerae
O139 multiple-drug resistance mediated by Ye rsinia pestis pIP1202-like
conjugative plasmids. Antimicrob Agents Chemother 52: 3829–3836.
50. Fricke WF, Welch TJ, McDermott PF, Mammel MK, LeClerc JE, et al. (2009)
Comparative genomics of the IncA/C multidrug resistance plasmid family.
J Bacteriol 191: 4750–4757.
51. Elton TC, Holland SJ, Frost LS, Hazes B (2005) F-like type IV secretion systems
encode proteins with thioredoxin folds that are putative DsbC homologues.
J Bacteriol 187: 8267–8277.
52. Beaber JW, Waldor MK (2004) Identification of operators and promoters that
control SXT conjugative transfer. J Bacteriol 186: 5945–5949.
53. Burrus V, Waldor MK (2004) Formation of SXT tandem arrays and SXT-R391
hybrids. J Bacteriol 186: 2636–2645.
54. Marrero J, Waldor MK (2007) The SXT/R391 family of integrative conjugative
elements is composed of two exclusion groups. J Bacteriol 189: 3302–3305.
55. Marrero J, Waldor MK (2005) Interactions between inner membrane proteins in
donor and recipient cells limit conjugal DNA transfer. Dev Cell 8: 963–970.
56. Juhala RJ, Ford ME, Duda RL, Youlton A, Hatfull GF, et al. (2000) Genomic
sequences of bacteriophages HK97 and HK022: pervasive genetic mosaicism in
the lambdoid bacteriophages. J Mol Biol 299: 27–51.
57. Filee J, Bapteste E, Susko E, Krisch HM (2006) A selective barrier to horizontal
gene transfer in the T4-type bacteriophages that has preserved a core genome
with the viral replication and structural genes. Mol Biol Evol 23: 1688–1696.
58. Hendrix RW, Hatfull GF, Smith MC (2003) Bacteriophages with tails: chasing
their origins and evolution. Res Microbiol 154: 253–257.
59. Hendrix RW, Smith MC, Burns RN, Ford ME, Hatfull GF (1999) Evolutionary
relationships among diverse bacteriophages and prophages: all the world’s a
phage. Proc Natl Acad Sci U S A 96: 2192–2197.
60. Fouts DE, Mongodin EF, Mandrell RE, Miller WG, Rasko DA, et al. (2005)
Major structural differences and novel potential virulence mechanisms from the
genomes of multiple campylobacter species. PLoS Biol 3: e15. doi:10.1371/
journal.pbio.0030015.
61. Myers EW, Sutton GG, Delcher AL, Dew IM, Fasulo DP, et al. (2000) A whole-
genome assembly of Drosophila. Science 287: 2196–2204.
62. Delcher AL, Harmon D, Kasif S, White O, Salzberg SL (1999) Improved
microbial gene identification with GLIMMER. Nucleic Acids Res 27:
4636–4641.
63. Thompson JD, Higgins DG, Gibson TJ (1994) CLUSTAL W: improving the
sensitivity of progres sive multipl e sequence alignment throug h sequence
weighting, position-specific gap penalties and weight matrix choice. Nucleic
Acids Res 22: 4673–4680.
64. Larkin MA, Blackshields G, Brown NP, Chenna R, McGettigan PA, et al. (2007)
Clustal W and Clustal X version 2.0. Bioinformatics 23: 2947–2948.
65. Datsenko KA, Wanner BL (2000) One-step inactivation of chromosomal genes
in Escherichia coli K-12 using PCR products. Proc Natl Acad Sci U S A 97:
6640–6645.
66. Burrus V, Quezada-Calvillo R, Marrero J, Waldor MK (2006) SXT-related
integrating conjugative element in New World Vibrio cholerae. Appl Environ
Microbiol 72: 3054–3057.
67. Das B, Halder K, Pal P, Bhadra RK (2007) Small chromosomal integration site
of classical CTX prophage in Mozambique Vibrio cholerae O1 biotype El Tor
strain. Arch Microbiol 188: 677–683.
68. Pearson MM, Sebaihia M, Churcher C, Quail MA, Seshasayee AS, et al. (2008)
Complete genome sequence of uropathogenic Proteus mirabilis, a master of both
adherence and motility. J Bacteriol 190: 4027–4037.
69. Nair GB, Faruque SM, Bhuiyan NA, Kamruz zaman M, Siddique AK, et al.
(2002) New variants of Vibrio cholerae O1 biotype El Tor with attributes of the
classical biotype from hospitalized patients with acute diarrhea in Bangladesh.
J Clin Microbiol 40: 3296–3299.
Comparative ICE Genomics
PLoS Genetics | www.plosgenetics.org 13 December 2009 | Volume 5 | Issue 12 | e1000786
... Our findings refine our phylogenetic knowledge of L. lactis ICEs, and corroborate that the evolution and diversification of ICEs is driven by the modular exchanges of genetic regions that encode homologous functions [24,25]. Our results show that, although there is conservation in composition of the core ICE clusters, there is intergenic variation present between these L.ICE_CGs. ...
... In the absence of verification of their mobilization activity in vivo, it is unclear how these additional core-genes influences ICE functions, and although they are likely to play a role in the ICE lifecycle it cannot be excluded that they belong to the ICE-cargo. Similar 'hotspots' for novel DNA acquisition in ICEs has been observed previously and supports the overall genetic plasticity of this family of MGEs [25]. ...
... The identifiers describe as followed: tcpA = coupling protein, mobT = relaxase, tcpC and tcpE two MPF associated genes, virB4 = ATPase function, int = integrase. The listed locus tags (first column) refer to the standardized locus tag numbering in the NCBI database, which differ from those in the original genome assembly, to facilitate their correspondence we provide the original locus tags for the first and last gene: LLKF_RS11295 = LLKF_2255, and LLKF_11180 = LLKF_2229 van der Els et al.BMC Genomics (2024) 25:Example of the phylogenetic trees and clade definitions generated for each of the 17 core L. lactis ICE genes. In this example is the phylogenetic tree of the MobT function and the E-values of each clade specific HMM profile on each entry. ...
Article
Full-text available
Lactococcus lactis is widely applied by the dairy industry for the fermentation of milk into products such as cheese. Adaptation of L. lactis to the dairy environment often depends on functions encoded by mobile genetic elements (MGEs) such as plasmids. Other L. lactis MGEs that contribute to industrially relevant traits like antimicrobial production and carbohydrate utilization capacities belong to the integrative conjugative elements (ICE). Here we investigate the prevalence of ICEs in L. lactis using an automated search engine that detects colocalized, ICE-associated core-functions (involved in conjugation or mobilization) in lactococcal genomes. This approach enabled the detection of 36 candidate-ICEs in 69 L. lactis genomes. By phylogenetic analysis of conserved protein functions encoded in all lactococcal ICEs, these 36 ICEs could be classified in three main ICE-families that encompass 7 distinguishable ICE-integrases and are characterized by apparent modular-exchangeability and plasticity. Finally, we demonstrate that phylogenetic analysis of the conjugation-associated VirB4 ATPase function differentiates ICE- and plasmid-derived conjugation systems, indicating that conjugal transfer of lactococcal ICEs and plasmids involves genetically distinct machineries. Our genomic analysis and sequence-based classification of lactococcal ICEs creates a comprehensive overview of the conserved functional repertoires encoded by this family of MGEs in L. lactis, which can facilitate the future exploitation of the functional traits they encode by ICE mobilization to appropriate starter culture strains.
... Therefore, the full set of genes required for conjugation has been experimentally defined for only a very limited number of prototype or representative plasmids/ICEs. [16][17][18][19] Rts1 is a self-transmissible kanamycin (Km)-resistance plasmid that is 217,182 bp in length and was originally found in a clinical isolate of Proteus vulgaris. 20 Rts1 is a prototype of the T-incompatibility group 21 and expresses pleiotropic 2 Full gene set for the conjugation of plasmid Rts1 thermosensitive phenotypes in autonomous replication, 22 conjugative transfer, 20 host cell growth, 23 and the restriction 13,14 of T-even phages. ...
... We re-performed a homology search of all predicted 300 ORFs of Rts1 ( Fig. 1) against the genes of functionally wellcharacterized conjugation elements and found homologs not only in plasmids F and R27 but also SXT, a 100-kb multidrug resistance-encoding ICE originally discovered in V. cholerae O139 31,32 whose conjugation system has also been functionally characterized 16,18,19,31,33,34 (Table 1, see Supplementary Table S3 for the revised annotation data of SXT). Of the conjugation systems of the three elements, the ORFs of Rts1 showed the highest similarity to those of SXT. ...
... 35 The MPF types of Rts1, SXT and R27 were F type (MFP F ). 36 The gene organization was also relatively well conserved between Rts1 and SXT (Fig. 2). We found that 19 ORFs of Rts1 (Orf196, Orf201, Orf202, Orf204, Orf207-Orf211, Orf213, Orf215-Orf218, Orf240-Orf242, Orf244, and Orf246) exhibited noticeable amino acid sequence similarity to the gene products required for the conjugation of SXT 19,31,33 (Table 1). Although Orf185 and Orf252 did not show noticeable homology to any gene products of SXT, the former showed noticeable homology to TrhP of R27, and the latter contained the MobI domain, suggesting its role in DNA transfer. ...
Article
Full-text available
While conjugation-related genes have been identified in many plasmids by genome sequencing, functional analyses have not yet been performed in most cases, and a full set of conjugation genes has been identified for only a few plasmids. Rts1, a prototype IncT plasmid, is a conjugative plasmid that was originally isolated from Proteus vulgaris. Here, we conducted a systematic deletion analysis of Rts1 to fully understand its conjugation system. Through this analysis along with complementation assays, we identified 32 genes that are required for the efficient conjugation of Rts1 from Escherichia coli to E. coli. In addition, the functions of the 28 genes were determined or predicted; 21 were involved in mating-pair formation, three were involved in DNA transfer and replication, including a relaxase gene belonging to the MOBH12 family, one was involved in coupling, and three were involved in transcriptional regulation. Among the functionally well-analysed conjugation systems, most of the 28 genes showed the highest similarity to those of the SXT element, which is an integrative conjugative element of Vibrio cholerae. The Rts1 conjugation gene set included all 23 genes required for the SXT system. Two groups of plasmids with conjugation systems nearly identical or very similar to that of Rts1 were also identified.
... SXT and R391 belong to the SXT/R391 family of ICEs, which were the first discovered members of this family (28)(29)(30). They share several common features such as the ability to integrate into a specific chromosomal integration site (prfC) in the host bacterium, to transfer horizontally between donor and recipient bacteria via conjugation, and their conserved gene structure (31,32). After the discovery of SXT and R391, many similar ICEs with comparable structure and function were found in other countries and strains (6,(33)(34)(35). ...
... The core gene region of most SXT/R391 ICEs consists of 52 core genes, forming a conserved structural backbone (32). In addition, the five hotspots (HS1-5) and five variable regions (VRI-VRV) are the most complex and studied areas of variation. ...
Article
Full-text available
Proteus mirabilis can transfer transposons, insertion sequences, and gene cassettes to the chromosomes of other hosts through SXT/R391 integrative and conjugative elements (ICEs), significantly increasing the possibility of antibiotic resistance gene (ARG) evolution and expanding the risk of ARGs transmission among bacteria. A total of 103 strains of P. mirabilis were isolated from 25 farms in China from 2018 to 2020. The positive detection rate of SXT/R391 ICEs was 25.2% (26/103). All SXT/R391 ICEs positive P. mirabilis exhibited a high level of overall drug resistance. Conjugation experiments showed that all 26 SXT/R391 ICEs could efficiently transfer to Escherichia coli EC600 with a frequency of 2.0 × 10 ⁻⁷ to 6.0 × 10 ⁻⁵ . The acquired ARGs, genetic structures, homology relationships, and conservation sequences of 26 (19 different subtypes) SXT/R391 ICEs were investigated by high-throughput sequencing, whole-genome typing, and phylogenetic tree construction. ICE Pmi ChnHBRJC2 carries erm (42 ), which have never been found within an SXT/R391 ICE in P. mirabilis , and ICE Pmi ChnSC1111 carries 19 ARGs, including clinically important cfr , bla CTX-M-65 , and aac(6')-Ib-cr , making it the ICE with the most ARGs reported to date. Through genetic stability, growth curve, and competition experiments, it was found that the transconjugant of ICE Pmi ChnSCNNC12 did not have a significant fitness cost on the recipient bacterium EC600 and may have a higher risk of transmission and dissemination. Although the transconjugant of ICE Pmi ChnSCSZC20 had a relatively obvious fitness cost on EC600, long-term resistance selection pressure may improve bacterial fitness through compensatory adaptation, providing scientific evidence for risk assessment of horizontal transfer and dissemination of SXT/R391 ICEs in P. mirabilis . IMPORTANCE The spread of antibiotic resistance genes (ARGs) is a major public health concern. The study investigated the prevalence and genetic diversity of integrative and conjugative elements (ICEs) in Proteus mirabilis , which can transfer ARGs to other hosts. The study found that all of the P. mirabilis strains carrying ICEs exhibited a high level of drug resistance and a higher risk of transmission and dissemination of ARGs. The analysis of novel multidrug-resistant ICEs highlighted the potential for the evolution and spread of novel resistance mechanisms. These findings emphasize the importance of monitoring the spread of ICEs carrying ARGs and the urgent need for effective strategies to combat antibiotic resistance. Understanding the genetic diversity and potential for transmission of ARGs among bacteria is crucial for developing targeted interventions to mitigate the threat of antibiotic resistance.
... OrbA was previously shown to protect ICP1 against the type I BREX system encoded by the integrative and conjugative element (ICE), SXT VchInd5 25 . The SXT family of ICEs was initially discovered for harboring resistance to sulfamethoxazole and trimethoprim (SXT) and has recently been shown to encode various phage defense systems in a region known to vary in genomic content, denoted as hotspot 5 25,30 . OrbA does not share predicted structural similarity with known inhibitors, suggesting it inhibits BREX through a different mechanism 25 . ...
Preprint
Full-text available
Unlabelled: Bacteria and phages are locked in a co-evolutionary arms race where each entity evolves mechanisms to restrict the proliferation of the other. Phage-encoded defense inhibitors have proven powerful tools to interrogate how defense systems function. A relatively common defense system is BREX (Bacteriophage exclusion); however, how BREX functions to restrict phage infection remains poorly understood. A BREX system encoded by the SXT integrative and conjugative element, Vch Ind5, was recently identified in Vibrio cholerae , the causative agent of the diarrheal disease cholera. The lytic phage ICP1 that co-circulates with V. cholerae encodes the BREX inhibitor OrbA, but how OrbA inhibits BREX is unclear. Here, we determine that OrbA inhibits BREX using a unique mechanism from known BREX inhibitors by directly binding to the BREX component BrxC. BrxC has a functional ATPase domain that, when mutated, not only disrupts BrxC function but also alters how BrxC multimerizes. Furthermore, we find that OrbA binding disrupts BrxC-BrxC interactions. We determine that OrbA cannot bind BrxC encoded by the distantly related BREX system encoded by the SXT Vch Ban9, and thus fails to inhibit this BREX system that also circulates in epidemic V. cholerae . Lastly, we find that homologs of the Vch Ind5 BrxC are more diverse than the homologs of the Vch Ban9 BrxC. These data provide new insight into the function of the BrxC ATPase and highlight how phage-encoded inhibitors can disrupt phage defense systems using different mechanisms. Importance: With renewed interest in phage therapy to combat antibiotic-resistant pathogens, understanding the mechanisms bacteria use to defend themselves against phages and the counter-strategies phages evolve to inhibit defenses is paramount. Bacteriophage exclusion (BREX) is a common defense system with few known inhibitors. Here, we probe how the vibriophage-encoded inhibitor OrbA inhibits the BREX system of Vibrio cholerae , the causative agent of the diarrheal disease cholera. By interrogating OrbA function, we have begun to understand the importance and function of a BREX component. Our results demonstrate the importance of identifying inhibitors against defense systems, as they are powerful tools for dissecting defense activity and can inform strategies to increase the efficacy of some phage therapies.
... ICEs are widely distributed in Gramnegative and Gram-positive bacteria (Delavat et al. 2017). Various cargo genes of ICEs have been reported, such as those responsible for antimicrobial resistance (Christie et al. 1987, Harada et al. 2010, Shoemaker et al. 1989, Whittle et al. 2002, Wozniak et al. 2009), biofilm formation (Carter et al. 2010), and metabolism of alternative carbon sources (Ravatn et al. 1998 3 3 1993 1 1 1994 3 3 1995 3 3 1999 0 2000 2 2 2001 1 1 1 3 2002 1 1 2003 2 1 3 2004 1 1 2005 1 4 5 2006 2 3 1 6 2007 1 3 1 5 2008 1 3 1 2 7 2009 2 1 3 6 2010 4 4 2011 1 7 2 10 2012 5 2 7 2013 7 2 5 14 2014 2 4 6 2015 6 4 10 2016 11 11 2017 1 1 Total 15 16 55 2 2 4 0 12 34 140 of 86 open reading frames (ORFs) were identified in ICEmST, and 24 ORFs were predicted to be heavy metal resistance genes (Arai et al. 2019). Among the 24 ORFs, 17 formed a heavy metal homeostasis/resistance island called the copper homeostasis and silver resistance island (CHASRI). ...
Article
Salmonella enterica subsp. enterica serovar Typhimurium and its monophasic variant (Salmonella 4,[5],12:i:-) are two of the most frequently isolated serovars in food animals worldwide. Especially, human and swine infections by Salmonella 4,[5],12:i:- have increased considerably in Western countries since the mid-1990s. Although bovine and swine salmonellosis by Salmonella 4,[5],12:i:- has been increasing in Japan, the genetic background of the serovar has been unclear. In our study, whole genome sequence (WGS)-based phylogenetic analysis was performed to reveal the characteristics of Salmonella Typhimurium and Salmonella 4,[5],12:i:- isolated in Japan. Nine distinctive clades were identified and clonal replacements were observed several times in the last four decades. Among the nine clades, clade 9, which is a multidrug-resistant and predominant clone after 2010, has similar characteristics to the European clone that has been a major threat to public and animal health worldwide in recent years. Furthermore, the two distinctive clades in clade 9 were identified by phylogenetic analysis. Each clade has diverged through microevolution mediated by the acquisition or deletion of mobile genetic elements such as plasmids, prophages, composite transposons, and integrative and conjugative elements. The present review focuses on WGS-based phylogenetic studies of Salmonella Typhimurium and Salmonella 4,[5],12:i:- isolated from food animals in Japan.
Preprint
Full-text available
In Bangladesh, Vibrio cholerae lineages are undergoing genomic evolution, causing pandemics and outbreaks with increased virulence, resistance, spreading ability and disease severity. However, our understanding of the genomic determinants influencing transmission and disease severity patterns, as well as their interplay, remains incomplete. Here, we developed a computational framework based on machine-learning, genome scale metabolic modelling (GSSM) and 3D structural analysis, to identify V. cholerae signatures genomic traits linked to lineage transmission dynamics and disease severity. We analysed isolates collected from in-patients across six regions in Bangladesh from 2015 to 2021, and uncovered a core set of accessory genes, coding, and intergenic SNPs uniquely present in the most recent dominant lineage and underlying lineage transmission, with virulence, motility, colonization, biofilm formation, acid tolerance and bacteriophage resistance functions. Furthermore, we uncovered the existence of a strong correlation between a core set of V. cholerae genomic determinants and disease severity patterns (diarrhoeal duration, number of stools, abdominal pain, vomit, and dehydration). A subset of these determinants overlapped with those driving lineage transmission dynamics. Through GSMM and 3D structure analysis, we inferred the mechanistic bases underlying their selection and unveiled a complex interplay between transcription regulation, protein interactions and stability, and metabolic networks leading to severe symptoms. These connections influence lifestyle adaptation, intestinal colonization, oxidative stress and acid tolerance through modulation of ribosome, fatty acid and peptide biosynthesis, bacterial efflux systems, virulence, and resistant genes. Our computational framework allows to uncover signature traits which can provide insights for advancing therapeutics and developing targeted interventions to mitigate cholera spread.
Article
Objective: The objective of this study was to characterize ICEAplChn2, a novel SXT/R391-related integration and conjugation element (ICE) carrying 19 drug resistance genes, in a clinical isolate of Actinobacillus pleuropneumoniae from swine. Methods: Whole genome sequencing (WGS) of A. pleuropneumoniae CP063424 strain was completed using a combination of third-generation PacBio and second-generation Illumina. The putative ICE was predicted by the online tool ICEfinder. ICEAplChn2 was analyzed by PCR, conjugation experiments, and bioinformatics tools. Results: A. pleuropneumoniae CP063424 strain exhibited high minimum inhibitory concentrations of clindamycin (1,024 mg/L). The WGS data revealed that ICEAplChn2, with a length of 167,870 bp and encoding 151 genes, including multiple antibiotic resistance genes such as erm(42), VanE, LpxC, dfrA1, golS, aadA3, EreA, dfrA32, tetR(C), tet(C), sul2, aph(3)″-lb, aph(6)-l, floR, dfrA, ANT(3″)-IIa, catB11, and VanRE, was found to be related to the SXT/R391 family on the chromosome of A. pleuronipneumoniae CP063424. The circular intermediate of ICEAplChn2 was detected by PCR, but conjugation experiments showed that it was not self-transmissible. Conclusions: To our knowledge, ICEAplChn2 is the longest member with the most resistance genes in the SXT/R391 family. Meanwhile, ATP-binding cassette superfamily was found to be inserted in the ICEAplChn2 and possessed a new insertion region, which is the first description in the SXT/R391 family.
Article
Introduction: Cholera lytic phages contribute to the genetic diversity and evolution of Vibrio cholerae. To protect against the phages, the pathogen has acquired various resistance mechanisms. Objective: To identify antiphage systems located on mobile genetic elements in V. cholerae serogroup O1 El Tor biotype strains. Materials and methods: Nucleotide sequences of complete genomes of 77 toxigenic V. cholerae O1 El Tor strains imported to the Russian Federation and neighboring countries in 1970–2014 were analyzed using the Blast NCBI GenBank algorithm and REALPHY online tool. Results: We observed that the examined strains contained two types of anti-phage systems in hotspot 5 of the ICE SXT element: BREX, common for ICE VchBan9, and BREX with Abi typical of ICE VchInd5. We established a direct relationship between the presence of the PLE4 antiphage island and the kappa phage. V. cholerae O1 El Tor strains containing PLE4, except for one isolate, have BREX ICE VchBan9 and are grouped into a separate cluster in phylogenetic analysis. Strains with ICE VchInd5 lacking PLE4 and kappa phage also form a separate group. Conclusions: The data obtained on the presence of antiphage systems in previously imported strains of V. cholerae O1 biotype El Tor expand knowledge of their genetic organization. The study of the structure of antiphage genes of hotspot 5 of the ICE SXT element makes it possible to reveal genetic differences between closely related strains of V. cholerae O1 biotype El Tor and to determine the type of ICE SXT element.
Article
Full-text available
Aquaculture has been recognized as a hotspot for the emergence and spread of antimicrobial resistance genes (ARGs) conferring resistance to clinically-important antibiotics. This review gives insights into studies investigating the prevalence of colistin and carbapenem resistance (CCR) among Gram-negative bacilli in aquaculture. Overall, a high incidence of CCR has been reported in aquatic farms in several countries, with CCR being more prevalent among opportunistic human pathogens such as Acinetobacter nosocomialis, Shewanella algae, Photobacterium damselae, Vibrio spp., Aeromonas spp., as well as members of Enterobacteriaceae family. A high proportion of isolates in these studies exhibited wide-spectrum profiles of antimicrobial resistance, highlighting their multidrug-resistance properties (MDR). Several mobile colistin resistance genes (including, mcr-1, mcr-1.1, mcr-2, mcr-2.1, mcr-3, mcr-3.1, mcr-4.1, mcr-4.3, mcr-5.1, mcr-6.1, mcr-7.1, mcr-8.1, and mcr-10.1) and carbapenemase encoding genes (including, blaOXA-48, blaOXA-55, blaNDM, blaKPC, blaIMI, blaAIM, blaVIM, and blaIMP) have been detected in aquatic farms in different countries. The majority of these were carried on MDR Incompatibility (Inc) plasmids including IncA/C, and IncX4, which have been associated with a wide host range of different sources. Thus, there is a risk for the possible spread of resistance genes between fish, their environments, and humans. These findings highlight the need to monitor and regulate the usage of antimicrobials in aquaculture. A multisectoral and transdisciplinary (One Health) approach is urgently needed to reduce the spread of resistant bacteria and/or resistance genes originating in aquaculture and avoid their global reach.
Article
Full-text available
Gallibacterium anatis (G. anatis), a member of the Pasteurellaceae family, normally inhabits the upper respiratory and lower genital tracts of poultry. However, under certain circumstances of immunosuppression, co-infection (especially with Escherichia coli or Mycoplasma), or various stressors, G. anatis caused respiratory, reproductive, and systemic diseases. Infection with G. anatis has emerged in different countries worldwide. The bacterium affects mainly chickens; however, other species of domestic and wild birds may get infected. Horizontal, vertical, and venereal routes of G. anatis infection have been reported. The pathogenicity of G. anatis is principally related to the presence of some essential virulence factors such as Gallibacterium toxin A, fimbriae, haemagglutinin, outer membrane vesicles, capsule, biofilms, and protease. The clinical picture of G. anatis infection is mainly represented as tracheitis, oophoritis, salpingitis, and peritonitis, while other lesions may be noted in cases of concomitant infection. Control of such infection depends mainly on applying biosecurity measures and vaccination. The antimicrobial sensitivity test is necessary for the correct treatment of G. anatis. However, the development of multiple drug resistance is common. This review article sheds light on G. anatis regarding history, susceptibility, dissemination, virulence factors, pathogenesis, clinical picture, diagnosis, and control measures.
Article
Full-text available
Bacterial plasmids are fragments of extrachromosomal double-stranded DNA that can contain a variety of genes that are beneficial to the host organism, like those responsible for antimicrobial resistance. The objective of this study was to characterize a collection of 437 Salmonella enterica isolates from different animal sources for their antimicrobial resistance phenotypes and plasmid replicon types and, in some cases, by pulsed-field gel electrophoresis (PFGE) in an effort to learn more about the distribution of multidrug resistance in relation to replicon types. A PCR-based replicon typing assay consisting of three multiplex PCRs was used to detect 18 of the 26 known plasmid types in the Enterobacteriaceae based on their incompatibility (Inc) replicon types. Linkage analysis was completed with antibiograms, replicon types, serovars, and Inc A/C. Inc A/C plasmids were prevalent in multidrug-resistant isolates with the notable exception of Salmonella enterica serovar Typhimurium. Cluster analysis based on PFGE of a subset of 216 isolates showed 155 unique types, suggesting a variable population, but distinct clusters of isolates with Inc A/C plasmids were apparent. Significant linkage of serovar was also seen with Inc replicon types B/O, I1, Frep, and HI1. The present study showed that the combination of Salmonella, the Inc A/C plasmids, and multiple-drug-resistant genes is very old. Our results suggest that some strains, notably serovar Typhimurium and closely related types, may have once carried the plasmid but that the resistance genes were transferred to the chromosome with the subsequent loss of the plasmid.
Data
Full-text available
Abstract Epidemics of cholera caused by Vibrio cholerae O1 occur regularly in Bangladesh, but until lately V cholerae non-O1 has been associated only with sporadic cases of diarrhoeal disease in many parts of the world, including Bangladesh. We describe a large epidemic of cholera-like disease in Bangladesh that is due to a V cholerae non-O1. The epidemic began in December, 1992, in southern Bangladesh and spread throughout the country. By the end of March 107,297 cases of diarrhoea and 1473 deaths had been reported. The disease is indistinguishable from cholera in clinical features and response to treatment, but most of the cases are in adults, which suggests that the population has no previous immunological experience of the organism. At two centres 375 (40%) of 938 and 236 (48%) of 492 rectal swabs were positive for V cholerae non-O1, as were 5 of 54 surface water samples. 55 isolates of V cholerae non-O1 were studied in detail. They resembled El Tor vibrios in being resistant to polymyxin B and positive for agglutination of chicken erythrocytes. The strain did not belong to any of the 138 known V cholerae serogroups; so a new serogroup O139, with the suggested name Bengal, is proposed. All the isolates studied produced large amounts of an enterotoxin apparently identical to cholera toxin. This strain seems to have pandemic potential. It is important that other countries in southeast Asia are aware of the strain's potential to cause severe morbidity and mortality.
Article
Full-text available
This paper introduces a range of methods used for data gathering and the results of their analysis in the evaluation of Joint Forest Management initiatives undertaken in India. There were 200 evaluation reports collected from different states in India, of which only 99 had reported on methods used, and issues addressed, in relation to socio-economic, institutional, ecological and gender and training aspects. At the national level 17 reports are available. Many studies were conducted in states where there was support from donor agencies. There were 33 reports addressing the socio-economic issues related to JFM, while only 15 reports addressed ecological issues. There were only 4 reports that addressed training needs of different stakeholders of JFM, while 12 related to gender issues. Based on the data collected so far, it was found that there were no reports that have focused on the monitoring of JFM, and all are evaluation reports. Monitoring is needed to understand the change over time with respect to vegetation, awareness amongst the community members, participation in various JFM activities, enhanced representation of women, incentives accrued and other matters. It is also important to document how the changes have taken place over time, the factors that determine change and its impact. Many projects have conducted mid-term evaluation, in order to enhance their performance through proper corrective measures. A significant finding in this paper is that, it is not clear at this stage, how many took the mid-term evaluation reports seriously and what review process was adopted to improve their future implementation process.
Article
Full-text available
In Vibrio cholerae, the second messenger bis-(3'-5')-cyclic dimeric guanosine monophosphate (c-di-GMP) increases exopolysaccharides production and biofilm formation and decreases virulence and motility. As such, c-di-GMP is considered an important player in the transition from the host to persistence in the environment. c-di-GMP level is regulated through a complex network of more than 60 chromosomal genes encoding predicted diguanylate cyclases (DGCs) and phosphodiesterases. Herein we report the characterization of two additional DGCs, DgcK and DgcL, encoded by integrating conjugative elements (ICEs) belonging to the SXT/R391 family. SXT/R391 ICEs are self-transmissible mobile elements that are widespread among vibrios and several species of enterobacteria. We found that deletion of dgcL increases the motility of V. cholerae, that overexpression of DgcK or DgcL modulates gene expression, biofilm formation and bacterial motility, and that a single amino acid change in the active site of either enzyme abolishes these phenotypes. We also show that DgcK and DgcL are able to synthesize c-di-GMP in vitro from GTP. DgcK was found to co-purify with non-covalently bound flavin mononucleotide (FMN). DgcL's enzymatic activity was augmented upon phosphorylation of its phosphorylatable response-regulator domain suggesting that DgcL is part of a two-component signal transduction system. Interestingly, we found orthologues of dgcK and dgcL in several SXT/R391 ICEs from two species of Vibrio originating from Asia, Africa and Central America. We propose that besides conferring usual antibiotic resistances, dgcKL-bearing SXT/R391 ICEs could enhance the survival of vibrios in aquatic environments by increasing c-di-GMP level.
Article
Full-text available
Multidrug resistance (MDR) plasmids belonging to the IncA/C plasmid family are widely distributed among Salmonella and other enterobacterial isolates from agricultural sources and have, at least once, also been identified in a drug-resistant Yersinia pestis isolate (IP275) from Madagascar. Here, we present the complete plasmid sequences of the IncA/C reference plasmid pRA1 (143,963 bp), isolated in 1971 from the fish pathogen Aeromonas hydrophila, and of the cryptic IncA/C plasmid pRAx (49,763 bp), isolated from Escherichia coli transconjugant D7-3, which was obtained through pRA1 transfer in 1980. Using comparative sequence analysis of pRA1 and pRAx with recent members of the IncA/C plasmid family, we show that both plasmids provide novel insights into the evolution of the IncA/C MDR plasmid family and the minimal machinery necessary for stable IncA/C plasmid maintenance. Our results indicate that recent members of the IncA/C plasmid family evolved from a common ancestor, similar in composition to pRA1, through stepwise integration of horizontally acquired resistance gene arrays into a conserved plasmid backbone. Phylogenetic comparisons predict type IV secretion-like conjugative transfer operons encoded on the shared plasmid backbones to be closely related to a group of integrating conjugative elements, which use conjugative transfer for horizontal propagation but stably integrate into the host chromosome during vegetative growth. A hipAB toxin-antitoxin gene cluster found on pRA1, which in Escherichia coli is involved in the formation of persister cell subpopulations, suggests persistence as an early broad-spectrum antimicrobial resistance mechanism in the evolution of IncA/C resistance plasmids.
Article
Full-text available
SXT is an integrative and conjugative element (ICE) that confers resistance to multiple antibiotics upon many clinical isolates of Vibrio cholerae. In most cells, this approximately 100 Kb element is integrated into the host genome in a site-specific fashion; however, SXT can excise to form an extrachromosomal circle that is thought to be the substrate for conjugative transfer. Daughter cells lacking SXT can theoretically arise if cell division occurs prior to the element's reintegration. Even though approximately 2% of SXT-bearing cells contain the excised form of the ICE, cells that have lost the element have not been detected. Here, using a positive selection-based system, SXT loss was detected rarely at a frequency of approximately 1 x 10(-7). As expected, excision appears necessary for loss, and factors influencing the frequency of excision altered the frequency of SXT loss. We screened the entire 100 kb SXT genome and identified two genes within SXT, now designated mosA and mosT (for maintenance of SXT Antitoxin and Toxin), that promote SXT stability. These two genes, which lack similarity to any previously characterized genes, encode a novel toxin-antitoxin pair; expression of mosT greatly impaired cell growth and mosA expression ameliorated MosT toxicity. Factors that promote SXT excision upregulate mosAT expression. Thus, when the element is extrachromosomal and vulnerable to loss, SXT activates a TA module to minimize the formation of SXT-free cells.
Article
Full-text available
A conjugative plasmid, pMRV150, which mediated multiple-drug resistance (MDR) to at least six antibiotics, including ampicillin, streptomycin, gentamicin, tetracycline, chloramphenicol, and trimethoprim-sulfamethoxazole, was identified in a Vibrio cholerae O139 isolate from Hangzhou, eastern China, in 2004. According to partial pMRV150 DNA sequences covering 15 backbone regions, the plasmid is most similar to pIP1202, an IncA/C plasmid in an MDR Yersinia pestis isolate from a Madagascar bubonic plague patient, at an identity of 99.99% (22,180/22,183 nucleotides). pMRV150-like plasmids were found in only 7.69% (1/13) of the O139 isolates tested during the early period of the O139 epidemic in Hangzhou (1994, 1996, and 1997); then the frequency increased gradually from 60.00% (3/5) during 1998 and 1999 to 92.16% (47/51) during 2000 to 2006. Most (42/51) of the O139 isolates bearing pMRV150-like plasmids were resistant to five to six antibiotics, whereas the plasmid-negative isolates were resistant only to one to three antibiotics. In 12 plasmid-bearing O139 isolates tested, the pMRV150-like plasmids ranged from approximately 140 kb to 170 kb and remained at approximately 1 or 2 copies per cell. High (4.50 x 10(-2) and 3.08 x 10(-2)) and low (0.88 x 10(-8) to 3.29 x 10(-5)) plasmid transfer frequencies, as well as no plasmid transfer (under the detection limit), from these O139 isolates to the Escherichia coli recipient were observed. The emergence of pMRV150-like or pIP1202-like plasmids in many bacterial pathogens and nonpathogens occupying diverse niches with global geographical distribution indicates an increasing risk to public health worldwide. Careful tracking of these plasmids in the microbial ecosystem is warranted.
Article
Full-text available
The Bacteroides conjugal tetracycline resistance (Tcr) elements appear not to be plasmids. In many cases, resistance to erythromycin (Emr) is cotransferred with Tcr. Using a newly constructed shuttle cosmid, pNJR1, we cloned 44 to 50 kilobase pairs of a conjugal Tcr Emr element on overlapping cosmid clones. Cosmid libraries were made in Escherichia coli with DNA from the original clinical Bacteroides thetaiotaomicron DOT strain containing Tcr Emr-DOT or from a Bacteroides uniformis Tcr Emr-DOT transconjugant strain. The cosmid clones were mobilized from E. coli into B. uniformis in groups of 10 to 20 per filter mating, with selection for Tcr or Emr transconjugants. The Tcr and Emr genes were cloned both separately and together on 30-kilobase-pair fragments. Several of the Tcr clones also contained transfer genes that permitted self-transfer of the cosmid from B. uniformis donors to E. coli or B. uniformis recipients. Neither the Tcr nor the Emr gene conferred resistance on E. coli, and the transfer-proficient clones did not self-transfer out of E. coli. Southern blot analysis was used to compare DNA from independently isolated Bacteroides strains carrying conjugal Tcr or Tcr Emr elements and their respective B. uniformis transconjugants. Results of these analyses indicate that there are large regions of homology, including regions outside the Tcr and Emr genes, but that the elements are not identical. Some Tcr clones contained a region which hybridized to chromosomal DNA from the wild-type B. uniformis recipient strain that did not carry the Tcr Emr-DOT element. This region of homology appeared not to be a junction fragment. It was not required in a Bacteroides recipient for successful transfer of the Tcr Emr element. Although we are not sure we have cloned a junction fragment between the Tcr Emr-DOT element and the B. uniformis chromosome, the preliminary function and restriction map appears to be linear.
Article
The covalently closed circular form of the conjugative transposon Tn916, which acts as an intermediate in transposition, is produced by a novel type of recombination. Excision of the element pairs noncomplementary base pairs, which flank the transposon in a heteroduplex, at the joint of a circular form. By a reversal of the excision process, the base pairs from the heteroduplex are inserted into the next target. We present a detailed molecular model for the movement of conjugative transposons that involves the initial formation of staggered nicks in the "coupling regions" that flank the inserted element. The different products of excision and insertion of Tn916 can be explained by this model.