Skip to main content
Advertisement
  • Loading metrics

A Phyletically Rare Gene Promotes the Niche-specific Fitness of an E. coli Pathogen during Bacteremia

  • Travis J. Wiles,

    Affiliation Division of Microbiology and Immunology, Pathology Department, University of Utah School of Medicine, Salt Lake City, Utah, United States of America

  • J. Paul Norton ,

    Contributed equally to this work with: J. Paul Norton, Sara N. Smith

    Affiliation Division of Microbiology and Immunology, Pathology Department, University of Utah School of Medicine, Salt Lake City, Utah, United States of America

  • Sara N. Smith ,

    Contributed equally to this work with: J. Paul Norton, Sara N. Smith

    Affiliation Department of Microbiology and Immunology, University of Michigan Medical School, Ann Arbor, Michigan, United States of America

  • Adam J. Lewis,

    Affiliation Division of Microbiology and Immunology, Pathology Department, University of Utah School of Medicine, Salt Lake City, Utah, United States of America

  • Harry L. T. Mobley,

    Affiliation Department of Microbiology and Immunology, University of Michigan Medical School, Ann Arbor, Michigan, United States of America

  • Sherwood R. Casjens,

    Affiliation Division of Microbiology and Immunology, Pathology Department, University of Utah School of Medicine, Salt Lake City, Utah, United States of America

  • Matthew A. Mulvey

    mulvey@path.utah.edu

    Affiliation Division of Microbiology and Immunology, Pathology Department, University of Utah School of Medicine, Salt Lake City, Utah, United States of America

Abstract

In bacteria, laterally acquired genes are often concentrated within chromosomal regions known as genomic islands. Using a recently developed zebrafish infection model, we set out to identify unique factors encoded within genomic islands that contribute to the fitness and virulence of a reference urosepsis isolate—extraintestinal pathogenic Escherichia coli strain CFT073. By screening a series of deletion mutants, we discovered a previously uncharacterized gene, neaT, that is conditionally required by the pathogen during systemic infections. In vitro assays indicate that neaT can limit bacterial interactions with host phagocytes and alter the aggregative properties of CFT073. The neaT gene is localized within an integrated P2-like bacteriophage in CFT073, but was rarely found within other proteobacterial genomes. Sequence-based analyses revealed that neaT homologues are present, but discordantly conserved, within a phyletically diverse set of bacterial species. In CFT073, neaT appears to be unameliorated, having an exceptionally A+T-rich composition along with a notably altered codon bias. These data suggest that neaT was recently brought into the proteobacterial pan-genome from an extra-phyletic source. Interestingly, even in G+C-poor genomes, as found within the Firmicutes lineage, neaT-like genes are often unameliorated. Sequence-level features of neaT homologues challenge the common supposition that the A+T-rich nature of many recently acquired genes reflects the nucleotide composition of their genomes of origin. In total, these findings highlight the complexity of the evolutionary forces that can affect the acquisition, utilization, and assimilation of rare genes that promote the niche-dependent fitness and virulence of a bacterial pathogen.

Author Summary

Bacterial pathogens, even those belonging to the same species, can be incredibly diverse with regard to the genes they carry. However, the design of vaccines and antibiotics typically relies upon identification of general molecular features shared by the targeted organisms. Thus, we have traditionally focused on broadly conserved characteristics of pathogenic bacteria, often ignoring the genes that account for their individuality. In this article we report the discovery of a unique gene, neaT, that promotes the fitness of a pathogenic Escherichia coli isolate in zebrafish and mouse models of systemic blood infections. Surprisingly, neaT is rarely found in other related strains of E. coli and appears to have been recently acquired from distant lineages of bacteria via a process known as ‘lateral gene transfer’ that is used by microbes to swap genetic material. Expression of the neaT gene appears to help pathogens avoid interactions with host immune cells, possibly by altering bacterial surface structures. This work provides an interesting example of how the lateral acquisition of a rare gene can impact the niche-specific virulence properties of a pathogen, shedding light on the mechanisms that drive pathogen evolution and diversity.

Introduction

As a species, Escherichia coli is best known for colonizing the lower intestine of humans and other warm-blooded vertebrates [1], [2]. The contingent exit from the intestinal tract presents strains of E. coli with a multitude of secondary habitats, including host-associated and free-living niches [2], [3], [4], [5], [6]. A subset of E. coli designated extraintestinal pathogenic E. coli (ExPEC) excels at colonizing host-associated extraintestinal environments, resulting in an array of human diseases including urinary tract infections, bacteremia, and meningitis [7]. ExPEC strains also exhibit an impressive zoonotic capacity, being able to persist and cause disease in a variety of domesticated animals [8], [9], [10], [11]. Collectively, ExPEC-related diseases represent daunting medical, agricultural, and economic burdens that threaten to worsen as antibiotic-resistant strains become more prevalent [8], [12], [13]. The evolutionary forces that underlie the emergence and niche tropisms of ExPEC have yet to be completely defined. Considering gene content, substantial intra-specific variation often exists between bacterial isolates, particularly among strains of pathogenic E. coli. Key questions regarding the origin of this heterogeneity and its impact on the fitness of virulent strains remain unanswered.

Bacteria are proficient at rapidly developing innovative, selectable traits to maintain fitness within complex environments—a property known as ‘evolvability’ [14], [15], [16], [17], [18]. Despite being largely asexual organisms that multiply by binary fission, bacteria engage in a genetically promiscuous behavior known as ‘lateral gene transfer’ (LGT). Laterally acquired genes can provide context-specific functions, such as the ability to metabolize atypical substrates [19], adhere to a variety of surfaces [7], neutralize antibiotics and other toxic compounds [5], or participate in niche construction [20]. Bacteria have several means of obtaining potentially beneficial elements through LGT: direct acquisition from the environment (transformation), transfer through cell-to-cell mating (conjugation), and acquisition from bacterial viruses known as bacteriophages (transduction) [14], [21], [22], [23], [24]. It has been estimated that ∼81% of all genes within a bacterial chromosome have been involved in LGT at some point, suggesting that this behavior is not just an anomalous event, but that over time it is a foundational component of bacterial evolution [25].

The genomes of E. coli are laden with the signatures of past LGT events. Since the first genome sequencing projects it has been apparent that E. coli chromosomes are highly mosaic [26], [27], [28]. In part, this chromosomal architecture results from the presence of ‘genomic islands’ (GI) that intermittently disrupt synteny [29], [30], [31], [32], [33], [34]. Many GIs exhibit clear signs of having been involved in past LGT events as they are often in proximity to mobile elements, such as transposons, or are themselves integrated phages or plasmids [35]. Accompanying this interchangeable chromosomal arrangement is a vast superset of genes defined as the pan-genome [32], [36], [37]. Whereas an average E. coli genome contains about 4,700 genes, the pan-genome of this species is estimated to be over 17,000 genes. Most E. coli strains share a subset of the pan-genome, which encodes vertically inherited genes that dictate the fundamental cellular properties of the lineage. This core genome surprisingly accounts for only 40–50% of the genetic makeup of any particular isolate. The rest of the chromosome contains strain-specific combinations of genes that are infused throughout the core genome and encode a variety of accessory functions that can provide unique selective advantages [32], [37], [38].

With this information in mind, we systematically screened GIs of a urosepsis ExPEC isolate for laterally acquired genes that affect virulence in a surrogate zebrafish host model. We identified a novel gene—designated neaT (nomadically evolved acyltransferase)—that is required during blood-borne, but not localized, infections in both zebrafish and mice. The neaT locus was unexpectedly rare in the genomes of closely related E. coli strains and other Proteobacteria, suggesting that it was obtained from outside the contemporary E. coli pan-genome. Proteobacterial neaT homologues, in general, exhibit a high degree of allelic variance, have reduced guanine and cytosine (G+C) content, and are often localized within the integrated genomes of unrelated bacteriophages. These observations indicate that neaT-like alleles may have been recently acquired on multiple occasions by the proteobacterial supraspecies pan-genome. Together, our results provide molecular and bioinformatic evidence that the acquisition of unique genes like neaT during the evolution of particular ExPEC isolates can significantly impact bacterial fitness and virulence within specific host environments. Possible evolutionary forces that generate the observed sequence-level features of neaT and the role that bacterial individuality plays in pathogenesis are considered.

Results

The P2-like prophage b0847 promotes the fitness and virulence of ExPEC strain CFT073 during systemic infection of zebrafish embryos

The ExPEC strain CFT073 was isolated from the blood of a human patient with acute pyelonephritis (kidney infection) [26], [39]. This urosepsis isolate is versatile, with the apparent ability to traverse several host microenvironments to reach the bloodstream, and has a relatively large genome of 5,369 protein-coding genes and several GIs. In previous work, we found that CFT073 is exceptionally lethal in an infection model that uses zebrafish embryos as surrogate hosts for the high-throughput analysis of ExPEC virulence [40]. At 48 h post-fertilization (hpf), zebrafish possess an innate immune system composed primarily of phagocytic cells and antimicrobial peptides [41], [42], [43], [44]. These defenses mirror those employed by mammalian hosts to combat ExPEC.

To identify GI-associated virulence factors carried by CFT073, we screened 11 previously described deletion mutants that each lack a specific GI (Table 1 and Figure 1A) [45]. In blinded assays, 48 hpf zebrafish embryos were infected with 1,000 to 2,000 colony-forming units (CFU) of either wild type CFT073 or one of the 11 GI mutants. Bacteria were delivered into one of two injection sites: the fluid-filled sac surrounding the heart referred to as the pericardial cavity (PC), which mimics a localized tissue infection, and the circulation valley (CV), which facilitates rapid dispersal of bacteria into the bloodstream [40]. Each of these sites likely challenges the pathogen with different nutrient limitations, receptor availability, and host defenses.

thumbnail
Figure 1. The φb0847 island is important for CFT073 pathogenicity during systemic infection in zebrafish embryos.

(A) Diagram of GIs and their location within the CFT073 chromosome that were screened in the zebrafish host. Magenta indicates island mutants that had no observable defects, while green denotes island mutants that displayed significant attenuation. (B) The pericardial cavity (PC, top row) and blood (bottom row) of 48 hpf embryos were inoculated with 1,000–2,000 CFU. Fish were scored for death at 0, 24, 48, and 72 h post-inoculation (hpi). Data are presented as Kaplan-Meier survival plots and p values were calculated using a log-rank (Mantel-Cox) test (sample sizes for each curve are listed in Table 1). (C) Equal numbers (1,000–2,000 CFU total) of wild type and CFT073Δφb0847 were inoculated into the bloodstream of embryos. Fish were sacrificed and bacterial loads enumerated at the indicated times by differential plating (n = 10 to 20 embryos). Data are represented as competitive indices, where negative values indicate a reduction in fitness of the mutant strain. (D) Bacteria were prepared as in (C) and inoculated into the PC or yolk. Fish were sacrificed at 18 hpi and bacterial numbers determined (n = 9–10 embryos). Data from blood infections is the same as in (C), provided as a reference. P values were determined using two-tailed Mann-Whitney t tests. Median values are indicated by bars in (C) and (D).

https://doi.org/10.1371/journal.ppat.1003175.g001

thumbnail
Table 1. Summary of GI screen in zebrafish infection sites.

https://doi.org/10.1371/journal.ppat.1003175.t001

In this infection model, increased growth of ExPEC is associated with decreased survival of the host [40]. All 11 GI mutants, with the exception of ΔGI-aspV, grew equally well in broth culture at 28.5°C and 37°C (data not shown). Following inoculation into the PC, only deletion of the 123 kb GI PAI-pheV [I] resulted in a significant decrease in virulence relative to wild type CFT073 (Figure 1B, top). This was not surprising as PAI-pheV [I] encodes the notable ExPEC-associated virulence factors α-hemolysin (pore-forming toxin), SAT (vacuolating toxin), P pili (adhesive organelles), aerobactin (iron acquisition system), and K2 capsule (immune evasion). The ability of the ΔPAI-pheV mutant to still kill approximately half of the embryos suggests that additional factors with overlapping roles in virulence within the PC are encoded outside of PAI-pheV and the 10 other GIs tested.

The ΔPAI-pheV mutant was also attenuated following inoculation of the CV to initiate systemic infection, as were the GI mutants ΔGI-selC, ΔGI-cobU, and Δφb0847 (Figure 1B, bottom). In addition to several hypothetical genes, the selC and cobU islands harbor genes that appear to be components of polyamine and iron transport systems, respectively. Both polyamines and iron acquisition systems are known to be important mediators of ExPEC fitness in mouse models of infection [46], [47], [48]. Although the ΔGI-cobU mutant exhibited only a modest reduction in virulence in these assays using inoculation doses of 1,000–2,000 CFU/embryo, with slightly higher doses between 2,000 to 3,000 CFU/embryo this mutant displayed more dramatic and significant (p<0.05) attenuation (Table S1). This observation supports previous findings indicating that the inoculation dose can markedly influence the discernibility of some mutant phenotypes in the zebrafish host [40].

The remaining GI showing a phenotype in our screen is composed of an intact integrated phage genome (prophage) named ‘φb0847’ (Figure 1B) [45]. This prophage is 33 kb in length and contains 48 predicted open reading frames, most of which encode recognizable phage proteins that share homology with genes of tailed phages belonging to the order Caudovirales. More specifically, φb0847 carries genes involved in regulation, replication, and virion assembly that are related to and syntenic with the genes of phage P2 and its relatives (Figure 2). From this analysis, it is clear that the φb0847 prophage is a member of the P2-like phage group and likely represents a fully functional phage genome complete with all the essential genes associated with P2-like phages [49]. Aside from the ΔpheV mutant, with its fairly well characterized assortment of virulence genes, Δφb0847 displayed the most pronounced defect of the island mutants examined. Therefore, the φb0847 GI became the primary focus of our investigation.

thumbnail
Figure 2. Alignment of φb0847 genome to other P2-like bacteriophage.

Related P2-like prophages are aligned relative to their respective integration sites (att). Size is measured in kilobase pairs (Kbp). E. coli phages P2 and 186 and Salmonella phages Fels-2 and SopE5 were previously characterized. HS2 is an uncharacterized prophage contained within the genome of the commensal E. coli strain HS. Our unpublished analysis indicates that P2, 186, and Fels-2 represent three different “sequence types” based on virion proteins, which are typically >85% identical within each of these three groups and 50–70% identical between the groups. Bracketed numbers below φb0847 indicate positions of the deletion mutants generated to assess the functionality of genes within broad regions of φb0847. The neaT gene is distinguished by a red open reading frame in moron position 2.

https://doi.org/10.1371/journal.ppat.1003175.g002

To further define the contribution of φb0847 to the virulence and fitness of CFT073, we carried out competitive assays in which a one-to-one mixture of wild type and mutant bacteria were injected into the CV of zebrafish embryos (Figure 1C). At the indicated time points, the infected embryos were homogenized and bacteria present were enumerated by dilution plating on selective agar. Δφb0847 carries a kanamycin resistance cassette that was used to distinguish wild type and mutant strains. No differences between wild type CFT073 and the Δφb0847 mutant were observed until 6 h post-inoculation (hpi), when Δφb0847 titers began to decline (Figure 1C). These results indicate that the φb0847 island is dispensable during initial stages of a systemic infection, but enhances bacterial fitness at later time points, coordinate with the upregulation of host inflammatory responses engage. The Δφb0847 mutant displayed more modest, though still significant, decreases in fitness during competitive assays against wild type CFT073 within the PC and yolk sac at 18 hpi (Figure 1D). Phagocytes are recruited into the PC en masse in response to infection with ExPEC [40], possibly contributing to the competitive disadvantage of the Δφb0847 mutant within this niche. On the other hand, the yolk is a rich source of nutrients for bacteria and is mostly free of phagocytes and other immunosurveillance mechanisms. However, the yolk does contain maternally inherited antimicrobial compounds that could account for the slight reduction in fitness of Δφb0847 within this host environment [50]. Competitive experiments in broth culture did not reveal appreciable differences between the wild type and mutant strains (data not shown).

φb0847 harbors a multigenic region that contributes to fitness

To identify genes within φb0847 that, when deleted, recapitulate the attenuated phenotypes of Δφb0847, we constructed partial deletion mutants lacking one of three nearly equal-sized regions of the prophage island (designated Δ1–2, Δ2–3, and Δ3–4, as indicated along the φb0847 genome in Figure 2). In competitive assays, the Δ1–2 and Δ2–3 mutants were significantly more fit than the full Δφb0847 mutant at 12 hpi (Figure 3A). Analysis at 12 hpi allowed time for selection to take place, while limiting artifacts due to bacterial replication at later time points in dead and dying hosts where selective pressures are presumably weaker. In these assays, only the Δ3–4 mutant phenocopied the complete φb0847 island deletion mutant (Figure 3A). Lethality of this mutant variant was also significantly reduced in comparison to wild type CFT073 and the Δ1–2 and Δ2–3 mutants in independent challenges (Figure 3B). These results indicate that one or more genes within the terminal 3–4 region of the φb0847 prophage enhances both the fitness and virulence of CFT073 during systemic infections within the zebrafish host.

thumbnail
Figure 3. φb0847 harbors multiple loci that contribute to the fitness of CFT073 during systemic challenge.

(A) Equal numbers (1,000–2,000 CFU total) of wild type CFT073 and each mutant derivative indicated were inoculated into the bloodstream of embryos. Fish were sacrificed and bacterial loads enumerated at 12 hours post inoculation (hpi) by differential plating (n>28). Data are presented as competitive indices with negative values indicating a reduction in fitness of the mutant. P values were determined using two-tailed Mann-Whitney t tests; bars indicate median values. (B), (C), and (D) 1,000–2,000 CFU of wild type CFT073, the indicated mutant, or recombinant derivative were each inoculated into the blood of 48 hpf embryos. Fish were scored for death every 6 h starting at 18 hpi until 48 hpi (n = 40 or more embryos for each curve). pGEN-mcs in (D) serves as an empty vector control for pGEN-neaTPnative. Data in (B), (C), and (D) are presented as Kaplan-Meier survival plots. A log-rank (Mantel-Cox) test was used to determine p values; ns = not significant.

https://doi.org/10.1371/journal.ppat.1003175.g003

Temperate prophage genomes like φb0847 can carry ‘lysogenic conversion’ genes that affect the bacterial host but are not essential for lytic phage growth. To avoid disruption of critical phage processes, the integration of this genetic material is generally tolerated only in certain regions of the prophage genome. These added sequences are known as ‘morons’, because bacteriophages with such insertions have more DNA [51], [52]. Moron genes typically contain their own regulatory elements and vary among individual phage genomes. They often alter the surface structure or physiology of the bacterial host and can benefit the phage by making its host refractory to competing parasites or otherwise promoting bacterial survival and growth [52], [53].

The P2-like phages appear to have at least two variable moron positions (Figure 2). Using phage P2 as a reference, the location of moron position 1 is between the DNA replication gene A and head assembly gene Q, and moron position 2 is between the tail fiber gene G and tail sheath gene FI (Figure 2) [49]. In φb0847 within CFT073, the second moron site, which is absent from the Δ3–4 mutant, contains one open reading frame that is oriented in the opposite transcriptional direction to the flanking tail genes. This gene, which we named neaT for reasons described later, encodes a putative acyltransferase (Pfam:PF01757). This gene is not conserved among P2-like phages and is likely not critical for lytic replication of φb0847.

In light of this information, neaT, the immediately proximal gene yfdK (homologous to P2 phage tail gene G), and the collection of distal tail genes (FI through D) were individually deleted from the φb0847 prophage in CFT073. All three mutant derivatives—ΔyfdK, ΔneaT, and ΔFI-D—were attenuated in their ability to kill zebrafish embryos after injection into the blood via the CV (Figure 3C). Despite the significantly reduced virulence of these mutants, no defects in fitness were observed in competitive assays with wild type CFT073 (data not shown). The lack of any discernable fitness defects in competition assays may 1) reflect the ability of wild type CFT073 to trans-complement the mutant strains in vivo and/or 2) indicate that there is cooperative interplay among the yfdK, neaT, and FI-D loci. Of note, disruption of loci flanking neaT did not appreciably alter its expression in broth culture (Figure S1). Furthermore, we found no evidence that the neaT mutant could be complemented in vivo during competition assays by acquiring φb0847 sequences from the wild type strain (Figure S2). Interestingly, a yfdK homologue was recently shown to aid the survival of a K-12 laboratory strain of E. coli in acidic environments [54], but no mechanism for this effect is known, and to the authors' knowledge, yfdK homologues have not been implicated in pathogenesis.

The neaT gene restores virulence to the φb0847 island deletion mutant

The in vivo assays presented in Figure 3C and bioinformatic analyses described below highlight neaT as a gene of potential importance to the fitness and virulence of CFT073. To test this possibility, the neaT locus, including an upstream promoter region of 211 bp, was amplified from the CFT073 chromosome and cloned into the high-retention plasmid pGEN-mcs, yielding pGEN-neaTPnative. Semi-quantitative reverse transcription polymerase chain reaction (RT-PCR) indicated that neaT transcript levels made from the pGEN-neaTPnative vector in broth culture were about 1.7-fold higher than those observed in wild type CFT073 (Figure S3). Complementation experiments were performed comparing the lethality of wild type CFT073/pGEN-mcs, Δφb0847/pGEN-mcs, and Δφb0847/pGEN-neaTPnative in zebrafish embryos after inoculation of the CV (Figure 3D). The complete prophage deletion mutant Δφb0847 carrying the empty vector pGEN-mcs exhibited a significant delay in killing relative to either the wild type strain CFT073/pGEN-mcs or the complemented mutant Δφb0847/pGEN-neaTPnative. In total, these experiments identify neaT as a virulence determinant contained within the φb0847 island of CFT073; therefore, the uncharacterized neaT gene became the focal point for the remainder of our investigation.

neaT is required for fitness during systemic, but not localized infections in a mammalian host

To extend our observations made using zebrafish, we employed a murine model to further define the requirement for neaT during localized and systemic infections. For localized challenges, we took advantage of a well-characterized mouse model of urinary tract infection. Wild type CFT073 and the ΔneaT mutant were mixed at a one-to-one ratio and inoculated via transurethral catheterization into adult female CBA/J mice. After 3 days, animals were sacrificed and bacterial titers within the bladders and kidneys were enumerated, revealing no outright competitive advantage for wild type CFT073 over the ΔneaT mutant in either organ (Figure 4A). To appraise the requirement for neaT during systemic infections, we utilized a recently described sub-lethal bacteremia model in which CBA/J mice were injected with a one-to-one mixture of wild type and mutant bacteria via the tail vein [55]. At 24 hpi the ΔneaT mutant was recovered at significantly reduced levels from the spleen and liver compared to wild type CFT073 (Figure 4B). These results confirm and extend our findings in the zebrafish host, demonstrating that neaT provides niche-specific advantages to CFT073 during systemic infections.

thumbnail
Figure 4. neaT enhances the fitness of CFT073 in a murine model of bacteremia.

(A) Equal numbers (107 CFU total) of wild type CFT073 and CFT073ΔneaT were transurethrally inoculated into the bladder of CBA/J female mice. Mice were sacrificed, organs harvested, and bacterial loads enumerated at 3 d post inoculation. (B) Equal numbers (106 CFU total) of wild type and CFT073ΔneaT were inoculated into the bloodstream of CBA/J female mice via tail vein injection and bacterial titers present in the spleen and liver were enumerated 24 h later. Data are shown as competitive indices, where negative values indicate a reduction in the fitness of CFT073ΔneaT. Bars indicate median values for each group; n≥9 mice. P values determined using Wilcoxon-matched paired signed rank; ns = not significant.

https://doi.org/10.1371/journal.ppat.1003175.g004

Diversity and phage association of NeaT homologues

There are no closely related homologues of NeaT in E. coli. Only four matches were found in the current NCBI collection of 170 RefSeq E. coli genomes (as of June 2012) that produce an alignment E value<10−6 with similarity over >50% of the NeaT protein length. A PCR-based survey for the presence of neaT within various clinical E. coli isolates corroborated our in silico observation that neaT is rare among this taxon (Figure S4). Out of 21 randomly chosen isolates, none carried the CFT073 neaT allele. Homologues of neaT are also rarely detected in P2-like phage genomes; among 45 randomly chosen P2-like phages and prophages in E. coli, Salmonella, Shigella, and Enterobacter that we examined, only φb0847 carries a neaT-like gene. The closest match to NeaT in the NCBI database is encoded by locus Ent638_2581 of Enterobacter sp. 638, whose protein product is only about 33% identical to NeaT. We note that several more distantly related neaT homologues are present in the genomes of other temperate phages and prophages of the bacterial family Enterobacteriaceae (Table 2). They are found, for example, in the Shigella flexneri phage Sf6 genome and several uncharacterized prophages of S. flexneri, in E. coli phage φV10 and a nearly identical prophage in the Shiga toxin-producing E. coli isolate DEC4D, and in a putative prophage within Citrobacter rodentium strain ICC168. The above Enterobacter homologue Ent638_2581 is also carried within a putative prophage that is similar to Shigella phage SfV. Each of these phage-associated neaT homologues is un-ameliorated with respect to its bacterial host genome (see below), and each lies within a known moron position in its phage genome. Because neaT homologues differ substantially in sequence conservation and are found in a variety of tailed-phages, neaT-like genes may have been laterally acquired by Enterobacteriaceae lineages on several occasions, possibly via phage. Multiple neaT acquisition events would indicate that this gene has an underlying evolutionary importance to either the phages themselves or their hosts. In considering its putative function (see Figures S5, S6, and Text S1), its apparent lateral acquisition, and its allelic variance within the proteobacterial lineage, this gene was named ‘neaT’—nomadically evolved acyltransferase. In the following sections we explore the evolutionary history of this gene by analyzing the diversity and distribution of neaT-like genes in more detail.

Phyletic distribution of neaT

To investigate the evolutionary source of E. coli neaT genes, we assessed the phyletic distribution of its homologues. BLASTp alignments were performed on the publically available NCBI database using NeaT from CFT073 as a probe for the search sets of Proteobacteria, Firmicutes, Bacteroidetes, Actinobacteria, Spirochaetes, and Fusobacteria [56]. Sequences were declared to be homologous if they had an alignment significance (E value) of <10−6 over >50% of their lengths [57]. These searches retrieved a total of 317 non-paralogous NeaT-like sequences. The distribution of phyla containing these sequences is depicted in Figure 5A (left), revealing that the majority of neaT homologues are from the Firmicutes and Bacteroidetes rather than Proteobacteria. In an attempt to control for the inherent bias of NCBI databases, we plotted the number of available gene sequences for each phylum represented in Figure 5A (right). This plot demonstrates that the high number of neaT homologues identified among non-proteobacterial phyla is not due to a skew in sequence abundances. To the contrary, total proteobacterial gene sequences overshadow those from other phyla and therefore underscore the relative rarity of neaT alleles in this taxon.

thumbnail
Figure 5. neaT is discordantly conserved.

(A) Left: Phyletic distribution of neaT homologues among genomes retrieved from NCBI (n = 317). Right: Number of gene sequences deposited in NCBI for each phylum as of November 2011. (B) Left: Pie graph showing distribution of neaT homologues among each phylum represented in the custom database (n = 21 non-paralogous neaT genes). Right: Theoretical distribution of neaT homologues within phyla present in the custom database based on random chance. P values for the observed versus theoretical phyletic abundance of neaT homologues were calculated by .score analysis. (C) Upper y-axis: bar graph depicts total number of non-paralogous homologues retrieved from the custom database (DB) for each gene encoded within φb0847 (plotted along the x-axis with respect to its position within the prophage genome). The neaT open reading frame is indicated in red. Lower y-axis: bar graph showing the percent of proteobacterial homologuesfound in the homologue set for each φb0847 gene. Sequences unique to CFT073 were assigned 100% proteobacterial conservation.

https://doi.org/10.1371/journal.ppat.1003175.g005

To quantify the phyletic distribution of NeaT homologues with greater statistical confidence, we performed bi-directional alignments of NeaT using BLASTp with a manually assembled database of open reading frames from a representative, yet broad, assortment of 165 phylogenetically classified bacterial genomes and associated plasmids obtained from NCBI (Table S3). This analysis confirmed that, compared to random chance, neaT homologues are significantly enriched among species belonging to the phyla Firmicutes and Bacteroidetes (Figure 5B). Moreover, many of the neaT homologues were detected in notable plant and animal pathogens, including Erwinia spp., Bacillus spp., Staphylococcus aureus, Streptococcus oralis, Clostridium botulinum, and Porphyromonas spp.

Results from similar alignments of neaT and all other φb0847-encoded genes are presented graphically in Figure 5C. For each prophage gene, the number of non-paralogous matches found in the custom database are represented as bars (upper axis) and the percent of those hits that are harbored within proteobacterial genomes (lower axis) are plotted against the position of the gene within φb0847 (x-axis). Given the host range of known P2-like phages, it is not unexpected that the majority of genes within φb0847 were exclusive to the proteobacterial phylum. Exceptions, in addition to neaT, include homologues of φb0847 genes encoding the phage integrase and a Dam methylase. However, neaT is unique among the φb0847 prophage genes in that over 75% of its matches (16 of 21) were from outside the Proteobacteria (Figure 5C and Table S4). The discordant conservation of neaT highlights its likely extra-phyletic origin.

neaT displays signatures of recent lateral transfer

If a gene has origins outside its immediate genome, it would carry sequence-level vestiges of its previous host until it adopts the characteristics of the current host—a process known as ‘amelioration’ [27], [58]. Commonly used parameters that distinguish laterally transferred and unameliorated genes are atypical codon usage and guanine-cytosine (G+C) content [57], [59], [60]. We analyzed these features of neaT in the context of the CFT073 genome and φb0847 prophage. Using all 5,369 protein-coding genes of CFT073, the frequency with which specific codons are used for each amino acid was calculated (Table S5). Each gene was then assigned a ‘codon deviation score’ representing how often it uses atypical codons (Methods and Table S6). Scoring correlates with conformity; genes scoring low have a more typical codon usage, whereas poorly conformed genes score high. This analysis shows that neaT possesses a significantly abnormal codon usage compared to the rest of the CFT073 genome (p = 0.0260) (Figure 6A, left panel). The neaT gene was also observed to be G+C-poor (29.84%), making it a significant outlier from the CFT073 genome-wide median of 51.5% (p = 0.0001) (Figure 6A, right panel). We also analyzed the codon deviation score (Figure 6B, upper axis) and nucleotide composition (Figure 6B, lower axis) of neaT with respect to the genome of φb0847. Most genes withinφb0847 conform to the codon usage and G+C content of CFT073. This is expected for a parasite that has been co-evolving with proteobacterial hosts over an extensive period of evolutionary time [59]. Thus, the aberrant codon usage and nucleotide composition of neaT is not simply an inherited trait of φb0847. Because of its relatively low G+C content and poorly conformed codon usage, we conclude that neaT is a relatively recent acquisition by both φb0847 and the genome of CFT073.

thumbnail
Figure 6. neaT is maintained in an un-ameliorated state.

(A) Left: distribution of codon deviation scores assigned to the 5,369 protein-coding genes of CFT073. Right: distribution of %GC content of each protein-coding gene of CFT073. Bar and whiskers indicate median and interquartile ranges. Red ‘X’ marks position of neaT within each distribution. (B) Upper y-axis: bar graph depicting codon deviation score for each φb0847 gene plotted with respect to position within the prophage (x-axis cartoon, with the neaT gene highlighted in red.). Lower y-axis: line graph representing fluctuations in %GC content across φb0847 (window = 182 bp, step = 182 bp). (C) %GC content of neaT alleles found across phyla (color-coded) plotted against the %GC content of each respective genome (x-axis). Inset shows the same sort of analysis for the poxB allele as a comparison. Points falling on the dashed lines represent alleles that are completely ameliorated with respect to their host genomes. (D) Graph shows the ratio of %GC content of neaT alleles and total genomic %GC content for individual isolates within the indicated phyla. Bars indicate medians, and p values were determined using two-tailed Mann-Whitney t tests (n = 18 (Proteobacteria), 51 (Firmicutes), 22 (Bacteroidetes)).

https://doi.org/10.1371/journal.ppat.1003175.g006

To determine if the apparently unameliorated state of neaT in CFT073 is unique or if it is hinting at a more widespread phenomenon, we plotted the G+C content of a representative subset of neaT homologues identified in Figure 5 against the G+C content of their respective genomes (Figure 6C). As a control, we also plotted the G+C content of poxB, which encodes the metabolic enzyme pyruvate oxidase and exists in an ameliorated state within several phyla (Figure 6C, inset). Most proteobacterial neaT genes were significantly less ameliorated than those found in the genomes of Bacteroidetes and many Firmicutes (Figure 6C and D). Interestingly, even though Firmicutes genomes generally have a low G+C content, neaT-like genes within this lineage are still relatively G+C-poor, at least in a major fraction of Firmicute species (Figure 6C). Cumulatively, these results indicate that, at least among the three phyla compared here, neaT-like genes have likely been associated with Bacteroidetes the longest, whereas acquisition by the Proteobacteria was a more recent event.

Discussion

Summary and impact of findings

Presented here are the results from a screen conducted using the ExPEC isolate CFT073 and a high-throughput zebrafish surrogate host model of infection. We screened GIs for novel virulence genes, which were expected to have a history of lateral gene transfer. Three loci within the P2-like prophage φb0847 were found to contribute to the virulence of CFT073 during systemic infection. A previously uncharacterized gene—designated here as neaT—was discovered to augment the virulence capacity of CFT073, independent of other prophage components (Figure 3D). We demonstrated that neaT is conditionally required for maximal fitness during bacteremic infections of both zebrafish and mice, suggesting that CFT073 has potentially co-opted this phage-borne gene for specific virulence behaviors. By tracing the evolutionary history of the neaT gene, we found that it is relatively rare and has sequence-based features suggesting that it was recently absorbed into the proteobacterial supraspecies pan-genome. Signs of its novelty are typified by high allelic variance—possibly a result of multiple entries into the Proteobacteria lineage via phage—and its mostly unameliorated state within proteobacterial genomes.

We also investigated the putative function(s) of NeaT in vitro. The NeaT protein shares homology with several characterized acyltransferases encoded within a variety of non-E. coli genomes. These putative membrane-localized enzymes can modify components of the bacterial cell wall, particularly peptidoglycan [61], [62], [63], [64]. Alteration of this macromolecule can often provide bacterial pathogens with protection from host antimicrobial peptides and enzymes such as lysozyme. However, deletion of neaT had no effect on the sensitivity of CFT073 to lysozyme, the antimicrobial cationic peptide polymyxin B, or antibacterial factors present in human serum (see accompanying supplemental Text S1). Interestingly, expression of neaT did alter the behavior of CFT073 in swarming assays and induced bacterial aggregation on swim plates (Figure S5A–C, Text S1)—phenotypes that may be attributable to NeaT-mediated modification of components within the bacterial envelope. We also found that expression of recombinant NeaT can inhibit production of surface structures like curli and cellulose in some strain backgrounds (Fig. S5D–E, Text S1), supporting the notion that NeaT can affect salient properties of the bacterial surface and thereby alter bacterial group behavior.

The apparent capacity of NeaT to modulate bacterial aggregation (Fig. S5C) is especially intriguing in light of a recent work demonstrating that aggregate formation can promote bacterial survival within the bloodstream of infected mice [65]. Building on these observations, we found that expression of the neaT gene from a low copy number plasmid significantly decreased the capacity of CFT073 to associate with murine macrophages, suggesting that NeaT serves as an immune evasion factor (Fig. S6). The specific mechanism(s) by which NeaT promotes bacterial fitness during systemic infections, as well as the environmental cues that control neaT expression, require further investigation. As it stands, this work contributes to the idea that ExPEC isolates do not all share the same set of virulence factors, which are likely dictated by the distinct evolutionary trajectory and particular niche tropism of each strain.

neaT-based models for evolution of laterally acquired genes

Our analysis defines neaT as a recently acquired locus of the Proteobacteria. Evidence for this is drawn from its discordant conservation, abnormal codon usage, and low G+C content. In large part, the unameliorated state of neaT-like genes in Proteobacteria and Firmicutes genomes suggests that there is a general phenomenon accounting for its relative A+T-rich composition beyond having originated in an A+T-rich genome, as previously suggested [59]. We posit that the observed A+T-richness of laterally transferred genes can be, to some extent, accounted for by an ‘exploratory mechanism’ [16]. Upon introgression of a foreign gene, its retention depends on its adaption to the host's genetic and cellular machinery, a process that can take several millions of years [66]. During this time the gene may fall under relaxed selection whereby mutations accrue until a beneficial allele is ‘discovered’ and acted upon by selection. Connecting relaxed selection to reduced G+C content is the observation that there is a universal mutation bias for G/C to A/T transitions in bacterial genomes [67], [68], [69]. It then follows that immediately after a gene is acquired, it will initially accumulate A+T-rich character until a selectable version is ameliorated. From the findings presented here, we speculate that the neaT variant in CFT073 is an example of a newfound allele that is being used to promote bacterial fitness in pathogenic contexts.

Arguably, neaT may represent an ancient gene that has simply failed to fix within the proteobacterial lineage. Therefore, an alternative hypothesis is that the conditional requirement for neaT by CFT073 within different environments may have driven its current evolved state. We observed in two vertebrate model systems that neaT contributes significantly to pathogen fitness primarily during systemic infections. Considering the ecology of many bacterial pathogens, a question often left unaddressed is: what are the evolutionary forces that act on niche-specific genes in the absence of selective pressure? Particularly for E. coli, which has a complex multi-niche life cycle, the evolutionary consequences resulting from time outside selective environments on genes like neaT are not clear.

Work directly addressing this question is scarce. However, insight into this issue is provided by findings that genes under relaxed constraint have increased variance at the sequence level [17], [70], [71], [72], [73]. In contrast to relaxed selection, which occurs when purifying selection is alleviated, as discussed above, ‘relaxed constraint’ refers to a limitation in the exposure of a particular gene to selection. For example, eukaryotic genes with expression patterns that are sex-restricted are effectively ‘hidden’ from selection in half of the population. This is the case for the Drosophila spp. maternal-effect gene bicoid, which is maternally-restricted and critical for the embryonic development of fruit flies [70]. The bicoid gene was found to have a 2-fold higher heterozygosity compared to zygotically-expressed genes. Similarly, genes with caste-biased expression (i.e., queen versus worker) in the social insects Solenopsis invicta (fire ant) and Apis mellifera (honey bee) were shown to be evolving more rapidly than genes expressed among all castes [71]. For both of these situations, the higher mutation rate observed for contextually expressed genes was concluded to be due to relaxed constraint. Further investigation into the exploratory mechanism and relaxed constraint hypotheses of neaT evolution is required and must be considered in parallel with other processes and factors, including, for example, the susceptibility of laterally transferred genes to endogenous restriction enzymes [59].

ExPEC individuality and virulence

There exists an enormous amount of genetic heterogeneity among Eubacteria lineages. Genome sequencing and bioinformatic analyses have underscored this extensively. Perhaps the most intriguing aspect of this diversity is that even closely related members of the same species can differ greatly with respect to their gene contents. Strikingly, any two E. coli genomes can differ by up to 20–30% of their respective gene contents—in sharp contrast to the relatively minor difference of 1% that exists between, for example, the mouse and human species [32], [74]. Decades worth of epidemiological and experimental studies have focused on the identification of genes that define the pathogenic behavior of ExPEC [7], [32]. However, it appears that a single, ubiquitous genetic identifier of ExPEC, such as a gene encoding a particular toxin or adhesin, does not exist and, rather, what actually binds these pathogens is more qualitative and multigenic in nature [32], [75].

In support of this view, we recently demonstrated that the toxin α-hemolysin, shared among many ExPEC isolates, is differentially required for virulence depending on strain background [40]. Similarly, we found that the pathogenicity of particular ExPEC isolates depends on another toxin, cytotoxic necrotizing factor, while other equally virulent strains naturally lack this gene. Coupled with the work presented here, these observations suggest that there exists a spectrum of only partially overlapping virulence gene requirements among ExPEC, reflecting the idea that these pathogens have emerged from distinct evolutionary trajectories driven by LGT [76], [77]. Accordingly, we found that the expression of NeaT from plasmid pGEN-neaTPnative in other E. coli strains, including Nissle 1917 (gut isolate), F11 (cystitis isolate), and S88 (meningitis isolate), did not augment virulence in the zebrafish infection model (data not shown). These findings suggest that the ability of a rare gene like neaT to affect fitness and virulence is dependent upon the genetic background of individual bacterial strains. The beneficial effects of neaT, and its potential to sweep through bacterial populations, is therefore likely linked to the presence, or coordinate acquisition, of other as-yet undefined bacterial factor(s). The identification, characterization, and continued monitoring of rare genes like neaT will be important to our understanding of ExPEC evolution. As a case in point, we note that the sasX gene, originally defined as rare among strains of methicillin resistant Staphylococcus aureus (MRSA), increased in prevalence among MRSA isolates between 2003 and 2011 and is now considered an emerging virulence determinant [78]. Interestingly, like neaT, sasX is also maintained within a prophage and can affect bacterial interactions with phagocytes. At this point, it is difficult to predict if neaT will sweep ExPEC populations in the future, but work presented here along with recent findings concerning sasX underscore how laterally acquired genes can alter the virulence potential of bacterial pathogens, continually challenging the development of broad spectrum therapeutics.

Going forward, as we continue to characterize the composition of pan-genomic elements of ExPEC and other pathogens, it will be important to consider the evolutionary context of their virulence genes. Identifying the spatial and temporal parameters that govern the lateral acquisition of virulence genes from distant lineages will need to be reconciled. Genome compatibility (codon and tRNA usage) and ecology are thought to be influential in the success of LGT events between bacteria [79], [80], [81]. In light of this, several interesting questions arise. How did neaT come to be in the proteobacterial gene pool? How does residence of neaT within a prophage impact its evolution? What conditions fostered the assimilation of neaT into the virulence regulon of its host? Using neaT as a stepping-stone, it will be informative to resolve the amount of strain-specific innovation that goes into producing and fine-tuning pathogen genomes. By understanding the mechanisms of chromosome assembly and the sources of individual genetic components, unrealized patterns may emerge that could prove useful for future diagnostics and disease mitigation.

Methods

Ethics statement

Animals used in this study were handled in accordance with IACUC protocols approved at either the University of Utah or the University of Michigan Medical School following standard guidelines as described at www.zfin.org and in the Guide for the Care and Use of Laboratory Animals, 8th Edition [55], [82].

Bacterial strains and plasmids

All bacterial strains and plasmids used in this study are listed in Table 3. Unless specified otherwise, bacteria were cultured statically at 37°C for 24 h in 20 ml of a defined M9 minimal medium (6 g/l Na2HPO4, 3 g/l KH2PO4, 1 g/l NH4Cl, 0.5 g/l NaCl, 1 mM MgSO4, 0.1 mM CaCl2, 0.1% glucose, 0.0025% nicotinic acid, 0.2% casein amino acids, and 16.5 mg/ml thiamine in H2O). Antibiotics (kanamycin or ampicillin) were added to the growth medium when necessary to maintain recombinant plasmids or select for mutants.

Targeted gene knockouts were generated in the ExPEC isolate CFT073 using the lambda Red-mediated linear transformation system [83], [84]. Briefly, a kanamycin resistance cassette was amplified using polymerase chain reaction (PCR) from pKD4 with 40-base pair overhangs specific to the 5′ and 3′ ends of each targeted locus. PCR products were introduced via electroporation into CFT073 carrying pKM208, which encodes an IPTG (isopropyl-β-D-thiogalactopyranoside)-inducible lambda red recombinase. Knockouts were confirmed by PCR. Primer sets used are listed in Table S8.

Cloning and construction of neaT expression constructs were done using standard molecular techniques employing the high-retention plasmid pGEN-mcs [85]. For native regulation, neaT (locus tag: c0970), plus 211 bp of upstream sequences, were amplified from the chromosome of CFT073 and TA-cloned into pCR2.1-TOPO vector per manufacture's protocol (Invitrogen). Subsequently, the cloned fragment was isolated using BamHI and NotI restriction enzymes (New England Biosciences) and ligated into pGEN-mcs using the same sites, yielding pGEN-neaTPnative. For construction of pGEN-neaTPlac, a synthetic ribosome binding sequence was introduced upstream of neaT within the 5′ PCR primer, and the resulting PCR product was ligated via an engineered NdeI restriction site with the lac promoter amplified from pGFPmut3.1 (Clonetech). The ligated Plac-neaT product was amplified and TA-cloned into the pCR2.1-TOPO vector. Using BamHI and NcoI restrictions sites, the Plac controlled neaT variant was then sub-cloned into pGEN-mcs. All experiments involving pGEN-neaTPlac were performed without IPTG induction. Primer sequences used to generate these plasmids are listed in Table S8.

Zebrafish embryos

*AB wild-type zebrafish embryos were collected from a laboratory-breeding colony that was maintained on a 14-h/10-h light/dark cycle. Embryos were grown at 28.5°C in E3 medium (5 mM NaCl, 0.17 mM KCl, 0.4 mM CaCl2, 0.16 mM MgSO4) containing 0.000016% methylene blue as an anti-fungal agent.

Infection of zebrafish embryos

One ml from each 24 h bacterial culture was pelleted, washed once with 1 ml sterile PBS (Hyclone) and re-suspended in 1 ml PBS to obtain appropriate bacterial densities for microinjection. Prior to injection, 48 hpf embryos were manually dechorionated, briefly anesthetized using 0.77 mM ethyl 3- aminobenzoate methanesulfonate salt (tricaine) (Sigma-Aldrich), and embedded in 0.8% low-melt agarose (MO BIO Laboratories) without tricaine. Approximately 1 nl of bacteria was injected directly into the pericardial cavity or the blood via the circulation valley located ventral to the yolk sac using a YOU-1 micromanipulator (Narishige), a Narishige IM-200 microinjector, and a JUN-AIR model 3-compressor setup. For each experiment, average CFU introduced per injection were determined by adding 10 drops of each inoculum into 1 ml 0.7% NaCl, which was then serially diluted and plated on Luria-Bertani (LB) agar plates. For co-challenge experiments, input doses were plated on LB agar+/−kanamycin (50 µg/ml) to determine relative numbers of the wild type and mutant strains present. After injection, embryos were carefully extracted from the agar and placed individually into wells of a 96-well microtiter plate (Nunc) containing E3 medium lacking both tricaine and methylene blue. For lethality assays, fish were examined at indicated times over the course of a 48 or 72 h period and scored for “death”, defined here as the complete absence of heart rhythm and blood flow. Survival graphs depict total pooled results from two or more independent experiments in which groups of 10 to 20 embryos were injected. To quantify bacterial numbers during the course of co-challenge experiments, embryos were homogenized at the indicated time points in 500 µL PBS containing 0.5% Triton X-100 using a mechanical PRO 250 homogenizer (PRO Scientific). Homogenates were then serially diluted and plated on LB agar+/−kanamycin (50 µg/ml) to determine relative numbers of wild type and mutant bacteria.

Mouse infections

For co-challenge during urinary tract infection, seven- to nine-week old female CBA/J mice (Jackson Labs) mice were anesthetized using isoflurane inhalation and inoculated via transurethral catheterization with 50 µl of a 1∶1 wild type to mutant bacterial suspension containing a total of 107 bacteria suspended in PBS. Bladders and kidneys were recovered 3 days later and each was weighed and homogenized in 1 ml containing 0.025% Triton X-100. Homogenates were serially diluted and plated on LB agar+/−kanamycin (50 µg/ml) to determine number of both wild type and mutant bacteria. Mouse experiments were repeated at least twice.

For systemic infections, female CBA/J mice (Jackson Labs) aged 6 to 8 weeks were restrained using a Universal Restrainer (Braintree Scientific, Braintree, MA) and inoculated via the tail vein over a 30 s period with a 100 µl bacterial suspension, delivering 106 CFU/mouse. The inoculum was prepared by re-suspending overnight cultures in PBS and diluting them to 1×107 CFU/ml. For co-challenges, wild type and mutant suspensions were mixed 1∶1 before inoculation. Perfusion was performed on euthanized animals by cutting a small hole in the right cardiac ventricle and infusing the left ventricle slowly with 40 ml 0.9% sterile saline before organ removal. Blanching of the organs occurred with the first 20 ml of sterile saline. Excised spleens and livers were homogenized in 3 ml PBS using a mechanical homogenizer (Omni International, Marietta, GA), and homogenates were plated using an Autoplate 4000 (Spiral Biotech, Norwood, MA) onto LB agar+/−kanamycin (50 µg/ml) to differentiate wild type and mutant strains.

Statistical analysis of zebrafish and mouse infections

Kaplan-Meier survival and scatter plots were generated using GraphPad Prism 5. For Kaplan-Meier survival plots (independent challenges), the log-rank (Mantel-Cox) test was used to determine statistical differences between datasets. For competitive assays (co-challenges), numbers of wild type and mutant bacteria present in the inoculum and recovered from host tissues were determined as described above and a competitive index was calculated using the following equation where wt represents numbers wild type bacteria:Negative values obtained using the competitive index equation indicate a reduction in mutant fitness. To determine statistical significance, the Wilcoxon signed-rank test (with a hypothetical value of 0) on log-transformed competitive index values was used for co-challenges and the two-tailed Mann-Whitney statistical analysis was performed to determine significant differences between samples in non-competitive assays.

Bioinformatic analyses

Homology searches and phyletic enrichment of homologue sets.

A custom database of 165 genomes and associated plasmids was assembled using the compilation of protein coding genes (.faa files downloaded from ftp://ftp.ncbi.nih.gov/genomes/Bacteria/) of each isolate listed in Table S3. BLASTp (v. 2.2.20, [56]) was used to run bidirectional protein alignments between the φb0847 genome and the database to identify homologues. Two sequences were considered homologous if they aligned along >50% of their lengths with an E value of <10−6. To identify phyletic enrichment, sets of non-paralogous homologues for each φb0847 gene were analyzed for relative contributions made by each phylum. Then, based on the number genes in each homologue set, the same number of genomes was randomly sampled from the genome list in Table S3. In this way, we could determine the significance of the phyletic contributions to each homologue set that was observed compared to a theoretical random sampling. With custom software written in Python using SciPy, p values were generated from a Z score. Standard scores were calculated using the equation below, where x is the observed proportion contributed by a single phylum, μ is the theoretical average contribution by the same phylum (n = 1000 random samplings), and σ is the standard deviation: z = (x−μ)/σ

Sequences used for comparisons between the GC content of neaT from CFT073 and homologues in other bacteria (see Figure 6C) were retrieved manually from NCBI for downstream analysis. Genome GC compositions were obtained from NCBI Genomes (http://www.ncbi.nlm.nih.gov/genomes/MICROBES/microbial_taxtree.html)

Nucleotide composition analysis.

For nucleotide composition analysis, the nucleotide sequences of protein coding genes of CFT073 (.ffn files downloaded from ftp://ftp.ncbi.nih.gov/genomes/Bacteria/Escherichia_coli_CFT073_uid57915/) were used to calculate codon deviation scores and GC content using custom software written in Python with SciPy or NumPy. For codon deviation scores, genome-wide protein coding nucleotide sequences were analyzed for codon usage frequencies on a per amino acid basis. The resulting table (Table S5) was then used to determine differences between specific codon frequencies contained within a particular gene and the genome-wide frequency. The absolute values of differences in frequency were summed over a single gene to obtain the final codon deviation score. Statistical significance was determined by Z-score analysis using the genome-wide mean codon deviation score and standard deviation. Table S6 lists all deviation scores and p values for the CFT073 genome. GC content of genes was determined by counting the proportion of guanines and cytosines over the length of a given locus, and Z-score analysis was again implemented to determine the position of each gene within the genome-wide distribution (Table S7).

Supporting Information

Figure S1.

Expression of the neaT gene in various mutant backgrounds. RNA was extracted from the indicated strains after overnight growth in M9 medium and used to generate cDNA libraries by reverse transcription (+RT). To control for genomic DNA contamination, a set of samples was prepared in parallel without reverse transcriptase (−RT). Wild type CFT073, CFT073ΔneaT, CFT073ΔyfdK, and CFT073ΔFI-D were used to determine the relative expression levels of neaT in each genetic background. Three µg of each cDNA library was used as a template for PCR amplification (30 cycles) of an internal 218 bp fragment of neaT. Equal amounts of each PCR reaction were resolved using 1% agarose gels.

https://doi.org/10.1371/journal.ppat.1003175.s001

(TIF)

Figure S2.

Determination of in vivo lateral transfer of the neaT gene. Zebrafish were inoculated with a one-to-one mixture of wt CFT073 and CFT073ΔneaT. Infections progressed for ∼12 h post-inoculation prior to homogenization and recovery of bacteria by plating on LB agar+/−kanamycin. Bacterial colonies recovered from 5 separate fish were used for colony PCR to detect presence of either the kanamycin resistance gene (lane 1 control, ∼1,500 bp) or neaT (lane 2 control, 218 bp internal fragment). No double positive colonies were detected. Primers used to amplify the kanamycin gene are specific to the priming regions of the pKD4 template plasmid. neaT was amplified using neaT RT forward/reverse (Table S8).

https://doi.org/10.1371/journal.ppat.1003175.s002

(TIF)

Figure S3.

Plasmid-based neaT expression analysis. RNA was extracted from the indicated strains after overnight growth in LB broth and used to generate cDNA libraries by reverse transcription (+RT). To control for genomic DNA contamination, a set of samples was prepared in parallel without reverse transcriptase (−RT). Wild type (WT) CFT073 or HS were used to reference basal neaT message levels. CFT073ΔneaT or HS carrying pGEN-mcs (empty vector, EV), pGEN-neaTPnative (native promoter, NP), or pGEN-neaTPlac (over-expressing, OE) were used to determine the relative expression levels of pGEN-neaT variants in each genetic background. Three µg of each cDNA library was used as a template for PCR amplification (28 cycles) of an internal 218 bp fragment of neaT. Equal amounts of each PCR reaction were resolved using 1% agarose gels. Graph shows average levels of neaT transcripts ± SD normalized to 16S rRNA (not shown). Data are presented relative to WT CFT073, n = 3.

https://doi.org/10.1371/journal.ppat.1003175.s003

(TIF)

Figure S4.

Survey of clinical isolates for presence of the neaT gene. Various clinical E. coli isolates were surveyed for presence of the neaT gene using polymerase chain reaction. Primers used in (A) amplified a 218 bp region internal to neaT (Table S8). (B) Shows amplification of the 16s ribosomal RNA gene as a control. Isolates are described as: strain (clinical disease presentation).

https://doi.org/10.1371/journal.ppat.1003175.s004

(TIF)

Figure S5.

neaT contributes to multicellular behaviors. (A) Swarm motility of wild type (wt) CFT073 and its mutant derivatives on 0.25% Eiken agar plates following overnight incubation at 37°C. (B) Complementation of swarm defect of CFT073ΔneaT by introduction of pGEN-neaTPlac. The empty vector pGEN-mcs and pGEN-neaTPnative did not complement the ΔneaT mutant. (C) Swim motility of indicated CFT073 derivatives following incubations at 37°C for times indicated. Red arrows indicate advancing swim fronts and insets show magnified bright field images of the center region of each plate. (D) Images of single Nissle 1917 colonies carrying pGEN-mcs or pGEN-neaTPlac grown for 48 h at 37°C on agar plates containing 0.001% Congo red dye to stain curli fibers. (E) Streaks of Nissle 1917 carrying pGEN-mcs, pGEN-neaTPnative, or pGEN-neaTPlac grown overnight at 37°C on 1.2% LB agar containing 50 µg/ml Fluorescent Brightener 28 to visulalize cellulose production. Image was captured under ultraviolet light.

https://doi.org/10.1371/journal.ppat.1003175.s005

(TIF)

Figure S6.

NeaT limits bacterial interactions with murine macrophages. (Left) The indicated bacterial strains were added to bone marrow derived macrophage (BMDM) monolayers at a multiplicity of infection of 10. After a 1-h incubation at 37C, total viable bacteria remaining in the wells were enumerated. (Right) Alternatively, monolayers were washed at the 1-h time point with PBS, prior to lysis, in order to determine numbers of macrophage-associated bacteria. Bars represent the means ± SD of three independent experiments performed in triplicate. *p<0.05, **p<0.01; as determined by Student's t test.

https://doi.org/10.1371/journal.ppat.1003175.s006

(TIF)

Table S1.

Summary of results obtained from initial GI screen using different dose ranges.

https://doi.org/10.1371/journal.ppat.1003175.s007

(XLSX)

Table S2.

Summary of inconclusive in vitro experiments.

https://doi.org/10.1371/journal.ppat.1003175.s008

(XLSX)

Table S3.

List of strains contained within the custom 165 genome database.

https://doi.org/10.1371/journal.ppat.1003175.s009

(XLSX)

Table S4.

List of strains/genomes from the 165 genome database that contain at least one neaT homologue.

https://doi.org/10.1371/journal.ppat.1003175.s010

(XLSX)

Table S5.

Codon usage frequency for each amino acid based on the 5,369 protein-coding genes of CFT073.

https://doi.org/10.1371/journal.ppat.1003175.s011

(XLSX)

Table S6.

Codon deviation score for each gene in CFT073.

https://doi.org/10.1371/journal.ppat.1003175.s012

(XLSX)

Table S7.

GC content for each gene in CFT073.

https://doi.org/10.1371/journal.ppat.1003175.s013

(XLSX)

Table S8.

Primers used in this study to generate recombinant strains and plasmids.

https://doi.org/10.1371/journal.ppat.1003175.s014

(XLSX)

Text S1.

Expression of NeaT may alter bacterial group behavior.

https://doi.org/10.1371/journal.ppat.1003175.s015

(DOCX)

Acknowledgments

We thank Dr. Nels Elde (University of Utah) for his enlightening discussions and Dr. Kael Fischer (University of Utah) for his guidance throughout the bioinformatic analyses.

Author Contributions

Conceived and designed the experiments: TJW SRC MAM. Performed the experiments: TJW JPN SNS SRC AJL. Analyzed the data: TJW SRC MAM. Contributed reagents/materials/analysis tools: TJW JPN SNS HLTM SRC MAM. Wrote the paper: TJW JPN SRC MAM.

References

  1. 1. Ley RE, Peterson DA, Gordon JI (2006) Ecological and evolutionary forces shaping microbial diversity in the human intestine. Cell 124: 837–848.
  2. 2. Savageau MA (1983) Escherichia-Coli Habitats Cell Types and Molecular Mechanisms of Gene Control. American Naturalist 122: 732–744.
  3. 3. Walk ST, Alm EW, Calhoun LM, Mladonicky JM, Whittam TS (2007) Genetic diversity and population structure of Escherichia coli isolated from freshwater beaches. Environmental microbiology 9: 2274–2288.
  4. 4. Winfield MD, Groisman EA (2003) Role of nonhost environments in the lifestyles of Salmonella and Escherichia coli. Applied and environmental microbiology 69: 3687–3694.
  5. 5. Fricke WF, Wright MS, Lindell AH, Harkins DM, Baker-Austin C, et al. (2008) Insights into the environmental resistance gene pool from the genome sequence of the multidrug-resistant environmental isolate Escherichia coli SMS-3-5. Journal of bacteriology 190: 6779–6794.
  6. 6. Luo C, Walk ST, Gordon DM, Feldgarden M, Tiedje JM, et al. (2011) Genome sequencing of environmental Escherichia coli expands understanding of the ecology and speciation of the model bacterial species. Proceedings of the National Academy of Sciences of the United States of America 108: 7200–7205.
  7. 7. Wiles TJ, Kulesus RR, Mulvey MA (2008) Origins and virulence mechanisms of uropathogenic Escherichia coli. Experimental and molecular pathology 85: 11–19.
  8. 8. Ewers C, Janssen T, Wieler LH (2003) [Avian pathogenic Escherichia coli (APEC)]. Berliner und Munchener tierarztliche Wochenschrift 116: 381–395.
  9. 9. Shpigel NY, Elazar S, Rosenshine I (2008) Mammary pathogenic Escherichia coli. Current opinion in microbiology 11: 60–65.
  10. 10. Tan C, Xu Z, Zheng H, Liu W, Tang X, et al. (2011) Genome sequence of a porcine extraintestinal pathogenic Escherichia coli strain. Journal of bacteriology 193: 5038.
  11. 11. Carvallo FR, Debroy C, Baeza E, Hinckley L, Gilbert K, et al. (2010) Necrotizing pneumonia and pleuritis associated with extraintestinal pathogenic Escherichia coli in a tiger (Panthera tigris) cub. Journal of veterinary diagnostic investigation: official publication of the American Association of Veterinary Laboratory Diagnosticians, Inc 22: 136–140.
  12. 12. Foxman B, Brown P (2003) Epidemiology of urinary tract infections: transmission and risk factors, incidence, and costs. Infectious disease clinics of North America 17: 227–241.
  13. 13. Johnson JR, Johnston B, Clabots C, Kuskowski MA, Castanheira M (2010) Escherichia coli sequence type ST131 as the major cause of serious multidrug-resistant E. coli infections in the United States. Clinical infectious diseases: an official publication of the Infectious Diseases Society of America 51: 286–294.
  14. 14. Wiedenbeck J, Cohan FM (2011) Origins of bacterial diversity through horizontal genetic transfer and adaptation to new ecological niches. FEMS microbiology reviews 35: 957–976.
  15. 15. Stokes HW, Gillings MR (2011) Gene flow, mobile genetic elements and the recruitment of antibiotic resistance genes into Gram-negative pathogens. FEMS microbiology reviews 35: 790–819.
  16. 16. Kirschner M, Gerhart J (1998) Evolvability. Proceedings of the National Academy of Sciences of the United States of America 95: 8420–8427.
  17. 17. Pigliucci M (2008) Is evolvability evolvable? Nature reviews Genetics 9: 75–82.
  18. 18. Gogarten JP, Doolittle WF, Lawrence JG (2002) Prokaryotic evolution in light of gene transfer. Molecular biology and evolution 19: 2226–2238.
  19. 19. Hehemann JH, Correc G, Barbeyron T, Helbert W, Czjzek M, et al. (2010) Transfer of carbohydrate-active enzymes from marine bacteria to Japanese gut microbiota. Nature 464: 908–912.
  20. 20. Marchetti M, Capela D, Glew M, Cruveiller S, Chane-Woon-Ming B, et al. (2010) Experimental evolution of a plant pathogen into a legume symbiont. PLoS biology 8: e1000280.
  21. 21. Thomas CM, Nielsen KM (2005) Mechanisms of, and barriers to, horizontal gene transfer between bacteria. Nature reviews Microbiology 3: 711–721.
  22. 22. Medigue C, Rouxel T, Vigier P, Henaut A, Danchin A (1991) Evidence for horizontal gene transfer in Escherichia coli speciation. Journal of molecular biology 222: 851–856.
  23. 23. Canchaya C, Fournous G, Brussow H (2004) The impact of prophages on bacterial chromosomes. Molecular microbiology 53: 9–18.
  24. 24. Casjens S, Hendrix RW (2005) Bacteriophages and the bacterial genome. Bacterial Chromosome. Washington: Amer Soc Microbiology. pp. 39–52.
  25. 25. Dagan T, Artzy-Randrup Y, Martin W (2008) Modular networks and cumulative impact of lateral transfer in prokaryote genome evolution. Proceedings of the National Academy of Sciences of the United States of America 105: 10039–10044.
  26. 26. Welch RA, Burland V, Plunkett G 3rd, Redford P, Roesch P, et al. (2002) Extensive mosaic structure revealed by the complete genome sequence of uropathogenic Escherichia coli. Proceedings of the National Academy of Sciences of the United States of America 99: 17020–17024.
  27. 27. Lawrence JG, Ochman H (1998) Molecular archaeology of the Escherichia coli genome. Proceedings of the National Academy of Sciences of the United States of America 95: 9413–9417.
  28. 28. Perna NT, Plunkett G 3rd, Burland V, Mau B, Glasner JD, et al. (2001) Genome sequence of enterohaemorrhagic Escherichia coli O157:H7. Nature 409: 529–533.
  29. 29. Belda-Ferre P, Cabrera-Rubio R, Moya A, Mira A (2011) Mining virulence genes using metagenomics. PloS one 6: e24975.
  30. 30. Langille MG, Hsiao WW, Brinkman FS (2010) Detecting genomic islands using bioinformatics approaches. Nature reviews Microbiology 8: 373–382.
  31. 31. Juhas M, van der Meer JR, Gaillard M, Harding RM, Hood DW, et al. (2009) Genomic islands: tools of bacterial horizontal gene transfer and evolution. FEMS microbiology reviews 33: 376–393.
  32. 32. Touchon M, Hoede C, Tenaillon O, Barbe V, Baeriswyl S, et al. (2009) Organised genome dynamics in the Escherichia coli species results in highly diverse adaptive paths. PLoS genetics 5: e1000344.
  33. 33. Kao JS, Stucker DM, Warren JW, Mobley HL (1997) Pathogenicity island sequences of pyelonephritogenic Escherichia coli CFT073 are associated with virulent uropathogenic strains. Infection and immunity 65: 2812–2820.
  34. 34. Gal-Mor O, Finlay BB (2006) Pathogenicity islands: a molecular toolbox for bacterial virulence. Cellular microbiology 8: 1707–1719.
  35. 35. Dorman CJ (2009) Regulatory integration of horizontally-transferred genes in bacteria. Frontiers in bioscience: a journal and virtual library 14: 4103–4112.
  36. 36. Tettelin H, Masignani V, Cieslewicz MJ, Donati C, Medini D, et al. (2005) Genome analysis of multiple pathogenic isolates of Streptococcus agalactiae: implications for the microbial “pan-genome”. Proceedings of the National Academy of Sciences of the United States of America 102: 13950–13955.
  37. 37. Tettelin H, Riley D, Cattuto C, Medini D (2008) Comparative genomics: the bacterial pan-genome. Current opinion in microbiology 11: 472–477.
  38. 38. Nakamura Y, Itoh T, Matsuda H, Gojobori T (2004) Biased biological functions of horizontally transferred genes in prokaryotic genomes. Nature genetics 36: 760–766.
  39. 39. Mobley HL, Green DM, Trifillis AL, Johnson DE, Chippendale GR, et al. (1990) Pyelonephritogenic Escherichia coli and killing of cultured human renal proximal tubular epithelial cells: role of hemolysin in some strains. Infection and immunity 58: 1281–1289.
  40. 40. Wiles TJ, Bower JM, Redd MJ, Mulvey MA (2009) Use of zebrafish to probe the divergent virulence potentials and toxin requirements of extraintestinal pathogenic Escherichia coli. PLoS pathogens 5: e1000697.
  41. 41. Trede NS, Langenau DM, Traver D, Look AT, Zon LI (2004) The use of zebrafish to understand immunity. Immunity 20: 367–379.
  42. 42. Li X, Wang S, Qi J, Echtenkamp SF, Chatterjee R, et al. (2007) Zebrafish peptidoglycan recognition proteins are bactericidal amidases essential for defense against bacterial infections. Immunity 27: 518–529.
  43. 43. Jault C, Pichon L, Chluba J (2004) Toll-like receptor gene family and TIR-domain adapters in Danio rerio. Molecular immunology 40: 759–771.
  44. 44. Lieschke GJ, Oates AC, Crowhurst MO, Ward AC, Layton JE (2001) Morphologic and functional characterization of granulocytes and macrophages in embryonic and adult zebrafish. Blood 98: 3087–3096.
  45. 45. Lloyd AL, Henderson TA, Vigil PD, Mobley HL (2009) Genomic islands of uropathogenic Escherichia coli contribute to virulence. Journal of bacteriology 191: 3469–3481.
  46. 46. Bower JM, Gordon-Raagas HB, Mulvey MA (2009) Conditioning of uropathogenic Escherichia coli for enhanced colonization of host. Infection and immunity 77: 2104–2112.
  47. 47. Bower JM, Mulvey MA (2006) Polyamine-mediated resistance of uropathogenic Escherichia coli to nitrosative stress. Journal of bacteriology 188: 928–933.
  48. 48. Garcia EC, Brumbaugh AR, Mobley HL (2011) Redundancy and specificity of Escherichia coli iron acquisition systems during urinary tract infection. Infection and immunity 79: 1225–1235.
  49. 49. Nilsson AS, Haggard-Ljungquist E (2007) Evolution of P2-like phages and their impact on bacterial evolution. Research in microbiology 158: 311–317.
  50. 50. Wang Z, Zhang S, Wang G, An Y (2008) Complement activity in the egg cytosol of zebrafish Danio rerio: evidence for the defense role of maternal complement components. PloS one 3: e1463.
  51. 51. Juhala RJ, Ford ME, Duda RL, Youlton A, Hatfull GF, et al. (2000) Genomic sequences of bacteriophages HK97 and HK022: pervasive genetic mosaicism in the lambdoid bacteriophages. Journal of molecular biology 299: 27–51.
  52. 52. Hendrix RW, Lawrence JG, Hatfull GF, Casjens S (2000) The origins and ongoing evolution of viruses. Trends in microbiology 8: 504–508.
  53. 53. Waldor MK, Friedman DI, Adhya SL (2005) Phages: their role in bacterial pathogenesis and biotechnology. Washington, D.C.: ASM Press. pp. 37–65.
  54. 54. Wang X, Kim Y, Ma Q, Hong SH, Pokusaeva K, et al. (2010) Cryptic prophages help bacteria cope with adverse environments. Nature communications 1: 147.
  55. 55. Smith SN, Hagan EC, Lane MC, Mobley HL (2010) Dissemination and systemic colonization of uropathogenic Escherichia coli in a murine model of bacteremia. mBio 1.
  56. 56. Altschul SF, Madden TL, Schaffer AA, Zhang J, Zhang Z, et al. (1997) Gapped BLAST and PSI-BLAST: a new generation of protein database search programs. Nucleic acids research 25: 3389–3402.
  57. 57. Papanikolaou N, Trachana K, Theodosiou T, Promponas VJ, Iliopoulos I (2009) Gene socialization: gene order, GC content and gene silencing in Salmonella. BMC genomics 10: 597.
  58. 58. Ochman H, Lawrence JG, Groisman EA (2000) Lateral gene transfer and the nature of bacterial innovation. Nature 405: 299–304.
  59. 59. Daubin V, Lerat E, Perriere G (2003) The source of laterally transferred genes in bacterial genomes. Genome biology 4: R57.
  60. 60. Kuo CH, Ochman H (2009) The fate of new bacterial genes. FEMS microbiology reviews 33: 38–43.
  61. 61. Berck S, Perret X, Quesada-Vincens D, Prome J, Broughton WJ, et al. (1999) NolL of Rhizobium sp. strain NGR234 is required for O-acetyltransferase activity. Journal of bacteriology 181: 957–964.
  62. 62. Bera A, Herbert S, Jakob A, Vollmer W, Gotz F (2005) Why are pathogenic staphylococci so lysozyme resistant? The peptidoglycan O-acetyltransferase OatA is the major determinant for lysozyme resistance of Staphylococcus aureus. Molecular microbiology 55: 778–787.
  63. 63. Yoshida Y, Yang J, Peaker PE, Kato H, Bush CA, et al. (2008) Molecular and antigenic characterization of a Streptococcus oralis coaggregation receptor polysaccharide by carbohydrate engineering in Streptococcus gordonii. The Journal of biological chemistry 283: 12654–12664.
  64. 64. Vollmer W (2008) Structural variation in the glycan strands of bacterial peptidoglycan. FEMS microbiology reviews 32: 287–306.
  65. 65. Thornton MM, Chung-Esaki HM, Irvin CB, Bortz DM, Solomon MJ, et al. (2012) Multicellularity and Antibiotic Resistance in Klebsiella pneumoniae Grown Under Bloodstream-Mimicking Fluid Dynamic Conditions. The Journal of infectious diseases 206: 588–595.
  66. 66. Lercher MJ, Pal C (2008) Integration of horizontally transferred genes into regulatory interaction networks takes many million years. Molecular biology and evolution 25: 559–567.
  67. 67. Lind PA, Andersson DI (2008) Whole-genome mutational biases in bacteria. Proceedings of the National Academy of Sciences of the United States of America 105: 17878–17883.
  68. 68. Hershberg R, Petrov DA (2010) Evidence that mutation is universally biased towards AT in bacteria. PLoS genetics 6.
  69. 69. Van Leuven JT, McCutcheon JP (2012) An AT mutational bias in the tiny GC-rich endosymbiont genome of Hodgkinia. Genome biology and evolution 4: 24–27.
  70. 70. Barker MS, Demuth JP, Wade MJ (2005) Maternal expression relaxes constraint on innovation of the anterior determinant, bicoid. PLoS genetics 1: e57.
  71. 71. Hunt BG, Ometto L, Wurm Y, Shoemaker D, Yi SV, et al. (2011) Relaxed selection is a precursor to the evolution of phenotypic plasticity. Proceedings of the National Academy of Sciences of the United States of America 108: 15936–15941.
  72. 72. Van Dyken JD, Wade MJ (2010) The genetic signature of conditional expression. Genetics 184: 557–570.
  73. 73. Whitlock MC (1996) The red queen beats the jack-of-all-trades: The limitations on the evolution of phenotypic plasticity and niche breadth. American Naturalist 148: S65–S77.
  74. 74. Waterston RH, Lindblad-Toh K, Birney E, Rogers J, Abril JF, et al. (2002) Initial sequencing and comparative analysis of the mouse genome. Nature 420: 520–562.
  75. 75. Johnson JR, Porter SB, Zhanel G, Kuskowski MA, Denamur E (2012) Virulence of Escherichia coli Clinical Isolates in a Murine Sepsis Model in Relation to Sequence Type ST131 Status, Fluoroquinolone Resistance, and Virulence Genotype. Infection and immunity
  76. 76. Wirth T, Falush D, Lan R, Colles F, Mensa P, et al. (2006) Sex and virulence in Escherichia coli: an evolutionary perspective. Molecular microbiology 60: 1136–1151.
  77. 77. Reid SD, Herbelin CJ, Bumbaugh AC, Selander RK, Whittam TS (2000) Parallel evolution of virulence in pathogenic Escherichia coli. Nature 406: 64–67.
  78. 78. Li M, Du X, Villaruz AE, Diep BA, Wang D, et al. (2012) MRSA epidemic linked to a quickly spreading colonization and virulence determinant. Nature medicine 18: 816–819.
  79. 79. Tuller T, Girshovich Y, Sella Y, Kreimer A, Freilich S, et al. (2011) Association between translation efficiency and horizontal gene transfer within microbial communities. Nucleic acids research 39: 4743–4755.
  80. 80. Smillie CS, Smith MB, Friedman J, Cordero OX, David LA, et al. (2011) Ecology drives a global network of gene exchange connecting the human microbiome. Nature 480: 241–244.
  81. 81. Andam CP, Gogarten JP (2011) Biased gene transfer in microbial evolution. Nature reviews Microbiology 9: 543–555.
  82. 82. Mulvey MA, Lopez-Boado YS, Wilson CL, Roth R, Parks WC, et al. (1998) Induction and evasion of host defenses by type 1-piliated uropathogenic Escherichia coli. Science 282: 1494–1497.
  83. 83. Murphy KC, Campellone KG (2003) Lambda Red-mediated recombinogenic engineering of enterohemorrhagic and enteropathogenic E. coli. BMC molecular biology 4: 11.
  84. 84. Datsenko KA, Wanner BL (2000) One-step inactivation of chromosomal genes in Escherichia coli K-12 using PCR products. Proceedings of the National Academy of Sciences of the United States of America 97: 6640–6645.
  85. 85. Lane MC, Alteri CJ, Smith SN, Mobley HL (2007) Expression of flagella is coincident with uropathogenic Escherichia coli ascension to the upper urinary tract. Proceedings of the National Academy of Sciences of the United States of America 104: 16669–16674.
  86. 86. Grozdanov L, Raasch C, Schulze J, Sonnenborn U, Gottschalk G, et al. (2004) Analysis of the genome structure of the nonpathogenic probiotic Escherichia coli strain Nissle 1917. Journal of bacteriology 186: 5432–5441.
  87. 87. Rasko DA, Rosovitz MJ, Myers GS, Mongodin EF, Fricke WF, et al. (2008) The pangenome structure of Escherichia coli: comparative genomic analysis of E. coli commensal and pathogenic isolates. Journal of bacteriology 190: 6881–6893.
  88. 88. Kulesus RR, Diaz-Perez K, Slechta ES, Eto DS, Mulvey MA (2008) Impact of the RNA chaperone Hfq on the fitness and virulence potential of uropathogenic Escherichia coli. Infection and immunity 76: 3019–3026.