Next Article in Journal
Generation of PHB from Spent Sulfite Liquor Using Halophilic Microorganisms
Next Article in Special Issue
Staphylococcus aureus in Animals and Food: Methicillin Resistance, Prevalence and Population Structure. A Review in the African Continent
Previous Article in Journal
Does the Gut Microbiota Contribute to Obesity? Going beyond the Gut Feeling
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Tackling Drug Resistant Infection Outbreaks of Global Pandemic Escherichia coli ST131 Using Evolutionary and Epidemiological Genomics

School of Biotechnology, Faculty of Science and Health, Dublin City University, Dublin 9, Ireland
Submission received: 16 February 2015 / Revised: 28 April 2015 / Accepted: 30 April 2015 / Published: 20 May 2015
(This article belongs to the Special Issue Antibiotic Resistance Mechanisms)

Abstract

:
High-throughput molecular screening is required to investigate the origin and diffusion of antimicrobial resistance in pathogen outbreaks. The most frequent cause of human infection is Escherichia coli, which is dominated by sequence type 131 (ST131)—a set of rapidly radiating pandemic clones. The highly infectious clades of ST131 originated firstly by a mutation enhancing conjugation and adhesion. Secondly, single-nucleotide polymorphisms occurred enabling fluoroquinolone-resistance, which is near-fixed in all ST131. Thirdly, broader resistance through beta-lactamases has been gained and lost frequently, symptomatic of conflicting environmental selective effects. This flexible approach to gene exchange is worrying and supports the proposition that ST131 will develop an even wider range of plasmid and chromosomal elements promoting antimicrobial resistance. To stop ST131, deep genome sequencing is required to understand the origin, evolution and spread of antimicrobial resistance genes. Phylogenetic methods that decipher past events can predict future patterns of virulence and transmission based on genetic signatures of adaptation and gene exchange. Both the effect of partial antimicrobial exposure and cell dormancy caused by variation in gene expression may accelerate the development of resistance. High-throughput sequencing can decode measurable evolution of cell populations within patients associated with systems-wide changes in gene expression during treatments. A multi-faceted approach can enhance assessment of antimicrobial resistance in E. coli ST131 by examining transmission dynamics between hosts to achieve a goal of pre-empting resistance before it emerges by optimising antimicrobial treatment protocols.

Graphical Abstract

1. The Advent of Genome-Based Escherichia coli Monitoring

The rate of infectious disease outbreaks globally increased significantly during 1980–2013 [1,2]. Understanding the transmission dynamics of antimicrobial resistance (AMR) necessitates deep molecular screening [3,4]. The most informative tools for epidemiological investigation of AMR exploit emerging high-throughput technologies [5], of which genome sequencing is becoming standard for resolving the origins of outbreaks over timespans ranging from centuries to days [6]. Only such tools can trace the transmission of infection outbreaks: other genetic profiling methods such as multi-locus sequencing typing (MLST) of housekeeping genes or pulse-field gel electrophoresis (PFGE) are not sufficiently discriminatory [7].
Although PFGE is still used internationally for reconstruction the evolution of bacterial outbreaks, its reproducibility between labs can be inconsistent due to variability in genome-wide restriction enzyme site frequency [8]. This can be reduced with use of standardised protocols, materials, and interpretive criteria. Although PFGE has revealed different pulsotypes for Escherichia coli (such as 170 for ST131) [9], the limited genetic diversity examined by this approach indicates other methods are superior [10].
MLST has moderate power to determine sequence type (ST) given sufficient elapsed time [11] but is limited by selecting few genes from the genome [12] (Table 1). The Pasteur MLST system numbers eight genes: DNA polymerase (dinB), isocitrate dehydrogenase (icdA), p-aminobenzoate synthase (pabB), polymerase PolII (polB), proline permease (putP), tryptophan synthase subunits A and B (trpA/trpB) and beta-glucuronidase (uidA) (www.pasteur.fr/recherche/genopole/PF8/mlst/EColi.html). This differs from the Achtman scheme, which uses seven housekeeping genes: adenylate kinase (adk), fumarate hydratase (fumC), DNA gyrase subunit B (gyrB), icd, malate dehydrogenase (mdh), adenylosuccinate dehydrogenase (purA), and DNA repair enzyme RecA (recA) [13]. Over shorter timescales, MLST is limited [14]: genome sequencing can be considered simply as an extension of MLST assessing all genes and thus finer resolution, which is critical for evaluating rapid clonal expansions.
The comprehensive profiling provided by genome sequencing has enabled the production of new antimicrobials, diagnostics and surveillance systems, and is a fundamental shift in practice in infection control that enables more accurate pathogen tracing [15]. An example is the enteroaggregative ESBL-producing E. coli O104:H4 summer 2011 outbreak that infected thousands and killed many in France and Germany [16]. Its source was initially misidentified (but later clarified as fenugreek seeds). The incubation time in patients was eight days, and progression from diarrhoea to haemolytic-uremic syndrome a mere five days [17]: this indicates the short timespan required for the medical application of molecular technologies. The draft genome was sequenced within three days and a consensus genome sequence was reconstructed within two further days [18], allowing validation of discovered AMR markers within 16 days [19]. In other examples, genome sequencing successfully determined the number of origins of ST131 CMT-X-15-positive infections in a neonatal unit [20] and in enterohaemorrhagic O157:H7 isolates [21]. Similar strategies have worked in clinical settings for other bacterial species like Staphylococcus aureus and Clostridium difficile [22].
Table 1. Genes associated with virulence or ST131 typing. Genes encoding virulence factors (adhesion, capsules, siderophores, antimicrobial resistance, also known as AMR, toxins) and used to define Extraintestinal E. coli, H30-Rx ST131 and the Pasteur/Achtman multi-locus sequencing typing (MLST) schemes are highlighted.
Table 1. Genes associated with virulence or ST131 typing. Genes encoding virulence factors (adhesion, capsules, siderophores, antimicrobial resistance, also known as AMR, toxins) and used to define Extraintestinal E. coli, H30-Rx ST131 and the Pasteur/Achtman multi-locus sequencing typing (MLST) schemes are highlighted.
GeneGene ProductRole in VirulenceTyping Use
H7 fliCFlagellin variantMultiple
issIncreased serum survivalMultiple
malXPathogenicity island markerMultiple
ompTOuter membrane protease TMultiple
traTSerum resistance-associatedMultiple
uspUropathogenic-specific proteinMultiple
K1/K2/K5/K100Group 2 capsule variantsCapsule-related
kpsM IIkpsM II group 2 capsuleCapsule-relatedExtraintestinal E. coli definition
kpsMT IIIGroup 3 capsuleCapsule-relatedExtraintestinal E. coli definition
afaAfa adhesinAdhesionExtraintestinal E. coli definition
draBCDr-binding adhesinAdhesionExtraintestinal E. coli definition
fimHType 1 fimbriaeAdhesionExtraintestinal E. coli definition
papC/papE/papF/papGP fimbriae operonAdhesionExtraintestinal E. coli definition
sfa/focF1C fimbriae or SAdhesionExtraintestinal E. coli definition
F10 papAP fimbriae subunit variantAdhesion
focGF1C fimbriae adhesinAdhesion
hraHeat-resistant agglutininAdhesion
ihaAdhesion siderophoreAdhesion
tshTemperature sensitive hemagglutininAdhesion
iutAAerobactin receptorSiderophoreExtraintestinal E. coli definition
fyuAYersiniabactin receptorSiderophore
ireASiderophore receptorSiderophore
iroNSalmochelin receptorSiderophore
stbBPlasmid stabilityplasmid stability
aacAminoglycoside acetyltransferasesAMR (multiple)
aadA5Aminoglycoside 3′-adenylyltransferaseAMR (streptomycin)
acrRRepressor of acrAB genesAMR (multiple)
catB4Chloramphenicol acetyltransferaseAMR (chloramphenicol)
dfrA7Dihydrofolate reductase type VIIAMR (trimethoprim)
gyrADNA gyrase subunit AAMR (fluoroquinolone)
marRRepressor of the marRAB operonAMR (multiple)
mphAMacrolide 2′-phosphotransferase IAMR (macrolide)
parCDNA topoisomerase 4 subunit AAMR (fluoroquinolone)
rpsL105Ribosomal gene S12AMR (multiple)
sulIDihydropteroate synthaseAMR (sulfonamide)
tetATetracycline efflux pumpAMR (tetracycline)
Tn10Transposon-tetracycline resistanceAMR (tetracycline)
astAArginine succinyltransferaseToxin
cnf1Cytotoxic necrotizing factorToxin
hlyDAlpha-HemolysinToxin
picSerine proteaseToxin
satSecreted autotransporter toxinToxin
vatVacuolating toxinToxin
dinBDNA polymerasePasteur MLST system
pdbBP-aminobenzoate synthasePasteur MLST system
polBPolymerase PolIIPasteur MLST system
putPProline permeasePasteur MLST system
trpA/trpBTryptophan synthase subunits A, BPasteur MLST system
uidABeta-glucuronidasePasteur MLST system
icd/icdAIsocitrate dehydrogenasePasteur/Achtman MLST system
adkAdenylate kinaseAchtman MLST system
fumCFumarate hydrataseAchtman MLST system
gyrBDNA gyrase subunit BAchtman MLST system
mdhMalate dehydrogenaseAchtman MLST system
purAAdenylosuccinate dehydrogenaseAchtman MLST system
recAATP/GTP binding motifAchtman MLST system
Recent methodological advances have dramatically improved the accuracy and costs of high-throughput sequencing technologies, primarily for genome sequence contiguity through better library preparation and genome assembly. Genomes can be amplified from just 20+ pg DNA [23] or even single colonies [24], and up to 384 samples can be sequenced simultaneously [25]. Sequencing single bacterial cells does not require prior cultivation, and therefore could test bacteria from environmental surfaces [26]. This removes biases associated with growing cells to obtain sufficient DNA, which favours those best-suited to media. Sequence depth and repetitive sequence biases can be reduced by using microwell displacement amplification systems, which have produced assemblies of over 90% of an E. coli genome [27]. De novo assembly of E. coli (and other bacterial) genomes as a single unit using long sequence reads is now viable using single-molecule real-time sequencing [28,29]. This approach can differentiate carbapenemase-producing elements in E. coli [30], determine the precise origins of the linked mobile genetic elements (MGEs) like insertion sequences (ISs) [31], and also create gapless enterohaemorrhagic O157:H7 EDL933 genome with no uncertain mutation alleles [32]. Other strategies like nanopore sequencing have produced high-resolution E. coli genomes [33].

2. Escherichia coli ST131 Is a Major Global Health Issue

E. coli is the most frequent cause of acute bacterial infections [4], particularly blood stream infections (BSIs) and urinary tract infections (UTIs) [34]. E. coli causing UTIs and BSIs are not genetically distinct and can transfer between distinct anatomical environments [35]. Uropathogenic E. coli cause 80% of UTIs [36] and infect 20% of adult women at least once in their lifetime [37]. Approximately half of hospital-acquired (HA) and 70%–95% of community-acquired (CA) UTIs are caused by E. coli, and these are suspected to originate from intestinal colonisers [37,38]. There are multiple potential sources of E. coli UTIs and BSIs because it contaminates food [39,40,41] such as fruit [42] and meat [43], and infects companion animals [44] like cats and dogs [45].
E. coli strains are categorised in terms of their anatomical niches into three main groups [46]: the first are commensals, which typically belong to phylogenetic groups A or B1, and are not normally pathogenic when outside the intestinal tract because they lack virulence factors. The second are intestinal E. coli that are generally absent in healthy hosts, are virulent and come from a variety of phylogenetic groups. The third affect extraintestinal areas (e.g., UTIs): these are genetically distinct from intestinal and commensal E. coli, and ordinarily come from phylogenetic groups B2 or D. They may exist as asymptomatic gut commensals (see for example [47]), but can efficiently colonise regions outside the intestine.
E. coli are the most frequent extended spectrum beta-lactamase (ESBL)-producing bacterial species. Sequence type O25b:H4-B2-ST131 account for 61% of ESBL-producing E. coli in Spain [48], 41% in Israel [49], 41% in Japan [50], 47% [51] to 56% [52] in the USA, 78% in Canada [53] and 91% in Ireland [54]. ST131 is a set of pandemic clones from phylogroup B2 that the most significant threat among ESBL-producing E. coli and is common worldwide [55,56,57]. B2 is associated with extraintestinal infection, but is also present asymptomatically in humans [58,59]. There is evidence that B2 diverged early among E. coli groups and is composed of nine subclasses, including ST131 [60]. O25b ST131 has been detected in many countries, and serotypes O16:H5, O (not typeable):H4, and O157 have also been observed in Australia [61], China [62], Denmark [63], Japan [64], Spain [48], the USA [65] and the UK [55,66].
Even though it was first reported in 2002 [67], ST131 is not new. ST131 isolates taken from humans, animals and environmental locations evaluated using PFGE include samples from 1967, 1982, 1983, 1985, 1986, and 22 from 1990 to 1999 [9]. The earliest other ST131 are from Sweden and Britain and date to 1968, with others from the 1970s (7), 1980s (9) and 1990s (16) [68]—as well as one from 1985 [65].
Bacterial AMR is a major threat to public health [69]: AMR in E. coli is evident worldwide [4] and encompasses many compounds [24,70]. AMR in E. coli is becoming more frequent [71] and is driven by environmental exposure [72], such as effluent waste water [73,74] and soil treated with manure from antibiotic-treated livestock [75]. Antibiotic treatment failure jumped by ~12% in the UK from 1991 to 2012, and was more frequent against certain second-line antimicrobials (such as quinolones and cephalosporins) than in first-line ones (penicillins, macrolides, flucloxacillins) [76]. 75% of E. coli from 1997 to 2007 in Irish hospitals displayed AMR to eight or more of 16 antimicrobials [77]. ESBL-producing E. coli are resistant to cephalosporins and fluoroquinolones, leaving carbapenems as the sole last-resort antimicrobial [4]. However, multi-AMR ESBL-producing ST131 with carbapenem-resistance are now frequent [78] and in Algeria [79], China [80] and Ireland [81].
ST131 represents a universal problem whose evolutionary epidemiology needs deeper study. Not enough is known about how it evolves and spreads at a local scale [82,83]. As a fraction of total E. coli infections, ST131 infection rates are proportionally higher in long-term care facilities (LTCFs) (76%) than hospitals (49%) and the public (15%) [84]. This was originally driven by the acquisition of fluoroquinolone-resistance sometime prior to 2000, and has subsequently been amplified by the gain of CTX-M (cefotaximase) elements, which are beta-lactamases (bla) that hydrolyse beta-lactam rings [85]. UTIs caused by blaCTX-M-positive ST131 are now widespread in LTCFs [86], hospitals [87] and the community [88,89,90], to the point where all ESBL-producing E. coli in certain LTCFs are ST131 [91]. Notably, in the latter study 93% of these belonged to just a single PFGE pulsotype. Current hypotheses on the sources of E. coli have identified LTCFs as a closed microenvironment in which AMR bacteria evolve [92], and subsequently they diffuse into the wider community, spreading AMR alleles [93].

3. The Genomic Landscape of Antimicrobial Resistance in E. coli

Commonly E. coli genomes are separated into the largely chromosomal highly-conserved core genome and the accessory genomes. Notably, the core genome varies between studies [94] because it is composed of genes present in all samples for that analysis [95]. The accessory genome typically has much lower sequence conservation and encodes non-essential traits associated with virulence and antimicrobial metabolism [96]. It represents a major component of microbial variation and includes MGEs like plasmids, transposable elements (TEs), pathogenicity islands, and prophages.
E. coli undergo extensive horizontal gene transfer (HGT) [97]: HGT accounts for ~31% of genome-wide variation in ST131 [98]. Genes arising by HGT are frequently associated with duplications, which can permit new functions among the two gene copies [99]. HGT typically occurs through three processes: the primary one is transduction via bacteriophages after cell lysis and invasion via homologous recombination. ST131 acquire AMR through extrachromosomal MGEs from bacteriophages [100], or the integration of prophage DNA into chromosomes [101]. Prophage elements account for over 35% of E. coli CDS variation [102], and encode genes linked to virulence [103], growth during nutrient deficit [104], transcriptional regulation [105], AMR [106] and biofilms [107]. HGT segment length is a function of phage type during generalised transduction. Other options for HGT are the transformation of local DNA from dead cells, or conjugation mediated by plasmids, TEs, integrons or integrating conjugative elements (ICEs). Homologous recombination between ICEs increase their diversity [108]. TEs such as ISs have a high frequency of integration and excision, and mediate both AMR gene composition and their expression rates. Although HGT is generally a function of sequence homology, TEs do not need high levels to insert.
More detailed knowledge of AMR mutations and gene expression rates can help optimise antimicrobial treatments [109,110]. A large number of ST131 genomes have been published. The first was NA114 from an Indian UTI, encompassing a 4936 Kb genome of which 88.4% was coding sequence (CDS) [111]. In comparison, genome JJ1886 from a UTI was 5,130 Kb and genetically more representative of the super-spreading subgroups within ST131 [112]. Notably, JJ1886 contained a chromosomal CTX-M-15 element due to a partial insertion of a Tn3 element into a lambda-like prophage locus. The third was isolate EC958 from the UK was also blaCTX-M-15-positive and provided a 5110 Kb genome somewhat distinct from NA114 and JJ1886 [113], perhaps due to its earlier isolation date. The fourth was carbapenem-resistant H30-Rx sample MNCRE44 from the USA in 2012: it was assembled using long-read sequencing technology as a 5011 Kb genome with six plasmids, two of which encoded multiple beta-lactamase and other AMR genes [114]. It is important to note that the number of genes, tRNAs and rRNAs in NA114 (4875; 67; 3), JJ1886 (5086; 88; 22), EC958 (4982; 89; 7) and MNCRE44 (5407; 86; 22) likely differ due to differences in sequence library preparation and genome assembly: not all genomes are created equal.

4. Antimicrobial Resistance Genetically Defines ST131

The most closely related clonal complexes (ST1876 and ST95) provide a genetic definition for ST131 (Figure 1). ST131 possesses a variety of genes associated with virulence [48,115] and pathogenesis [116] (Table 1). A key one is the fimH gene encoding a type 1 fimbrial adhesion protein expressed on the cell membrane. ST131 encompasses firstly serotype O16:H5 with a fimH41 allele: this is equivalent to Clade A from Petty et al [117]. ST131 secondly includes serotypes O25b:H4, some of which possess fimH22 allele, corresponding to Clade B [117]. Genomes from 1967, 1983 (avian infection), 1985, 1990 (avian), 1992, 1995 (chicken) and 1997 were all fluoroquinolone-sensitive fimH22 (Clade B) ST131 [98]. Genome sequencing and the Pasteur MLST system can distinguish O16:H5 from O25b:H4, but not the Achtman MLST scheme.
Figure 1. ST131 genetic groups and acquisition of drug-resistance changes. This cladogram shows ST1876 and ST95 as the most closely related clonal complexes to ST131 (with published genomes). H41 is in serotype O16:H5 and is Clade A (H41) [117]. H22 is a set of O25b:H4 ST131 called Clade B [117]. H30 is a subgroup of O25b:H4 ST131 defined by their acquisition of the fimH30 allele (dark green) [98]: this is Clade C. H30-R is a subset of H30 that are nearly always fluoroquinolone-resistant (emerald green) due to alleles 1AB in gyrA and 1aAB in parC. Within this group, H30-Rx have blaCTX-M-15 (bright green) [98] but not always [117]. This diagram does not mean that H30, H30-R or H30-Rx are monophyletic [68].
Figure 1. ST131 genetic groups and acquisition of drug-resistance changes. This cladogram shows ST1876 and ST95 as the most closely related clonal complexes to ST131 (with published genomes). H41 is in serotype O16:H5 and is Clade A (H41) [117]. H22 is a set of O25b:H4 ST131 called Clade B [117]. H30 is a subgroup of O25b:H4 ST131 defined by their acquisition of the fimH30 allele (dark green) [98]: this is Clade C. H30-R is a subset of H30 that are nearly always fluoroquinolone-resistant (emerald green) due to alleles 1AB in gyrA and 1aAB in parC. Within this group, H30-Rx have blaCTX-M-15 (bright green) [98] but not always [117]. This diagram does not mean that H30, H30-R or H30-Rx are monophyletic [68].
Microorganisms 03 00236 g001
The remaining ST131 belong to a group named H30 (or Clade C [117]) defined by their acquisition of a fimH30 allele: the earliest date of appearance of H30 is at least 1997 [118]. Fluoroquinolone-sensitive ST131 was common prior to 2000, but is now rare. This is due mutations in two genes: one encodes DNA gyrase subunit A (gyrA) with mutation D87N (from 2000) on a background of S83L (from 1997) [118]. DNA gyrases are the primary targets of quinolones. Notably, D87N-S83L (known as 1AB) is fluoroquinolone-resistant whereas S83L alone (1A) is sensitive. The other gene encodes DNA topoisomerase 4 subunit A (parC) and mutated sometime before 2000 to fluoroquinolone-resistant allele 1aAB (S80I and E84V). Nested within H30 is H30-R, which accounts for 58% of ST131 and is becoming more prevalent [117]. H30-R is nearly always fluoroquinolone-resistant—but this can be lost [98].
A subgroup of H30-R named H30-Rx is a recent radiation dating to at least 2002 and is nearly always blaCTX-M-positive [98,117,118,119,120]. Just three genome-wide SNPs distinguish H30-Rx from other H30-R. It is the main driver of increasing rates of blaCTX-M-15, which is the most frequent H30-Rx CTX-M element type and originated from Kluyvera in the early 2000s [121], such that 49% of E. coli LTCF samples from 2004 to 2006 were blaCTX-M-15-positive ST131 [122]. Other older fluoroquinolone-resistant blaCTX-M-15-positive H30-Rx with published genomes predating 2006 are EC958 (March 2005) [113], several from 2000 to 2005 isolated in Canada [117], and from 2002 again in Canada, but also 2003 in Korea and 2005 in Portugal [98]. Such isolates could be essential for determining the origins of H30-Rx. BlaCTX-M-15 is also found less commonly in ST38, ST405 and ST648 [120].

5. The Key Antimicrobial Resistance Elements in ST131

Numerous core genome mutations are implicated in ST131’s AMR [123]. These include regulators of drug efflux, repressor of the marRAB (multiple antibiotic resistance) operon (marR), and repressor of acrAB (achromobactin outer membrane receptor) genes (acrR), that respond to drugs associated with high fitness costs—for example AcrAB helps non-aminoglycoside drug efflux from the cell. Others are L42R in ribosomal gene S12 (rpsL105); a transposon mediating resistance to tetracycline (Tn10); parC1aAB to fluoroquinolones; ciprofloxacin and nalidixate thanks to changes at gyrA (most frequently allele 1AB) [124]. Chromosomal ampC-like cephalosporinases have been found in E. coli but are rare [49].
ST131’s virulence genes are used to categorise it into distinct virotypes [68]. The aerobactin receptor (iutA), group 2/3 capsule synthesis (kpsM II/III), Afa adhesion (afa), Dr-binding adhesin (draBC), P fimbriae operon genes, F1C fimbriae or S (sfa/foc) and fimH genes are part of a system for defining extraintestinal E. coli [125,126]. ST131 virulence factor genes are typically acquired or lost rapidly [115,117], and encode toxins, adhesions, lipopolysaccharides, polysaccharide capsules, proteases, and invasins (Table 1). Other genes universal in ST131 include secreted autotransporter toxin (sat), yersiniabactin receptor (fyuA), uropathogen-specific protein (usp), pathogenicity island marker (malX), adhesion receptor (iha), outer membrane receptor (ompT), aerobactin (iucD) and serum resistance associated (traT) [127].
ST131 has different plasmids from multiple incompatibility groups (Inc): IncF, IncI1, IncN and IncA/C have been detected so far [128]. IncF are limited to Enterobacteriaceae and are the most frequent type in ST131. IncF plasmids are typically > 100 kb long, are maintained at a low copy number, and typically possess multiple AMR, beta-lactamase, virulence, toxin and antitoxin elements. Conjugation of IncFII plasmids has facilitated the spread of blaCTX-M elements, especially blaCTX-M-15 (such as blaCTX-M-15-positive pEC958 in strain EC958). In one study, a majority of isolated ST131 had IncFII pEK516 plasmids, which has 103 genes including blaCTX-M-15, blaOXA-1, blaTEM-1, chloramphenicol acetyltransferase (catB4), tetracycline efflux pump (tetA) and plasmid stability (stbB) for stable plasmid inheritance [129]. This plasmid (pEK516) also contained aminoglycoside acetyltransferase genes (aac), allowing resistance to aminoglycosides and ciprofloxacin (both via aac(6’)-Ib-cr) as well as gentamicin (via aac(3)-II).
Plasmids contain genes encoding various beta-lactamases that are grouped into Ambler classes A (blaCTX-M, blaTEM, blaKPC, blaSHV); B (blaVIM, blaNDM, blaIMP); C (blaCMY, ampC-related like blaDHA); D (blaOXA). Classes A/C/D use serine for beta-lactam hydrolysis, whereas B are metallo-beta-lactamases that use divalent zinc ions [130]. BlaCTX-M in class A are the most frequent ESBL in E. coli, followed next by TEM beta-lactamases. The latter are exchanged extensively between species: for example, TEM-52C in E. coli is mediated by an IncI1-g plasmid on a Tn3 transposon [131] that been found in Salmonella. SHV beta-lactamases are less common in E. coli, but are abundant in K. pneumoniae, from which ST131 has acquired other beta-lactamases [132].
AMR driven by KPC-2 (K. pneumoniae carbapenemase, Class A) has spread from K. pneumoniae to E. coli ST131 [133,134]. Other carbapenemases expressed by ST131 include class B metallo-beta-lactamases like Verona integron-encoded metallo-beta-lactamase (VIM), which has been found in Italy [135], and blaIMP-8 in Taiwan [136]. Class B NDM-1 (New Delhi metallo-beta-lactamase)-positive ST131 has been found in India [137] where it may have originated by in vivo conjugation with K. pneumoniae in 2009 or earlier [132]. BlaNDM-1-positive ST131 is now common in northern India (6% in Varanasi [138]), southern India (7% in Mumbai [139]) and northern Pakistan (15% [140]). BlaNDM-1 has been found on fast-spreading IncFII plasmids, along with Class D carbapenemase blaOXA-1 (oxacillinases) [141]. Lastly, class C includes plasmid-mediated ampC beta-lactamases blaDHA-1, which probably was obtained by an ancestor of ST131 from Salmonella enterica [142], and blaCMY-4 [48] and blaCMY-2 [49]—the latter is a cephamycinase originally transferred from Citrobacter.

6. How Did ST131 Adapt to Be so Successful?

Bacterial infectious disease is driven in part by rapid mutation rates (~10-5 per genome per generation in E. coli) [143], short generation times [144], and non-lethal antimicrobial doses due to non-adherence [145]. Remarkably, AMR may have no fitness cost and may even increase fitness in E. coli [146,147]. ST131 is highly resistant without any fitness cost [128], even more so in CA compared to HA ST131 isolates [148]. As outlined, the three key changes define H30-Rx ST131: the acquisition of fimH30 [117], fluoroquinolone-resistance mutations and beta-lactamase element gains [98,117]. In contrast to the single acquisitions of fimH30 and fluoroquinolone-resistance, the type of beta-lactamase varies, and they are lost and recovered frequently. Consequently, antimicrobial treatments have inadvertently created the H30-Rx superbug, and associated fitness costs may be linked with beta-lactamases, but less so for fimH30, gyrA1AB and parC1aAB. As acknowledged above, there are rare cases of fimH30 loss and mutation of gyrA/parC within H30-Rx.

6.1. Extensive Recombination and Horizontal Gene Transfer

Extraintestinal E. coli have higher rates of homologous recombination than commensals: this is positively correlated with the number of virulence factors and may assist with replication of segments arising by HGT [149]. ST131 is restricted largely to UTIs and BSIs [150], so this extensive recombination may have originated to cope with these local environments [149]. Yet in spite of this pervasive recombination and HGT, DNA exchange with other bacterial species is rare because of this restriction to UTIs and BSIs, affecting just ~0.4% of the core genome conserved across the E. coli genus [151]. UTI-causing ESBL-positive ST131 acquire resistance through HGT events with other ST131 colonising the same individual, but rarely from other E. coli phylogroups [55]. ST131 in poultry retains a distinct set of blaCTX-M and blaTEM-52 elements despite high overall genetic similarity [152,153]. Naively, this suggests previous HGT with Kluyvera [119] and K. pneumoniae [132] is rare. However, examples of plasmids derived from K. pneumoniae in ST131 [154] is consistent with a proposal of alarming potential genetic flexibility in ST131.

6.2. Maintenance of a Broad Resistome

ST131 regulates the activity of its resistome: this is its entire complement of its AMR genes. There is evidence of functional robustness and redundancy in the resistance mechanisms of E. coli’s “proto-resistome” comprising penicillin-binding proteins, cell wall modifying enzymes and cell division genes [155]. Expression of 61 genes is associated with small increases in tolerance to 86 antimicrobial-related compounds [156]. Many potential AMR genes remain unknown in E. coli [157], even though their activities vary even within putatively clonal cell populations [158] and the mechanism of action of many antimicrobials remains unclear [159]. A sequencing approach could assess this in a comprehensive manner.

6.3. Regulatory Fine-Tuning of Gene Expression

Regulatory sequences alter AMR through gene expression levels [160] including bursting [161] and transitions between pathogenic and non-pathogenic states [127]. Promoters evolved to alter gene expression rates in response to rapidly changing environmental conditions [162], which can affect AMR phenotypes by up to 106-fold [163]. Promoter gene expression regulation is higher at non-essential genes with lower sequence conservation, and their activity rates are more variable than those of essential genes [164]. During multi-drug exposure, 30% of E. coli expression variation is attributed to promoter mutations [165]—for example, a mutation T32A at a beta-lactamase ampC gene promoter elevates expression rates genome-wide [166].

6.4. Cross-Antimicrobial Resistance and Compensatory Mutations

ST131’s success may be enhanced by synergy between different antimicrobials [167]—however, antagonism is frequent, even to the point that a Harvey Effect of simultaneous synergy and antagonism is possible. Some compensatory mutations are specific to individual antimicrobials but others promote cross-AMR simultaneously [168]. Feedback-based cross-AMR in response to combination treatments is a function of the number of unlinked (positive) gene regulatory networks [169]. AMR is reduced by negative epistasis during cyclical treatment, which decreases resistance to both drugs more effectively than to single drugs—nearly as well as dual drug therapies [170]. Certain compensatory mutations alter AMR without being clearly associated with the phenotype—for instance, blaCTX-M-linked carbapenem-resistance requires water channel (porin) gene loss to mitigate fitness costs [171]. Additionally, structural rearrangements at ISs are associated with rapid increases in fitness during long-term E. coli evolution [172].

6.5. Resistance through Cell Growth Arrest

Antimicrobials limit E. coli growth [173], which is controlled by gene expression [174]. Growth depends on environmental conditions and is controlled by the concentration of transcription-associated proteins [175]. Even without a specific stress, cells spontaneously enter and exit a quiescent non-growth state during both log and stationary growth phases [176]. Lowered expression is more strongly associated with non-essential genes like toxins [164]. There are (so far) 22 known toxin-antitoxin systems in E. coli K12 that can be recruited during environmental stress to regulate growth, affecting biofilm formation [177]. For example, in the hipBA toxin-antitoxin system, antitoxin HipB represses serine-threonine-protein kinase HipA but has a short half-life. When HipB levels fall, HipA phosphorylates a glutamyl tRNA synthetase to mimic starvation [178], reducing translation and slowing growth, resulting in a non-growing phenotype [179].

7. Stopping ST131

7.1. Evolutionary Phylogenomics and Fitness Measurement

Pathogens resist antimicrobials by evolving: consequently, analysis of evolution is needed [180] which should operate on a genome-wide basis for outbreaks like ST131 to more finely resolve of evolutionary patterns at all potential AMR factors [181]. A direct application of this would to investigate the evolution of fluoroquinolone-resistance, because D87G (rather than D87N in ST131) in gyrA in Salmonella promotes sensitivity to other non-quinolone compounds, perhaps due to DNA coiling and this gene expression effects [182]. Genome-based evolutionary analysis and real-time diagnostic evaluation is now as cheap and time-effective as conventional approaches [183] and informative bioinformatics strategies underpin improvements in clinical microbiology [184].
Variation arising by recombination between strains can be deduced from the genome-wide DNA mutation density distribution [181]: signatures of HGT have much higher polymorphism rates, reflecting the mutational time spent in a different species. For H30-Rx, this includes MGEs, integrated phage DNA and capsule-related genes [185]. Comparing the most likely genealogy at a given panel of SNPs (haplotype) to the mean signal across the genome can identify ancestral and recent recombination between and within subpopulations [186] and determine the phylogeographic origin of specific recombination events in a coalescent framework [187]. Phylogenetic and genealogical histories at regions arising via HGT deviate substantially from the mean genome-wide phylogenetic reconstruction [188,189].
New beneficial variants increase in frequency faster than average [190] and retain a more recent time of origin [191]. Phylogenetic branching topology can measure relative fitness assuming a persistent source of antimicrobial pressure causing mutations with small fitness effects [192]. The latter is valid for E. coli because many genes regulate AMR and virulence [156,193]. No variation in relative fitness would mean low variability in the rates of coalescence of descendent nodes compared with ancestral ones. In contrast, a high difference in relative fitness would mean highly fit ancestors will produce many more descendent nodes. This would appear as a radiating population: such as H30 within ST131, H30-R within H30, and H30-Rx with H30-R. So what will follow H30-Rx?

7.2. Modelling Historical and Future Epidemiology

Systematic monitoring of infection and transmission in human populations can be used to infer sources of ST131, including from potential reservoirs like livestock and intestinal colonisers [61]. Current hypotheses pinpoint the Indian subcontinent as a possible original reservoir for ST131. During the discovery of ST131 in 2008, there was a high rate of recent travel to this region in those with H30-Rx ST131 infection in New Zealand [194] and Canada [195,196]. Whether this was caused by both a single H30-Rx clone and through plasmid-mediated HGT is unclear [197] due to the limited sampling of the Indian subcontinent prior to 2002. Phylogenomic reconstruction of the evolutionary relationships would address this puzzle but requires extensive geographic sampling including samples during the acquisition of fimH30 in the 1990s and fluoroquinolone-resistance sometime before 2002.
Genome sequencing can help address the rise of AMR [38] by predicting future resistance. The most likely future dominant subtype of a pathogen is typically the most fit one [198]: this can be applied to ST131 based on de novo and known mutations at key AMR determinants using evolutionary methods developed for viruses [199]. The relative rate of allele frequency change reflects its association with antimicrobial selection pressure [200], and can be measured from time-series genome data [201] to distinguish this effect from random drift [202]. Such sampling over extended periods allows the identification of genes implicated in adaptation based on the mutation’s age, functional effects, and ancestral phylogenetic position [203].
Phylodynamic models determine future evolutionary and epidemiological patterns based on phylogenetic tree shape [204]. Such schemes have revealed missed events during transmission of tuberculosis [205], pin-pointed an asymptomatic spreader of methicillin-resistant S. aureus between babies [206], assessed the potential for multiple simultaneous outbreaks to originate from the same common ancestor [207], and examined evidence for multiple origins of infection [208]. Although such methods have not yet been applied to E. coli, they can enhance our understanding of the inter-host spread of AMR. For instance, gyrA1AB E. coli with a marR gene deletion are much more ciprofloxacin tolerant if they express a plasmid qnrS1 (quinolone resistance) gene as well in vitro [209].

7.3. Assessing Cell Growth Arrest

Stochastic phenotype switching is common to all domains of life, is accentuated by antimicrobial stress, and in bacteria is observed as changes in gene expression. This growth-dormancy bistability is an evolutionary bet-hedging strategy that results in a mix of cells with varying gene expression levels [210]. It arises due to gene regulatory network structure, leading to rare but occasional switches from the growing wild-type (WT) state to a non-growing dormant one. Cell-cell interactions may affect variation in gene expression rates [211], allowing this randomness to adaptively drive regulatory changes.
Cell dormancy causes large shifts in metabolic gene expression [212]. This contrasts with ribosomal gene expression that has buffering mechanisms [174,213] and so scales linearly with cell growth [214]. E. coli alter their carbon metabolism rapidly [215] by co-regulating enzymes to cope with rapidly changing environmental conditions [216]. However, metabolic flux is regulated extensively [217,218] so differing concentrations can produce the same net metabolic effects [219,220] or affect other pathway components [221]. E. coli optimise gene expression by moderating ribosome production [222]: this means shifts in metabolic gene expression can be compared to those for ribosomal RNA, which should match the antimicrobial toxicity and bacterial AMR level. AMR-driven changes in gene expression vary non-monotonically with dose and are a function of the complexity of the mechanism: antimicrobials inactivating targets may elicit higher target expression, whereas drugs causing gain-of-function changes may cause mutation [110]. This means the relative rates of metabolic and ribosomal gene activity provide sufficient information to deduce AMR levels and types.

7.4. Avoiding Partial Treatments and Cell Population Heterogeneity

Suboptimal treatment regimens [34], poor compliance [223] and drug pharmacokinetics [224] lead to spatial structure in antimicrobial concentrations. This is frequent in human and animals both at an individual level [225] and in groups of treated and untreated patients [226]. At low antimicrobial doses (such as a minimal selective concentration [227]), WT and resistant cells grow equally well, whereas only resistant cells would survive when exposed to dose greater than the minimal inhibitory concentration (MIC). Partial treatment regimens accelerate AMR for mechanisms requiring numerous mutational steps with small fitness costs by offering a sub-MIC sanctuary where resistant cells can evolve further despite transiently lowered fitness [226]. However, partial regimens can slow AMR if it requires few mutational steps or substantial temporary falls in fitness [228]. In addition, the mutational time required for resistance emergence may be a function of cell density [229].
Spatial or temporal variation in treatment application allows WT cells more time to accrue new mutations, causing faster AMR evolution [230]. These genetically resistant cells are known as type 1 persisters to differentiate them from type 2 persisters that are resistant due to gene expression control causing non-growth [231]. Although E. coli type 2 persister gene activity depends on adjacent cells [165], they revert to growth after dormancy ends, and attain exponential growth rates once resistant mutations occur when all WT cells have died [232]. These type 2 persisters may require a genetic predisposition [231] and initially comprise just 0.001% of cells [233].

7.5. Dissecting Measurably Evolving Infections

A single ST131 infection may become a genetically diverse cell community descended from a recent common ancestor driven by antimicrobial selective pressure. Multiple populations could exist: for example S. aureus measurably evolves (30 SNPs and 4 indels in 16 months) to magnify virulence [234]. Furthermore, distinct Burkholderia dolosa lineages co-exist within individual cystic fibrosis patients for years and have undergone extensive clonal interference during exposure to antimicrobials [235]. Soft selective sweeps during host adaptation of Pseudomonas aeruginosa are prevalent, indicating different loss-of-function mutations may target the same gene within a heterogeneous cell community [236].
Consequently, deep genome sequencing is required to grapple with mutations in a cell population [237]. Such population mixes could be described using birth-death [181], logistic growth [238] and two-level population models within a single infection [239] or a structured environment [240] to optimise antibiotic treatment protocols [241]. However, there are caveats: high-dose antimicrobial treatments may give rise to a single dominant variant within the host [110], rare variants may become more advantageous at differing rates as the host microenvironment changes [190], and mutants may be pre-existing rather than de novo [124].

8. Conclusion: Future Avenues for ST131 Infection Genomics

Genome sequencing can predict virulence, toxicity and AMR phenotypes in E. coli ST131 [242]. It is a pivotal tool for infection control because it facilitates the decoding of molecular mechanisms of treatment resistance [243] and can dissect outbreaks of monomorphic bacteria [244]. It can decipher transmission routes and inter-host contact [181], and evaluate genetic diversity within hosts over time [245]. This evolutionary approach should become the basis for analysing E. coli ST131 outbreaks. A closing point is that there are several other possibilities for genome-based AMR control not discussed here: E. coli post-transcriptional processing disturbs expected correlations between mRNA and protein levels [246]—this includes AMR-driven regulatory changes by small RNAs [247]. Additionally, epigenetic heterogeneity within E. coli cell populations screened using long-read sequencing methylation data [248] is associated with extensive regulatory control of phenotypes [249]. Finally, activity changes in a small set of genes can predict AMR, such that integration of genome-wide and transcriptome profiling is an effective approach to understand AMR levels [250].

Acknowledgments

I thank Cathal Seoighe, Martin Cormican, Dearbhàile Morris and Catherine Ludden (all NUI Galway) for discussions. This work was supported by the National University of Ireland Galway (Ireland) and USA National Science Foundation (Grant No. NSF PHY11-25915).

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Centers for Disease Control and Prevention. In Antibiotic resistance threats in the United States, 2013; US Department of Health and Human Services, 2013. Available online: http://www.cdc.gov/drugresistance/threat-report-2013/pdf/ar-threats-2013–508.pdf (accessed on 9 September 2014).
  2. Smith, K.F.; Goldberg, M.; Rosenthal, S.; Carlson, L.; Chen, J.; Chen, C.; Ramachandran, S. Global rise in human infectious disease outbreaks. J. R Soc. Interf. 2014, 11. [Google Scholar] [CrossRef]
  3. WHO. The Evolving Threat of Antimicrobial Resistance—Options for Action. Available online: http://whqlibdoc.who.int/publications/2012/9789241503181_eng.pdf (accessed on 29 August 2014).
  4. WHO. Antimicrobial Resistance: Global Report on Surveillance 2014. Available online: http://apps.who.int/iris/bitstream/10665/112642/1/9789241564748_eng.pdf (accessed on 2 September 2014).
  5. Reuter, S.; Ellington, M.J.; Cartwright, E.J.; Köser, C.U.; Török, M.E.; Gouliouris, T.; Harris, S.R.; Brown, N.M.; Holden, M.T.; Quail, M.; et al. Rapid bacterial whole-genome sequencing to enhance diagnostic and public health microbiology. JAMA Intern. Med. 2013, 173, 1397–1404. [Google Scholar] [CrossRef] [PubMed]
  6. Achtman, M. Insights from genomic comparisons of genetically monomorphic bacterial pathogens. Philos. Trans. R. Soc. Lond. B Biol. Sci. 2012, 367, 860–867. [Google Scholar] [CrossRef] [PubMed]
  7. Palmer, A.C.; Kishony, R. Understanding, predicting and manipulating the genotypic evolution of antibiotic resistance. Nat. Rev. Genet 2013, 14, 243–248. [Google Scholar] [CrossRef] [PubMed]
  8. Noller, A.C.; McEllistrem, M.C.; Pacheco, A.G.; Boxrud, D.J.; Harrison, L.H. Multilocus variable-number tandem repeat analysis distinguishes outbreak and sporadic Escherichia coli O157:H7 isolates. J. Clin. Microbiol. 2003, 41, 5389–5397. [Google Scholar] [CrossRef] [PubMed]
  9. Johnson, J.R.; Nicolas-Chanoine, M.H.; Debroy, C.; Castanheira, M.; Robicsek, A.; Hansen, G.; Weissman, S.; Urban, C.; Platell, J.; Trott, D.; et al. MASTER Investigators. Comparison of Escherichia coli ST131 pulsotypes, by epidemiologic traits, 1967–2009. Emerg. Infect. Dis. 2012, 18, 598–607. [Google Scholar] [CrossRef] [PubMed]
  10. Mathers, A.J.; Peirano, G.; Pitout, J.D. Escherichia coli ST131: The Quintessential Example of an International Multiresistant High-Risk Clone. Adv. Appl. Microbiol. 2015, 90, 109–154. [Google Scholar] [PubMed]
  11. Sullivan, C.B.; Diggle, M.A.; Clarke, S.C. Multilocus sequence typing: Data analysis in clinical microbiology and public health. Mol. Biotechnol. 2005, 29, 245–254. [Google Scholar] [CrossRef] [PubMed]
  12. Coughlan, S.; Barreira, S.; Seoighe, C.; Downing, T. Chapter 3: Genome-wide variant discovery using sequence assembly, mapping and population-wide analysis. In Bioinformatics and Data Analysis in Microbiology, 1st ed.; Bishop, O.T., Ed.; Caister Academic Press: Norfolk, UK, 2014; pp. 51–80. [Google Scholar]
  13. Escherichia coli MLST Database. Available online: http://mlst.warwick.ac.uk/mlst/dbs/Ecoli (accessed on 23 August 2014).
  14. Bryant, J.; Chewapreecha, C.; Bentley, S.D. Developing insights into the mechanisms of evolution of bacterial pathogens from whole-genome sequences. Future Microbiol. 2012, 7, 1283–1296. [Google Scholar] [CrossRef] [PubMed]
  15. Aarestrup, F.M.; Brown, E.W.; Detter, C.; Gerner-Smidt, P.; Gilmour, M.W.; Harmsen, D.; Hendriksen, R.S.; Hewson, R.; Heymann, D.L.; Johansson, K.; et al. Integrating genome-based informatics to modernize global disease monitoring, information sharing, and response. Emerg. Infect. Dis. 2012, 18. [Google Scholar] [CrossRef]
  16. Gault, G.; Weill, F.X.; Mariani-Kurkdjian, P.; Jourdan-da Silva, N.; King, L.; Aldabe, B.; Charron, M.; Ong, N.; Castor, C.; Mace, M.; et al. Outbreak of haemolytic uraemic syndrome and bloody diarrhoea due to Escherichia coli O104:H4, south-west France, June 2011. Euro Surveill. 2011, 16. pii: 19905. [Google Scholar] [PubMed]
  17. Frank, C.; Werber, D.; Cramer, J.P.; Askar, M.; Faber, M.; an der Heiden, M.; Bernard, H.; Fruth, A.; Prager, R.; Spode, A.; et al. Epidemic profile of Shiga-toxin-producing Escherichia coli O104:H4 outbreak in Germany. N. Engl. J. Med. 2011, 365, 1771–1780. [Google Scholar] [CrossRef] [PubMed]
  18. Mellmann, A.; Harmsen, D.; Cummings, C.A.; Zentz, E.B.; Leopold, S.R.; Rico, A.; Prior, K.; Szczepanowski, R.; Ji, Y.; Zhang, W.; et al. Prospective genomic characterization of the German enterohemorrhagic Escherichia coli O104:H4 outbreak by rapid next generation sequencing technology. PLoS ONE 2011, 6, e22751. [Google Scholar] [CrossRef] [PubMed]
  19. Rasko, D.A.; Webster, D.R.; Sahl, J.W.; Bashir, A.; Boisen, N.; Scheutz, F.; Paxinos, E.E.; Sebra, R.; Chin, C.S.; Iliopoulos, D.; et al. Origins of the E. coli strain causing an outbreak of hemolytic-uremic syndrome in Germany. N. Engl. J. Med. 2011, 365, 709–717. [Google Scholar] [CrossRef] [PubMed]
  20. Sherry, N.L.; Porter, J.L.; Seemann, T.; Watkins, A.; Stinear, T.P.; Howden, B.P. Outbreak investigation using high-throughput genome sequencing within a diagnostic microbiology laboratory. J. Clin. Microbiol. 2013, 51, 1396–1401. [Google Scholar] [CrossRef] [PubMed]
  21. Underwood, A.P.; Dallman, T.; Thomson, N.R.; Williams, M.; Harker, K.; Perry, N.; Adak, B.; Willshaw, G.; Cheasty, T.; Green, J.; et al. Public health value of next-generation DNA sequencing of enterohemorrhagic Escherichia coli isolates from an outbreak. J. Clin. Microbiol. 2013, 51, 232–237. [Google Scholar] [CrossRef] [PubMed]
  22. Eyre, D.W.; Golubchik, T.; Gordon, N.C.; Bowden, R.; Piazza, P.; Batty, E.M.; Ip, C.L.; Wilson, D.J.; Didelot, X.; O’Connor, L.; et al. A pilot study of rapid benchtop sequencing of Staphylococcus aureus and Clostridium difficile for outbreak detection and surveillance. BMJ Open 2012, 2, e001124. [Google Scholar] [CrossRef] [PubMed]
  23. Parkinson, N.J.; Maslau, S.; Ferneyhough, B.; Zhang, G.; Gregory, L.; Buck, D.; Ragoussis, J.; Ponting, C.P.; Fischer, M.D. Preparation of high-quality next-generation sequencing libraries from picogram quantities of target DNA. Genome Res. 2012, 22, 125–133. [Google Scholar] [CrossRef] [PubMed]
  24. Koser, C.U.; Fraser, L.J.; Ioannou, A.; Becq, J.; Ellington, M.J.; Holden, M.T.; Reuter, S.; Török, M.E.; Bentley, S.D.; Parkhill, J.; et al. Rapid single-colony whole-genome sequencing of bacterial pathogens. J. Antimicrob. Chemother. 2014, 69, 1275–1281. [Google Scholar] [CrossRef] [PubMed]
  25. Quail, M.A.; Smith, M.; Jackson, D.; Leonard, S.; Skelly, T.; Swerdlow, H.P.; Gu, Y.; Ellis, P. SASI-Seq: Sample assurance Spike-Ins, and highly differentiating 384 barcoding for Illumina sequencing. BMC Genomics 2014, 15. [Google Scholar] [CrossRef] [PubMed]
  26. Lasken, R.S.; McLean, J.S. Recent advances in genomic DNA sequencing of microbial species from single cells. Nat. Rev. Genet 2014, 15, 577–584. [Google Scholar] [CrossRef] [PubMed]
  27. Gole, J.; Gore, A.; Richards, A.; Chiu, Y.J.; Fung, H.L.; Bushman, D.; Chiang, H.I.; Chun, J.; Lo, Y.H.; Zhang, K. Massively parallel polymerase cloning and genome sequencing of single cells using nanoliter microwells. Nat. Biotechnol. 2013, 31, 1126–1132. [Google Scholar] [CrossRef] [PubMed]
  28. Kim, K.E.; Peluso, P.; Baybayan, P.; Yeadon, P.J.; Yu, C.; Fisher, W.; Chin, C.S.; Rapicavoli, N.A.; Rank, D.R.; Li, J.; et al. Long-read, whole genome shotgun sequence data from five model organisms. Nat. Sci. Data 2014. [Google Scholar] [CrossRef]
  29. Chin, C.S.; Alexander, D.H.; Marks, P.; Klammer, A.A.; Drake, J.; Heiner, C.; Clum, A.; Copeland, A.; Huddleston, J.; Eichler, E.E.; et al. Nonhybrid, finished microbial genome assemblies from long-read SMRT sequencing data. Nat. Methods 2013, 10, 563–569. [Google Scholar] [CrossRef] [PubMed]
  30. Conlan, S.; Thomas, P.J.; Deming, C.; Park, M.; Lau, A.F.; Dekker, J.P.; Snitkin, E.S.; Clark, T.A.; Luong, K.; Song, Y.; et al. Single-molecule sequencing to track plasmid diversity of hospital-associated carbapenemase-producing Enterobacteriaceae. Sci. Transl. Med. 2014, 6. [Google Scholar] [CrossRef]
  31. Beatson, S.A.; Walker, M.J. Microbiology: Tracking antibiotic resistance. Science 2014, 345, 1454–1455. [Google Scholar] [CrossRef] [PubMed]
  32. Latif, H.; Li, H.J.; Charusanti, P.; Palsson, B.Ø.; Aziz, R.K. A Gapless, Unambiguous Genome Sequence of the Enterohemorrhagic Escherichia coli O157:H7 Strain EDL933. Genome Announc. 2014, 2, e00821-14. [Google Scholar] [CrossRef] [PubMed]
  33. Quick, J.; Quinlan, A.R.; Loman, N.J. A reference bacterial genome dataset generated on the MinION™ portable single-molecule nanopore sequencer. Gigascience 2014, 3, 22. [Google Scholar] [CrossRef] [PubMed]
  34. Trecarichi, E.M.; Cauda, R.; Tumbarello, M. Detecting risk and predicting patient mortality in patients with extended-spectrum beta-lactamase-producing Enterobacteriaceae bloodstream infections. Future Microbiol. 2012, 7, 1173–1189. [Google Scholar] [CrossRef] [PubMed]
  35. Salipante, S.J.; Roach, D.J.; Kitzman, J.O.; Snyder, M.W.; Stackhouse, B.; Butler-Wu, S.M.; Lee, C.; Cookson, B.T.; Shendure, J. Large-scale genomic sequencing of extraintestinal pathogenic Escherichia coli strains. Genome Res. 2015, 25, 119–128. [Google Scholar] [CrossRef] [PubMed]
  36. Russo, T.A.; Johnson, J.R. Proposal for a new inclusive designation for extraintestinal pathogenic isolates of Escherichia coli: ExPEC. J. Infect. Dis. 2000, 181, 1753–1754. [Google Scholar] [CrossRef] [PubMed]
  37. Foxman, B. The epidemiology of urinary tract infection. Nat. Rev. Urol. 2010, 7, 653–660. [Google Scholar] [CrossRef] [PubMed]
  38. Didelot, X.; Meric, G.; Falush, D.; Darling, A.E. Impact of homologous and non-homologous recombination in the genomic evolution of Escherichia coli. BMC Genomics 2012, 13, 256. [Google Scholar] [CrossRef] [PubMed]
  39. Wiles, T.J.; Kulesus, R.R.; Mulvey, M.A. Origins and virulence mechanisms of uropathogenic Escherichia coli. Exp. Mol. Pathol. 2008, 85, 11–19. [Google Scholar] [CrossRef] [PubMed]
  40. Wu, G.; Day, M.J.; Mafura, M.T.; Nunez-Garcia, J.; Fenner, J.J.; Sharma, M.; Rodríguez, I.; van Essen-Zandbergen, A.; Dierikx, C.; Kadlec, K.; et al. Comparative analysis of ESBL-positive Escherichia coli isolates from animals and humans from the UK, The Netherlands and Germany. PLoS ONE 2013, 8, e75392. [Google Scholar] [CrossRef]
  41. Tango, C.N.; Mansur, A.R.; Oh, D.H. Fumaric Acid and Slightly Acidic Electrolyzed Water Inactivate Gram Positive and Gram Negative Foodborne Pathogens. Microorganisms 2015, 3, 34–46. [Google Scholar] [CrossRef]
  42. Vincent, C.; Boerlin, P.; Daignault, D.; Dozois, C.M.; Dutil, L.; Galanakis, C.; Reid-Smith, R.J.; Tellier, P.P.; Tellis, P.A.; Ziebell, K.; et al. Food reservoir for Escherichia coli causing urinary tract infections. Emerg. Infect. Dis. 2010, 16, 88–95. [Google Scholar] [CrossRef] [PubMed]
  43. Lupindu, A.M.; Olsen, J.E.; Ngowi, H.A.; Msoffe, P.L.; Mtambo, M.M.; Scheutz, F.; Dalsgaard, A. Occurrence and characterization of Shiga toxin-producing Escherichia coli O157:H7 and other non-sorbitol-fermenting E. coli in cattle and humans in urban areas of Morogoro, Tanzania. Vector Borne Zoonotic Dis. 2014, 14, 503–510. [Google Scholar] [CrossRef] [PubMed]
  44. Ewers, C.; Grobbel, M.; Stamm, I.; Kopp, P.A.; Diehl, I.; Semmler, T.; Fruth, A.; Beutlich, J.; Guerra, B.; Wieler, L.H.; et al. Emergence of human pandemic O25:H4-ST131 CTX-M-15 extended-spectrum-beta-lactamase-producing Escherichia coli among companion animals. J. Antimicrob. Chemother. 2010, 65, 651–660. [Google Scholar] [CrossRef] [PubMed]
  45. Nebbia, P.; Tramuta, C.; Odore, R.; Nucera, D.; Zanatta, R.; Robino, P. Genetic and phenotypic characterisation of Escherichia coli producing cefotaximase-type extended-spectrum beta-lactamases: First evidence of the ST131 clone in cats with urinary infections in Italy. J. Feline Med. Surg. 2014, 16, 966–971. [Google Scholar] [CrossRef] [PubMed]
  46. Köhler, C.D.; Dobrindt, U. What defines extraintestinal pathogenic Escherichia coli? Int. J. Med. Microbiol. 2011, 301, 642–647. [Google Scholar] [CrossRef] [PubMed]
  47. Madigan, T.; Johnson, J.R.; Clabots, C.; Johnston, B.D.; Porter, S.B.; Slater, B.S.; Banerjee, R. Extensive household outbreak of urinary tract infection and intestinal colonization due to extended-spectrum beta lactamase (ESBL)-producing Escherichia coli sequence type 131 (ST131). Clin. Infect. Dis. 2015. [Google Scholar] [CrossRef]
  48. Dahbi, G.; Mora, A.; López, C.; Alonso, M.P.; Mamani, R.; Marzoa, J.; Coira, A.; García-Garrote, F.; Pita, J.M.; Velasco, D.; et al. Emergence of new variants of ST131 clonal group among extraintestinal pathogenic Escherichia coli producing extended-spectrum beta-lactamases. Int. J. Antimicrob. Agents 2013, 42, 347–351. [Google Scholar] [CrossRef] [PubMed]
  49. Izdebski, R.; Baraniak, A.; Fiett, J.; Adler, A.; Kazma, M.; Salomon, J.; Lawrence, C.; Rossini, A.; Salvia, A.; Vidal Samso, J.; et al. Clonal structure, extended-spectrum beta-lactamases, and acquired AmpC-type cephalosporinases of Escherichia coli populations colonizing patients in rehabilitation centers in four countries. Antimicrob. Agents Chemother. 2013, 57, 309–316. [Google Scholar] [CrossRef] [PubMed]
  50. Matsumura, Y.; Yamamoto, M.; Higuchi, T.; Komori, T.; Tsuboi, F.; Hayashi, A.; Sugimoto, Y.; Hotta, G.; Matsushima, A.; Nagao, M.; et al. Prevalence of plasmid-mediated AmpC beta-lactamase-producing Escherichia coli and spread of the ST131 clone among extended-spectrum beta-lactamase-producing E. coli in Japan. Int. J. Antimicrob. Agents 2012, 40, 158–162. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. Johnson, J.R.; Urban, C.; Weissman, S.J.; Jorgensen, J.H.; Lewis, J.S.; Hansen, G.; Edelstein, P.H.; Robicsek, A.; Cleary, T.; Adachi, J.; et al. Molecular epidemiological analysis of Escherichia coli sequence type ST131 (O25, H4) and blaCTX-M-15 among extended-spectrum-beta-lactamase-producing E. coli from the United States, 2000 to 2009. Antimicrob. Agents Chemother. 2012, 56, 2364–2370. [Google Scholar] [CrossRef] [PubMed]
  52. Chen, L.F.; Freeman, J.T.; Nicholson, B.; Keiger, A.; Lancaster, S.; Joyce, M.; Woods, C.W.; Cook, E.; Adcock, L.; Louis, S.; et al. Widespread dissemination of CTX-M-15 genotype extended-spectrum-beta-lactamase-producing enterobacteriaceae among patients presenting to community hospitals in the southeastern United States. Antimicrob. Agents Chemother. 2014, 58, 1200–1202. [Google Scholar] [CrossRef] [PubMed]
  53. Johnson, J.R.; Menard, M.; Johnston, B.; Kuskowski, M.A.; Nichol, K.; Zhanel, G.G. Epidemic clonal groups of Escherichia coli as a cause of antimicrobial-resistant urinary tract infections in Canada, 2002 to 2004. Antimicrob. Agents Chemother. 2009, 53, 2733–2739. [Google Scholar] [CrossRef] [PubMed]
  54. Fennell, J.; Vellinga, A.; Hanahoe, B.; Morris, D.; Boyle, F.; Higgins, F.; Lyons, M.; O’Connell, K.; Keady, D.; Cormican, M. Increasing prevalence of ESBL production among Irish clinical Enterobacteriaceae from 2004 to 2008: An observational study. BMC Infect. Dis. 2012, 12, 116. [Google Scholar] [CrossRef] [PubMed]
  55. Clark, G.; Paszkiewicz, K.; Hale, J.; Weston, V.; Constantinidou, C.; Penn, C.; Achtman, M.; McNally, A. Genomic analysis uncovers a phenotypically diverse but genetically homogeneous Escherichia coli ST131 clone circulating in unrelated urinary tract infections. J. Antimicrob. Chemother. 2012, 67, 868–877. [Google Scholar] [CrossRef] [PubMed]
  56. Rogers, B.A.; Sidjabat, H.E.; Paterson, D.L. Escherichia coli O25b-ST131: A pandemic, multiresistant, community-associated strain. J. Antimicrob. Chemother. 2011, 66, 1–14. [Google Scholar] [CrossRef] [PubMed]
  57. Woodford, N.; Ward, M.E.; Kaufmann, M.E.; Turton, J.; Fagan, E.J.; James, D.; Johnson, A.P.; Pike, R.; Warner, M.; Cheasty, T.; et al. Community and hospital spread of Escherichia coli producing CTX-M extended-spectrum beta-lactamases in the UK. J. Antimicrob. Chemother. 2004, 54, 735–743. [Google Scholar] [CrossRef] [PubMed]
  58. Escobar-Paramo, P.; Grenet, K.; Le Menac’h, A.; Rode, L.; Salgado, E.; Amorin, C.; Gouriou, S.; Picard, B.; Rahimy, M.C.; Andremont, A.; et al. Large-scale population structure of human com- mensal Escherichia coli isolates. Appl. Environ. Microbiol. 2004, 70, 5698–5700. [Google Scholar] [CrossRef] [PubMed]
  59. Escobar-Paramo, P.; Le Menac’h, A.; Le Gall, T.; Amorin, C.; Gouriou, S.; Picard, B.; Skurnik, D.; Denamur, E. Identification of forces shaping the commensal Escherichia coli genetic structure by comparing animal and human isolates. Environ. Microbiol. Rep. 2006, 8, 1975–1984. [Google Scholar] [CrossRef] [Green Version]
  60. Le Gall, T.; Clermont, O.; Gouriou, S.; Picard, B.; Nassif, X.; Denamur, E.; Tenaillon, O. Extraintestinal virulence is a coincidental by-product of commensalism in B2 phylogenetic group Escherichia coli strains. Mol. Biol. Evol. 2007, 24, 2373–2384. [Google Scholar] [CrossRef] [PubMed]
  61. Platell, J.L.; Johnson, J.R.; Cobbold, R.N.; Trott, D.J. Multidrug-resistant extraintestinal pathogenic Escherichia coli of sequence type ST131 in animals and foods. Vet. Microbiol. 2011, 153, 99–108. [Google Scholar] [CrossRef] [PubMed]
  62. Zong, Z.; Yu, R. blaCTX-M-carrying Escherichia coli of the O25b ST131 clonal group have emerged in China. Diagn. Microbiol. Infect. Dis. 2011, 69, 228–231. [Google Scholar] [CrossRef] [PubMed]
  63. Olesen, B.; Hansen, D.S.; Nilsson, F.; Frimodt-Møller, J.; Leihof, R.F.; Struve, C.; Scheutz, F.; Johnston, B.; Krogfelt, K.A.; Johnson, J.R. Prevalence and characteristics of the epidemic multiresistant Escherichia coli ST131 clonal group among extended-spectrum beta-lactamase-producing E. coli isolates in Copenhagen, Denmark. J. Clin. Microbiol. 2013, 51, 1779–1785. [Google Scholar] [CrossRef] [PubMed]
  64. Matsumura, Y.; Yamamoto, M.; Nagao, M.; Hotta, G.; Matsushima, A.; Ito, Y.; Takakura, S.; Ichiyama, S. Emergence and spread of B2-ST131-O25b, B2-ST131-O16 and D-ST405 clonal groups among extended-spectrum-beta-lactamase-producing Escherichia coli in Japan. J. Antimicrob. Chemother. 2012, 67, 2612–2620. [Google Scholar] [CrossRef] [PubMed]
  65. Johnson, J.R.; Johnston, B.; Clabots, C.; Kuskowski, M.A.; Castanheira, M. Escherichia coli sequence type ST131 as the major cause of serious multidrug-resistant E. coli infections in the United States 2007. Clin. Infect. Dis. 2010, 51, 286–294. [Google Scholar] [CrossRef] [PubMed]
  66. Lau, S.H.; Kaufmann, M.E.; Livermore, D.M.; Woodford, N.; Willshaw, G.A.; Cheasty, T.; Stamper, K.; Reddy, S.; Cheesbrough, J.; Bolton, F.J.; et al. UK epidemic Escherichia coli strains A-E, with CTX-M-15 beta-lactamase, all belong to the international O25:H4-ST131 clone. J. Antimicrob. Chemother. 2008, 62, 1241–1244. [Google Scholar] [CrossRef] [PubMed]
  67. Alobwede, I.; M’Zali, F.H.; Livermore, D.M.; Heritage, J.; Todd, N.; Hawkey, P.M. CTX-M extended-spectrum beta-lactamase arrives in the UK. J. Antimicrob. Chemother. 2003, 51, 470–471. [Google Scholar] [CrossRef] [PubMed]
  68. Olesen, B.; Frimodt-Møller, J.; Leihof, R.F.; Struve, C.; Johnston, B.; Hansen, D.S.; Scheutz, F.; Krogfelt, K.A.; Kuskowski, M.A.; Clabots, C.; et al. Temporal trends in antimicrobial resistance and virulence-associated traits within the Escherichia coli sequence type 131 clonal group and its H30 and H30-Rx subclones, 1968 to 2012. Antimicrob. Agents Chemother. 2014, 58, 6886–6895. [Google Scholar] [CrossRef] [PubMed]
  69. Huttner, A.; Harbarth, S.; Carlet, J.; Cosgrove, S.; Goossens, H.; Holmes, A.; Jarlier, V.; Voss, A.; Pittet, D. Antimicrobial resistance: A global view from the 2013 World Healthcare-Associated Infections Forum. Antimicrob. Resist. Infect. Control. 2013, 2, 31. [Google Scholar] [CrossRef] [PubMed]
  70. Gupta, K.; Scholes, D.; Stamm, W.E. Increasing prevalence of antimicrobial resistance among uropathogens causing acute uncomplicated cystitis in women. JAMA 1999, 281, 736–738. [Google Scholar] [CrossRef] [PubMed]
  71. Andersson, D.I.; Hughes, D. Evolution of antibiotic resistance at non-lethal drug concentrations. Drug Resist. Updat. 2012, 15, 162–172. [Google Scholar] [CrossRef] [PubMed]
  72. Oz, T.; Guvenek, A.; Yildiz, S.; Karaboga, E.; Tamer, Y.T.; Mumcuyan, N.; Ozan, V.B.; Senturk, G.H.; Cokol, M.; Yeh, P.; et al. Strength of selection pressure is an important parameter contributing to the complexity of antibiotic resistance evolution. Mol. Biol. Evol. 2014, 31, 2387–2401. [Google Scholar] [CrossRef] [PubMed]
  73. Harris, S.; Morris, C.; Morris, D.; Cormican, M.; Cummins, E. The effect of hospital effluent on antimicrobial resistant E. coli within a municipal wastewater system. Environ. Sci. Process. Impacts 2013, 15, 617–622. [Google Scholar] [CrossRef] [PubMed]
  74. Galvin, S.; Boyle, F.; Hickey, P.; Vellinga, A.; Morris, D.; Cormican, M. Enumeration and characterization of antimicrobial-resistant Escherichia coli bacteria in effluent from municipal, hospital, and secondary treatment facility sources. Appl. Environ. Microbiol. 2010, 76, 4772–4779. [Google Scholar] [CrossRef] [PubMed]
  75. Udikovic-Kolic, N.; Wichmann, F.; Broderick, N.A.; Handelsman, J. Bloom of resident antibiotic-resistant bacteria in soil following manure fertilization. Proc. Natl. Acad. Sci. USA 2014, 111, 15202–15207. [Google Scholar] [CrossRef] [PubMed]
  76. Currie, C.J.; Berni, E.; Jenkins-Jones, S.; Poole, C.D.; Ouwens, M.; Driessen, S.; de Voogd, H.; Butler, C.C.; Morgan, C.L. Antibiotic treatment failure in four common infections in UK primary care 1991–2012: Longitudinal analysis. BMJ 2014, 349, g5493. [Google Scholar] [CrossRef] [PubMed]
  77. Morris, D.; Boyle, F.; Buckley, V.; Xu, L.; Hanahoe, B.; Hawkey, P.; Cormican, M. CTX-M enzymes are the predominant extended-spectrum beta-lactamases produced by Enterobacteriaceae in Ireland. J. Antimicrob. Chemother. 2009, 64, 864–866. [Google Scholar] [CrossRef] [PubMed]
  78. Morris, D.; McGarry, E.; Cotter, M.; Passet, V.; Lynch, M.; Ludden, C.; Hannan, M.M.; Brisse, S.; Cormican, M. Detection of OXA-48 carbapenemase in the pandemic clone Escherichia coli O25b:H4-ST131 in the course of investigation of an outbreak of OXA-48-producing Klebsiella pneumoniae. Antimicrob. Agents Chemother. 2012, 56, 4030–4031. [Google Scholar] [CrossRef] [PubMed]
  79. Agabou, A.; Pantel, A.; Ouchenane, Z.; Lezzar, N.; Khemissi, S.; Satta, D.; Sotto, A.; Lavigne, J.P. First description of OXA-48-producing Escherichia coli and the pandemic clone ST131 from patients hospitalised at a military hospital in Algeria. Eur. J. Clin. Microbiol. Infect. Dis. 2014, 33, 1641–1646. [Google Scholar] [CrossRef] [PubMed]
  80. Cai, J.C.; Zhang, R.; Hu, Y.Y.; Zhou, H.W.; Chen, G.X. Emergence of Escherichia coli sequence type 131 isolates producing KPC-2 carbapenemase in China. Antimicrob. Agents Chemother. 2014, 58, 1146–1152. [Google Scholar] [CrossRef] [PubMed]
  81. Morris, D.; Boyle, F.; Ludden, C.; Condon, I.; Hale, J.; O’Connell, N.; Power, L.; Boo, T.W.; Dhanji, H.; Lavallee, C.; et al. Production of KPC-2 carbapenemase by an Escherichia coli clinical isolate belonging to the international ST131 clone. Antimicrob. Agents Chemother. 2011, 55, 4935–4936. [Google Scholar] [CrossRef] [PubMed]
  82. Riley, L.W. Pandemic lineages of extraintestinal pathogenic Escherichia coli. Clin. Microbiol. Infect. 2014, 20, 380–390. [Google Scholar] [CrossRef] [PubMed]
  83. Blanco, J.; Mora, A.; Mamani, R.; Lopez, C.; Blanco, M.; Dahbi, G.; Herrera, A.; Marzoa, J.; Fernandez, V.; de la Cruz, F.; et al. Four main virotypes among extended-spectrum-beta-lactamase-producing isolates of Escherichia coli O25b:H4-B2-ST131: Bacterial, epidemiological, and clinical characteristics. J. Clin. Microbiol. 2013, 51, 3358–3367. [Google Scholar] [CrossRef] [PubMed]
  84. Banerjee, R.; Johnston, B.; Lohse, C.; Porter, S.B.; Clabots, C.; Johnson, J.R. Escherichia coli sequence type 131 is a dominant, antimicrobial-resistant clonal group associated with healthcare and elderly hosts. Infect. Control. Hosp. Epidemiol. 2013, 34, 361–369. [Google Scholar] [CrossRef] [PubMed]
  85. Livermore, D.M.; Woodford, N. The beta-lactamase threat in Enterobacteriaceae, Pseudomonas and Acinetobacter. Trends Microbiol. 2006, 14, 413–420. [Google Scholar] [CrossRef] [PubMed]
  86. Pelly, H.; O’Connell, E.; Hanahoe, B.; Chambers, C.; Biernacka, K.; Gray, S.; Cormican, M. Outbreak of extended spectrum beta-lactamase producing E. coli in a nursing home in Ireland, May 2006. Euro Surveill. 2006, 11, E060831.1. [Google Scholar] [PubMed]
  87. Burke, L.; Humphreys, H.; Fitzgerald-Hughes, D. The revolving door between hospital and community: Extended-spectrum beta-lactamase-producing Escherichia coli in Dublin. J. Hosp. Infect. 2012, 81, 192–198. [Google Scholar] [CrossRef] [PubMed]
  88. Pitout, J.D.; Gregson, D.B.; Campbell, L.; Laupland, K.B. Molecular characteristics of extended-spectrum-beta-lactamase-producing Escherichia coli isolates causing bacteremia in the Calgary Health Region from 2000 to 2007: Emergence of clone ST131 as a cause of community-acquired infections. Antimicrob. Agents Chemother. 2009, 53, 2846–2851. [Google Scholar] [CrossRef] [PubMed]
  89. Karami, N.; Helldal, L.; Welinder-Olsson, C.; Ahrén, C.; Moore, E.R. Sub-typing of extended-spectrum-beta-lactamase-producing isolates from a nosocomial outbreak: Application of a 10-loci generic Escherichia coli multi-locus variable number tandem repeat analysis. PLoS ONE 2013, 8, e83030. [Google Scholar] [CrossRef] [PubMed]
  90. Shakir, S.M.; Goldbeck, J.M.; Robison, D.; Eckerd, A.M.; Chavez-Bueno, S. Genotypic and Phenotypic Characterization of Invasive Neonatal Escherichia coli Clinical Isolates. Am. J. Perinatol. 2014, 31, 975–982. [Google Scholar] [CrossRef] [PubMed]
  91. Ludden, C.; Cormican, M.; Vellinga, A.; Johnson, J.R.; Austin, B.; Morris, D. Colonisation with ESBL-producing and carbapenemase-producing Enterobacteriaceae, vancomycin-resistant enterococci, and meticillin-resistant Staphylococcus aureus in a long-term care facility over one year. BMC Infect. Dis. 2015, 15, 168. [Google Scholar] [CrossRef] [PubMed]
  92. Suetens, C. Healthcare-associated infections in European long-term care facilities: How big is the challenge? Euro Surveill. 2012, 17. pii. 20259. [Google Scholar] [PubMed]
  93. Peirano, G.; Pitout, J.D.D. Molecular epidemiology of Escherichia coli producing CTX-M β-lactamases: The worldwide emergence of clone ST131 O25:H4. Int. J. Antimicrob. Agents 2010, 35, 316–321. [Google Scholar] [CrossRef] [PubMed]
  94. Van Tonder, A.J.; Mistry, S.; Bray, J.E.; Hill, D.M.; Cody, A.J.; Farmer, C.L.; Klugman, K.P.; von Gottberg, A.; Bentley, S.D.; Parkhill, J.; et al. Defining the estimated core genome of bacterial populations using a Bayesian decision model. PLoS Comput. Biol. 2014, 10, e1003788. [Google Scholar] [CrossRef] [PubMed]
  95. Hooper, S.D.; Berg, O.G. Duplication is more common among laterally transferred genes than among indigenous genes. Genome Biol. 2003, 4, R48. [Google Scholar] [CrossRef] [PubMed]
  96. McNally, A.; Cheng, L.; Harris, S.R.; Corander, J. The evolutionary path to extraintestinal pathogenic, drug-resistant Escherichia coli is marked by drastic reduction in detectable recombination within the core genome. Genome Biol. Evol. 2013, 5, 699–710. [Google Scholar] [CrossRef] [PubMed]
  97. Rasko, D.A.; Rosovitz, M.J.; Myers, G.S.; Mongodin, E.F.; Fricke, W.F.; Gajer, P.; Crabtree, J.; Sebaihia, M.; Thomson, N.R.; Chaudhuri, R.; et al. The pangenome structure of Escherichia coli: comparative genomic analysis of Escherichia coli commensal and pathogenic isolates. J. Bacteriol. 2008, 190, 6881–6893. [Google Scholar] [CrossRef] [PubMed]
  98. Price, L.B.; Johnson, J.R.; Aziz, M.; Clabots, C.; Johnston, B.; Tchesnokova, V.; Nordstrom, L.; Billig, M.; Chattopadhyay, S.; Stegger, M.; et al. The epidemic of extended-spectrum-beta-lactamase-producing Escherichia coli ST131 is driven by a single highly pathogenic subclone, H30-Rx. mBio 2013, 4, e00377-13. [Google Scholar] [PubMed]
  99. Treangen, T.J.; Rocha, E.P. Horizontal transfer, not duplication, drives the expansion of protein families in prokaryotes. PLoS Genet 2011, 7, e1001284. [Google Scholar] [CrossRef] [PubMed]
  100. Bobay, L.M.; Touchon, M.; Rocha, E.P. Pervasive domestication of defective prophages by bacteria. Proc. Natl. Acad. Sci. USA 2014, 111, 12127–12132. [Google Scholar] [CrossRef] [PubMed]
  101. Näsvall, J.; Sun, L.; Roth, J.R.; Andersson, D.I. Real-time evolution of new genes by innovation, amplification, and divergence. Science 2012, 338, 384–387. [Google Scholar] [CrossRef] [PubMed]
  102. Bobay, L.M.; Rocha, E.P.; Touchon, M. The adaptation of temperate bacteriophages to their host genomes. Mol. Biol. Evol. 2013, 30, 737–751. [Google Scholar] [CrossRef] [PubMed]
  103. Waldor, M.K.; Friedman, D.I. Phage regulatory circuits and virulence gene expression. Curr. Opin. Microbiol. 2005, 8, 459–465. [Google Scholar] [CrossRef] [PubMed]
  104. Edlin, G.; Lin, L.; Bitner, R. Reproductive fitness of P1, P2, and Mu lysogens of Escherichia coli. J. Virol. 1977, 21, 560–564. [Google Scholar] [PubMed]
  105. Rabinovich, L.; Sigal, N.; Borovok, I.; Nir-Paz, R.; Herskovits, A.A. Prophage excision activates Listeria competence genes that promote phagosomal escape and virulence. Cell 2012, 150, 792–802. [Google Scholar] [CrossRef] [PubMed]
  106. Wang, X.; Kim, Y.; Ma, Q.; Hong, S.H.; Pokusaeva, K.; Sturino, J.M.; Wood, T.K. Cryptic prophages help bacteria cope with adverse environments. Nat. Commun. 2010, 1. [Google Scholar] [CrossRef]
  107. Gödeke, J.; Paul, K.; Lassak, J.; Thormann, K.M. Phage-induced lysis enhances biofilm formation in Shewanella oneidensis MR-1. ISME J. 2011, 5, 613–626. [Google Scholar] [CrossRef] [PubMed]
  108. Wozniak, R.A.; Fouts, D.E.; Spagnoletti, M.; Colombo, M.M.; Ceccarelli, D.; Garriss, G.; Déry, C.; Burrus, V.; Waldor, M.K. Comparative ICE genomics: insights into the evolution of the SXT/R391 family of ICEs. PLoS Genet 2009, 5, e1000786. [Google Scholar] [CrossRef] [PubMed]
  109. D’Agata, E.M.; Dupont-Rouzeyrol, M.; Magal, P.; Olivier, D.; Ruan, S. The impact of different antibiotic regimens on the emergence of antimicrobial-resistant bacteria. PLoS ONE 2008, 3, e4036. [Google Scholar] [CrossRef] [PubMed]
  110. Palmer, A.C.; Kishony, R. Opposing effects of target overexpression reveal drug mechanisms. Nat. Commun. 2014, 5. [Google Scholar] [CrossRef] [PubMed]
  111. Avasthi, T.S.; Kumar, N.; Baddam, R.; Hussain, A.; Nandanwar, N.; Jadhav, S.; Ahmed, N. Genome of multidrug-resistant uropathogenic Escherichia coli strain NA114 from India. J. Bacteriol. 2011, 193, 4272–4273. [Google Scholar] [CrossRef] [PubMed]
  112. Andersen, P.S.; Stegger, M.; Aziz, M.; Contente-Cuomo, T.; Gibbons, H.S.; Keim, P.; Sokurenko, E.V.; Johnson, J.R.; Price, L.B. Complete Genome Sequence of the Epidemic and Highly Virulent CTX-M-15-Producing H30-Rx Subclone of Escherichia coli ST131. Genome Announc 2013, 1, e00988-13. [Google Scholar] [CrossRef] [PubMed]
  113. Forde, B.M.; Zakour, B.; Stanton-Cook, M.; Phan, M.D.; Totsika, M.; Peters, K.M.; Chan, K.G.; Schembri, M.A.; Upton, M.; Beatson, S.A. The Complete Genome Sequence of Escherichia coli EC958: A High Quality Reference Sequence for the Globally Disseminated Multidrug Resistant E. coli O25b:H4-ST131 Clone. PLoS ONE 2014, 9, e104400. [Google Scholar] [CrossRef] [PubMed]
  114. Johnson, T.J.; Hargreaves, M.; Shaw, K.; Snippes, P.; Lynfield, R.; Aziz, M.; Price, L.B. Complete Genome Sequence of a Carbapenem-Resistant Extraintestinal Pathogenic Escherichia coli Strain Belonging to the Sequence Type 131 H30R Subclade. Genome Announc. 2015, 3. [Google Scholar] [CrossRef]
  115. Totsika, M.; Beatson, S.A.; Sarkar, S.; Phan, M.D.; Petty, N.K.; Bachmann, N.; Szubert, M.; Sidjabat, H.E.; Paterson, D.L.; Upton, M.; et al. Insights into a multidrug resistant Escherichia coli pathogen of the globally disseminated ST131 lineage: Genome analysis and virulence mechanisms. PLoS ONE 2011, 6, e26578. [Google Scholar] [CrossRef] [PubMed]
  116. Clermont, O.; Lavollay, M.; Vimont, S.; Deschamps, C.; Forestier, C.; Branger, C.; Denamur, E.; Arlet, G. The CTX-M-15-producing Escherichia coli diffusing clone belongs to a highly virulent B2 phylogenetic subgroup. J. Antimicrob. Chemother. 2008, 61, 1024–1028. [Google Scholar] [CrossRef] [PubMed]
  117. Petty, N.K.; Zakour, B.; Stanton-Cook, M.; Skippington, E.; Totsika, M.; Forde, B.M.; Phan, M.D.; Moriel, D.G.; Peters, K.M.; Davies, M.; et al. Global dissemination of a multidrug resistant Escherichia coli clone. Proc. Natl. Acad. Sci. USA 2014, 111, 5694–5699. [Google Scholar] [CrossRef] [PubMed]
  118. Johnson, J.R.; Tchesnokova, V.; Johnston, B.; Clabots, C.; Roberts, P.L.; Billig, M.; Riddell, K.; Rogers, P.; Qin, X.; Butler-Wu, S.; et al. Abrupt emergence of a single dominant multi-drug-resistant strain of Escherichia coli. J. Infect. Dis. 2013, 207, 919–928. [Google Scholar] [CrossRef] [PubMed]
  119. Peirano, G.; van der Bij, A.K.; Freeman, J.L.; Poirel, L.; Nordmann, P.; Costello, M.; Pitout, J.D.; Tchesnokova, V.L. Characteristics of Escherichia coli sequence type 131 isolates that produce extended-spectrum beta-lactamases: Global distribution of the H30-Rx sublineage. Antimicrob. Agents Chemother. 2014, 58, 3762–3767. [Google Scholar] [CrossRef] [PubMed]
  120. Peirano, G.; Pitout, J.D. Fluoroquinolone-resistant Escherichia coli sequence type 131 isolates causing bloodstream infections in a Canadian region with a centralized laboratory system: Rapid emergence of the H30-Rx sublineage. Antimicrob. Agents Chemother. 2014, 58, 2699–2703. [Google Scholar] [CrossRef] [PubMed]
  121. Cantón, R.; Coque, T.M. The CTX-M beta-lactamase pandemic. Curr. Opin. Microbiol. 2006, 9, 466–475. [Google Scholar] [CrossRef] [PubMed]
  122. Rooney, P.J.; O’Leary, M.C.; Loughrey, A.C.; McCalmont, M.; Smyth, B.; Donaghy, P.; Badri, M.; Woodford, N.; Karisik, E.; Livermore, D.M. Nursing homes as a reservoir of extended-spectrum beta-lactamase (ESBL)-producing ciprofloxacin-resistant Escherichia coli. J. Antimicrob. Chemother. 2009, 64, 635–641. [Google Scholar] [CrossRef] [PubMed]
  123. Banerjee, R.; Johnson, J.R. Escherichia coli ST131: Variations on a theme of clonal expansion. Enferm Infecc Microbiol. Clin. 2013, 31, 355–356. [Google Scholar] [CrossRef] [PubMed]
  124. Gullberg, E.; Cao, S.; Berg, O.G.; Ilbäck, C.; Sandegren, L.; Hughes, D.; Andersson, D.I. Selection of resistant bacteria at very low antibiotic concentrations. PLoS Pathog. 2011, 7, e1002158. [Google Scholar] [CrossRef] [PubMed]
  125. Adams-Sapper, S.; Diep, B.A.; Perdreau-Remington, F.; Riley, L.W. Clonal composition and community clustering of drug-susceptible and -resistant Escherichia coli isolates from bloodstream infections. Antimicrob. Agents Chemother. 2013, 57, 490–497. [Google Scholar] [CrossRef] [PubMed]
  126. Johnson, J.R.; Murray, A.C.; Gajewski, A.; Sullivan, M.; Snippes, P.; Kuskowski, M.A.; Smith, K.E. Isolation and molecular characterization of nalidixic acid-resistant extraintestinal pathogenic Escherichia coli from retail chicken products. Antimicrob. Agents Chemother. 2003, 47, 2161–2168. [Google Scholar] [CrossRef] [PubMed]
  127. Pitout, J.D. Extraintestinal Pathogenic Escherichia coli: A Combination of Virulence with Antibiotic Resistance. Front Microbiol. 2012, 3, 9. [Google Scholar] [CrossRef] [PubMed]
  128. Nicolas-Chanoine, M.H.; Bertrand, X.; Madec, J.Y. Escherichia coli ST131, an intriguing clonal group. Clin. Microbiol. Rev. 2014, 27, 543–574. [Google Scholar] [CrossRef] [PubMed]
  129. Dhanji, H.; Doumith, M.; Rooney, P.J.; O’Leary, M.C.; Loughrey, A.C.; Hope, R.; Woodford, N.; Livermore, D.M. Molecular epidemiology of fluoroquinolone-resistant ST131 Escherichia coli producing CTX-M extended-spectrum beta-lactamases in nursing homes in Belfast, UK. J. Antimicrob. Chemother. 2011, 66, 297–303. [Google Scholar] [CrossRef] [PubMed]
  130. Bush, K.; Jacoby, G.A. Updated functional classification of beta-lactamases. Antimicrob. Agents Chemother. 2010, 54, 969–976. [Google Scholar] [CrossRef] [PubMed]
  131. Jensen, L.B.; Hasman, H.; Agersø, Y.; Emborg, H.-D.; Aarestrup, F.M. First description of an oxyiminocephalosporin-resistant, ESBL-carrying Escherichia coli isolated from meat sold in Denmark. J. Antimicrob. Chemother. 2006, 57, 793–794. [Google Scholar] [CrossRef] [PubMed]
  132. Yong, D.; Toleman, M.A.; Giske, C.G.; Cho, H.S.; Sundman, K.; Lee, K.; Walsh, T.R. Characterization of a new metallo-beta-lactamase gene, bla(NDM-1), and a novel erythromycin esterase gene carried on a unique genetic structure in Klebsiella pneumoniae sequence type 14 from India. Antimicrob. Agents Chemother. 2009, 53, 5046–5054. [Google Scholar] [CrossRef] [PubMed]
  133. Naas, T.; Cuzon, G.; Gaillot, O.; Courcol, R.; Nordmann, P. When carbapenem-hydrolyzing beta-lactamase KPC meets Escherichia coli ST131 in France. Antimicrob. Agents Chemother. 2011, 55, 4933–4934. [Google Scholar] [CrossRef] [PubMed]
  134. Kim, Y.A.; Qureshi, Z.A.; Adams-Haduch, J.M.; Park, Y.S.; Shutt, K.A.; Doi, Y. Features of infections due to Klebsiella pneumoniae carbapenemase-producing Escherichia coli: Emergence of sequence type 131. Clin. Infect Dis. 2012, 55, 224–231. [Google Scholar] [CrossRef] [PubMed]
  135. Mantengoli, E.; Luzzaro, F.; Pecile, P.; Cecconi, D.; Cavallo, A.; Attala, L.; Bartoloni, A.; Rossolini, G.M. Escherichia coli ST131 producing extended-spectrum beta-lactamases plus VIM-1 carbapenemase: Further narrowing of treatment options. Clin. Infect. Dis. 2011, 52, 690–691. [Google Scholar] [CrossRef] [PubMed]
  136. Yan, J.-J.; Tsai, L.-H.; Wu, J.-J. Emergence of the IMP-8 metallo-beta-lactamase in the Escherichia coli ST131 clone in Taiwan. Int. J. Antimicrob. Agents 2012, 40, 281–282. [Google Scholar] [CrossRef] [PubMed]
  137. Peirano, G.; Schreckenberger, P.C.; Pitout, J.D. Characteristics of NDM-1-producing Escherichia coli isolates that belong to the successful and virulent clone ST131. Antimicrob. Agents Chemother. 2011, 55, 2986–2988. [Google Scholar] [CrossRef] [PubMed]
  138. Seema, K.; Ranjan Sen, M.; Upadhyay, S.; Bhattacharjee, A. Dissemination of the New Delhi metallo-beta-lactamase-1 (NDM-1) among Enterobacteriaceae in a tertiary referral hospital in north India. J. Antimicrob. Chemother. 2011, 66, 1646–1647. [Google Scholar] [CrossRef] [PubMed]
  139. Deshpande, P.; Rodrigues, C.; Shetty, A.; Kapadia, F.; Hedge, A.; Soman, R. New Delhi Metallo-beta lactamase (NDM-1) in Enterobacteriaceae: Treatment options with carbapenems compromised. J. Assoc. Physicians India 2010, 58, 147–149. [Google Scholar] [PubMed]
  140. Perry, J.D.; Naqvi, S.H.; Mirza, I.A.; Alizai, S.A.; Hussain, A.; Ghirardi, S.; Orenga, S.; Wilkinson, K.; Woodford, N.; Zhang, J.; et al. Prevalence of faecal carriage of Enterobacteriaceae with NDM-1 carbapenemase at military hospitals in Pakistan, and evaluation of two chromogenic media. J. Antimicrob. Chemother. 2011, 66, 2288–2294. [Google Scholar] [CrossRef] [PubMed]
  141. Bonnin, R.A.; Poirel, L.; Carattoli, A.; Nordmann, P. Characterization of an IncFII plasmid encoding NDM-1 from Escherichia coli ST131. PLoS ONE 2012, 7, e34752. [Google Scholar] [CrossRef] [PubMed]
  142. Gaillot, O.; Clement, C.; Simonet, M.; Philippon, A. Novel transferable beta-lactam resistance with cephalosporinase characteristics in Salmonella enteritidis. J. Antimicrob. Chemother. 1997, 39, 85–87. [Google Scholar] [CrossRef] [PubMed]
  143. Sousa, A.; Magalhães, S.; Gordo, I. Cost of antibiotic resistance and the geometry of adaptation. Mol. Biol. Evol. 2012, 29, 1417–1428. [Google Scholar] [CrossRef] [PubMed]
  144. LeClerc, J.E.; Li, B.G.; Payne, W.L.; Cebula, T.A. High mutation frequencies among Escherichia coli and Salmonella pathogens. Science 1996, 274, 1208–1211. [Google Scholar] [CrossRef] [PubMed]
  145. Olofsson, S.K.; Geli, P.; Andersson, D.I.; Cars, O. Pharmacodynamic model to describe the concentration-dependent selection of cefotaxime-resistant Escherichia coli. Antimicrob. Agents Chemother. 2005, 49, 5081–5091. [Google Scholar] [CrossRef] [PubMed]
  146. Silva, R.F.; Mendonça, S.C.; Carvalho, L.M.; Reis, A.M.; Gordo, I.; Trindade, S.; Dionisio, F. Pervasive sign epistasis between conjugative plasmids and drug-resistance chromosomal mutations. PLoS Genet 2011, 7, e1002181. [Google Scholar] [CrossRef] [PubMed]
  147. Trindade, S.; Sousa, A.; Xavier, K.B.; Dionisio, F.; Ferreira, M.G.; Gordo, I. Positive epistasis drives the acquisition of multidrug resistance. PLoS Genet 2009, 5, e1000578. [Google Scholar] [CrossRef] [PubMed]
  148. Morgand, M.; Vimont, S.; Bleibtreu, A.; Boyd, A.; Thien, H.V.; Zahar, J.R.; Denamur, E.; Arlet, G. Extended-spectrum beta-lactamase-producing Escherichia coli infections in children: Are community-acquired strains different from nosocomial strains? Int. J. Med. Microbiol. 2014, 304, 970–976. [Google Scholar] [CrossRef] [PubMed]
  149. Rodríguez-Beltrán, J.; Tourret, J.; Tenaillon, O.; López, E.; Bourdelier, E.; Costas, C.; Matic, I.; Denamur, E.; Blázquez, J. High Recombinant Frequency in Extraintestinal Pathogenic Escherichia coli Strains. Mol. Biol. Evol. 2015. [Google Scholar] [CrossRef]
  150. Leflon-Guibout, V.; Blanco, J.; Amaqdouf, K.; Mora, A.; Guize, L.; Nicolas-Chanoine, M.H. Absence of CTX-M enzymes but high prevalence of clones, including clone ST131, among fecal Escherichia coli isolates from healthy subjects living in the area of Paris, France. J. Clin. Microbiol. 2008, 46, 3900–3905. [Google Scholar] [CrossRef] [PubMed]
  151. Angst, D.C.; Hall, A.R. The cost of antibiotic resistance depends on evolutionary history in Escherichia coli. BMC Evol. Biol. 2013, 13, 163. [Google Scholar] [CrossRef] [PubMed]
  152. Kawamura, K.; Goto, K.; Nakane, K.; Arakawa, Y. Molecular epidemiology of extended-spectrum beta-lactamases and Escherichia coli isolated from retail foods including chicken meat in Japan. Foodborne Pathog. Dis. 2014, 11, 104–110. [Google Scholar] [CrossRef] [PubMed]
  153. Randall, L.P.; Clouting, C.; Horton, R.A.; Coldham, N.G.; Wu, G.; Clifton-Hadley, F.A.; Davies, R.H.; Teale, C.J. Prevalence of Escherichia coli carrying extended-spectrum beta-lactamases (CTX-M and TEM-52) from broiler chickens and turkeys in Great Britain between 2006 and 2009. J. Antimicrob. Chemother. 2011, 66, 86–95. [Google Scholar] [CrossRef] [PubMed]
  154. Kutumbaka, K.K.; Han, S.; Mategko, J.; Nadala, C.; Buser, G.L.; Cassidy, M.P.; Beldavs, Z.G.; Weissman, S.J.; Morey, K.E.; Vega, R.; et al. Draft Genome Sequence of blaNDM-1-Positive Escherichia coli O25b-ST131 Clone Isolated from an Environmental Sample. Genome Announc. 2014, 2. [Google Scholar] [CrossRef] [PubMed]
  155. Morar, M.; Wright, G.D. The genomic enzymology of antibiotic resistance. Annu. Rev. Genet 2010, 44, 25–51. [Google Scholar] [CrossRef] [PubMed]
  156. Soo, V.W.; Hanson-Manful, P.; Patrick, W.M. Artificial gene amplification reveals an abundance of promiscuous resistance determinants in Escherichia coli. Proc. Natl. Acad. Sci. USA 2011, 108, 1484–1489. [Google Scholar] [CrossRef] [PubMed]
  157. Rodrigo, G.; Baji´c, D.; Elola, I.; Poyatos, J.F. Dual autogenous control of the multiple antibiotic resistance phenotype in Escherichia coli. Available online: http://biorxiv.org/content/early/2014/08/19/008169 (accessed on 3 September 2014).
  158. Ravva, S.V.; Cooley, M.B.; Sarreal, C.Z.; Mandrell, R.E. Fitness of Outbreak and Environmental Strains of Escherichia coli O157:H7 in Aerosolizable Soil and Association of Clonal Variation in Stress Gene Regulation. Pathogens 2014, 3, 528–548. [Google Scholar] [CrossRef] [PubMed]
  159. Imming, P.; Sinning, C.; Meyer, A. Drugs, their targets and the nature and number of drug targets. Nat. Rev. Drug Discov. 2006, 5, 821–834. [Google Scholar] [CrossRef] [PubMed]
  160. Wilkinson, D.J. Stochastic modelling for quantitative description of heterogeneous biological systems. Nat. Rev. Genet 2009, 10, 122–133. [Google Scholar] [CrossRef] [PubMed]
  161. Ferguson, M.L.; le Coq, D.; Jules, M.; Aymerich, S.; Radulescu, O.; Declerck, N.; Royer, C.A. Reconciling molecular regulatory mechanisms with noise patterns of bacterial metabolic promoters in induced and repressed states. Proc. Natl. Acad. Sci. USA 2012, 109, 155–160. [Google Scholar] [CrossRef] [PubMed]
  162. Wolf, L.; Silander, O.K.; van Nimwegen, E.J. Expression noise facilitates the evolution of gene regulation. Available online: http://biorxiv.org/content/early/2014/07/18/007237 (accessed on 25 July 2014).
  163. Cheng, A.A.; Dinge, H.; Lu, T.K. Enhanced killing of antibiotic-resistant bacteria enabled by massively parallel combinatorial genetics. Proc. Natl. Acad. Sci. USA 2014, 111, 12462–12467. [Google Scholar] [CrossRef] [PubMed]
  164. Silander, O.K.; Nikolic, N.; Zaslaver, A.; Bren, A.; Kikoin, I.; Alon, U.; Ackermann, M. A genome-wide analysis of promoter-mediated phenotypic noise in Escherichia coli. PLoS Genet 2012, 8, e1002443. [Google Scholar] [CrossRef] [PubMed]
  165. Bollenbach, T.; Kishony, R. Resolution of gene regulatory conflicts caused by combinations of antibiotics. Mol. Cell 2011, 42, 413–425. [Google Scholar] [CrossRef] [PubMed]
  166. Caroff, N.; Espaze, E.; Gautreau, D.; Richet, H.; Reynaud, A. Analysis of the effects of −42 and −32 ampC promoter mutations in clinical isolates of Escherichia coli hyperproducing ampC. J. Antimicrob. Chemother. 2000, 45, 783–788. [Google Scholar] [CrossRef] [PubMed]
  167. Teichmann, S.A.; Babu, M.M. Gene regulatory network growth by duplication. Nat. Genet 2004, 36, 492–496. [Google Scholar] [CrossRef] [PubMed]
  168. Toprak, E.; Veres, A.; Michel, J.B.; Chait, R.; Hartl, D.L.; Kishony, R. Evolutionary paths to antibiotic resistance under dynamically sustained drug selection. Nat. Genet 2011, 44, 101–105. [Google Scholar] [CrossRef] [PubMed]
  169. Thomas, R. Laws for the dynamics of regulatory networks. Int. J. Dev. Biol. 1998, 42, 479–485. [Google Scholar] [PubMed]
  170. Kim, S.; Lieberman, T.D.; Kishony, R. Alternating antibiotic treatments constrain evolutionary paths to multidrug resistance. Proc. Natl. Acad. Sci. USA 2014, 111, 14494–14499. [Google Scholar] [CrossRef] [PubMed]
  171. Adler, M.; Anjum, M.; Andersson, D.I.; Sandegren, L. Influence of acquired beta-lactamases on the evolution of spontaneous carbapenem resistance in Escherichia coli. J. Antimicrob. Chemother. 2013, 68, 51–59. [Google Scholar] [CrossRef] [PubMed]
  172. Raeside, C.; Gaffé, J.; Deatherage, D.E.; Tenaillon, O.; Briska, A.M.; Ptashkin, R.N.; Cruveiller, S.; Médigue, C.; Lenski, R.E.; Barrick, J.E.; et al. Large chromosomal rearrangements during a long-term evolution experiment with Escherichia coli. mBio 2014, 5, e01377-14. [Google Scholar] [CrossRef] [PubMed]
  173. Aldridge, B.B.; Fernandez-Suarez, M.; Heller, D.; Ambravaneswaran, V.; Irimia, D.; Toner, M.; Fortune, S.M. Asymmetry and aging of mycobacterial cells lead to variable growth and antibiotic susceptibility. Science 2012, 335, 100–104. [Google Scholar] [CrossRef] [PubMed]
  174. Klumpp, S.; Zhang, Z.; Hwa, T. Growth rate-dependent global effects on gene expression in bacteria. Cell 2009, 139, 1366–1375. [Google Scholar] [CrossRef] [PubMed]
  175. Elowitz, M.B.; Levine, A.J.; Siggia, E.D.; Swain, P.S. Stochastic gene expression in a single cell. Science 2002, 297, 1183–1186. [Google Scholar] [CrossRef] [PubMed]
  176. Balaban, N.Q.; Merrin, J.; Chait, R.; Kowalik, L.; Leibler, S. Bacterial persistence as a phenotypic switch. Science 2004, 305, 1622–1625. [Google Scholar] [CrossRef] [PubMed]
  177. Fasani, R.A.; Savageau, M.A. Molecular mechanisms of multiple toxin-antitoxin systems are coordinated to govern the persister phenotype. Proc. Natl. Acad. Sci. USA 2013, 110, E2528–E2537. [Google Scholar] [CrossRef] [PubMed]
  178. Fridman, O.; Goldberg, A.; Ronin, I.; Shoresh, N.; Balaban, N.Q. Optimization of lag time underlies antibiotic tolerance in evolved bacterial populations. Nature 2014, 513, 418–421. [Google Scholar] [CrossRef] [PubMed]
  179. Lou, C.; Li, Z.; Ouyang, Q. A molecular model for persister in E. coli. J. Theor. Biol. 2008, 255, 205–209. [Google Scholar] [CrossRef] [PubMed]
  180. Grundmann, H. Towards a global antibiotic resistance surveillance system: A primer for a roadmap. Upsala J. Med. Sci. 2014, 119, 87–95. [Google Scholar] [CrossRef] [PubMed]
  181. Li, L.M.; Grassly, N.C.; Fraser, C. Genomic analysis of emerging pathogens: Methods, application and future trends. Genome Biol. 2014, 15, 541. [Google Scholar] [CrossRef] [PubMed]
  182. Webber, M.A.; Ricci, V.; Whitehead, R.; Patel, M.; Fookes, M.; Ivens, A.; Piddock, L.J. Clinically relevant mutant DNA gyrase alters supercoiling, changes the transcriptome, and confers multidrug resistance. mBio 2013, 4, e00273-13. [Google Scholar] [CrossRef] [PubMed]
  183. Joensen, K.G.; Scheutz, F.; Lund, O.; Hasman, H.; Kaas, R.S.; Nielsen, E.M.; Aarestrup, F.M. Real-time whole-genome sequencing for routine typing, surveillance, and outbreak detection of verotoxigenic Escherichia coli. J. Clin. Microbiol. 2014, 52, 1501–1510. [Google Scholar] [CrossRef] [PubMed]
  184. Caboche, S.; Audebert, C.; Hot, D. High-Throughput Sequencing, a Versatile Weapon to Support Genome-Based Diagnosis in Infectious Diseases: Applications to Clinical Bacteriology. Pathogens 2014, 3, 258–279. [Google Scholar] [CrossRef] [PubMed]
  185. Alqasim, A.; Scheutz, F.; Zong, Z.; McNally, A. Comparative genome analysis identifies few traits unique to the Escherichia coli ST131 H30Rx clade and extensive mosaicism at the capsule locus. BMC Genomics 2014, 15, 830. [Google Scholar] [CrossRef] [PubMed]
  186. Yahara, K.; Furuta, Y.; Oshima, K.; Yoshida, M.; Azuma, T.; Hattori, M.; Uchiyama, I.; Kobayashi, I. Chromosome painting in silico in a bacterial species reveals fine population structure. Mol. Biol. Evol. 2013, 30, 1454–1464. [Google Scholar] [CrossRef] [PubMed]
  187. Yahara, K.; Didelot, X.; Ansari, M.A.; Sheppard, S.K.; Falush, D. Efficient inference of recombination hot regions in bacterial genomes. Mol. Biol. Evol. 2014, 31, 1593–1605. [Google Scholar] [CrossRef] [PubMed]
  188. Shimodaira, H. An approximately unbiased test of phylogenetic tree selection. Syst. Biol. 2002, 51, 492–508. [Google Scholar] [CrossRef] [PubMed]
  189. Schierup, M.H. Consequences of recombination on traditional phylogenetic analysis. Genetics 2000, 156, 879–891. [Google Scholar] [PubMed]
  190. Barrick, J.E.; Kauth, M.R.; Strelioff, C.C.; Lenski, R.E. Escherichia coli rpoB mutants have increased evolvability in proportion to their fitness defects. Mol. Biol. Evol. 2010, 27, 1338–1347. [Google Scholar] [CrossRef] [PubMed]
  191. Pybus, O.G.; Rambaut, A.; Belshaw, R.; Freckleton, R.P.; Drummond, A.J.; Holmes, E.C. Phylogenetic evidence for deleterious mutation load in RNA viruses and its contribution to viral evolution. Mol. Biol. Evol. 2007, 24, 845–852. [Google Scholar] [CrossRef] [PubMed]
  192. Neher, R.A.; Russell, C.A.; Shraiman, B.I. Predicting evolution from the shape of genealogical trees. Elife 2014, 3. [Google Scholar] [CrossRef]
  193. Dobrindt, U. (Patho-)Genomics of Escherichia coli. Int. J. Med. Microbiol. 2005, 295, 357–371. [Google Scholar] [CrossRef] [PubMed]
  194. Freeman, J.T.; McBride, S.J.; Heffernan, H.; Bathgate, T.; Pope, C.; Ellis-Pegler, R.B. Community-onset genitourinary tract infection due to CTX-M-15-Producing Escherichia coli among travelers to the Indian subcontinent in New Zealand. Clin. Infect. Dis. 2008, 47, 689–692. [Google Scholar] [CrossRef] [PubMed]
  195. Laupland, K.B.; Church, D.L.; Vidakovich, J.; Mucenski, M.; Pitout, J.D. Community-onset extended-spectrum beta-lactamase (ESBL) producing Escherichia coli: Importance of international travel. J. Infect. 2008, 57, 441–448. [Google Scholar] [CrossRef] [PubMed]
  196. Pitout, J.D.; Campbell, L.; Church, D.L.; Wang, P.W.; Guttman, D.S.; Gregson, D.B. Using a commercial DiversiLab semiautomated repetitive sequence-based PCR typing technique for identification of Escherichia coli clone ST131 producing CTX-M-15. J. Clin. Microbiol. 2009, 47, 1212–1215. [Google Scholar] [CrossRef] [PubMed]
  197. Van der Bij, A.K.; Pitout, J.D. The role of international travel in the worldwide spread of multiresistant Enterobacteriaceae. J. Antimicrob. Chemother. 2012, 67, 2090–2100. [Google Scholar] [CrossRef] [PubMed]
  198. Haigh, J. The accumulation of deleterious genes in a population—Muller’s Ratchet. Theor. Popul. Biol. 1978, 14, 251–267. [Google Scholar] [CrossRef] [PubMed]
  199. Landis, M.J.; Bedford, T. Phylowood: Interactive web-based animations of biogeographic and phylogeographic histories. Bioinformatics 2014, 30, 123–124. [Google Scholar] [CrossRef] [PubMed]
  200. Illingworth, C.J.; Mustonen, V. A method to infer positive selection from marker dynamics in an asexual population. Bioinformatics 2012, 6, 831–837. [Google Scholar] [CrossRef]
  201. Skoglund, P.; Sjödin, P.; Skoglund, T.; Lascoux, M.; Jakobsson, M. Investigating Population History Using Temporal Genetic Differentiation. Mol. Biol. Evol. 2014, 31, 2516–2527. [Google Scholar] [CrossRef] [PubMed]
  202. Foll, M.; Poh, Y.P.; Renzette, N.; Ferrer-Admetlla, A.; Bank, C.; Shim, H.; Malaspinas, A.S.; Ewing, G.; Liu, P.; Wegmann, D.; et al. Influenza virus drug resistance: a time-sampled population genetics perspective. PLoS Genet 2014, 10, e1004185. [Google Scholar] [CrossRef] [PubMed]
  203. Luksza, M.; Lässig, M. A predictive fitness model for influenza. Nature 2014, 507, 57–61. [Google Scholar] [CrossRef] [PubMed]
  204. Rasmussen, D.A.; Volz, E.M.; Koelle, K. Phylodynamic inference for structured epidemiological models. PLoS Comput. Biol. 2014, 10, e1003570. [Google Scholar] [CrossRef] [PubMed]
  205. Török, M.E.; Reuter, S.; Bryant, J.; Köser, C.U.; Stinchcombe, S.V.; Nazareth, B.; Ellington, M.J.; Bentley, S.D.; Smith, G.P.; Parkhill, J.; et al. Rapid whole-genome sequencing for investigation of a suspected tuberculosis outbreak. J. Clin. Microbiol. 2013, 51, 611–614. [Google Scholar] [CrossRef] [PubMed]
  206. Harris, S.R.; Cartwright, E.J.; Török, M.E.; Holden, M.T.; Brown, N.M.; Ogilvy-Stuart, A.L.; Ellington, M.J.; Quail, M.A.; Bentley, S.D.; Parkhill, J.; et al. Whole-genome sequencing for analysis of an outbreak of meticillin-resistant Staphylococcus aureus: A descriptive study. Lancet Infect. Dis. 2013, 13, 130–136. [Google Scholar] [CrossRef] [PubMed]
  207. Gardy, J.L.; Johnston, J.C.; Ho Sui, S.J.; Cook, V.J.; Shah, L.; Brodkin, E.; Rempel, S.; Moore, R.; Zhao, Y.; Holt, R.; et al. Whole-genome sequencing and social-network analysis of a tuberculosis outbreak. N. Engl. J. Med. 2011, 364, 730–739. [Google Scholar] [CrossRef] [PubMed]
  208. Machuca, J.; Briales, A.; Labrador, G.; Díaz-de-Alba, P.; López-Rojas, R.; Docobo-Pérez, F.; Martínez-Martínez, L.; Rodríguez-Baño, J.; Pachón, M.E.; Pascual, A.; et al. Interplay between plasmid-mediated and chromosomal-mediated fluoroquinolone resistance and bacterial fitness in Escherichia coli. J. Antimicrob. Chemother. 2014, 69, 3203–3215. [Google Scholar] [CrossRef] [PubMed]
  209. Török, M.E.; Harris, S.R.; Cartwright, E.J.; Raven, K.E.; Brown, N.M.; Allison, M.E.; Greaves, D.; Quail, M.A.; Limmathurotsakul, D.; Holden, M.T.; et al. Zero tolerance for healthcare-associated MRSA bacteraemia: Is it realistic? J. Antimicrob. Chemother. 2014, 69, 2238–2245. [Google Scholar] [CrossRef] [PubMed]
  210. Bull, J.J. Evolution of phenotypic variance. Evolution 1987, 41, 303–315. [Google Scholar] [CrossRef]
  211. Vander Heiden, M.G.; Cantley, L.C.; Thompson, C.B. Understanding the Warburg effect: The metabolic requirements of cell proliferation. Science 2009, 324, 1029–1033. [Google Scholar] [CrossRef] [PubMed]
  212. Munsky, B.; Neuert, G.; van Oudenaarden, A. Using gene expression noise to understand gene regulation. Science 2012, 336, 183–187. [Google Scholar] [CrossRef] [PubMed]
  213. Yun, H.S.; Hong, J.; Lim, H.C. Regulation of ribosome synthesis in Escherichia coli: Effects of temperature and dilution rate changes. Biotechnol. Bioeng. 1996, 52, 615–624. [Google Scholar] [CrossRef] [PubMed]
  214. Zaslaver, A.; Kaplan, S.; Bren, A.; Jinich, A.; Mayo, A.; Dekel, E.; Alon, U.; Itzkovitz, S. Invariant distribution of promoter activities in Escherichia coli. PLoS Comput. Biol. 2009, 5, e1000545. [Google Scholar] [CrossRef] [PubMed]
  215. Sabarly, V.; Bouvet, O.; Glodt, J.; Clermont, O.; Skurnik, D.; Diancourt, L.; de Vienne, D.; Denamur, E.; Dillmann, C. The decoupling between genetic structure and metabolic phenotypes in Escherichia coli leads to continuous phenotypic diversity. J. Evol. Biol. 2011, 24, 1559–1571. [Google Scholar] [CrossRef] [PubMed]
  216. Kiviet, D.J.; Nghe, P.; Walker, N.; Boulineau, S.; Sunderlikova, V.; Tans, S.J. Stochasticity of metabolism and growth at the single-cell level. Nature 2014, 514, 376–379. [Google Scholar] [CrossRef] [PubMed]
  217. Rodriguez, M.; Good, T.A.; Wales, M.E.; Hua, J.P.; Wild, J.R. Modeling allosteric regulation of de novo pyrimidine biosynthesis in Escherichia coli. J. Theor. Biol. 2005, 234, 299–310. [Google Scholar] [CrossRef] [PubMed]
  218. Hart, Y.; Madar, D.; Yuan, J.; Bren, A.; Mayo, A.E.; Rabinowitz, J.D.; Alon, U. Robust control of nitrogen assimilation by a bifunctional enzyme in E. coli. Mol. Cell 2011, 41, 117–127. [Google Scholar] [CrossRef] [PubMed]
  219. Levine, E.; Hwa, T. Stochastic fluctuations in metabolic pathways. Proc. Natl. Acad. Sci. USA 2007, 104, 9224–9229. [Google Scholar] [CrossRef] [PubMed]
  220. Wilson, W.A.; Roach, P.J.; Montero, M.; Baroja-Fernández, E.; Muñoz, F.J.; Eydallin, G.; Viale, A.M.; Pozueta-Romero, J. Regulation of glycogen metabolism in yeast and bacteria. FEMS Microbiol. Rev. 2010, 34, 952–985. [Google Scholar] [PubMed]
  221. El-Mansi, E.M.; Holms, W.H. Control of carbon flux to acetate excretion during growth of Escherichia coli in batch and continuous cultures. J. Gen Microbiol. 1989, 135, 2875–2883. [Google Scholar] [PubMed]
  222. Bremer, H.; Dennis, P.P. Modulation of chemical composition and other parameters of the cell at different exponential growth rates. In EcoSal—Escherichia coli and Salmonella, Cellular and Molecular Biology; Curtiss, R., III, Ed.; ASM Press: Washington, DC, U.S.A., 2008. [Google Scholar]
  223. Baguley, D.; Lim, E.; Bevan, A.; Pallet, A.; Faust, S.N. Prescribing for children—Taste and palatability affect adherence to antibiotics: A review. Arch. Dis. Childhood 2012, 97, 293–297. [Google Scholar] [CrossRef]
  224. Ambrose, P.G.; Bhavnani, S.M.; Ellis-Grosse, E.J.; Drusano, G.L. Pharmacokinetic-pharmacodynamic considerations in the design of hospital-acquired or ventilator-associated bacterial pneumonia studies: look before you leap! Clin. Infect. Dis. 2010, 51, S103–S110. [Google Scholar] [CrossRef] [PubMed]
  225. Baquero, F.; Negri, M.C. Selective compartments for resistant microorganisms in antibiotic gradients. Bioessays 1997, 19, 731–736. [Google Scholar] [CrossRef] [PubMed]
  226. Hermsen, R.; Deris, J.B.; Hwa, T. On the rapidity of antibiotic resistance evolution facilitated by a concentration gradient. Proc. Natl. Acad. Sci. USA 2012, 109, 10775–10780. [Google Scholar] [CrossRef]
  227. Andersson, D.I.; Hughes, D. Persistence of antibiotic resistance in bacterial populations. FEMS Microbiol. Rev. 2011, 35, 901–911. [Google Scholar] [CrossRef] [PubMed]
  228. Greulich, P.; Waclaw, B.; Allen, R.J. Mutational pathway determines whether drug gradients accelerate evolution of drug-resistant cells. Phys. Rev. Lett. 2012, 109. [Google Scholar] [CrossRef] [PubMed]
  229. Levin, B.R.; Concepcion-Acevedo, J.; Udekwu, K. Persistence: A copacetic and parsimonious hypothesis for the existence of non-inherited resistance to antibiotics. Curr. Opin. Microbiol. 2014, 21, 18–21. [Google Scholar] [CrossRef] [PubMed]
  230. Moreno-Gamez, S.; Hill, A.L.; Rosenbloom, D.I.S.; Petrov, D.A.; Nowak, M.A.; Pennings, P. Imperfect drug penetration leads to spatial monotherapy and rapid evolution of multi-drug resistance. Available online: http://biorxiv.org/content/early/2014/12/19/013003 (accessed on 9 January 2015).
  231. Kussell, E.; Leibler, S. Phenotypic diversity, population growth, and information in fluctuating environments. Science 2005, 309, 2075–2078. [Google Scholar] [CrossRef] [PubMed]
  232. Deris, J.B.; Kim, M.; Zhang, Z.; Okano, H.; Hermsen, R.; Groisman, A.; Hwa, T. The innate growth bistability and fitness landscapes of antibiotic-resistant bacteria. Science 2013, 342. [Google Scholar] [CrossRef]
  233. Moyed, H.S.; Bertrand, K.P. hipA, a newly recognized gene of Escherichia coli K-12 that affects frequency of persistence after inhibition of murein synthesis. J. Bacteriol. 1983, 155, 768–775. [Google Scholar] [PubMed]
  234. Young, B.C.; Golubchik, T.; Batty, E.M.; Fung, R.; Larner-Svensson, H.; Votintseva, A.A.; Miller, R.R.; Godwin, H.; Knox, K.; Everitt, R.G.; et al. Evolutionary dynamics of Staphylococcus aureus during progression from carriage to disease. Proc. Natl. Acad. Sci. USA 2012, 109, 4550–4555. [Google Scholar] [CrossRef] [PubMed]
  235. Lieberman, T.D.; Flett, K.B.; Yelin, I.; Martin, T.R.; McAdam, A.J.; Priebe, G.P.; Kishony, R. Genetic variation of a bacterial pathogen within individuals with cystic fibrosis provides a record of selective pressures. Nat. Genet 2013, 46, 82–87. [Google Scholar] [CrossRef] [PubMed]
  236. Chung, J.C.; Becq, J.; Fraser, L.; Schulz-Trieglaff, O.; Bond, N.J.; Foweraker, J.; Bruce, K.D.; Smith, G.P.; Welch, M. Genomic variation among contemporary Pseudomonas aeruginosa isolates from chronically infected cystic fibrosis patients. J. Bacteriol. 2012, 194, 4857–4866. [Google Scholar] [CrossRef] [PubMed]
  237. Folkvardsen, D.B.; Svensson, E.; Thomsen, V.Ø.; Rasmussen, E.M.; Bang, D.; Werngren, J.; Hoffner, S.; Hillemann, D.; Rigouts, L. Can molecular methods detect 1% isoniazid resistance in Mycobacterium tuberculosis? J. Clin. Microbiol. 2013, 51, 1596–1599. [Google Scholar] [CrossRef] [PubMed]
  238. Korolev, K.S.; Müller, M.J.; Karahan, N.; Murray, A.W.; Hallatschek, O.; Nelson, D.R. Selective sweeps in growing microbial colonies. Phys. Biol. 2012, 9. [Google Scholar] [CrossRef]
  239. Coward, C.; Restif, O.; Dybowski, R.; Grant, A.J.; Maskell, D.J.; Mastroeni, P. The effects of vaccination and immunity on bacterial infection dynamics in vivo. PLoS Pathog. 2014, 10, e1004359. [Google Scholar] [CrossRef] [PubMed]
  240. Webb, G.F.; D’Agata, E.M.; Magal, P.; Ruan, S. A model of antibiotic-resistant bacterial epidemics in hospitals. Proc. Natl. Acad. Sci. USA 2005, 102, 13343–13348. [Google Scholar] [CrossRef] [PubMed]
  241. Ankomah, P.; Levin, B.R. Exploring the collaboration between antibiotics and the immune response in the treatment of acute, self-limiting infections. Proc. Natl. Acad. Sci. USA 2014, 111, 8331–8338. [Google Scholar] [CrossRef] [PubMed]
  242. Stoesser, N.; Batty, E.M.; Eyre, D.W.; Morgan, M.; Wyllie, D.H.; Del Ojo Elias, C.; Johnson, J.R.; Walker, A.S.; Peto, T.E.; Crook, D.W. Predicting antimicrobial susceptibilities for Escherichia coli and Klebsiella pneumoniae isolates using whole genomic sequence data. J. Antimicrob. Chemother. 2013, 68, 2234–2244. [Google Scholar] [PubMed]
  243. Köser, C.U.; Ellington, M.J.; Peacock, S.J. Whole-genome sequencing to control antimicrobial resistance. Trends Genet 2014, 30, 401–407. [Google Scholar] [CrossRef] [PubMed]
  244. Raoult, D. Editorial: emerging clones of bacterial epidemics in the genomic area. Clin. Microbiol. Infect. 2014, 20, 371–372. [Google Scholar] [CrossRef] [PubMed]
  245. Grad, Y.H.; Lipsitch, M. Epidemiologic data and pathogen genome sequences: A powerful synergy for public health. Genome Biol. 2014, 15, 538. [Google Scholar] [CrossRef] [PubMed]
  246. Taniguchi, Y.; Choi, P.J.; Li, G.W.; Chen, H.; Babu, M.; Hearn, J.; Emili, A.; Xie, X.S. Quantifying E. coli proteome and transcriptome with single-molecule sensitivity in single cells. Science 2010, 329, 533–538. [Google Scholar] [CrossRef] [PubMed]
  247. Lalaouna, D.; Eyraud, A.; Chabelskaya, S.; Felden, B.; Massé, E. Regulatory RNAs involved in bacterial antibiotic resistance. PLoS Pathog. 2014, 10, e1004299. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  248. Beaulaurier, J.; Zhu, S.; Sebra, R.; Zhang, X.S.; Rosenbluh, C.; Deikus, G.; Shen, N.; Munera, D.; Waldor, M.K.; Blaser, M.; et al. Single molecule-level detection and long read-based phasing of epigenetic variations in bacterial methylomes. Available online: http://biorxiv.org/content/early/2014/09/16/007823 (accessed on 25 September 2014).
  249. Fang, G.; Munera, D.; Friedman, D.I.; Mandlik, A.; Chao, M.C.; Banerjee, O.; Feng, Z.; Losic, B.; Mahajan, M.C.; Jabado, O.J.; et al. Genome-wide mapping of methylated adenine residues in pathogenic Escherichia coli using single-molecule real-time sequencing. Nat. Biotechnol. 2012, 30, 1232–1239. [Google Scholar] [CrossRef] [PubMed]
  250. Suzuki, S.; Horinouchi, T.; Furusawa, C. Prediction of antibiotic resistance by gene expression profiles. Nat. Commun. 2014, 5. [Google Scholar] [CrossRef] [PubMed]

Share and Cite

MDPI and ACS Style

Downing, T. Tackling Drug Resistant Infection Outbreaks of Global Pandemic Escherichia coli ST131 Using Evolutionary and Epidemiological Genomics. Microorganisms 2015, 3, 236-267. https://0-doi-org.brum.beds.ac.uk/10.3390/microorganisms3020236

AMA Style

Downing T. Tackling Drug Resistant Infection Outbreaks of Global Pandemic Escherichia coli ST131 Using Evolutionary and Epidemiological Genomics. Microorganisms. 2015; 3(2):236-267. https://0-doi-org.brum.beds.ac.uk/10.3390/microorganisms3020236

Chicago/Turabian Style

Downing, Tim. 2015. "Tackling Drug Resistant Infection Outbreaks of Global Pandemic Escherichia coli ST131 Using Evolutionary and Epidemiological Genomics" Microorganisms 3, no. 2: 236-267. https://0-doi-org.brum.beds.ac.uk/10.3390/microorganisms3020236

Article Metrics

Back to TopTop